Abstract
In the 50 years that succeeded Richard Feynman’s exposition of the idea that there is “plenty of room at the bottom“ for manipulating individual atoms for the synthesis and manufacturing processing of materials, the materials-by-design paradigm is being developed gradually through synergistic integration of experimental material synthesis and characterization with predictive computational modeling and optimization. This paper reviews how this paradigm creates the possibility to develop materials according to specific, rational designs from the molecular to the macroscopic scale. We discuss promising techniques in experimental small-scale material synthesis and large-scale fabrication methods to manipulate atomistic or macroscale structures, which can be designed by computational modeling. These include recombinant protein technology to produce peptides and proteins with tailored sequences encoded by recombinant DNA, self-assembly processes induced by conformational transition of proteins, additive manufacturing for designing complex structures, and qualitative and quantitative characterization of materials at different length scales. We describe important material characterization techniques using numerous methods of spectroscopy and microscopy. We detail numerous multi-scale computational modeling techniques that complements these experimental techniques: DFT at the atomistic scale; fully atomistic and coarse-grain molecular dynamics at the molecular to mesoscale; continuum modeling at the macroscale. Additionally, we present case studies that utilize experimental and computational approaches in an integrated manner to broaden our understanding of the properties of two-dimensional materials and materials based on silk and silk-elastin-like proteins.
Keywords: Materials physics, design, resilience, multiscale, modeling, materiomics
1. Introduction
In 1959, Richard Feynman expressed his vision around the concept that there exists “plenty of room at the bottom“ for manipulating individual atoms for the synthesis and manufacturing processing of materials1. More than 50 years after this visionary idea, recent advances in nanotechnology, materials characterization, additive manufacturing, synthetic procedures, and supercomputing have made it possible to develop innovative materials with sophisticated structures and multiple material functions, in accordance with specific rational designs, from the molecular to the macroscopic scale. This new materials-by-design paradigm, deeply rooted in Feynman’s belief, resulted in a paradigm shift towards a powerful integrated approach that connects theoretical design at the atomic level, modeling of materials at the nano-meso-micro-scale, and additive manufacturing, synthetic chemistry, and process engineering at the macroscale (Figure 1). Achieving such a precise level of control over materials manufacturing strongly impacts our society as it contributes to unprecedented advances in the development of more sustainable technologies, energy sources, miniaturization, and countless fields of science and engineering. The potential of this paradigm arises due to the convergence of several important factors: new manufacturing technologies; more developed material characterization techniques; new modeling methods; increased computational power; new optimization algorithms; and artificial intelligence.
Figure 1.

The ‘materials-by-design’ paradigm that integrates numerous aspects of computation, synthesis, and characterization. List of acronyms: CVD – Chemical Vapor Deposition; DNA – Deoxyribonucleic Acid; FEM – Finite Element Method; CG – Coarse-grained; MD – Molecular Dynamics; DFT – Density Functional Theory; XRD – X-ray Diffraction; AFM – Atomic Force Microscopy; TEM – Transmission Electron Microscopy; IR – Infra-Red.
The resulting material-by-design paradigm is highly influential in current developments of biomaterials, proteins, polymers and nanomaterials. Accordingly, in this article we provide an overview of different techniques of experimental synthesis, characterization, and computational modeling that combine to give rise to such a paradigm shift of manipulating materials on the atomic scale, together with some examples of biological, biomimetic, or nature-inspired materials where that interdisciplinary approach is already being proven successful by incorporating designs from atoms to structures. These studies will show that the “room at the bottom“ dreamed by Feynman is already being filled up.
2. Materials-by-Design: Synthesis & Characterization (Experimentation)
Materials synthesis and their corresponding characterization must be tightly coupled and developed simultaneously to push the boundaries of bottom-up materials-by-design. Clearly, a great multitude of organic and inorganic material synthesis, fabrication, and characterization technologies are in existence, such as gas condensation and sol-gel techniques. However, not all such methods can synergize well with computational design. Therefore, in this section, we will examine several pertinent and promising methods of synthesizing or fabricating biological, biomimetic, or nature-inspired materials from atoms to structures: recombinant DNA technology, protein self-assembly, and additive manufacturing. Correspondingly, we will also discuss the techniques for characterizing these materials through spectroscopy and microscopy.
2.1. Material Synthesis and Fabrication Techniques
Recombinant DNA Technology
Recombinant protein technology is a biosynthesis method that produces peptides and proteins with tailored sequences encoded by recombinant DNA2,3. The core methodology of such recombinant protein methods is the recombinant DNA technology, which constructs defined DNA molecules using genetic recombination methods (such molecular cloning) that integrate multiple genetic materials into one sequence, leading to the expression of any mutant protein4. This biosynthesis technique is a de novo design strategy to produce peptides and proteins with tailored sequences, molecular weight, conformation, and even complex tertiary structures5. Figure 2a summarizes the four critical steps to obtain reconstructed protein via recombinant DNA technology.
Figure 2.

The typical routes of protein recombination and silk fibroin self-assembly. a) Schematic of the processing steps of recombinant protein technology5. Adapted with permission from Ref [5]. Copyright (2017) American Chemical Society. b) Pathways of silk fibroin self-assembly8,9. Adapted from Ref [6] licensed under CC BY-NC 4.0 from AAAS, and also adapted from Ref [7] with permission of The Royal Society of Chemistry.
The first step is the design and assembly of the protein or peptide genes into genetic “cassettes”. The sequences used for building recombinant DNA molecules can originate from any species or sources, and can even be computationally designed. Smaller monomeric gene sequences can be synthesized to be as short as single-stranded oligonucleotides (up to 100 base pairs) by commercial oligonucleotide synthesis, while large oligomeric sequences are usually constructed by using concatemerization, step-by-step directional approach, and recursive ligation6.
In the second step, the designed DNA segments are inserted into a DNA vector, which can be derived from plasmids or viruses. This is usually a small DNA molecule with all necessary genetic signals for replication in the living cell. A series of methods, such as the Gibson assembly method that joins DNA fragments in a single, isothermal reaction, and restriction enzyme/ligase cloning, can be applied to connect DNA segments with vectors. The resultant recombinant DNA vectors are transferred into the host organism for further expression. A series of heterologous host systems have been explored to express the recombinant proteins, such as bacteria, yeast, insects, plants, and even mammalian cell lines and transgenic animals6. Among these host systems, the bacteria, Escherichia coli, is the most well-established biosynthesis system due to its low cost, short generation time, and easier manipulation and scalability7. Not all of the expressed organisms will contain the reconstructed DNA vectors; these organisms must be selected and purified to isolate those that definitely contain the desired DNA inserts.
The final step is protein purification to isolate the designed proteins. Protein purification typically consists of a series of steps to isolate the desired proteins from the complex mixtures derived from the cells or organisms. These separation processes are usually distinct depending on the proteins being isolated, since they are directly related to the protein size, biological activity, binding affinity, and physicochemical properties.
Protein Self-assembly
Compared to other synthetic polymers, a unique feature of proteins is their inherent tendency to assemble into sophisticated hierarchical structures10,11. The self-assembly processes can be initiated or triggered by an external stimuli, i.e., pH, ions, heat, chemical species, or any other environmental factors able to induce a conformational transition of the protein11. Protein self-assembly has become a fundamental strategy to construct advanced functional protein materials. In particular, its ability to position molecules or nano-scale entities with nanometric spatial resolution poses unprecedented opportunities for the de novo synthesis of one- to three-dimensional nano-architectures with exceptional structure and functions tailored for applications in biomedicine and nanotechnology12.
Although there is a vast array of different proteins and self-assembling mechanisms, this review focuses on two self-assembly regulation routes of silk fibroin8,12,13, which is an abundant protein in nature and has a wide range of applications in biomedicine, water treatment, and optical and electronic devices9,14–18. The two main methods, ethanol-induced self-assembly and heat-induced self-assembly, induce the silk fibroin to self-assemble into silk nanofibrils in a simple and environmentally harmless procedure (Figure 2b). Although these methods are discussed in the context of silk fibroin assembly, they are suitable for other protein systems as well, such as amyloid fibrils19 and soy proteins20. The ethanol-induced self-assembly method uses silk fibroin aqueous solution as starting material (seed) with ethanol solution being added. The heat-induced self-assembly method uses the same starting materials without any ethanol while being incubated at a mild temperature (60-80 °C for example). Furthermore, the self-assembly process can be tuned by regulating the processing conditions. The silk nanofibrils obtained from these two self-assembly methods are necklace-like topological structures connected by homogeneous nanoglobules. Compared to other silk fibroin-based materials, these silk nanofibrils typically have higher β-sheet content (~50%), and their moduli are comparable with amyloid and collagen fibrils. The diameters of single nanofibrils are around of 3-5 nm, while their lengths can reach up to tens of micrometers8. Similar to other protein-based nanomaterials such as amyloid fibrils and collagen fibrils, silk nanofibrils have several unique advantages for biomedical and nanotechnology applications. For example, silk nanofibril is an all-protein nanomaterial that have similar structural and mechanical properties as collagen fibrils, thus they can serve as in-vitro extracellular matrix materials9. Furthermore, silk nanofibrils can be directly used for making porous nanomaterials, due to their fine dimensions and high specific surface area. Our previous works have indicated that the pore size of silk nanofibril material are tunable in the range of several to hundreds of nanometers17, which is a pore size range that other silk-based materials cannot attain. Thus, silk nanofibrils are good candidates for nano/ultra-filtration or other cutting-edge technologies, such as batteries, biosensors, or optical devices.
Additive Manufacturing
Additive manufacturing has revolutionized component and system-level materials fabrication by enabling the design of complex structures that consist of a miscellany of internal voids, multiple materials, and irregular shapes21. Using different manufacturing methods, traditional engineering materials such as polymers, metals (e.g. aluminum, titanium), and ceramics can be formed into arbitrary or carefully-designed shapes that were infeasible with conventional synthesis tools and traditional manufacturing techniques such as mills, lathes, and molds21–23. Among all different techniques and emerging technologies within the field of additive manufacturing, there are seven different types of additive processes that we can emphasize as being mostly used nowadays: Vat polymerization, material jetting, binder jetting, material extrusion, powder bed fusion, sheet lamination, and direct energy deposition (DED)24. A brief overview of these additive manufacturing technologies are presented below; a more detailed description is given by Gibson et al.24,25.
Vat polymerization processes involve using a moving laser beam to build models layer-by-layer by tracing the shape of the printed object with mostly photopolymer resins, which cure or harden upon contact with UV light. This method provides the best surface finish, part accuracy, and feature resolution compared to other additive manufacturing processes. Material jetting, in some cases known as PolyJet™ technology, uses a printing head to jet liquid photopolymers from a nozzle onto a build platform, where they are then cured with UV light in situ. This method offers high resolution printing and can produce materials that consist of multiple colors and properties by jetting out multiple materials simultaneously, providing near perfect adhesion. Additionally, this process allows different materials to be mixed to create other materials with intermediate mechanical properties. A disadvantage for this method is the limited number of materials available to use for the drop deposition process. The binder jetting process involves a powder based material and a binder, where the binder acts as an adhesive between powder layers. The print head deposits alternating layers of the build material and binding material. Fused deposition modeling (FDM) is a technique that consists of depositing semi-molten thermoplastics in ultra-fine beads along the extrusion path as indicated by the computer model. Compared to other 3Dprinting processes, FDM differs in the addition of the material, which is added through a nozzle under constant pressure and in a continuous stream. The powder bed fusion process includes the following common printing techniques: Electron beam melting (EBM), selective laser melting (SLM), and selective laser sintering (SLS). This method either uses a laser or electron beam to melt and fuse material powder together. SLS synthesizes structures using an expensive, high power laser to selectively sinter polymer powders, making it a very costly procedure. SLS has limitations when it comes to surface finish and feature resolution but can process a wide array of materials such as nylon, glass, ceramics, and metals. SLM uses a high-power laser beam to fully melt metallic powders such as aluminum and titanium into 3D objects. EBM uses a computer controlled electron beam under high vacuum to melt metallic powder at high temperatures. Sheet lamination processes include laminated objective manufacturing (LOM), which fuses adhesive-coated layers of paper, plastic, or metal laminates under high heat and pressure. Afterwards, the object is shaped by using a laser cutter or knife. A DED machine consists of a nozzle that deposits melted material onto a surface where it solidifies thereafter. The process can be used with polymers, ceramics, and metals. Overall, the selection of the additive method depends on the applications and user needs, which implies choosing the most suitable and cost-effective for the design at hand.
Limitations in Additive Manufacturing
Additive manufacturing has demonstrated its utility and effectiveness across many fields with applications in medicine, aerospace, and automotive industries, to name a few, yet certain limitations exist. These limitations pose substantial hurdles in establishing additive manufacturing as a primary, large-scale, industrial, processing method21–23. While a complete review on the state-of-the-art and future prospective of additive manufacturing is beyond the scope of this work, a few key hindrances are outlined in this section. An immediate practical limitation is the cleaning process of the support material, which is especially difficult and sometimes infeasible when dealing with complicated structures with voids. Additionally, additive manufacturing also suffers from the limited range of materials that can be printed, which precludes the design space of optimized structures. When multiple materials are printed, however, interfaces are created and printing across material boundaries is very limited, constraining the design of heterogeneous objects. Of course, time and space also constrain additive manufacturing, although these are controlled by the type of process and the build tray size. Finally, there is tradeoff in printing resolution. Otherwise stated, printers that print at the nanoscale need a significantly long time printing at the macroscale. Commercial 3D printers still have these limitations and much research is underway to improve additive manufacturing.
Applications in Material Design
Many engineering materials possess inherently conflicting mechanical properties, such as toughness and strength26. For instance, metals are very tough but not very stiff, whereas, ceramics are very stiff but not very tough. Natural materials often circumvent these intrinsic tradeoffs through their intricate, hierarchical structures. As a result, many research groups strive to mimic nature’s complex micro- and nano-architectures to improve the mechanical properties in engineering materials27–29. Due to the complexity of hierarchical and heterogeneous geometries, additive manufacturing is an excellent method to fabricate these bio-inspired designs30.
Additive manufacturing provides a tool for understanding how the nature-designed architectures overcome the tradeoffs for better mechanical properties. Buehler et al. have investigated various aspects of bio-inspired design by using a multimaterial printer called Connex 3. This printer uses polyjet technology that can print up to three materials simultaneously, allowing contrasting materials to be mixed to create materials with intermediate mechanical properties. Dimas et al. designed composites inspired by bone and nacre, 3D-printed them, and tested them to compare with simulations (Figure 3a)31. The generated composites have a soft matrix phase with stiff reinforcements and they discovered that these nature-inspired designs can deflect cracks better than their homogeneous counterparts, and as a result, have higher toughness. Mirzaeifar et al. expanded on the work and studied the effect of adding hierarchy to similar geometries (Figure 3b)32. Through the use of simulation and experiments with 3D-printed samples, they showed that increasing the level of hierarchy improved the defect tolerance of the composites. These two studies are largely concerned with nacre-like designs and the brick-and-mortar scheme. However, current literature shows that structural features such as mineral bridges also contribute to its high toughness and strength. Due to manufacturing difficulties, work on the effects of mineral bridges is limited to a small range of volume fractions and number of mineral bridges. Gu et al. systematically evaluated the effects of structural parameters, such as volume fraction of mineral phase and material hierarchy, on the mechanical response of nacre-inspired additive manufactured composites (Figure 3c)33. Their results demonstrate that it is possible to tune composite properties by controlling the size and content of structural features in a heterogeneous material.
