Fact-checked by Grok 2 weeks ago

Elastic modulus

The elastic modulus is a fundamental mechanical property that quantifies a material's , defined as the ratio of applied to the resulting within the deformation range, where the material returns to its original shape upon removal of the load. This property arises from the and molecular of materials and is crucial for understanding how substances respond to forces without undergoing permanent damage. There are three primary types of elastic moduli, each corresponding to different deformation modes: (E), which measures resistance to linear extension or under uniaxial and is calculated as E = / axial ; (G), which quantifies resistance to angular distortion under and is given by G = / ; and (K), which describes resistance to uniform volumetric and is defined as K = - (pressure change) / (relative volume change). is the most commonly referenced elastic modulus in contexts, often simply called the modulus of elasticity, and its value indicates how much a material will elongate or shorten per unit . For example, has a Young's modulus around 200 GPa, making it significantly stiffer than rubber, which is about 0.01–0.1 GPa. These moduli are essential in fields like , , and for predicting structural behavior, designing load-bearing components, and selecting materials that balance strength and flexibility. Values are typically expressed in pascals (Pa) or gigapascals (GPa) in the , reflecting the material's inherent resistance to deformation rather than its size or shape. Experimental determination involves applying controlled loads and measuring deformations, often using machines to generate stress-strain curves from which the moduli are derived.

Basic Concepts

Stress and Strain

In solid mechanics, is defined as the internal force per unit area acting within a , quantifying the intensity of forces distributed over a surface. The , introduced by in the 1820s, provides a complete mathematical description of this force distribution in the deformed configuration, relating the traction on any surface to its normal through the second-order tensor σ, where the normal component σ_n = t · n and arises from the tangential component. For a simple case, normal σ is given by σ = F / A, where F is the applied force perpendicular to the cross-sectional area A. Stress manifests in different types depending on the direction and nature of the force: tensile stress occurs when forces pull the material apart, elongating it along the force direction; compressive stress arises from forces pushing the material together, shortening it; and shear stress results from forces acting parallel to the surface, causing layers to slide relative to one another. These types are fundamental to analyzing material behavior under load, with tensile and compressive stresses being normal to the surface and shear stresses tangential. Strain, conversely, measures the relative deformation or distortion of a from its original , serving as the kinematic counterpart to . Normal ε quantifies linear deformation as ε = ΔL / L_0, where ΔL is the change in length and L_0 is the original length, positive for extension (tensile) and negative for (compressive). γ describes angular distortion as γ = Δx / L, where Δx is the transverse and L is the over which the acts, often approximated as the of the shear angle for small deformations. A key distinction exists between engineering strain and true strain: engineering strain uses the original undeformed length L_0 in the denominator, suitable for small deformations, while true (or logarithmic) strain employs the instantaneous length, defined as ε_true = ∫ (dL / L) = ln(L / L_0), providing a more accurate measure for large deformations where cross-sectional area changes significantly. For small deformations in , the infinitesimal strain tensor ε_ij = (1/2)(∂u_i/∂x_j + ∂u_j/∂x_i) captures the symmetric part of the displacement gradient, neglecting higher-order terms and assuming rotations do not contribute to strain. The foundational concepts of stress and strain trace back to early continuum mechanics, with Leonhard Euler and developing initial ideas on deformation and internal forces in the , while Cauchy's work in the 1820s formalized the stress tensor and clarified strain as a deformation measure independent of rigid-body motions.