Figure 3.

Biomimicry leads to materials with exceptional mechanical properties. a) Simulation, additive manufacturing, and mechanical testing form the framework used to characterize complex natural material microstructures such as those seen in nacre and bone31. Adapted from Ref [28] with permission from John Wiley and Sons. b) Using the same framework, the effects of structural hierarchy levels (1H, 2H, 3H) on defect tolerance of nacre and bone-like materials are examined32. Adapted with permission from Ref [29]. Copyright (2015) American Chemical Society. c) Systematic exploration of small structural features of nacre, such as mineral bridges (left), is possible using additive manufacturing. Simulation of nacre-like composites with mineral bridges (right)33. Adapted from Ref [30], Copyright (2017), with permission from Elsevier. d) Impact testing of nacre-like composites in simulation are compared to experiments on additive manufactured samples35. e) Testing and fabrication of bone-inspired composites using additive manufacturing show toughening mechanisms in real bone can be recovered in synthetic bone-like materials34. Adapted from Ref [32] with permission from John Wiley and Sons. f) Designs inspired by conch shells (left) show superior impact resistant properties (middle) with crack arresting mechanisms gained from their complex architectures (right)36. Adapted from Ref [33] with permission from John Wiley and Sons.
Not constraining the discussion of natural materials to nacre, Libonati et al. designed bone-like additive manufactured composites by mimicking the fundamental microstructural features of cortical bone such as osteons and showed that the toughening mechanisms demonstrated in their synthetic composites are very similar to those seen in real cortical bone (Figure 3e)34. These toughening mechanisms include crack deflection and branching, micro-cracking, and fibril bridging. Expanding on similar work, Gu et al. combined finite element simulations and drop-tower experiments to show how nacre-like synthetic materials are superior to their constituent materials in impact resistance (Figure 3d)35. They showed that the nacre-like 3D-printed designs were able to bring the energy of the impact down to a zero-impact velocity, while the stiff homogeneous material was only able to reduce the energy of the impact by half. The authors further studied the conch shell, which is reported to have ten times higher toughness than nacre.
Most biomimetic literature focused on nacre and bone because the three-tiered microstructure in conch shells proved to be a significant modeling challenge to study and emulate. With the use of additive manufacturing, the authors created the first 3D prototype and finite element model of conch shells (Figure 3f). They studied the effects of added levels of hierarchy on impact properties and discovered that geometries demonstrating second level hierarchy absorbed 70% more energy than geometries with first level hierarchy and 85% more than the monolithic (no hierarchy) constituent material36. Future work in this field will be to optimize these structural features for increased impact resistance. The advent of additive manufacturing offers the opportunity to do so by offering freedom in geometrical design.
2.2. Characterization
Qualitative and quantitative characterization of materials at different length scales is critical for understanding the relation between material composition, structure, property, and function. Since this characterization is based on the physical and chemical interaction between the samples and the characterization tool, characterization also provides opportunities for improvement and control over the synthesized samples. The most demanding feature of characterization is to provide accuracy to improve our understanding, synthesis, and production of materials. Currently, a plethora of experimental techniques are available not only to characterize but also to simultaneously interact in situ with the structure of the materials. This level of characterization has enabled the existence of nanoscience and nanotechnology as we understand them today, as well as the development of the “materials-by-design” paradigm that is expounded in this review. Describing in detail all the techniques available will require an entire monographic work, and such an endeavor goes beyond the purpose of the present review. However, we will briefly describe some techniques that have major roles in nano- and biomaterials science.
X-ray spectroscopy is a family of different spectroscopy techniques for investigating the local atomic structure of materials using X-ray radiation (Figure 4b). It is mostly based on X-ray emission, although X-ray absorption spectroscopy (XAS) is also applicable in materials characterization. X-ray radiation spectra are analyzed by different techniques, mainly energy-dispersive X-ray spectroscopy (EDS) for elemental analysis and wavelength dispersive X-ray spectroscopy (WDS), which provides diffraction patterns for a single wavelength. From an historical perspective, Wilhelm Conrad Rontgen received the first Nobel Prize in Physics in 1901 for discovering the X-ray, an electromagnetic radiation having a wavelength ranging from 0.01 to 10 nanometers. Subsequently, Max von Laue, William Henry Bragg, and William Lawrence Bragg were awarded Nobel Prizes in Physics in 1914 and 1915 for their contribution to X-ray spectroscopy, where diffraction patterns are generated from injection of X-ray to lattices of atoms in crystalline formations. It is also common knowledge that the helical structure of deoxyribonucleic acid (DNA)37, was first reported in 1953 with X-ray spectroscopy performed by Watson and Crick for which they were awarded the Nobel Prize in Physiology or Medicine in 1962. Since then, technologies based on X-rays have contributed to numerous fields of science and were crucial in the award of more than 20 Nobel Prizes in Medicine and Chemistry. In our daily life, X-ray radiation is most commonly known for its application in the medical field, in which radiography is a common technology that provides images of the insides of the human body.
Figure 4.

a) Schematic of Raman spectroscopy and measured Raman shift for different layers of MoS2 from 488 nm laser line, showing two distinct shifts: E12g ~ 383 cm−1 and A1g ~ 408 cm−1 shifts according to the number of layers38. Adapted from Ref [34] with permission from John Wiley and Sons. b) Schematic of X-ray diffraction and a sample of small-angle X-ray diffraction (SAXD) pattern for elastin molecule. Powered wide-angle X-ray diffraction (WAXD) pattern for graphene, graphene oxide (GO), and reduced graphene oxide (rGO)39, adapted from Ref [35] with permission of The Royal Society of Chemistry, and nanostructure of tropoelastin40 obtained from SAXD, adapted from Ref [36] licensed under CC BY-NC 4.0 from AAAS. c) Schematic of an optical tweezer and a sample from collagen and the force-extension curve for a single type II collagen molecule41. Adapted from [37], Copyright (2004), with permission from Elsevier. d) Schematic of atomic force microscopy (AFM) and the force-extension behavior obtained from silk. The force-extension curve for the unfolding of pS(4+1) silk protein shows several rupture peak forces which are fitted with worm-like-chain (WLC) polymer model curves42. Adapted from [38] with permission from the National Academy of Science, USA.
As a spectroscopic technique, Raman spectroscopy is based on the interaction between matter and electromagnetic radiation (Figure 4a). While X-ray spectroscopy mainly utilizes elastic scattering, also known as Rayleigh scattering, where the high-energy electromagnetic radiation is scattered without changing in energy, Raman spectroscopy makes use of inelastic scattering where the low frequency of the emitted light changes due to the interaction between photons and molecules. The incident photon changes the vibrational, rotational, and any other low-frequency modes of the molecular system, thus generating excited states. The energy difference between the two states of molecules results in shifts in the frequency. When the light loses its energy to the vibration of molecules, it is called Stokes scattering and it is responsible for the more common spectra, while anti-Stokes scattering refers to the light gaining extra energy from molecules. This technique can provide information on chemical compounds that are present in the system based on their vibration because each bond type has a corresponding specific Raman shift, which cannot be easily characterized in other spectroscopy technology.
Transmission electron microscopy (TEM), a technique where electron beams are transmitted into the samples to visualize their images, was developed in early 1930s. TEM utilizes the short wavelength of electrons rather than light or UV, which allows much higher resolutions than optical microscopy because the wavelength of electrons is much shorter than visible light or UV. The electron beam can generate diffraction patterns from crystals (selected area diffraction patterns), which is a useful tool for studying the crystallinity of materials43. Various versions of TEM have been developed for higher resolutions and better image quality, such as scanning TEM, cryo-TEM, and aberration corrected (AC)-TEM. Currently, sub-Angstrom resolutions are achievable, thus playing a critical role in studying atomic structures and charge redistribution of chemical bonds.
Unlike previously described microscopy techniques, atomic force microscopy (AFM)44 actually touches the surface of materials by utilizing a cantilever beam and a probe tip (Figure 4d). AFM can provide three-dimensional surface profiles in contrast to the two-dimensional profiles from other microscopy techniques. Although the scanning speed is limited, AFM can perform well in ambient conditions and near-physiological conditions45 in contrast to TEM, and this allows the probing of biological materials in ambient conditions such as hydrated environments. Furthermore, Steven Chu developed the technique of optically trapping of particles in water46. Called optical tweezers (Figure 4c), a focused laser can trap nano- to micro-scale dielectric particles for studying the behaviors and forces associated with biological molecules47,48. The pulling force can be much smaller than AFM, from 0.1 to 100 pN. The mechanical properties of biomolecules can also be directly compared with MD simulations49. These two techniques significantly expanded the capabilities beyond merely visualizing atomic structures as the mechanical properties can be accurately determined in a variety of environments.
3. Materials-by-Design: Modeling & Optimization (Computation)
A critical driver of rapid advances in the materials-by-design paradigm in biological, biomimetic, or nature-inspired materials is the combination of multiscale computational models and the extensive utilization high-performance supercomputing with different techniques of experimental synthesis and characterization. This partnership between theory, computational simulations, and experiments is the key catalyst that enables the high-throughput design and detailed manipulation of materials from atoms to structures. In this section, multiscale computational techniques will be reviewed: density functional theory; molecular dynamics modeling such as atomistic and coarsegrained simulations and how these can be validated with experimental techniques of WAXS, SAXS, Raman, and infrared spectroscopy; continuum modeling; and numerical optimization.
3.1. Density Functional Theory
Density functional theory (DFT) is probably one of the most successful approaches to compute the electronic structure of matter. In its original formulation, it provides the ground-state energy and the electron density of atoms and molecules. In quantum mechanics, the wave function that arises from solving the Schrodinger equation contains all the information available about a given molecular system, and its square provides the electron density. Since it is impossible to solve the Schrodinger equation analytically for a system of more than two interacting particles, successive approximations have been implemented in computational chemistry to provide useful solutions. The first and most basic of these approximations is the Bom-Oppenheimer approximation, which assumes that the motion of the nuclei is much slower than that of the electrons since the nuclei are significantly more massive in comparison to the electrons. Thus, the electrons are assumed to respond instantaneously to the motion of the nuclei and the nuclei are fixed to solve the movement of the electrons. Within this approximation, DFT states that the energy of electrons is a functional, i.e. a function of a function, of the electron density in the homogenized electron gas, which allows us to approximate the problem of N-N interacting electrons (for a system consisting of N electrons) as a simpler system of one electron interacting with a homogenous external and effective potential. Therefore, the ground state properties of a A-electron system with 3N spatial coordinates are uniquely determined by an electron density that depends only on 3 spatial coordinates with the effective electron potential. In short, the functional constitutes the total energy that is determined from the electron density, which accounts for electron exchange effects coming from Pauli’s exclusion principle and electron correlation effects coming from homogenization. However, the exact form of the exchange-correlation function has not been identified. Hence, the implementation of exchange-correlation functionals in DFT generates a plethora of different functional flavors that correspond to different approaches and subsequent approximations to describe the behavior of the electrons. Some examples are the local density approximation (LDA), generalized gradient approximation (GGA), meta-GGA, hybrid functionals. In LDA, the functional depends only on the (local) density at a given point, an example of which is the SVWN functional, while the functional in GGA depends on the local density and its gradient, such as the PW91 and LYP and the B88 exchange functional. In meta-GGA, the functional depends on the local density, its gradient, and its second derivative, for instance in the M06-L functional. Finally, in hybrid functionals, a mix of Hartree-Fock exchange with either LDA, GGA, or meta-GGA, depending on the class, is implemented. Among the last family, B3LYP, which is a hybrid GGA, and M05-2X and M06-2X, which are hybrid meta-GGAs, are the most extensively employed.
Despite the challenges associated to its implementation, DFT provides a sound basis for developing several computational strategies for obtaining information about the energetics, structure, and properties of molecular systems at much lower costs than traditional ab initio wave function techniques. Nowadays, the original formulation has been generalized to deal with many different situations: spin polarized systems, time-dependent phenomena, reactivity descriptors and excited states, just to mention a few.
Within DFT, conceptual DFT is a branch of DFT that quantitatively define chemical concepts like chemical potential, electronegativity, or chemical hardness, within the formal formulation of DFT. Conceptual DFT provides a family of reactivity descriptors that have been proven useful in the design of new materials from the molecular scale. However, one of the greatest challenges in this context is to implement the conceptual DFT descriptor for usage in periodic systems and bulk materials.
3.2. Molecular Dynamics Modeling
Atomistic Molecular Dynamics Modeling
While extremely detailed characteristics of atoms can be determined through ab initio quantum chemical simulations or DFT calculations, this level of detail comes at a significant computational cost that will be insurmountable for systems that extends beyond hundreds of atoms. Molecular dynamics simulations simplified the interactions between atoms to be simple functions and the point masses are moving according to classical Newtonian equations of motions50,51. While the greatly-reduced computational cost allows systems with millions of atoms to be simulated, a major downside is the loss of chemical or electronic details such that the accurate determination of chemical reactions or charge transport is not possible. However, MD simulations are useful for solving classical many-body problems which generally lack analytical solutions. Although MD is underpinned theoretically by Newton’s laws of motions in its simplest form involving punctual particles, increasing complexity of problems mandates the incorporation of Euler equations, Hamilton’s quaternions, and Lagrange’s method of geometric constraints to formulate the equations of motion in a manner that enable greater ease of obtaining numerical solutions52.
By averaging out the motion of the electrons, the Hamiltonian of a system of atoms can be expressed as a function of their nuclear variables through the Born-Oppenheimer approximation. Assuming a classical description of a system is adequate, the Hamiltonian, H, of a system of N number of particles can be expressed as a sum of its potential, U, and kinetic, K energy functions. K and U in turn are functions of the set of coordinates qi and momenta pi of each atom i, where the variables in bold refer to vectors. From the Hamiltonian, equations for position and momentum can be derived, and trajectories are calculated by numerically integrating equations53.
A wide variety of MD potentials have been parameterized to describe a great multitude of interatomic interactions and to produce numerous material properties. One of the most fundamental of these potentials is the Lennard-Jones (LJ) potential, which typically takes the form
| (Equation 1) |
where ε is the potential depth, σ is the distance where the potential value is zero, and rij is the distance between two particles. Although its mathematical formulation is comparatively simple, it can accurately model the properties of several low-density gases, such as argon 53,54. More significantly, the LJ potential can be coupled to other interatomic potentials, providing dispersive interactions to model a much broader range of materials55–58. To approximate interactions between chemically dissimilar atoms, the Lorentz-Berthelot mixing rules50 can be applied.
However, greater complexity is demanded when modeling chemical species that have complex bonding structures, such as those in carbon and silicon compounds. For instance, in the case studies presented in the subsequent sections, highly detailed mathematical formulations have been derived, optimized, and validated against DFT and experimental results. These include many body potentials, e.g., Adaptive Intermolecular Reactive Empirical Bond Order (AIREBO) potential55, the Tersoff potential59, and classical potentials like those belonging to the large family of the Chemistry at HARvard Macromolecular Mechanics (CHARMM) potentials with harmonic forces60, just to name a few. These advanced potentials can account for accurate interatomic bond, angle, and dihedrals, bond breaking and formation, and Coulombic interactions. An example of such an advanced mathematical form exemplifying the above can be given by
| (Equation 2) |
where l is the bond length; θ the bond angle; ω the dihedral angle; and q the atomic charge, which accounts for the Coulombic interactions. The subscript 0 denotes the reference or equilibrium values. Furthermore, these potentials can be parameterized from, or validated against, structural data obtained from WAXS/SAXS or Raman and IR spectroscopy.
WAXS and SAXS
The diffraction patterns from X-ray or electrons include information of the materials’ structures. Conventional MD simulations utilize a vast number of atoms in the system, which complicate the analysis of structures and the validation of the models. The powdered samples’ diffraction can be obtained based on Debye’s equation61:
| (Equation 3) |
where Q, rij andf are the scattering vector, distance between two atoms i and j, and atomic scattering factor, respectively. The empirically obtained scattering factors depend on the atom type and scattering vector Q. The scattering intensity of graphene, graphene oxide, and graphite as a function of scattering angle are shown in Figure 4b. This virtual diffraction calculation is critical for validating the structural evolution of materials obtained from MD simulations, such as self-assembly and crystal growth62. The angle dependence is included in the equation of the scattering vector, Q = 4π sinθ/λ, where θ is a half-angle of the diffraction and λ is the wavelength of the X-ray. Typically, if Bragg’s peaks of crystal structures appear in the range of wide angles, the X-ray diffraction pattern is usually termed as wide-angle X-ray scattering (WAXS). In contrast, small angle X-ray scattering (SAXS) typically looks at the angles between 0.1 to 10 degrees, which can be applicable to non-crystallized systems such as macromolecules. The scattering information basically provides the size distributions of the constituent particles or molecules. Also, it can provide the overall shapes of unknown structures without crystallization. As an example, a SAXS-derived structure of tropoelastin is shown in Figure 4b. More comparisons between calculations and structures are found in the work by Hura et al63.
Raman and Infrared spectroscopy
Raman and IR spectra depend on the vibrational mode of atoms in the systems. IR spectroscopy observes the absorption of light due to molecular vibration while Raman spectroscopy observes the scattered light due to the vibration of molecules. There is a rule of mutual exclusions for a molecule with a center of symmetry: the normal mode has either Raman or IR active but not both. In general, for more complex molecules, a strong IR intensity of a molecule corresponds to weak Raman intensity. As described previously, the parameters for MD potentials are fitted to their ground states using simple analytical functions such as harmonic potentials. The frequency of the normal mode is sensitive to the interatomic potential shape, which limits the ability of classical MD to obtain correct normal modes, and thus most calculations for obtaining the Raman and IR spectra rely on ab-initio or first principles calculations. Due to limitations in computational costs, only small molecules or unit cell systems can be applied to compare the spectra to experiments but theory could provide valuable tools to characterize the states of the systems, which conventional electron micro spectroscopy is hardly able to provide. For example, strain-induced Raman shifts of monolayer MoS2 can be characterized by calculating the changes in the frequency of normal modes with the strains64, and also with different stacks of MoS2 and WSe265 as shown in Figure 4a.
Coarse-grained Molecular Dynamics Modeling
Coarse-grain (CG) models are efficient methods for simulating and investigating systems in which the response, property, and/or behavior of interest is intrinsically at a spatiotemporal scale that are inaccessible to numerical methods that capture phenomena on a much smaller scale. Furthermore, through hierarchical “handshaking” between the various scales, progressive multiscale transitions from atomic to mesoscale to macroscale structure-property-function relationships can be determined synergistically. This “finer-trains-coarser” approach immensely fortify the ability to design materials from the bottom-up by incorporating a comprehensive theoretical foundation for the prediction, characterization, and optimization of multiple properties in these novel materials.
Methods of coarse-graining in mesoscale numerical simulations of biological and polymeric materials can be broadly classified in three distinct categories: physics-based, knowledge-based, or structure-based approaches66. Physics-based approaches are the most widely used, where forcefields are constructed based on experimental data or quantum-mechanical calculations, in a similar manner as fully atomistic MD simulations. As a result of this close resemblance to MD, physics-based forcefields can be conveniently incorporated into existing MD software, which provide greater portability and broader usage. One of the most widely used CG forcefield for biological materials is the MARTINI forcefield67,68. Initially, it is parameterized to model lipids only69. Subsequent developments standardized the methodology to systematically reproduce the free energies of hydration, vaporization, and partitioning of polar and apolar phases of numerous chemical compounds including hexadecane, chloroform, ether, and octanol67. Atoms are clustered into four main types of polar (P), nonpolar (N), apolar (C), and charged (Q) CG beads and four heavy atoms with their corresponding hydrogen atoms are mapped into a single bead. Remarkably, this relatively simple representation of molecules has been parameterized for a whole host of materials, including proteins70,71, DNA72, numerous polymers and block co-polymers such as PEG (polyethylene glycol) and PEO (polyethylene oxide)73, and polysaccharides74 (Figure 5a). The only major downside is the inability to capture conformational changes in proteins, as the secondary structures have to be constrained via an elastic network. Therefore, only tertiary structures can be determined from this MARTINI CG method.
Figure 5.

Coarse-grained MD simulations can be performed on a wide variety of material and biological systems, such as a) lipid membranes and transmembrane proteins with the MARTINI forcefield75, adapted from Ref [63] licensed under CC BY from Nature Publishing Group. b) DPD simulations of polymer aggregation in inks for 3D printing76, adapted from Ref [64] with permission from John Wiley and Sons, and c) hydrocarbons with the VOTCA methodologies77. Adapted with permission from Ref [65]. Copyright (2009) American Chemical Society.
Another physics-based method of coarse-graining is dissipative particle dynamics (DPD)78,79 where the CG beads are typically bonded with soft Hookean springs or the finitely extensible nonlinear elastic (FENE) model80. In DPD simulations, the total force exerted on the CG beads consists of three components: a dissipative force, a random force, and a conservative linear repulsive force. The dissipative and random forces serve as a thermostat for maintaining the system’s temperature while the conservative force mimics excluded-volume interactions to confer each CG bead with a corresponding chemical identity. Similar to the MARTINI forcefield, DPD simulations have been widely applied to polymers76,81, proteins27,82,83, biomembranes84, and complex fluids85. For instance, critical parameters for the processing of bio-inspired silk fibres have been examined with DPD in order to predictively design hierarchical materials27. Optimal ratios of the hydrophobic and hydrophilic domains in the copolymer and their ideal lengths are determined, which led to the formation of fibres with excellent mechanical properties, and these are verified experimentally. This work is extended further to examine how terminal modification of the peptides affects fibrillation83. It is found that increasing the molecular weight at the terminals while decreasing the hydrophobicity will enhance the alignment of the peptide chains under shear. The mechanical strength and elasticity will also be improved at the cost of diminished extensibility and toughness. The properties of ultraviolet inks for 3D inkjet printing in nanoscale additive manufacturing have also been modelled with DPD76 and it is shown that sodium dodecyl sulphate is an effective surfactant that can decrease the average size and enhance the uniformity of agglomerates of polyethylene glycol and polystyrene (Figure 5b).
Many other methods are also gaining prominence, particularly those which are integrated into the Versatile Object-oriented Toolkit for Coarse-graining Applications (VOTCA)77,86,87 (Figure 5c). These are a family of structure-based CG methods. One such method is the iterative Boltzmann inversion (IBI) approach88 which aims to reproduce experimentally or numerically determined radial distribution functions (RDFs) by iteratively improving a coarse-grained pair potential that has the mathematical form
| (Equation 4) |
Successful convergence for a given RDF, ρtarget(rij), ensures a unique CG pair potential. The inverse Monte Carlo (IMC) method89 is also implemented in VOTCA which, in a similar manner as IBI methods, aims to reproduce a coarse-grained Hamiltonian through iterative convergence of a linear combination of forcefield terms to a target RDF. A third method of relative entropy minimization90 is that a CG ensemble can mimic a target fully atomistic simulation by optimizing the CG potential parameters such that the relative entropy between the CG and atomistic systems is minimized. This is possible because the overlap between the configurational phase-space of two molecular ensembles can be characterized by their relative entropy. A final notable method of coarse-graining is through force-matching where CG pairwise forces are derived from atomistic forces calculated from several reference sites91.
3.3. Continuum Modeling
Continuum modeling and simulations can be harnessed to probe material phenomena that are beyond the mesoscale, due to limitations in computational resources that are capable of determining the interactions between atoms or CG beads that number more than several orders of magnitude larger than those discussed in previous sections. This problem is circumvented by modeling the systems of interest as continuous media, in contrast to the discrete particles modeled in DFT, MD, or CGMD, and it is assumed that the infinitesimal subdivision of the material still retains the same material properties as those of the bulk. In a typical macroscopic continuum model, a set of coupled partial differential equations (PDEs) are constructed from fundamental physical laws, including conservation of mass, momentum, and energy, together with information regarding the materials’ properties, known as constitutive relations. These equations can then be solved with specific boundary conditions. However, a purely analytical solution to this set of equations can usually only be found for relatively simple systems. Therefore, numerical methods that approximate the solutions to these equations are needed.
One such numerical method is the finite element method (FEM). FEM has been widely used to solve a large number of engineering problems in structural optimization, solid mechanics, heat transfer, and fluid flow, just to name a few. True to its name, FEM entails the division of the system of interest into simpler domains (finite elements) that are connected through nodes. This representation of finite elements and nodes is also known as the mesh and it enables the continuum analysis of systems with complex geometries. In conventional FE analyses of the mechanics in linear systems, a system of linear equations is constructed based on the distribution of nodes and the nodal forces, displacement, and stiffnesses. This system of equations can then be solved using quite a number of different methods such as Gauss elimination, LU decomposition, or advanced multigrid method. As FEM is an extremely popular computational method that has been covered in a large number of publications, including undergraduate and graduate level textbooks92–96, a detailed exposition of FEM is beyond the scope of this review. As the FEM analysis of macroscopic material properties is extremely flexible and rapid, it can be combined very effectively with additive manufacturing to design and create novel bio-inspired materials. This is discussed extensively in Section 1 and 3.4 of this review.
However, FEM has its drawbacks, particularly due to the need for a properly predefined mesh to connect the nodes97. The meshing process consumes significant amounts of resources and time as an analyst needs to customize the mesh to suit the problem, especially when there are complex three-dimensional geometries. Conventional FEM formulations also assume that the displacement field is piecewise continuous, resulting in discontinuous stresses at the interfaces of the elements. This leads to inaccurate prediction of stresses, requiring special techniques to circumvent this issue. FEM also encounters significant difficulties in modeling systems undergoing large deformations or experiencing fracture98. These problems arise from the fundamental need for a mesh of the system, therefore a broad class of modeling techniques, known as meshfree methods, seeks to remove this need for a user-predefined mesh to discretize the computational models99. This is achieved by dispersing nodes both within and on the boundary of the domain of interest, thereby representing, in contrast to discretizing, the problem domain without forming a mesh. This means that the relationship between the nodes need not be known a priori. The main downside of meshfree methods is that they are relatively new in comparison to FEM and many categories of meshfree methods are under active development to solve broader classes of problems. There is also a lack of broadly distributed software that enables the rapid application of meshfree methods to generalized problems100. Despite these limitations, meshfree methods have been applied to numerous specific problems in structural and fracture mechanics101.
While numerous computational methods can be applied at different discretized spatiotemporal scales, the hierarchical variations in material properties across these scales may not be as obvious. A simple illustration of this is the fact that a smaller object has a larger ratio of surface area to volume compared to another object that has the same geometry and material but with greater dimensions. For instance, the surface area of a sphere scales according to the square of its radius but the volume scales with the cube of the radius. However, similar scaling in material properties become much more difficult to derive analytically for objects that have complex geometries and non-uniform material distribution. Therefore, computational modeling can aid in the understanding of how these material properties vary according to size by approximating the underlying scaling laws, thereby aiding the design of the same material at different length scales.
For example, scaling laws related to density have been derived for the mechanical and thermal properties of three-dimensional (3D) porous graphene that has gyroidal geometry102,103. On the one hand, it was demonstrated that higher densities led to better mechanical properties and this was verified experimentally through mechanical tests of 3D-printed polymeric gyroid structures102. On the other hand, gyroid graphene’s thermal conductivity did not significant variation with density, in contrast to typical porous materials such as silica aerogels104–106. These studies provide critical insights into design rules for light, stiff, and thermally insulating 3D porous graphene structures at different length scales with varying hierarchical features.
3.4. Optimization
The designs of natural materials follow a form-function relation in which structure dictates properties. At the same time, requisite properties defined by the environment and needs for survival carve out efficient material structures through natural selection. Many of nature’s materials are optimized for their own survival needs; these needs may include self-cleaning and hunting, for example107,108. Engineering materials, on the other hand, need to meet certain requirements for their applications, which may include stiffness, strength, toughness, and low weight. Work on optimization of various materials for high performing mechanical properties is prevalent in literature. Many works are related to optimization of composite laminate designs using genetic algorithms, which are based on evolutionary algorithms and concepts109,110. Topology optimization groups strive to optimize properties such as structural compliance based on known loads and boundary conditions, showing how an initial starting geometry slowly forms an optimized truss-like structure111–115. Researchers have also studied how to optimize composite topologies with three materials for thermal conductivity properties116. All of these works lead to an emerging trend of materials-by-design in which properties of materials can be tuned to meet desired functions and applications.
The structural complexity of optimized solutions has often served as a barrier to manufacturing. As a result, most of the optimization work, a lot of which are topology, in the past has been limited to theoretical work117,118. Recent advances in additive manufacturing have enabled researchers to physically create specimens with designs obtained from optimization algorithms119–127. Gaynor et al. used topology optimization to optimize multiple materials for compliance and used additive manufacturing to fabricate the designs and experiments to gauge their performance119. Recently, the authors developed an optimization algorithm to tune the distribution of stiff and soft phases, optimized for toughness, in a composite material with a given crack128. Applications for this work include designing materials to alleviate high stress concentrations at crack tips and preventing a pre-existing crack from propagating. A representative design obtained from the optimization algorithm with 12.5% volume fraction of soft material is shown in Figure 6b, where the grey zones represent a relatively stiff material compared to the black zones, which represent the softer material. It can be seen that the soft material surrounds the crack tip area to protect it and forms a wing-like shape around the crack tip. The strain field is shown in Figure 6b and it can be seen that the strain is no longer localized at the crack tip, but rather more distributed where the highest strained areas incorporate soft material that can handle more strain. Details of the simulation and parameters used can be found in the publication by Gu et al.128. What is more astonishing is that the design, obtained purely from mathematics and modeling, looks very similar to the design of dragonfly wings near the nodus point where cracks are most likely to initiate. Figure 6a shows a model representation of scanning electron microscope images of the dorsal and ventral structures near the nodus point129. It can be seen that at the nodus point in the ventral side there is a wing-like shape of soft material, similar to the shape obtained from optimization. This suggests that natural materials, exemplifying the same mathematical conclusion, have crack deflecting mechanisms that minimize damage. With multi-material 3D-printing, researchers can now study algorithmically optimized designs and develop future engineering materials with crack-deflecting mechanisms that perform as well as dragonfly wings, for example. In this work, optimization was carried out using an algorithm that constantly searches for better designs through iteration.