Hooke's Law

Hooke's law describes the fundamental linear relationship between stress and strain in elastic materials. In the simplest uniaxial case, it states that the normal stress \sigma is directly proportional to the corresponding normal strain \epsilon, expressed as \sigma = E \epsilon, where E is a material constant known as the elastic modulus. This relation implies that the deformation produced by a force is recoverable upon removal of the load, provided the material remains within its elastic limit. For more complex loading in three dimensions, Hooke's law generalizes to a tensor form, \sigma_{ij} = C_{ijkl} \epsilon_{kl}, where \sigma_{ij} is the stress tensor, \epsilon_{kl} is the strain tensor, and C_{ijkl} is the fourth-rank stiffness tensor that encapsulates the material's elastic properties; summation over repeated indices k and l is implied. The law originated with , a 17th-century English , who first hinted at the proportionality in 1676 through a Latin ("ceiiinosssttuv") in a letter to the Royal Society, and fully published it in 1678 in his work De potentia restitutiva, or of Spring. Hooke's empirical observation stemmed from experiments with springs and wires, capturing the restorative force in elastic bodies. In the early 1820s, French mathematicians and advanced this into a rigorous framework, deriving the general equations of by incorporating stress-strain relations into the balance of linear momentum, thus enabling analysis of deformable solids beyond simple one-dimensional cases. Hooke's law relies on key assumptions to ensure its linearity: deformations must be small (typically strains less than 0.1–1%, depending on the material) to maintain geometric linearity and avoid nonlinear effects; the stress-strain response must be linear, meaning no higher-order terms in the constitutive relation; and the behavior is time-independent, neglecting rate-dependent phenomena like or . These conditions align the law with reversible, elastic deformations in homogeneous materials. However, the law has limitations in its applicability. It holds only up to the proportional limit or yield point, beyond which the stress-strain curve deviates from , leading to permanent deformation rather than full recovery. Additionally, the simple forms often assume material , where properties are direction-independent; anisotropic materials require the full tensorial description without such simplification. A practical example illustrates the uniaxial application: consider a cylindrical wire of original length L, cross-sectional area A, subjected to a tensile F. The axial is \sigma = F / A, and by , the axial strain is \epsilon = \sigma / E = F / (A E). The resulting elongation (displacement) is then \Delta L = \epsilon L = F L / (A E), showing how the load induces a proportional extension that vanishes when the is removed, assuming operation within the elastic regime. This derivation directly ties the applied load to measurable displacement, forming the basis for calculations in members.

Definition and Properties

Definition of Elastic Modulus

The elastic modulus is a fundamental material property that quantifies the relationship between and within the linear elastic regime, serving as the constant of proportionality in . This law posits that, for small deformations, the applied \sigma is directly proportional to the resulting \epsilon, expressed mathematically as: E = \frac{\sigma}{\epsilon} where E denotes the elastic modulus. The SI unit of elastic modulus is the pascal (Pa), equivalent to newtons per square meter (N/m²), though values are commonly reported in gigapascals (GPa) for practical engineering contexts due to their typical magnitude. Physically, the elastic modulus represents a material's stiffness, or its resistance to deformation under an applied load, indicating how much stress is required to produce a unit strain. This measure applies specifically to elastic deformation, which is reversible and occurs when atomic bonds stretch temporarily without permanent rearrangement; in contrast, exceeding the elastic limit leads to plastic deformation, where the material undergoes irreversible changes in shape. For more complex loading scenarios beyond simple uniaxial , the elastic modulus concept generalizes to a collection of elastic constants that fully describe the material's response, relating the full stress tensor to the strain tensor in multidimensional space. To illustrate the range of across materials, exhibits an elastic modulus of approximately 200 GPa, enabling it to withstand significant loads with minimal deformation, whereas rubber has a much lower value of about 0.01 GPa, allowing substantial elastic stretching.

Distinction Between Elastic Constants and Moduli

In materials science, elastic constants denote the components of the fourth-rank stiffness tensor C_{ijkl}, which linearly relates the stress tensor \sigma_{ij} to the strain tensor \epsilon_{kl} via the generalized form of Hooke's law: \sigma_{ij} = C_{ijkl} \epsilon_{kl}. For a fully anisotropic material with triclinic symmetry, symmetries inherent to the stress and strain tensors (such as C_{ijkl} = C_{jikl} = C_{ijlk} = C_{klij}) reduce the potential 81 components to 21 independent elastic constants, fully characterizing the material's stiffness in all directions. Elastic moduli, on the other hand, are specific scalar measures derived from these constants that quantify a material's resistance to particular deformation modes under controlled stress conditions. Examples include E, which describes uniaxial extension or compression; the G, which governs resistance to shear deformation; and the K, which measures volumetric response to hydrostatic pressure. These moduli simplify analysis for engineering applications but apply directly only to scenarios matching their defined deformation type. In isotropic materials, where properties are uniform in all directions, the distinction blurs into equivalence, as the tensor reduces to just two independent constants—such as the \lambda (relating to volumetric changes) and \mu (the )—from which all scalar moduli can be expressed through algebraic relations. This reduction stems from the material's , eliminating directional dependence and allowing a complete description with minimal parameters. Historically, the concept originated with Thomas Young's 1807 introduction of the "modulus of elasticity" as a scalar quantity for longitudinal stiffness in beams and wires, reflecting early focus on simple deformations in isotropic contexts. Over time, as understanding of crystal anisotropy advanced in the 19th and 20th centuries, the terminology expanded to encompass the full tensorial framework of elastic constants, accommodating complex directional behaviors in crystalline solids. For example, isotropic materials like polycrystalline metals require only two constants to predict all elastic responses, whereas triclinic crystals, lacking higher symmetry, demand the full 21 independent constants to accurately model deformations in arbitrary directions.