Figure 6.

Design by optimization. a) A model representation of scanning electron microscope (SEM) images of the dorsal and ventral structures of dragonfly wings near the nodus point, where cracks are most likely to initiate129. Adapted from Ref [115], Copyright (2014), with permission from Elsevier. b) A representative design obtained from the optimization algorithm with 12.5% volume fraction of soft material, where the grey zones represent the stiffer material, and the black zones represent the softer material. It can be seen that the soft material surrounds the crack tip area to protect it and forms a wing-like shape around the crack tip. The strain field shows that strain is no longer localized at the crack tip, but rather more distributed where the highest strained areas coincide with the soft material enabling it to handle more strain128. Designs obtained purely from mathematics and from modeling look similar to the wing-like shape of soft material near the nodus point of the dragonfly wing, highlighted in the blue dotted areas.
Another approach to design new composites is using machine learning. Machine learning is a branch of artificial intelligence that uses a variety of techniques that allows computers to learn from past experience to detect patterns that hard to discern in large complex data sets. Machine learning is useful for problems that have a pattern associated with it, have lots of data, and are difficult to solve mathematically. Recently, Gu et al. applied machine learning to explore new geometric patterns with stiff and soft distributions of materials130. The study showed that machine learning was able to accurately predict composite material properties, such as fracture and strength, and learn the patterns for geometries that performed the best. Additionally, machine learning was able to accelerate the exploration processes compared to finite element methods and other traditional evaluation models. In fact, the study showed that machine learning was orders of magnitudes faster than a finite element approach when it comes to predicting material properties. Therefore, machine learning can be an instrumental tool for designing new materials for future engineering applications.
4. Materials-by-Design: Case Studies and Impacts
The materials-by-design paradigm is strongly driven by the powerful integrated connection between experimental techniques in synthetic chemistry, additive manufacturing, and process engineering at the macroscale, as discussed in Section 2, with theoretical design at the atomic level and modeling of materials at the nano-meso-micro-scale, as detailed in Section 3. To provide concrete examples of this integration, three case studies are presented here. Firstly, for the design of 2D materials, we discuss how their electronic band structures can be analyzed and tailored through DFT calculations for advanced implementations in electronic devices. The mechanical properties of 3D structures derived from these 2D materials can also be determined with MD simulations to inform experimental synthesis and characterization. Secondly, we examine the interplay between in silico prediction and the de novo design of a versatile ultrafiltration membrane derived from silk proteins. Thirdly, we detail the integration of recombinant protein technology with MD simulations for the production of novel stimuli-responsive hydrogels based on silk-elastin-like proteins.
4.1. Two-dimensional Materials Design
Graphene and Other Two-dimensional Materials
Atomically thin two-dimensional materials and their unique mechanical, electrical and optical properties make them ideal for functional bio-mimicry designs such as in material toughening, electrical and optical sensing, and smart active designs. After the initial development of transistors in 1948 at the Bell Laboratory131, electronic devices have changed our life very significantly. At the fundamental level, most electronic devices are designed to control system or process information by controlling electrical currents through the applied external voltage to open a current channel in semiconducting materials. Therefore, the electronic properties of materials, such as the band structures, density of states, electron transport, resistance, and conductivity, are key factors that have to be considered in the manufacture and implementation of electronic devices. The applied voltage experiences a process of screening that is inherent to the bulk structure of the constituent material. This screening is diminished if the thickness is reduced and the electrons are confined in at least one of the spatial directions. Accordingly, thin structured materials are ideal for the development of electronic devices. This effect is maximized in atomically thin two-dimensional (2D) materials. The confinement helps to reduce the screening of the applied voltage. Among the 2D materials, graphene is specially known for its extremely intriguing properties, such as high mobility of electrons, which were predicted theoretically132,133. Informed by previous theoretical research, graphene was isolated and characterized for the first time in 2004134 by Andre Geim and Konstantin Novoselov who were awarded the Nobel Prize in Physics in 2010 for their innovative way of obtaining graphene samples from graphite135. Since graphite is an allotrope of carbon that consists of stacked graphene layers by π–π interactions, Geim and Novoselov were able to exfoliate single-layered graphene from graphite by using Scotch tape. This simple but innovative technique can produce high quality micron-sized samples while also being equally applicable to other 2D materials that are multilayered bulk systems. In the case of graphene, its mechanical, electrical, and thermal properties are found to be exceptional: the thermal conductivity is estimated to be 5,000 W/mK136; the electron mobility is approximately 15,000 cm2/V·s at room temperature134; the high elastic modulus is around 1 TPa and its high intrinsic strength is around 130 GPa137. These properties have placed graphene as one of the leading materials for innovation in the field of condensed matter physics and materials science.
High-quality, single-crystalline graphene can be obtained from physical exfoliation. However, it has limited ability to produce large-area graphene in sufficient quantities for practical applications. Chemical vapor deposition (CVD) is one of the most successful way for large scale production141. The nucleation of carbon atoms on a metallic substrate (Ni or Cu) initiates the growth of graphene flake into large domains. Due to the grain boundaries of the metal substrates, the graphene produced by CVD also has grain boundaries where the boundaries are well-stitched with pentagon and heptagon defects142,143. The grain boundary affects the various properties of graphene based on their atomistic arrangements. However, the grain size obtained from CVD has improved significantly to produce inch-sized single-crystalline graphene144.
Although graphene is well known for its ballistic transport and high electron mobility, it has no band gap. Possessing a band gap is vital for various electronic applications, such as transistors145. It is possible to open the band gap of graphene by processing them in the form of nano-ribbons146 or applying uniaxial mechanical strain147. However, the mobility drops significantly to around 100 ~ 200 cm2/V·s for a 150 meV bandgap because the high mobility of electrons in graphene is a direct consequence of not having a band gap. Thus, researchers have expanded their attention to other two-dimensional materials, such as molybdenum disulfide (MoS2) and their transition metal dichalcogenide (TMD) family, hexagonal boron nitride (h-BN), black phosphorous, and many other 2D materials. In particular, monolayer MoS2 has attracted significant attention due to its excellent electron transport with intrinsic direct bandgap (1.8 eV)148. Figure 7a shows some representative two-dimensional materials of graphene, MoS2, h-BN, and their band structures.
Figure 7.

a) The atomic structures of 2D materials, including graphene, MoS>2, and h-BN, and their band structures. Band structures are obtained with DFT calculations in the Quantum-Espresso package138 using the Perdew-Burke-Emzerhof (PBE) functional139 and norm-conserving type pseudopotential140. b) Heterostructures with van der Waals interaction: hBN-Graphene-MoS2-Graphene-hBN. c) Lateral heterostructure of MoS2-WSe2.
The properties of graphene devices are strongly affected by their substrates as they are atomically thin, which limits their integration with conventional silicon-based devices. Utilizing h-BN as a substrate for graphene devices improved the mobility by approximately an order of magnitude as compared to devices with SiO2 substrates149. This research inspired other studies of new material systems with van der Waals heterostructures, including graphene, h-BN, MoS2, and WSe2150,151 as shown in Figure 7b152, which are emerging candidates for the next generation of electronic devices. Moreover, wafer-scale heterostructures of WS2, MoS2, and MoS2 were reported153. Alternatively, instead of weak binding by stacking vertically, laterally bonded structures between h-BN and graphene show great promise for atomically thin circuits154. Subsequently, in combination with the TMD family, various lateral heterostructures were proposed155–157. Through epitaxial growth, atomically sharp interfaces between MoS2-WSe2 lateral junction was also reported and this is illustrated in Figure 7c158. Through permutations of these atomically thin graphene and other two-dimensional materials with lateral and vertical heterostructures, highly efficient and flexible devices can be realized.
We briefly introduced that graphene and two-dimensional materials have shown great promise as materials for electronic devices with their astonishingly excellent properties. Another exceptional property of graphene is its mechanical strength that is known to be 200 times stronger than steel. However, it is simply not possible to utilize graphene’s mechanical advantages for large-scale construction materials due to its atomistic thickness. Thus, researchers have tried to transfer the properties of 2D graphene to 3D bulk materials. There are well-known examples of different dimensionalities from 2D graphene building blocks. The main feature of graphene as a carbon allotrope is its chemical bonding type: the sp2 bond. Wrapping graphene into 0D or rolling into ID structures result in fullerenes or carbon nanotubes respectively, which preserves the sp2 bonds. Stacking graphene into 3D structures is graphitic in nature. The van der Waals interactions required for stacking graphene is extremely weak, and this limits graphene’s application in 3D bulk systems. To overcome this problem, triply periodic minimal surfaces can be utilized for 3D graphene assembly, where the gyroid shape is proposed to be a good candidate for light and strong graphene materials102.
Recently, by building 3D structures inspired by origami and kirigami, the traditional Asian arts of paper folding and cutting, stretchable transistors can be constructed159 (Figure 8a). It might be counterintuitive to fold graphene into 3D origami structures because the 3D structures are not stable with low bending stiffness. However, it was revealed that the static ripples in the graphene membrane increased graphene’s effective thickness and thus served to stiffen the membrane. Moreover, controlling the crumpling and unfolding of large-area graphene on a polymer substrate, which are reversible due to its flexibility and robustness, was reported and tunable properties such as wettability and transmittance were demonstrated161. Self-assembling graphene ribbons from self-tearing and peeling from the substrate was observed at micrometer scale, which were activated by thermal fluctuation and may lead to self-assembled patterning in the future160 (Figure 8b). The large area re-foldable crumpling of graphene generated by relaxation on pre-stretched polymers showed multiple functionality, such as tunable hydrophobicity, transparency, and resistance (Figure 8c).
Figure 8.

a) Stretchable graphene electrode. Kirigami in-plane spring with (i) paper and (ii) graphene, and (iii) the graphene kirigami spring159 being stretched by 70%. Adapted by permission from Nature Publishing Group: Nature, Ref [157], copyright (2015). b) A series of snapshots of graphene ribbon self-assembly with self-tearing by van der Waals interaction and thermal energy160. Adapted by permission from Nature Publishing Group: Nature, Ref [158], copyright (2016). c) Multifunctional re-stretchable crumpled graphene on polymer161. (i) Schematic of the deformation of graphene on the pre-stretched biaxial polydimenthylsiloxane (PDMS). (ii) Delamination and buckled geometry of graphene after uniaxial relaxation of the substrate. (iii) Crumples after the biaxial relaxation of the substrate. Adapted by permission from Nature Publishing Group: Nature Materials, Ref [159], copyright (2013).
All these observed behaviors indicate that there are many possible ways to build 3D structures from graphene, and it is crucial to understand the role of the atomic thickness in modulating the properties of the generated structures.
Computational Modeling and Design
In this section, we provide more specific details on how computational modeling can assist in understanding and tailoring the properties of 2D materials. In the case of graphene-like nanomaterials, theoretical band structures of graphite with zero band gap were reported seventy years ago162. Other theoretical modeling and simulations of graphene studied the Dirac fermion, spin-orbit coupling, and quantum hall effects163. Currently, large-area graphene can be produced albeit with polycrystalline features. The grain boundaries include topological defects, i.e., pentagon and heptagon features, which are known to degrade the electrical, thermal, and mechanical properties drastically. These pentagon and heptagon rings are mostly located at the boundaries based on the relative rotational angles between grains. Modeling, such as with tight-binding (TB) methods or DFT calculations, allows us to study the differences in electronic properties due to specific atomic arrangements. For example, DFT calculations show how the band gap can be controlled by tailored periodic grain boundaries164, which opens the gap to ~1 eV. Due to relatively long-range strain field from these defects, DFT cannot handle the system sizes in many cases to determine mechanical and thermal properties. As discussed in Section 3.2, reactive forcefields with MD simulations can model bond breaking and formation in larger system sizes compared to those achievable by DFT or TB methods.
From the experimental point of view, AFM nano-indentation of suspended CVD graphene showed that their mechanical strength did not degrade as much as the electronic and thermal properties with increasing numbers of grain boundaries in comparison with single crystalline graphene. It was also demonstrated that crack propagates into the grains165. Their statistical results showed that smaller grain sizes had lower strength but broader distributions166, which implied that the strength of polycrystalline graphene depends on randomized defect distribution. MD simulations with realistic randomly oriented grains also reported different trends with the grain sizes167. The trend that the Young’s modulus increases as the grain size increases is consistent. However, the trend for the strength of polycrystalline graphene varied significantly. Kotaki and Yang reported size-independent strength168,169 while Sha and Chen reported increasing strength as the grain size increased. Both models were built in a similar way, however models that showed size-independent strengths had small voids at the boundaries. The regions that had voids can be considered as having an initial crack, and the pertinent property that must be considered in the presence of a crack is the fracture toughness. Based on the Griffith criteria22, the fracture toughness of CVD graphene was measured to be 4.0 ± 0.6 MPa/m1/2. The MD study revealed that there was enhancement of fracture toughness in polycrystalline graphene as the stress-concentration was delocalized (Figure 9a). The study showed that the mechanism was activated with the out-of-plane deformation170. Pentagon-heptagon defects causes out-of-plane deformation, where the elastic modulus is lower and becomes more stretchable locally in comparison with the regions of pristine, hexagonal lattices. Thus, it can be considered to be a composite of soft and stiff materials with enhanced fracture toughness as discussed with 3D printed polymer composites, implying that complexities in the mechanical behaviors of polycrystalline graphene can arise not only from the defects themselves but also from their out-of-plane deformation.
Figure 9.

Molecular dynamics (MD) simulations with a reactive interatomic forcefield: constructing the models, deformation under mechanical loads, and stress-strain behaviors. a) Polycrystalline graphene170. Adapted from Ref [169], Copyright (2015), with permission from Elsevier, (i) The 5-7 ring defects at the grain boundary caused intrinsic out-of-plane deformations. (ii) The spatial stress distribution of polycrystalline graphene at the failure points in panel a-(iii). (iii) The stress-strain curves of the polycrystalline and pristine graphene. b) 3D graphene assembly102. (i) 11 × 11 × 11 nm 3D graphene assembly fused via randomly distributed graphene flakes with spherical inclusions. The structures under tension and compression with εx = −0.5, 0.0, and 0.6 for (ii) to (iv) respectively. (v) Stress-strain curve of the assembly under compression and tension. c) Gyroid porous graphene102. (i) Representative geometry of 3D gyroid graphene. The deformation and fracture under compression with εx = 0.0, 0.1, and 0.15 for (ii) to (iv) respectively. (v) Stress-strain curves of 3D gyroid graphene with different porosity sizes from 3 nm to 20 nm102. (b-d) are adapted from Ref [100] licensed under CC BY-NC 4.0 from AAAS.
Furthermore, graphene aerogels are 3D porous materials that utilize graphene’s excellent mechanical strength, high electrical conductivity, and low thermal resistance171. Recently, newly developed methods to produce these aerogels by 3D printing were also proposed172,173. Although these 3D structures possessed light weight and good mechanical stability overall, the reported quality and properties had a wide range of variations. It is intriguing to determine the limit of the properties of 3D graphene assemblies. The gyroid shape was utilized to idealize the 3D graphene structure in MD simulations with all bonds being sp2 hybridized and having minimal surfaces102,103. The results showed that the mechanical strength could be greater than 10 times that of mild steel at a mere 4.6% of the density of mild steel102, and the thermal properties is low and density-insensitive103. These studies demonstrate that computational modeling can help to predict superior structures for designing 3D graphene materials (Figure 9b and c).
Due to its atomic thickness and high crystallinity, 2D materials are flexible and easily deformable by mechanical loading. Theoretically, strain engineering is expected to enable control over the electronic properties of TMD materials through mechanical strains174,175. It was confirmed experimentally that strain induced tunable band gaps in WSe2 and MoS2176,177, indicating that strain engineering is an effective method for tuning the optical and electronic properties of TMD materials. Polycrystalline MoS2 was synthesized with CVD where triangular grains was grown with MoO3 and S precursors178. Unlike graphene, many types of defect structures at the grain boundaries of polycrystalline MoS2 were reported179,180, and these can be studied systematically with modeling in the future.
Characterization (in-situ, and quantitative comparison to modeling)
In the previous sections, we briefly explored the history and modeling of graphene and other 2D materials. Here, we explain different ways of integrating experimental characterization and computational modeling. Transmission electron microscopy (TEM) is one of the most fundamental methods of imaging an atomic structure. The hexagonal atomic structure of the suspended monolayer graphene was observed with TEM in a previous study181. Also, the electron diffraction patterns in this study revealed the natural corrugation of graphene without substrates. Furthermore, high-resolution transmission electron microscopy (HRTEM) allowed the analysis of not only the atomic structures but also the charge redistribution due to chemical bonds such as nitrogen-doped point defects in graphene, which could be directly compared with the results of DFT182. Interestingly, the electron beam utilized in aberration-corrected transmission electron microscopy (AC-TEM) enabled the in situ manipulation and monitoring of atomic configurations, such as well-designed defects183 and monitoring behaviors of dislocation pairs in graphene184. Recently, the dynamics of the propagation of an atomically sharp crack tip was tracked in situ at the atomic level for two-dimensional materials (MoS2) with the AC-TEM technique (Figure 10a). The interaction between the crack tip and defects in the system were observed directly and compared with MD modeling, thus paving the way towards understanding the fundamental behaviors of materials fracture and enabling the bottom-up design of materials with enhanced fracture toughness185.
Figure 10.

in-situ characterization with AC-TEM. a) A crack propagating through monolayer MoS2. (i) in-situ AC-TEM images. (ii) Geometries obtained from MD simulation185. Adapted with permission from Ref [184]. Copyright(2016) American Chemical Society. b) SEM images of (i) 1D MoS2 channel embedded in WSe2 from (ii) an atomically sharp MoS2-WSe2 hetero junction. (iii) MD simulations of the channel growth. The 5-7 ring play a catalytic role in promoting 1D MoS2 channel growth186. Adapted by permission from Nature Publishing Group: Nature Materials, Ref [185], copyright (2017).
Likewise, scanning transmission electron microscopy (STEM) with annular dark field (ADF) imaging can distinguish atoms in MoS2-WSe2 lateral junctions based on the fact that the ADF signal is proportional to the atomic number (~Z1.7). This property enabled the qualities of the sharp interface junctions to be characterized and mapped directly to the atomic structures158. Recently, a MoS2 channel that had a width of a few nanometers was embedded in WSe2 by utilizing the 5|7 dislocation catalyst at the junction, and the difference between MoS2 and WSe2 was also clearly observed with ADF-TEM186 (Figure 10b). For measuring strains in graphene or other 2D materials, Raman spectroscopy is a sensitive and convenient technique that can also be compared with DFT calculations187. Moreover, multiple layers of graphene or MoS2 also produce uniquely identifiable Raman spectra38,188. Thus, precise characterization of the number of layers in 2D materials is possible with this technique. Furthermore, the quality of 3D printed graphene aerogel was quantified by looking at the 2D and G band from Raman spectroscopy172, where the G and 2D bands are the main spectra of graphene and they vary sensitively according to the number of layers. More distinct bands indicated the high quality of graphene without functional groups such as oxides.
4.2. Silk Ultrafiltration Membrane Design
Water purification membranes received significant attention in recent years with the aim of addressing global challenges in water pollution and the scarcity of potable water. Compared to other water-purification techniques, such as coagulation, flocculation, and sedimentation, membrane-based methods are more energy- and cost-efficient for removing molecular-level contaminants. However, it remains a significant challenge to fabricate low-cost water purification membranes while retaining excellent mechanical strength and high purification performance. A series of studies demonstrated that nanomaterial-based multilayer filtration membranes, such as graphene, graphene oxide, CNTs, tungsten disulfide, and molybdenum disulfide, can solve this problem189,190, due to their unique advantages of having a low-pressure drop, high molecular loading capacity, enhanced throughput, and high filtration efficiency. However, there is still a large room for improvement in the filtration-efficiency of these membranes since nano-porous structures can only be developed through complex processing and water can only permeate between gaps and interlayer spaces between the nanosheets during filtration.
Ling et al. produced de novo designs of a versatile ultrafiltration membrane by integrating computational modeling and experimental fabrication8. The first step for such a de novo design was to use computational modeling to predict the conditions under which systems that contain protein nanofibrils and nano-sized flakes of minerals can be assembled into multilayer structures. The authors established coarse-grained models of protein nanofibrils and mineral plates and simulated their self-assembly behaviors during vacuum filtration. The simulation results revealed that the multilayer structure of these two components could only form with weak interactions between nanofibrils and mineral plates. From this in silico prediction, a series of protein/mineral systems were scanned and finally a system that contained silk nanofibrils and hydroxyapatite were selected for experimental fabrication. In good agreement with the in silico prediction, the silk nanofibril/hydroxyapatite membranes formed highly ordered nano-sized multilayer structures. Remarkably, the resultant membranes not only showed ultrafast water penetration that was over five times greater than other reported ultrathin membranes with similar thickness, but it also exhibited universal and high efficiency in removing and even reusing (in some cases) heavy metal ions, dyes, proteins, and nanoparticles. More importantly, this synergistic integration of computational modeling and experimental fabrication is not only suitable for this specific case of silk nanofibril and hydroxyapatite systems. It is also useful for constructing multilayer nanoporous structures from different natural protein-based engineered peptides (e.g., amyloid, collagen fibril, etc.), protein-based fibrils (e.g., silk-, silk-elastin-, silaffin-, apatite-binding-peptide-based peptides and proteins) and their corresponding biomineral systems (e.g., calcification, silicification, etc.). Therefore, this design and fabrication strategy for membrane synthesis can lead to the production of functional materials for water purification and other practical applications.
4.3. Designing Silk-Elastin-Like Proteins: Going Beyond Nature
The field of biochemistry underwent a dramatic revolution as the technology to synthesize recombinant proteins matured over the past few decades. This technology led to the ability to rapidly express and purify large volumes of proteins with customized sequences for a wide variety of uses in scientific research, industrial processing, and development of commercial products. In particular, the production of biomaterials based on proteins, DNA, and cells was a major beneficiary191. The various mechanisms of producing recombinant proteins were covered extensively in numerous textbooks192 and reviews3, so they will not be described extensively here. Instead, the design, production, and characterization of recombinantly synthesized silk-elastin-like proteins (SELPs) will be the focus in this case study. SELPs are block co-polymers of tandemly repeating units of elastin-like and silk-like protein sequences. The elastin-like portion is derived from the hydrophobic domain of tropoelastin and has the protein sequence GXGVP, where varying the residues in the X position determines the particular stimuli-response the SELP can undergo, such as structural transitions with changes in temperature, pH, or light. The silk-like portion has the protein sequence of GAGAGS and it is derived from the β-sheet forming heavy chain domain of the Bombyx Mori silkworm silk. It can serve as tunable crosslinks through β-sheet formation, thereby tuning the mechanical properties of the resulting biomaterials.
The precise protein sequence of SELPs can be customized through the construction of different expression vectors and standard protocols in molecular biology197, including the use of pET expression vectors with the T7 promoter195. The nomenclature of the synthesized SELPs depend on the ratio between the silk- and elastin-like blocks, as well as the residues in the X position, such that different SELP constructs are labeled as SmEnX. For instance, if a SELP construct has a silk-to-elastin ratio of 1:8 and a tyrosine at the X position, it will be labeled as SE8Y. As a result of this diversity and ease of tailoring SELPs on a molecular level, numerous engineering applications were proposed to take advantage of the stimuli-responsive nature of SELPs. For instance, SELPs were spun into fibers through various methods, such as wet-spinning194 (Figure 11b) and electro-spinning198. SELPs were also found to be capable of self-assembly (Figure 11c), such as the formation of nanofibers on mica surfaces199 or thermally triggered formation of micellar-like nanoparticles while interacting with hydrophobic drugs like doxorubicin196. Plasmonic nanodevices were developed and they responded to thermal stimuli193. These were made from functionalized gold nanoparticles and thermoresponsive SELPs (Figure 11a). Self-assembly was demonstrated through thermally reversible conformational switching in the elastin-like block while silk-like blocks stabilized the transition by forming β-sheets. Thermal cycling of these aggregates changed the UV-vis extinction spectra reversibly, showing their versatility for applications in plasmonic nanodevices. SELPs are also highly suitable for drug delivery vehicles and tissue engineering200, particularly when they are crosslinked to form hydrogels through horseradish peroxidase-catalyzed tyrosine crosslinking201.
Figure 11.

Applications of SELPs in a) plasmonic nanodevices193, adapted with permission from Ref [192]. Copyright (2014) American Chemical Society, b) wet-spun fibers194, adapted with permission from Ref [193]. Copyright (2009) American Chemical Society, and c) self-assembled nanoparticles195,196, adapted with permission from Ref [194], Copyright (2011) American Chemical Society, and also adapted from Ref [195], Copyright(2014) American Chemical Society.
SELPs can now be characterized with molecular precision through the incorporation of computational molecular modelling201–203, especially with fully atomistic and coarse-grained molecular dynamics. For example, through replica exchange molecular dynamics (REMD) simulations of SELP peptides, a transition temperature between 300K and 320K was determined for SELPs with valine as the X-residue, whereas this temperature transition was suppressed for a lysine residue, which corroborated experimental findings201. Further REMD studies were performed on SELPs with tyrosine at the X position to probe the underlying molecular mechanisms responsible for the structural transitions203. The incorporation of tyrosine residues conferred thermos-responsiveness to the resulting SELP hydrogel and SEgy constructs were experimentally determined to have transition temperatures around 21°C201. Above this temperature, the hydrogel shrank to less than 10% of its original volume. This was attributable to a decrease in the end-to-end lengths of each SELP molecule as well as diminishing radius of gyration, accompanied by increasing amounts of hydrogen bonds. Tensile tests were also performed through steered MD where unfolding mechanisms were found to be highly dependent on temperature and the free energy barrier is significantly lower at colder temperatures. It is anticipated that this synergistic interplay between computational simulations and experiments will eventually realize the prospect of high throughput molecular design of SELP constructs by accurately predicting the structural and mechanical properties computationally before experimental synthesis.
5. Conclusions
In this review, we detailed the development of the materials-by-design paradigm through synergistic integration of experimental material synthesis and characterization with computational modeling and optimization. This paradigm creates the possibility to develop materials according to specific, rational designs from the molecular to the macroscopic scale and we present case studies that utilize this paradigm for materials development. In experimental material synthesis and characterization, we discussed extremely promising techniques that span recombinant protein technology that produces peptides and proteins with tailored sequences encoded by recombinant DNA, self-assembly processes induced by conformational transition of proteins, additive manufacturing for designing complex structures, and qualitative and quantitative characterization of materials at different length scales. We also detailed numerous multi-scale computational modeling techniques that fully complements these experiments: Density Functional Theory at the atomistic scale; fully atomistic and coarse-grain molecular dynamics at the molecular to mesoscale; continuum modeling at the macroscale. These experimental and computational approaches are utilized in an integrated manner to broaden our understanding of the properties of two-dimensional materials and materials based on silk and silk-elastin-like proteins. As the materials-by-design paradigm starts to become a reality, many limitations will impede its progress. Therefore, the next generation of computational techniques capable of simulating larger and more realistic predictive models must be developed to access the mesoscale gap in multi-scale modeling of materials, a gap that lies between atomistic and continuum models that must be filled in the future by the development of faster computers and novel methods.
Looking forward, machine learning is expected to provide a new revolution in computer science which, together with quantum computing, can boost the theoretical framework in which materials are currently designed and simulated. The development of these new computing avenues will impact many different aspects of our society beyond materials, including opportunities in the Internet of things (IoT), cyber-infrastructure, and robotics. All these will have to be coupled with further improvements in characterization techniques, as well as synthetic procedures that are more scalable. In the case of additive manufacturing, substantial hurdles in establishing additive manufacturing as a primary, large-scale, industrial, and processing method need to be addressed. Better cleaning processes of support materials, larger numbers of materials to print, better control over interfaces and material boundaries are some of the challenges that are hindering the growth of additive manufacturing into an established manufacturing technique. Overall, the achievement of these new, revolutionary advances in materials synthesis will help to tackle many of the grand challenges defined by the National Academy of Engineering (NAE). These include making solar energy far more economical, restore and improve urban infrastructure, provide access to clean water, or develop carbon sequestration methods. More importantly, achieving success in many of these NAE challenges will contribute significantly to the greater goals for sustainable development globally, defined by UN to include clean water and desalination, affordable and clean energy, sustainable cities and communities, and climate action.
Acknowledgments
The authors acknowledge support from the US Department of Defense, Office of Naval Research (N00014-16-1-233) and the National Institutes of Health (U01 EB014976). G.X.G. acknowledges support from the National Defense Science and Engineering Graduate (NDSEG) Fellowship.