Moduli for Isotropic Materials

Young's Modulus

Young's modulus, often denoted as E, quantifies a material's stiffness under uniaxial loading and is defined as the ratio of applied stress \sigma to the resulting axial strain \epsilon within the linear elastic regime for both tensile and compressive conditions. This measure arises from Hooke's law applied to one-dimensional deformation, where E = \frac{\sigma}{\epsilon}, and it remains constant up to the material's proportional limit. The units of Young's modulus are typically gigapascals (GPa) in engineering contexts, reflecting its role in predicting how materials resist elongation or shortening without permanent deformation. Young's modulus is commonly determined through , following standards such as ASTM E111, which outlines procedures for measuring the in structural materials by applying controlled loads and recording load-displacement to compute the of the stress-strain in the region. For a simple cylindrical specimen or rod of L and cross-sectional area A subjected to an axial F, the resulting change in \Delta L is given by \Delta L = \frac{F L}{A E}. This equation enables direct calculation of E from experimental measurements of \Delta L. In practice, extensometers or strain gauges ensure precise strain assessment during testing to avoid errors from machine compliance. The value of Young's modulus exhibits temperature dependence, with examples in metallic alloys showing variations that influence material performance in thermal environments. Such behavior underscores the need to consider operating conditions in design, particularly for components exposed to temperature fluctuations. In engineering applications, Young's modulus is essential for analyzing beam deflection and structural integrity, where higher values indicate greater resistance to bending under load. For example, the used in the has a Young's modulus of approximately 190 GPa, enabling the structure to withstand wind and gravitational forces while minimizing deflection and ensuring stability. Young's modulus also connects to \nu, defined as the negative ratio of lateral to axial under uniaxial , which describes the material's transverse contraction during axial extension. This relation aids in predicting volumetric changes but is considered alongside E for complete elastic characterization.

Shear Modulus and Bulk Modulus

The , also known as the modulus of rigidity and denoted by G, quantifies a material's resistance to deformation, which involves angular distortion without significant volume change. It is defined as the ratio of \tau (the force per unit area acting parallel to a surface) to \gamma (the angular displacement or tangent of the angle), expressed as G = \frac{\tau}{\gamma}. This modulus is particularly relevant for analyzing torsional loading in structural components like shafts and beams, where materials undergo ./20%3A_Miscellaneous/20.03%3A_Shear_Modulus_and_Torsion_Constant) In torsion tests, the shear modulus governs the relationship between applied torque and resulting twist. For a circular shaft of length L subjected to torque T, the angle of twist \theta (in radians) is given by \theta = \frac{T L}{G J}, where J is the polar moment of inertia of the cross-section. This equation allows determination of G by measuring \theta, T, L, and calculating J. Torsion tests, often conducted using specialized machines that apply controlled torque while recording angular deflection, provide direct measurement of G for metals and polymers, ensuring the material remains in the elastic regime. The bulk modulus, denoted K, measures a material's resistance to uniform volumetric compression or expansion under hydrostatic pressure, focusing on changes in volume rather than shape. It is defined as K = -V \frac{\Delta P}{\Delta V}, where V is the initial volume, \Delta P is the infinitesimal change in hydrostatic pressure applied equally from all directions, and \Delta V is the resulting volume change; the negative sign accounts for volume decrease under increased pressure. The reciprocal of the bulk modulus, \frac{1}{K}, is the compressibility, indicating how readily a material's volume alters under pressure. Hydrostatic conditions ensure no shear components, isolating pure volumetric response, as in pressurized fluids or confined solids. The concept of bulk modulus emerged in the mid-19th century amid advancements in continuum mechanics and thermodynamics. Bulk modulus is commonly measured using ultrasonic techniques, which exploit the relationship between wave speed and elastic properties. The speed of longitudinal waves c_L in a material relates to K via c_L = \sqrt{\frac{K + \frac{4}{3}G}{\rho}}, where \rho is ; by measuring c_L with transducers and knowing G and \rho, K can be derived non-destructively. This method is ideal for fluids and solids under controlled , providing high-precision values at ambient or elevated conditions. For example, exhibits a bulk modulus of approximately 2.2 GPa, rendering it nearly incompressible and crucial for hydraulic applications. In contrast, rubber demonstrates a low modulus of about 0.0006 GPa (facilitating easy shape distortion under ) paired with a high bulk modulus around 2 GPa (resisting volume change), highlighting how materials can prioritize flexibility in while maintaining incompressibility volumetrically.