References
- 1.Feynman RP There’s plenty of room at the bottom. Engineering and science 23, 22–36 (1960). [Google Scholar]
- 2.Fothergill-Gilmore LA in Protein Biotechnology: Isolation, Characterization, and Stabilization (ed Felix Franks) 467–487 (Humana Press, 1993). [Google Scholar]
- 3.Rosano GL & Ceccarelli EA Recombinant protein expression in Escherichia coli: advances and challenges. Frontiers in microbiology 5 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Brown TA Gene cloning and DNA analysis: an introduction. (John Wiley & Sons, 2016). [Google Scholar]
- 5.Huang W et al. Synergistic Integration of Experimental and Simulation Approaches for the de Novo Design of Silk-Based Materials. Acc. Chem. Res. 50, 866–876, doi: 10.1021/acs.accounts.6b00616 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 6.Tokareva O, Michalczechen-Lacerda VA, Rech EL & Kaplan DL Recombinant DNA production of spider silk proteins. Microb. Biotechnol. 6, 651–663, doi: 10.1111/1751-7915.12081 (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Rosano GL & Ceccarelli EA Recombinant protein expression in Escherichia coli: advances and challenges. Frontiers in Microbiology 5, doi: 10.3389/fmicb.2014.00172 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Ling S et al. Design and function of biomimetic multilayer water purification membranes. Science Advances 3, doi: 10.1126/sciadv.1601939 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Liu Y, Ling S, Wang S, Chen X & Shao Z Thixotropic silk nanofibril-based hydrogel with extracellular matrix-like structure. Biomater. Sci. 2, 1338–1342, doi: 10.1039/C4BM00214H (2014). [DOI] [PubMed] [Google Scholar]
- 10.Knowles TPJ & Buehler MJ Nanomechanics of functional and pathological amyloid materials. Nat. Nanotechnol. 6, 469–479, doi: 10.1038/nnano.2011.102 (2011). [DOI] [PubMed] [Google Scholar]
- 11.Bai Y, Luo Q & Liu J Protein self-assembly via supramolecular strategies. Chem. Soc. Rev. 45, 2756–2767, doi: 10.1039/C6CS00004E (2016). [DOI] [PubMed] [Google Scholar]
- 12.Ling SJ et al. Directed Growth of Silk Nanofibrils on Graphene and Their Hybrid Nanocomposites. ACS Macro Lett. 3, 146–152, doi: 10.1021/mz400639y (2014). [DOI] [PubMed] [Google Scholar]
- 13.Ling S et al. Modulating Materials by Orthogonally Oriented β -Strands: Composites of Amyloid and Silk Fibroin Fibrils. Advanced Materials 26, 4569–4574, doi: 10.1002/adma.201400730 (2014). [DOI] [PubMed] [Google Scholar]
- 14.Ling S, Li C, Jin K, Kaplan DL & Buehler MJ Liquid Exfoliated Natural Silk Nanofibrils: Applications in Optical and Electrical Devices. Adv. Mater. , n/a-n/a, doi: 10.1002/adma.201601783 (2016). [DOI] [PubMed] [Google Scholar]
- 15.Ling S et al. Printing of stretchable silk membranes for strain measurements. Lab on a Chip 16, 2459–2466, doi: 10.1039/C6LC00519E (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 16.Li C et al. Regenerated silk materials for functionalized silk orthopedic devices by mimicking natural processing. Biomaterials 110, 24–33, doi: 10.1016/i.biomaterials.2016,09,014 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 17.Ling S, Jin K, Kaplan DL & Buehler MJ Ultrathin Free-Standing Bombyx mori Silk Nanofibril Membranes. Nano Lett 16, 3795–3800, doi: 10.1021/acs.nanolett.6b01195 (2016). [DOI] [PubMed] [Google Scholar]
- 18.Zhu Z et al. High-Strength, Durable All-Silk Fibroin Hydrogels with Versatile Processability toward Multifunctional Applications. Advanced Functional Materials, 1704757, doi: 10.1002/adfm.201704757 (2018). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 19.Adamcik J et al. Understanding amyloid aggregation by statistical analysis of atomic force microscopy images. Nat. Nanotechnol. 5, 423–428, doi:Doi 10.1038/Nnano.2010.59 (2010). [DOI] [PubMed] [Google Scholar]
- 20.Ling S et al. Soy protein-directed one-pot synthesis of gold nanomaterials and their functional conductive devices. J Mater Chem B 4, 3643–3650, doi: 10.1039/C6TB00616G (2016). [DOI] [PubMed] [Google Scholar]
- 21.Huang Y, Leu MC, Mazumder J & Donmez A Additive manufacturing: current state, future potential, gaps and needs, and recommendations. Journal of Manufacturing Science and Engineering 137, 014001 (2015). [Google Scholar]
- 22.Gao W et al. The status, challenges, and future of additive manufacturing in engineering. Computer-Aided Design 69, 65–89 (2015). [Google Scholar]
- 23.Guo N & Leu MC Additive manufacturing: technology, applications and research needs. Frontiers of Mechanical Engineering 8, 215–243 (2013). [Google Scholar]
- 24.Gibson I, Rosen DW & Stucker B Additive manufacturing technologies. Vol. 238 (Springer, 2010). [Google Scholar]
- 25.Gibson I, Rosen D & Stucker B Additive manufacturing technologies: 3D printing, rapid prototyping, and direct digital manufacturing. (Springer, 2014). [Google Scholar]
- 26.Ritchie RO The conflicts between strength and toughness. Nat. Mater. 10, 817–822, doi:papers3://publication/doi/10.1038/nmat3115 (2011). [DOI] [PubMed] [Google Scholar]
- 27.Lin S et al. Predictive modelling-based design and experiments for synthesis and spinning of bioinspired silk fibres. Nature Communications 6, 6892, doi:10.1038/ncomms789210.1038/ncomms7892http://www.nature.com/articles/ncomms7892-supplementary-informationhttp://www.nature.com/articles/ncomms7892-supplementary-information (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Studart AR Biological and bioinspired composites with spatially tunable heterogeneous architectures. Advanced Functional Materials 23, 4423–4436 (2013). [Google Scholar]
- 29.Chen C-T, Martin-Martinez FJ, Ling S, Qin Z & Buehler MJ Nacre-inspired design of graphene oxide–polydopamine nanocomposites for enhanced mechanical properties and multi-functionalities. Nano Futures 1, 011003 (2017). [Google Scholar]
- 30.Gu GX et al. Three-dimensional-printing of bio-inspired composites. Journal of biomechanical engineering 138, 021006 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Dimas LS, Bratzel GH, Eylon I & Buehler MJ Tough composites inspired by mineralized natural materials: computation, 3D printing, and testing. Advanced Functional Materials 23, 4629–4638 (2013). [Google Scholar]
- 32.Mirzaeifar R, Dimas LS, Qin Z & Buehler MJ Defect-tolerant bioinspired hierarchical composites: simulation and experiment. ACS Biomaterials Science & Engineering 1, 295–304 (2015). [DOI] [PubMed] [Google Scholar]
- 33.Gu GX, Libonati F, Wettermark S & Buehler MJ Printing Nature: Unraveling the Role of Nacre’s Mineral Bridges. Journal of the Mechanical Behavior of Biomedical Materials (2017). [DOI] [PubMed] [Google Scholar]
- 34.Libonati F, Gu GX, Qin Z, Vergani L & Buehler MJ Bone - Inspired Materials by Design: Toughness Amplification Observed Using 3D Printing and Testing. Advanced Engineering Materials 18, 1354–1363 (2016). [Google Scholar]
- 35.Gu GX, Takaffoli M, Hsieh AJ & Buehler MJ Biomimetic additive manufactured polymer composites for improved impact resistance. Extreme Mechanics Letters 9, 317–323 (2016). [Google Scholar]
- 36.Gu GX, Takaffoli M & Buehler MJ Hierarchically Enhanced Impact Resistance of Bioinspired Composites. Advanced Materials (2017). [DOI] [PubMed] [Google Scholar]
- 37.Watson JD & Crick FHC Molecular structure of nucleic acids. Nature 171, 737–738 (1953). [DOI] [PubMed] [Google Scholar]
- 38.Li H et al. From Bulk to Monolayer MoS2: Evolution of Raman Scattering. Adv. Funct. Mater. 22, 1385–1390, doi: 10.1002/adfm.201102111 (2012). [DOI] [Google Scholar]
- 39.Cui P, Lee J, Hwang E & Lee H One-pot reduction of graphene oxide at subzero temperatures. Chem. Commun. 47, 12370–12372, doi: 10.1039/ClCC15569E (2011). [DOI] [PubMed] [Google Scholar]
- 40.Yeo GC et al. Subtle balance of tropoelastin molecular shape and flexibility regulates dynamics and hierarchical assembly. Science Advances 2 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 41.Sun Y-L, Luo Z-P, Fertala A & An K-N Stretching type II collagen with optical tweezers. J. Biomech. 37, 1665–1669, doi: 10.1016/i.ibiomech.2004.02.028 (2004). [DOI] [PubMed] [Google Scholar]
- 42.Oroudjev E et al. Segmented nanofibers of spider dragline silk: Atomic force microscopy and single-molecule force spectroscopy. Proceedings of the National Academy of Sciences 99, 6460–6465 (2002). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 43.Jung GS & Buehler MJ Multiscale Modeling of Muscular-Skeletal Systems. Annu. Rev. Biomed. Eng. 19, 435–457, doi: 10.1146/annurev-bioeng-071516-044555 (2017). [DOI] [PubMed] [Google Scholar]
- 44.Binnig G, Quate CF & Gerber C Atomic Force Microscope. Phys. Rev. Lett. 56, 930–933 (1986). [DOI] [PubMed] [Google Scholar]
- 45.Rief M, Oesterhelt F, Heymann B & Gaub ΗE Single Molecule Force Spectroscopy on Polysaccharides by Atomic Force Microscopy. Science 275, 1295 (1997). [DOI] [PubMed] [Google Scholar]
- 46.Ashkin A, Dziedzic JM, Bjorkholm JE & Chu S Observation of a single-beam gradient force optical trap for dielectric particles. Opt. Lett. 11, 288–290, doi: 10.1364/OL.11.000288 (1986). [DOI] [PubMed] [Google Scholar]
- 47.Neuman KC & Nagy A Single-molecule force spectroscopy: optical tweezers, magnetic tweezers and atomic force microscopy. Nat Meth 5, 491–505 (2008). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Jung GS, Qin Z & Buehler MJ in Polymer Mechanochemistry Vol. 369 317–343 (Springer Berlin Heidelberg Springer International Publishing, 2015). [Google Scholar]
- 49.Izrailev S et al. in Computational Molecular Dynamics: Challenges, Methods, Ideas: Proceedings of the 2nd International Symposium on Algorithms for Macromolecular Modelling, Berlin, May 21–24, 1997 (eds Peter Deuflhard et al. ) 39–65 (Springer Berlin Heidelberg, 1999). [Google Scholar]
- 50.Allen MP & Tildesley DJ Computer simulation of liquids. (Oxford University Press, 1989). [Google Scholar]
- 51.Rapaport DC The art of molecidar dynamics simulation. 2nd edn, (Cambridge University Press, 2004). [Google Scholar]
- 52.Goldstein H, Poole CP & Safko JL Classical mechanics, 3rd ed., (Addison Wesley, 2002). [Google Scholar]
- 53.Verlet L Computer “Experiments” on Classical Fluids. I. Thermodynamical Properties of Lennard-Jones Molecules. Physical Review 159, 98–103 (1967). [Google Scholar]
- 54.Hansen J-P & Verlet L Phase Transitions of the Lennard-Jones System. Physical Review 184, 151–161 (1969). [Google Scholar]
- 55.Stuart SJ, Tutein AB & Harrison JA A reactive potential for hydrocarbons with intermolecular interactions. J Chem Phys 112, 6472–6486 (2000). [Google Scholar]
- 56.Qian D, Yu M-F, Ruoff RS, Wagner GJ & Liu WK Mechanics of carbon nanotubes. Applied Mechanics Reviews 55, 495–533, doi: 10.1115/1.1490129 (2002). [DOI] [Google Scholar]
- 57.Becker CA, Hoyt JJ, Buta D & Asia M Crystal-melt interface stresses: Atomistic simulation calculations for a Lennard-Jones binary alloy, Stillinger-Weber Si, and embedded atom method Ni. Physical Review E 75, 061610 (2007). [DOI] [PubMed] [Google Scholar]
- 58.Jang I, Phillips R & Sinnott SB Study of C3H5+ ion deposition on polystyrene and polyethylene surfaces using molecular dynamics simulations. JAppl Phys 92, 3363–3367, doi:doi:http://dx.doi.Org/l0.1063/1.1500788 (2002). [Google Scholar]
- 59.Tersoff J New empirical approach for the structure and energy of covalent systems. Physical Review B: Condensed Matter and Materials Physics 37, 6991–7000 (1988). [DOI] [PubMed] [Google Scholar]
- 60.Huang J et al. CHARMM36m: an improved force field for folded and intrinsically disordered proteins. Nature Methods 14, 71–73, doi: 10.1038/nmeth.4067 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 61.Debye P Zerstreuung von Rontgenstrahlen. Annalen der Physik 351, 809–823, doi: 10.1002/andp.19153510606 (1915). [DOI] [Google Scholar]
- 62.Shen Y, Jester SB, Qi T & Reed EJ Nanosecond homogeneous nucleation and crystal growth in shock-compressed SiO2. Nat. Mater. 15, 60–65, doi:10.1038/nmat444710.1038/nmat4447http://www.nature.com/nmat/journal/v15/n1/abs/nmat4447.html-supplementary-informationhttp://www.nature.com/nmat/journal/v15/n1/abs/nmat4447.html-supplementary-information (2016). [DOI] [PubMed] [Google Scholar]
- 63.Hura GL et al. Robust, high-throughput solution structural analyses by small angle X-ray scattering (SAXS). Nat. Methods 6, 606–612, doi: 10.1038/nmeth.1353 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Rice C et al. Raman-scattering measurements and first-principles calculations of strain-induced phonon shifts in monolayer MoS${}_{2}$. Phys. Rev. B 87, 081307 (2013). [Google Scholar]
- 65.Liang L & Meunier V First-principles Raman spectra of MoS2, WS2 and their heterostructures. Nanoscale 6, 5394–5401, doi: 10.1039/C3NR06906K (2014). [DOI] [PubMed] [Google Scholar]
- 66.Kmiecik S et al. Coarse-Grained Protein Models and Their Applications. Chemical Reviews 116, 7898–7936, doi: 10.1021/acs.chemrev.6b00163 (2016). [DOI] [PubMed] [Google Scholar]
- 67.Marrink SJ, Risselada HJ, Yefimov S, Tieleman DP & de Vries AH The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J Phys Chem B 111, 7812–7824, doi: 10.1021/jp071097f (2007). [DOI] [PubMed] [Google Scholar]
- 68.Marrink SJ & Tieleman DP Perspective on the Martini model. Chem Soc Rev 42, 6801–6822, doi: 10.1039/C3CS60093A (2013). [DOI] [PubMed] [Google Scholar]
- 69.Marrink SJ, De Vries AH & Mark AE Coarse grained model for semiquantitative lipid simulations. The Journal of Physical Chemistry B 108, 750–760 (2004). [Google Scholar]
- 70.Monticelli L et al. The MARTINI Coarse-Grained Force Field: Extension to Proteins. Journal of Chemical Theory and Computation 4, 819–834, doi: 10.1021/ct700324x (2008). [DOI] [PubMed] [Google Scholar]