Interrelationships Among Isotropic Moduli

For isotropic materials, the elastic response is fully characterized by two independent constants, such as E and \nu, from which the G, K, and Lamé constants \lambda and \mu (with \mu = G) can be derived. These interrelationships arise because the isotropic reduces the general to just two parameters, linking uniaxial, shear, and volumetric behaviors mathematically. The key relations include: E = 2G(1 + \nu), \quad G = \frac{E}{2(1 + \nu)}, \quad K = \frac{E}{3(1 - 2\nu)} and for the Lamé constants: \lambda = K - \frac{2}{3}G. These can be derived from the strain energy density function for isotropic materials, which takes the quadratic form U = \frac{1}{2} \lambda (\mathrm{tr} \epsilon)^2 + \mu \epsilon : \epsilon, ensuring positive definiteness for stability, or via coordinate transformations: for instance, rotating a uniaxial stress state to pure shear yields the E-G relation, while hydrostatic loading connects to K. Poisson's ratio \nu governs the coupling between longitudinal and transverse strains and must satisfy $0 < \nu < 0.5 for most stable solids to maintain positive moduli and thermodynamic consistency, as values outside this range could imply negative stiffness or instability under loading. Physical bounds, such as G > 0 and K > 0, further enforce inequalities like K > \frac{2}{3}G (equivalent to \nu > -\frac{1}{3}) for consistency in isotropic elasticity. As an illustrative example, aluminum has a G = 26 GPa and \nu = 0.35; substituting into the relation gives E = 2 \times 26 \times (1 + 0.35) = 70 GPa, demonstrating how measured shear properties predict tensile .

Moduli for Anisotropic Materials

Elastic Stiffness and Compliance Tensors

In anisotropic materials, the relationship between and is described by fourth-rank tensors known as the elastic tensor C_{ijkl} and the elastic compliance tensor S_{ijkl}. The tensor relates the components of the tensor \sigma_{ij} to the strain tensor \epsilon_{kl} through the generalized : \sigma_{ij} = C_{ijkl} \epsilon_{kl}, where summation over repeated indices is implied. This tensor encapsulates the directional dependence of response, arising from the material's internal structure, such as crystal lattice arrangements. The stiffness tensor possesses inherent symmetries due to the symmetry of the stress and strain tensors themselves, as well as the existence of a strain energy potential. Specifically, it exhibits minor symmetries C_{ijkl} = C_{jikl} = C_{ijlk} from \sigma_{ij} = \sigma_{ji} and \epsilon_{kl} = \epsilon_{lk}, and a major symmetry C_{ijkl} = C_{klij} derived from the requirement that the elastic response is conservative. These symmetries reduce the number of potentially independent components from 81 (for a general fourth-rank tensor) to a maximum of 21 in the most general anisotropic case, such as triclinic crystals. Crystal symmetries further constrain this number; for example, orthorhombic materials have 9 independent components. The compliance tensor S_{ijkl} is the inverse of the stiffness tensor, satisfying S_{ijkl} C_{klmn} = \delta_{im} \delta_{jn}, and relates to via \epsilon_{ij} = S_{ijkl} \sigma_{kl}. It shares the same symmetries as C_{ijkl} and is particularly useful for analyzing deformations under applied loads. In the context of hyperelastic materials, the U is expressed as U = \frac{1}{2} C_{ijkl} \epsilon_{ij} \epsilon_{kl}, which ensures the major symmetry and provides a thermodynamic foundation for the tensor's properties. For cubic crystals, which possess high symmetry, the stiffness tensor reduces to only 3 independent components: C_{11}, C_{12}, and C_{44}. This case represents a bridge to isotropic materials, where the tensor simplifies further to just 2 independent constants, such as the Lamé parameters.

Voigt Notation and Elastic Matrices

To represent the fourth-rank elastic stiffness tensor C_{ijkl} and compliance tensor S_{ijkl} in a more compact form for anisotropic materials, the Voigt notation employs a contracted index system that reduces the tensor components to elements of 6×6 matrices. This mapping assigns single indices to pairs of tensor indices as follows: 11 → 1, 22 → 2, 33 → 3, 23 (and 32) → 4, 13 (and 31) → 5, 12 (and 21) → 6. In this scheme, the stress components \sigma_i (for i = 1 to $6) relate to the strain components \epsilon_j (for j = 1 to $6) via \sigma_i = C_{ij} \epsilon_j, where the strains for shear components (i,j = 4,5,6) use engineering shear strains \gamma_{23} = 2\epsilon_{23}, \gamma_{13} = 2\epsilon_{13}, \gamma_{12} = 2\epsilon_{12} to maintain consistency with the factor of 2 in the stiffness tensor definition. The resulting stiffness matrix [C] is symmetric (C_{ij} = C_{ji}), reflecting the symmetry of the elastic tensor from thermodynamic considerations, and similarly for the compliance matrix [S] = [C]^{-1}. For an isotropic material, the Voigt notation simplifies the stiffness matrix to a form dependent on only two independent parameters, the Lamé constants \lambda and \mu (where \mu is the shear modulus). The non-zero elements are C_{11} = C_{22} = C_{33} = \lambda + 2\mu, C_{12} = C_{13} = C_{23} = \lambda, and C_{44} = C_{55} = C_{66} = \mu, with all other off-diagonal elements zero. This structure arises because isotropy imposes maximal symmetry, reducing the 21 potential independent components of the general anisotropic case to just 2. Material symmetry further reduces the number of independent elastic constants in the Voigt matrix. For triclinic symmetry (lowest symmetry), all 21 components are independent due to the absence of symmetry planes or axes. Hexagonal symmetry, common in materials like or certain alloys, reduces this to 5 independent constants by enforcing equality among certain off-diagonal elements related to the hexagonal axis. These reductions exploit the inherent symmetries of the crystal lattice, making the matrix sparser and computationally efficient. In practical applications, such as finite element analysis for simulating structural deformations in anisotropic components, the Voigt matrix form serves as standard input for numerical solvers, enabling efficient handling of the contracted tensor without full fourth-rank operations. This notation facilitates the integration of elastic properties into engineering software, where symmetry-imposed zeros and equalities directly inform mesh-based computations.