- 71.Gautieri A, Russo A, Vesentini S, Redaelli A & Buehler MJ Coarse-Grained Model of Collagen Molecules Using an Extended MARTINI Force Field. Journal of Chemical Theory and Computation 6, 1210–1218, doi: 10.1021/ctl00015v (2010). [DOI] [Google Scholar]
- 72.Uusitalo JJ, Ingólfsson HI, Akhshi P, Tieleman DP & Marrink SJ Martini coarse-grained force field: extension to DNA. Journal of chemical theory and computation 11, 3932–3945 (2015). [DOI] [PubMed] [Google Scholar]
- 73.Lee H, de Vries AH, Marrink S-J & Pastor RW A coarse-grained model for polyethylene oxide and polyethylene glycol: conformation and hydrodynamics. The journal of physical chemistry B 113, 13186–13194 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.López CA et al. Martini coarse-grained force field: extension to carbohydrates. Journal of Chemical Theory and Computation 5, 3195–3210 (2009). [DOI] [PubMed] [Google Scholar]
- 75.Van Eerden FJ, Melo MN, Frederix PW, Periole X & Marrink SJ Exchange pathways of plastoquinone and plastoquinol in the photosystem II complex. Nat Commun 8 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Aphinyan S, Geethalakshmi K, Yeo J, Shakouri A & Ng TY Numerical characterization of ultraviolet ink fluid agglomeration and the surfactant effect in nanoinkjet printing. Polymers for Advanced Technologies (2017). [Google Scholar]
- 77.Rühle V, Junghans C, Lukyanov A, Kremer K & Andrienko D Versatile object-oriented toolkit for coarse-graining applications. Journal of Chemical Theory and Computation 5, 3211–3223 (2009). [DOI] [PubMed] [Google Scholar]
- 78.Hoogerbrugge P & Koelman J Simulating microscopic hydrodynamic phenomena with dissipative particle dynamics. EPL (Europhysics Letters) 19, 155 (1992). [Google Scholar]
- 79.Espanol P & Warren P Statistical mechanics of dissipative particle dynamics. EPL (Europhysics Letters) 30, 191 (1995). [Google Scholar]
- 80.Kremer K & Grest GS Dynamics of entangled linear polymer melts: A molecular - dynamics simulation. The Journal of Chemical Physics 92, 5057–5086, doi:doi: 10.1063/1.458541 (1990). [DOI] [Google Scholar]
- 81.Moeendarbary E, Ng T & Zangeneh M Dissipative particle dynamics in soft matter and polymeric applications—A review. International Journal of Applied Mechanics 2, 161–190 (2010). [Google Scholar]
- 82.Peter EK, Lykov K & Pivkin IV A polarizable coarse-grained protein model for dissipative particle dynamics. Physical Chemistry Chemical Physics 17, 24452–24461, doi: 10.1039/C5CP03479E (2015). [DOI] [PubMed] [Google Scholar]
- 83.Jacobsen MM et al. Effect of Terminal Modification on the Molecular Assembly and Mechanical Properties of Protein-Based Block Copolymers. Macromolecular Bioscience, 1700095-n/a, doi: 10.1002/mabi.201700095. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 84.Guigas G, Morozova D & Weiss M Exploring membrane and protein dynamics with dissipative particle dynamics. Advances in protein chemistry and structural biology 85, 143–182 (2011). [DOI] [PubMed] [Google Scholar]
- 85.Moeendarbary E, Ng T & Zangeneh M Dissipative particle dynamics: introduction, methodology and complex fluid applications—a review. International Journal of Applied Mechanics 1, 737–763 (2009). [Google Scholar]
- 86.Rühle V & Junghans C Hybrid Approaches to Coarse - Graining using the VOTCA Package: Liquid Hexane. Macromolecular Theory and Simulations 20, 472-477 (2011). [Google Scholar]
- 87.Mashayak SY et al. Relative Entropy and Optimization-Driven Coarse-Graining Methods in VOTCA. Plos One 10, e0131754, doi: 10.1371/journal.pone.0131754 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Schommers W Pair potentials in disordered many-particle systems: A study for liquid gallium. Physical Review A 28, 3599 (1983). [Google Scholar]
- 89.Lyubartsev AP & Laaksonen A Calculation of effective interaction potentials from radial distribution functions: A reverse Monte Carlo approach. Physical Review E 52, 3730 (1995). [DOI] [PubMed] [Google Scholar]
- 90.Shell MS The relative entropy is fundamental to multiscale and inverse thermodynamic problems. The Journal of chemical physics 129, 144108 (2008). [DOI] [PubMed] [Google Scholar]
- 91.Ercolessi F & Adams JB Interatomic potentials from first-principles calculations: the force-matching method. EPL (Europhysics Letters) 26, 583 (1994). [Google Scholar]
- 92.Liu G-R & Quek SS The finite element method: a practical course. (Butterworth-Heinemann, 2013). [Google Scholar]
- 93.Fish J & Belytschko T A first course in finite elements. (2007). [Google Scholar]
- 94.Zienkiewicz OC, Taylor RL & Nithiarasu P The finite element method for fluid dynamics. (Oxford : Butterworth-Heinemann, [2014] Seventh edition., 2014). [Google Scholar]
- 95.Zienkiewicz OC, Taylor RL & Fox D The finite element method for solid and structural mechanics. (Oxford : Elsevier/Butterworth-Heinemann, 2014. 7th ed., 2014). [Google Scholar]
- 96.Zienkiewicz OC, Taylor RL & Zhu JZ The finite element method : its basis and fundamentals. (Oxford : Burlington, Mass: ; Butterworth-Heinemann, 2005. 6th ed. /, 2005). [Google Scholar]
- 97.Liu G & Gu Y An introduction to meshfree methods and their programming. . Springer Berlin. MEMS, NANO and Smart Systems 10, 403–408 (2005). [Google Scholar]
- 98.Buehler MJ, Abraham FF & Gao H Hyperelasticity governs dynamic fracture at a critical length scale. Nature 426, 141–146 (2003). [DOI] [PubMed] [Google Scholar]
- 99.Liu GR (CRC Press, Boca Raton, 2010). [Google Scholar]
- 100.Hsieh Y-M & Pan M-S ESFM: An Essential Software Framework for Meshfree Methods. Advances in Engineering Software 76, 133–147, doi: 10.1016/i.advengsoft.2014.06.006 (2014). [DOI] [Google Scholar]
- 101.Daxini SD & Prajapati JM A Review on Recent Contribution of Meshfree Methods to Structure and Fracture Mechanics Applications. The Scientific World Journal 2014, 13, doi: 10.1155/2014/247172 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 102.Qin Z, Jung GS, Kang MJ & Buehler MJ The mechanics and design of a lightweight three-dimensional graphene assembly. Science advances 3, e1601536 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 103.Jung GS, Yeo J, Tian Z, Qin Z & Buehler MJ Unusually low and density-insensitive thermal conductivity of three-dimensional gyroid graphene. Nanoscale, doi: 10.1039/C7NR04455K (2017). [DOI] [PubMed] [Google Scholar]
- 104.Yeo J, Liu Z & Ng T Enhanced thermal characterization of silica aerogels through molecular dynamics simulation. Modelling and Simulation in Materials Science and Engineering 21, 075004 (2013). [Google Scholar]
- 105.Ng TY, Yeo JJ & Liu ZS A molecular dynamics study of the thermal conductivity of nanoporous silica aerogel, obtained through negative pressure rupturing. J Non-Cryst Solids 358, 1350–1355, doi: 10.1016/j.jnoncrysol.2012.03.007 (2012). [DOI] [Google Scholar]
- 106.Ng TY, Joshi SC, Yeo J & Liu Z in Advances in Nanocomposites 97–126 (Springer International Publishing, 2016). [Google Scholar]
- 107.Fratzl P Biomimetic materials research: what can we really learn from nature’s structural materials? Journal of the Royal Society Interface 4, 637–642 (2007). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Wegst U & Ashby M The mechanical efficiency of natural materials. Philosophical Magazine 84, 2167–2186 (2004). [Google Scholar]
- 109.Almeida FS & Awruch AM Design optimization of composite laminated structures using genetic algorithms and finite element analysis. Compos Struct 88, 443–454, doi:DOI 10.1016/j.compstruct.2008.05.004 (2009). [DOI] [Google Scholar]
- 110.Naik GN, Gopalakrishnan S & Ganguli R. Design optimization of composites using genetic algorithms and failure mechanism based failure criterion. Compos Struct 83, 354–367, doi:DOI 10.1016/j.compstruct.2007.05.005 (2008). [DOI] [Google Scholar]
- 111.Rozvany GI A critical review of established methods of structural topology optimization. Structural and Multidisciplinary Optimization 37, 217–237 (2009). [Google Scholar]
- 112.Hassani B & Hinton E Homogenization and structural topology optimization: theory, practice and software. (Springer Science & Business Media, 2012). [Google Scholar]
- 113.Sigmund OA 99 line topology optimization code written in Matlab. Structural and multidisciplinary optimization 21, 120–127 (2001). [Google Scholar]
- 114.Hajela P, Lee E & Lin C-Y in Topology design of structures 117–133 (Springer, 1993). [Google Scholar]
- 115.Sigmund O & Petersson J Numerical instabilities in topology optimization: a survey on procedures dealing with checkerboards, mesh-dependencies and local minima. Structural optimization 16, 68–75 (1998). [Google Scholar]
- 116.Sigmund O & Torquato S Design of materials with extreme thermal expansion using a three-phase topology optimization method. JMech Phys Solids 45, 1037–1067, doi:Doi 10.1016/S0022-5096(96)00114-7 (1997). [DOI] [Google Scholar]
- 117.Eschenauer HA & Olhoff N Topology optimization of continuum structures: a review. Applied Mechanics Reviews 54, 331–390 (2001). [Google Scholar]
- 118.Bendsoe MP & Sigmund O Topology optimization: theory, methods, and applications. (Springer Science & Business Media, 2013). [Google Scholar]
- 119.Gaynor AT, Meisel NA, Williams CB & Guest JK Multiple-material topology optimization of compliant mechanisms created via PolyJet three-dimensional printing. Journal of Manufacturing Science and Engineering 136, 061015 (2014). [Google Scholar]
- 120.Rosen DW Computer-aided design for additive manufacturing of cellular structures. Computer-Aided Design and Applications 4, 585–594 (2007). [Google Scholar]
- 121.Rezaie R, Badrossamay M, Ghaie A & Moosavi H Topology optimization for fused deposition modeling process. Procedia CIRP 6, 521–526 (2013). [Google Scholar]
- 122.Horn TJ & Harrysson OL Overview of current additive manufacturing technologies and selected applications. Science progress 95, 255–282 (2012). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Chu J, Engelbrecht S, Graf G & Rosen DW A comparison of synthesis methods for cellular structures with application to additive manufacturing. Rapid Prototyping Journal 16, 275–283 (2010). [Google Scholar]
- 124.Gao W et al. The status, challenges, and future of additive manufacturing in engineering. Computer-Aided Design (2015). [Google Scholar]
- 125.Gu GX et al. Three-Dimensional-Printing of Bio-Inspired Composites. Journal of Biomechanical Engineering 138, 021006–021006, doi: 10.1115/1.4032423 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Jared BH et al. Additive manufacturing: Toward holistic design. ScriptaMaterialia 135, 141–147 (2017). [Google Scholar]
- 127.Gu GX, Wettermark S & Buehler MJ Algorithm-driven design of fracture resistant composite materials realized through additive manufacturing. Additive Manufacturing (2017). [Google Scholar]
- 128.Gu GX, Dimas L, Qin Z & Buehler MJ Optimization of composite fracture properties: Method, validation, and applications. Journal of Applied Mechanics 83, 071006 (2016). [Google Scholar]
- 129.Fauziyah S, Alam C, Soesilohadi R, Retnoaji B & Alam P Morphological and mechanical characterisation of the hindwing nodus from the Libellulidae family of dragonfly (Indonesia). Arthropod structure & development 43, 415–422 (2014). [DOI] [PubMed] [Google Scholar]
- 130.Gu GX, Chen C-T & Buehler MJ De novo composite design based on machine learning algorithm. Extreme Mechanics Letters (2017). [Google Scholar]
- 131.Brodtkorb AR, Dyken C, Hagen TR, Hjelmervik JM & Storaasli OO State-of-the-art in heterogeneous computing. Scientific Programming 18, 1–33 (2010). [Google Scholar]
- 132.Dresselhaus MS & Dresselhaus G Intercalation compounds of graphite. Adv. Phys. 51, 1–186, doi: 10.1080/00018730110113644 (2002). [DOI] [Google Scholar]
- 133.Shenderova OA, Zhirnov VV & Brenner DW Carbon Nanostructures. Crit. Rev. Solid State Mater. Sci. 27, 227–356, doi: 10.1080/10408430208500497 (2002). [DOI] [Google Scholar]
- 134.Novoselov KS et al. Electric Field Effect in Atomically Thin Carbon Films. Science 306, 666 (2004). [DOI] [PubMed] [Google Scholar]
- 135.Nobelprize.org. The Nobel Prize in Physics 2010, <http://www.nobelprize.org/nobel_prizes/physics/laureates/2010/> (
- 136.Balandin AA et al. Superior Thermal Conductivity of Single-Layer Graphene. Nano Lett. 8, 902–907, doi: 10.1021/nl0731872 (2008). [DOI] [PubMed] [Google Scholar]
- 137.Lee C, Wei X, Kysar JW & Hone J in Science Vol. 321 385–388 (American Association for the Advancement of Science, 2008). [DOI] [PubMed] [Google Scholar]
- 138.Giannozzi P et al. QUANTUM ESPRESSO: a modular and open-source software project for quantum simulations of materials. J. Phys.: Condens. Matter 21, 395502, doi:papers3://publication/doi/10.1088/0953-8984/21/39/395502 (2009). [DOI] [PubMed] [Google Scholar]
- 139.Perdew JP, Burke K & Ernzerhof M Generalized Gradient Approximation Made Simple. Phys. Rev. Lett. 77, 3865–3868, doi:papers3://publication/doi/10.1103/PhysRevLett.77.3865 (1996). [DOI] [PubMed] [Google Scholar]
- 140.Troullier N & Martins JL Efficient pseudopotentials for plane-wave calculations. Phys. Rev. B 43, 1993–2006, doi:papers3://publication/doi/10.1103/PhysRevB.43.1993 (1991). [DOI] [PubMed] [Google Scholar]
- 141.Luan VH et al. Synthesis of a highly conductive and large surface area graphene oxide hydrogel and its use in a supercapacitor. J. Mater. Chem. A 1, 208–211, doi: 10.1039/C2TA00444E (2013). [DOI] [Google Scholar]
- 142.Kim P Graphene: Across the border. Nat. Mater. 9, 792–793, doi:papers3://publication/doi/10.1038/nmat2862 (2010). [DOI] [PubMed] [Google Scholar]
- 143.Chatrchyan S et al. Precise mapping of the magnetic field in the CMS barrel yoke using cosmic rays. J Instrum 5, doi:Artn T03021 Doi 10.1088/1748-0221/5/03/T03021 (2010). [Google Scholar]
- 144.Wu T et al. Fast growth of inch-sized single-crystalline graphene from a controlled single nucleus on Cu-Ni alloys. Nat. Mater. 15, 43–47, doi:10.1038/nmat447710.1038/nmat4477http://www.nature.com/nmat/journal/v15/n1/abs/nmat4477.html-supplementary-informationhttp://www.nature.com/nmat/journal/v15/n1/abs/nmat4477.html-supplementary-information (2016). [DOI] [PubMed] [Google Scholar]
- 145.Geim AK & Novoselov KS The rise of graphene. Nat Mater 6, 183–191 (2007). [DOI] [PubMed] [Google Scholar]
- 146.Li X, Wang X, Zhang L, Lee S & Dai H Chemically Derived, Ultrasmooth Graphene Nanoribbon Semiconductors. Science 319, 1229 (2008). [DOI] [PubMed] [Google Scholar]
- 147.Ni ZH et al. Uniaxial Strain on Graphene: Raman Spectroscopy Study and Band-Gap Opening. ACS Nano 2, 2301–2305, doi: 10.1021/nn800459e (2008). [DOI] [PubMed] [Google Scholar]
- 148.Radisavljevic B, Radenovic A, Brivio J, Giacometti V & Kis A. Single-layer MoS2 transistors. Nat Nano 6, 147–150, doi:http://www.nature.com/nnano/journal/v6/n3/abs/nnano.2010.279.html-supplementary-information (2011). [DOI] [PubMed] [Google Scholar]
- 149.Dean CR et al. Boron nitride substrates for high-quality graphene electronics. Nat Nano 5, 722–726, doi:http://www.nature.com/nnano/journal/v5/n10/abs/nnano.2010.172.html-supplementary-information (2010). [DOI] [PubMed] [Google Scholar]
- 150.Geim AK & Grigorieva IV Van der Waals heterostructures. Nature 499, 419–425, doi: 10.1038/nature12385 (2013). [DOI] [PubMed] [Google Scholar]
- 151.Lee C-H et al. Atomically thin p-n junctions with van der Waals heterointerfaces. Nat Nano 9, 676–681, doi:10.1038/nnano.2014.15010.1038/nnano.2014.150http://www.nature.com/nnano/journal/v9/n9/abs/nnano.2014.150.html-supplementary-informationhttp://www.nature.com/nnano/journal/v9/n9/abs/nnano.2014.150.html-supplementary-information (2014). [DOI] [PubMed] [Google Scholar]
- 152.Novoselov KS, Mishchenko A, Carvalho A & Castro Neto AH 2D materials and van der Waals heterostructures. Science 353 (2016). [DOI] [PubMed] [Google Scholar]
- 153.Kang K et al. Layer-by-layer assembly of two-dimensional materials into wafer-scale heterostructures. Nature advance online publication, doi:10.1038/nature2390510.1038/nature23905http://www.nature.com/nature/journal/vaop/ncurrent/abs/nature23905.html-supplementary-informationhttp://www.nature.com/nature/journal/vaop/ncurrent/abs/nature23905.html-supplementary-information (2017). [DOI] [PubMed] [Google Scholar]
- 154.Levendorf MP et al. Graphene and boron nitride lateral heterostructures for atomically thin circuitry. Nature 488, 627–632, doi:http://www.nature.com/nature/journal/v488/n7413/abs/nature11408.html-supplementary-information (2012). [DOI] [PubMed] [Google Scholar]
- 155.Gong Y et al. Vertical and in-plane heterostructures from WS2/MoS2 monolayers. Nat. Mater. 13, 1135–1142, doi:10.1038/nmat409110.1038/nmat4091http://www.nature.com/nmat/journal/v13/n12/abs/nmat4091.html-supplementary-informationhttp://www.nature.com/nmat/journal/v13/n12/abs/nmat4091.html-supplementary-information (2014). [DOI] [PubMed] [Google Scholar]
- 156.Duan X et al. Lateral epitaxial growth of two-dimensional layered semiconductor heterojunctions. Nat Nano 9, 1024–1030, doi:10.1038/nnano.2014.22210.1038/nnano.2014.222http://www.nature.com/nnano/journal/v9/n12/abs/nnano.2014.222.html-supplementary-informationhttp://www.nature.com/nnano/journal/v9/n12/abs/nnano.2014.222.html-supplementary-information (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Huang C et al. Lateral heterojunctions within monolayer MoSe2–WSe2 semiconductors. Nat. Mater. 13, 1096–1101, doi: 10.1038/nmat406410.1038/nmat4064http://www.nature.com/nmat/journal/v13/n12/abs/nmat4064.html-supplementary-informationhttp://www.nature.com/nmat/journal/v13/n12/abs/nmat4064.html-supplementary-information (2014). [DOI] [PubMed] [Google Scholar]
- 158.Li M-Y et al. Epitaxial growth of a monolayer WSe2-MoS2 lateral p-n junction with an atomically sharp interface. Science 349, 524 (2015). [DOI] [PubMed] [Google Scholar]
- 159.Blees MK et al. Graphene kirigami. Nature 524, 204–207, doi:10.1038/nature1458810.1038/nature14588http://www.nature.com/nature/journal/v524/n7564/abs/nature14588.html-supplementary-informationhttp://www.nature.com/nature/journal/v524/n7564/abs/nature14588.html-supplementary-information (2015). [DOI] [PubMed] [Google Scholar]
- 160.Annett J & Cross GLW Self-assembly of graphene ribbons by spontaneous self-tearing and peeling from a substrate. Nature 535, 271–275, doi: 10.1038/naturel8304 http://www.nature.com/nature/journal/v535/n7611/abs/nature18304.html-supplementary-information (2016). [DOI] [PubMed] [Google Scholar]
- 161.Zang J et al. Multifunctionality and control of the crumpling and unfolding of large-area graphene. Nat Mater 12, 321–325, doi:http://www.nature.com/nmat/journal/v12/n4/abs/nmat3542.html-supplementary-information (2013). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 162.Wallace PR The Band Theory of Graphite. Phys. Rev. 71, 622–634 (1947). [Google Scholar]
- 163.Castro Neto AH, Guinea F, Peres NMR, Novoselov KS & Geim AK The electronic properties of graphene. Rev. Mod. Phys. 81, 109–162 (2009). [Google Scholar]
- 164.Yazyev OV & Louie SG Electronic transport in polycrystalline graphene. Nat. Mater. 9, 806–809, doi:http://www.nature.com/nmat/journal/v9/n10/abs/nmat2830.html-supplementary-information (2010). [DOI] [PubMed] [Google Scholar]
- 165.Lee GH et al. in Science Vol. 340 1073–1076 (American Association for the Advancement of Science, 2013).23723231 [Google Scholar]
- 166.Yazyev OV & Chen YP Polycrystalline graphene and other two-dimensional materials. Nat Nano 9, 755–767, doi: 10.1038/nnano.2014.166 (2014). [DOI] [PubMed] [Google Scholar]
- 167.Andreas I et al. Scaling properties of polycrystalline graphene: a review. 2D Materials 4, 012002 (2017). [Google Scholar]
- 168.Kotakoski J & Meyer JC Mechanical properties of polycrystalline graphene based on a realistic atomistic model. Phys. Rev. B 85, 195447 (2012). [Google Scholar]
- 169.Yang Z et al. Size-dependent deformation behavior of nanocrystalline graphene sheets. Materials Science and Engineering: B 198, 95–101, doi: 10.1016/i.mseb.2015.03.019 (2015). [DOI] [Google Scholar]
- 170.Jung GS, Qin Z & Buehler MJ Molecular mechanics of polycrystalline grapheme with enhanced fracture toughness. Extreme Mechanics Letters 2, 52–59, doi:papers3://publication/doi/10.1016/j.eml.2015.01.007 (2015). [Google Scholar]
- 171.George G & Costas G Graphene aerogels: a review. 2D Materials 4, 032001 (2017). [Google Scholar]
- 172.Zhang Q et al. 3D Printing of Graphene Aerogels. Small 12, 1702–1708, doi: 10.1002/smll.201503524 (2016). [DOI] [PubMed] [Google Scholar]
- 173.Sha J et al. Three-Dimensional Printed Graphene Foams. ACS Nano, doi: 10.1021/acsnano.7b01987 (2017). [DOI] [PubMed] [Google Scholar]
- 174.Johari P & Shenoy VB Tuning the Electronic Properties of Semiconducting Transition Metal Dichalcogenides by Applying Mechanical Strains. ACS Nano 6, 5449–5456, doi: 10.1021/nn301320r (2012). [DOI] [PubMed] [Google Scholar]
- 175.Peelaers H & Van de Walle CG Effects of strain on band structure and effective masses in MoS${}_{2}$. Phys. Rev. B 86, 241401 (2012). [Google Scholar]
- 176.He K, Poole C, Mak KF & Shan J Experimental Demonstration of Continuous Electronic Structure Tuning via Strain in Atomically Thin MoS2. Nano Lett. 13, 2931–2936, doi: 10.1021/nl4013166 (2013). [DOI] [PubMed] [Google Scholar]
- 177.Shen T, Penumatcha AV & Appenzeller J Strain Engineering for Transition Metal Dichalcogenides Based Field Effect Transistors. ACS Nano 10, 4712–4718, doi: 10.1021/acsnano.6b01149 (2016). [DOI] [PubMed] [Google Scholar]
- 178.van der Zande AM et al. Grains and grain boundaries in highly crystalline monolayer molybdenum disulphide. Nat. Mater. 12, 554–561, doi:10.1038/nmat363310.1038/nmat3633http://www.nature.com/nmat/journal/v12/n6/abs/nmat3633.html-supplementary-informationhttp://www.nature.com/nmat/journal/v12/n6/abs/nmat3633.html-supplementary-information (2013). [DOI] [PubMed] [Google Scholar]
- 179.Zhou W et al. Intrinsic Structural Defects in Monolayer Molybdenum Disulfide. Nano Lett. 13, 2615–2622, doi: 10.1021/nl4007479 (2013). [DOI] [PubMed] [Google Scholar]
- 180.Najmaei S et al. Vapour phase growth and grain boundary structure of molybdenum disulphide atomic layers. Nat. Mater. 12, 754–759, doi:10.1038/nmat367310.1038/nmat3673http://www.nature.com/nmat/journal/v12/n8/abs/nmat3673.html-supplementary-informationhttp://www.nature.com/nmat/journal/v12/n8/abs/nmat3673.html-supplementary-information (2013). [DOI] [PubMed] [Google Scholar]
- 181.Meyer JC et al. The structure of suspended graphene sheets. Nature 446, 60–63, doi:http://www.nature.com/nature/journal/v446/n7131/suppinfo/nature05545_S1.html (2007). [DOI] [PubMed] [Google Scholar]
- 182.Meyer JC et al. Experimental analysis of charge redistribution due to chemical bonding by high-resolution transmission electron microscopy. Nat. Mater. 10, 209–215, doi:http://www.nature.com/nmat/journal/v10/n3/abs/nmat2941.html-supplementary-information (2011). [DOI] [PubMed] [Google Scholar]
- 183.Robertson AW et al. Spatial control of defect creation in graphene at the nanoscale. 3, 1144, doi:10.1038/ncomms214110.1038/ncomms2141https://www.nature.com/articles/ncomms2141-supplementary-informationhttps://www.nature.com/articles/ncomms2141-supplementary-information (2012). [DOI] [PubMed] [Google Scholar]
- 184.Warner JH et al. Dislocation-Driven Deformations in Graphene. Science 337, 209 (2012). [DOI] [PubMed] [Google Scholar]
- 185.Wang S et al. Atomically Sharp Crack Tips in Monolayer MoS2 and Their Enhanced Toughness by Vacancy Defects. ACS Nano 10, 9831–9839, doi: 10.1021/acsnano.6b05435 (2016). [DOI] [PubMed] [Google Scholar]
- 186.Han Y et al. Sub-nanometre channels embedded in two-dimensional materials. Nat Mater (2017). [DOI] [PubMed] [Google Scholar]
- 187.Mohiuddin TMG et al. Uniaxial strain in graphene by Raman spectroscopy: $G$ peak splitting, Gr\“uneisen parameters, and sample orientation. Phys. Rev. B 79, 205433 (2009). [Google Scholar]
- 188.Ferrari AC et al. Raman Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett. 97, 187401 (2006). [DOI] [PubMed] [Google Scholar]
- 189.Ang EYM, Ng TY, Yeo J, Liu Z & Geethalakshmi KR Free-standing graphene slit membrane for enhanced desalination. Carbon 110, 350–355, doi: 10.1016/i.carbon.2016,09,043 (2016). [DOI] [Google Scholar]
- 190.Ang EY, Ng TY, Yeo J, Lin R & Geethalakshmi K Nanoscale Fluid Mechanics Working Principles of Transverse Flow Carbon Nanotube Membrane for Enhanced Desalination. International Journal of Applied Mechanics, 1750034 (2017). [Google Scholar]
- 191.LeDuc PR & Robinson DN Using Lessons from Cellular and Molecular Structures for Future Materials. Adv Mater 19, 3761–3770, doi: 10.1002/adma.200701286 (2007). [DOI] [Google Scholar]
- 192.Voet D, Voet JG & Pratt CW Fundamentals of biochemistry: life at the molecular level. (John Wiley & Sons , Inc., 2013). [Google Scholar]
- 193.Lin Y et al. Genetically Programmable Thermoresponsive Plasmonic Gold/Silk-Elastin Protein Core/Shell Nanoparticles. Langmuir 30, 4406–4414, doi: 10.1021/la403559t (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 194.Qiu W, Teng W, Cappello J & Wu X Wet-Spinning of Recombinant Silk-Elastin-Like Protein Polymer Fibers with High Tensile Strength and High Deformability. Biomacromolecules 10, 602–608, doi: 10.1021/bm801296r (2009). [DOI] [PubMed] [Google Scholar]
- 195.Xia X-X, Xu Q, Hu X, Qin G & Kaplan DL Tunable Self-Assembly of Genetically Engineered Silk–Elastin-like Protein Polymers. Biomacromolecules 12, 3844–3850, doi: 10.1021/bm201165h (2011). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Xia X-X, Wang M, Lin Y, Xu Q & Kaplan DL Hydrophobic Drug-Triggered Self-Assembly of Nanoparticles from Silk-Elastin-Like Protein Polymers for Drug Delivery. Biomacromolecules 15, 908–914, doi: 10.1021/bm4017594 (2014). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Megeed Z, Cappello J & Ghandehari H Genetically engineered silk-elastinlike protein polymers for controlled drug delivery. Advanced Drug Delivery Reviews 54, 1075–1091, doi: 10.1016/S0169-409X(02100063-7 (2002). [DOI] [PubMed] [Google Scholar]
- 198.Ner Y, Stuart JA, Whited G & Sotzing GA Electrospinning nanoribbons of a bioengineered silk-elastin-like protein (SELP) from water. Polymer 50, 5828–5836, doi: 10.1016/i.polymer.2009.09.017 (2009). [DOI] [Google Scholar]
- 199.Hwang W et al. Surface Induced Nanofiber Growth by Self-Assembly of a Silk-Elastin-like Protein Polymer. Langmuir 25, 12682–12686, doi: 10.102l/la9015993 (2009). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 200.Huang W, Rollett A & Kaplan DL Silk-elastin-like protein biomaterials for the controlled delivery of therapeutics. Expert Opinion on Drug Delivery 12, 779–791, doi: 10.1517/17425247.2015.989830 (2015). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Huang W et al. Design of Multistimuli Responsive Hydrogels Using Integrated Modeling and Genetically Engineered Silk–Elastin-Like Proteins. Advanced Functional Materials 26, 4113–4123, doi: 10.1002/adfm.201600236 (2016). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Tarakanova A, Huang W, Weiss AS, Kaplan DL & Buehler MJ Computational smart polymer design based on elastin protein mutability. Biomaterials 127, 49–60, doi: 10.1016/i.biomaterials.2017.01.041 (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
- 203.Tarakanova A, Huang W, Qin Z, Kaplan DL & Buehler MJ Modeling and Experiment Reveal Structure and Nanomechanics across the Inverse Temperature Transition in B. mori Silk-Elastin-like Protein Polymers. ACS Biomaterials Science & Engineering (2017). [DOI] [PMC free article] [PubMed] [Google Scholar]