Determination and Measurement

Experimental Techniques

Experimental techniques for measuring elastic moduli encompass both static and dynamic approaches, applicable to isotropic and anisotropic materials, to determine properties such as , , and through controlled mechanical loading or wave propagation. These methods rely on precise to apply and measure resulting or deformation, ensuring reproducibility under standardized conditions. Static methods involve quasi-static loading to directly assess stress-strain relationships. For , tensile or tests are standard, where a specimen is elongated or compressed along its axis using a , and the modulus is calculated as the slope of the initial linear portion of the stress-strain curve. These tests follow ASTM E8 for metallic materials, specifying specimen geometry and loading rates to ensure uniform deformation. determination employs torsion tests, twisting a cylindrical specimen to measure and , yielding the ratio of to shear strain. is obtained via hydrostatic in a cell, where volumetric change under uniform is monitored, often using piston-cylinder apparatuses for fluids or porous materials. Dynamic methods utilize vibrational or wave-based excitations for non-destructive evaluation, particularly useful for small samples or in-situ measurements. Ultrasonic pulse-echo techniques propagate longitudinal and transverse waves through the material, measuring travel times to compute velocities; for instance, the longitudinal wave speed is given by v_L = \sqrt{\frac{K + \frac{4}{3}G}{\rho}}, where K is the bulk modulus, G is the shear modulus, and \rho is density, allowing derivation of moduli from known density. Resonance ultrasound spectroscopy (RUS) excites the sample's natural modes of vibration and analyzes the frequency spectrum to extract all elastic constants simultaneously from a single rectangular parallelepiped specimen, offering high precision for both isotropic and anisotropic cases. For anisotropic materials, such as single crystals, measurements account for directionality by orienting specimens along specific crystallographic axes or using averaged polycrystalline responses. Orientation-dependent tests, like tensile loading on oriented single-crystal rods, reveal variations, while quantifies lattice strains under applied to infer single-crystal elastic constants from peak shifts in diffraction patterns. Synchrotron-based high-energy enhances resolution for in-situ studies under load. Standardized protocols, such as , minimize errors from specimen preparation and testing, but sources like , microstructural defects, and misalignment can introduce variability up to 10-20% in measured moduli for textured metals. For thin films and coatings, applies a indenter to probe local properties, deriving the reduced modulus E_r via the as E_r = \frac{\sqrt{\pi}}{2} \frac{S}{\sqrt{A}}, where S is the from the unloading curve and A is the projected contact area. This technique is particularly effective for nanoscale layers, isolating film contributions from substrate effects.

Computational Methods

Computational methods for determining elastic moduli rely on theoretical frameworks and simulations that predict from atomic-scale interactions, without direct experimental input. These approaches are essential for materials where measurements are challenging, such as or extreme environments. Primary techniques include methods, empirical , and simulations, each offering trade-offs in accuracy, computational cost, and applicability to different classes. Recent advancements also encompass models trained on large DFT datasets to predict elastic tensors rapidly for high-throughput materials screening, achieving accuracies comparable to calculations for diverse inorganic compounds as of 2025. Ab initio calculations, particularly density functional theory (DFT), compute elastic constants by minimizing the ground-state energy of a crystal under applied strains or stresses. In this approach, the elastic stiffness tensor components C_{ijkl} are derived from the second derivatives of the total energy with respect to strain, or equivalently from stress-strain relationships obtained by deforming the unit cell and fitting quadratic responses. This method is implemented in codes like VASP and Quantum ESPRESSO, which solve the Kohn-Sham equations to obtain electron densities and energies. For example, GGA-DFT calculations for silicon yield C_{11} \approx 155 GPa, approximating the experimental value of \approx 166 GPa with about 7% error and demonstrating the technique's reliability for covalent solids. Empirical potentials, such as the embedded atom method (EAM), approximate interatomic forces for efficient simulations of metallic systems. Developed from density functional concepts, EAM models the total energy as a sum of pairwise interactions and embedding energies dependent on local , allowing derivation of elastic moduli from the of force constants or by computing stress responses to deformations. These potentials are parameterized against experimental or data, enabling rapid calculations for large metallic structures while capturing many-body effects crucial for moduli in alloys. Molecular dynamics (MD) simulations extend these methods to finite temperatures and larger systems by integrating with . Elastic moduli are obtained through finite strain techniques, where uniform deformations are applied to the simulation cell, and stresses are averaged over equilibrium ensembles to yield the stiffness tensor components. This approach is particularly useful for disordered or nanoscale materials, providing statistical averages that account for . Recent formulations incorporate non-affine displacements for improved accuracy in anisotropic cases. Despite their strengths, these methods have limitations. Standard DFT functionals often underestimate moduli in van der Waals-bound systems due to poor treatment of forces, necessitating corrections like DFT-D3. Advancements in , such as HSE06 introduced in the 2000s and refined post-2010, enhance accuracy by incorporating exact Hartree-Fock exchange, though at higher computational expense. The resulting elastic tensors can then inform applications, linking microscopic interactions to macroscopic properties.

Factors Influencing Elastic Moduli

Temperature and Pressure Dependence

The elastic moduli of solids generally decrease with increasing temperature, primarily due to and anharmonic effects in the , which soften the vibrations. This softening arises because higher temperatures excite modes with greater amplitudes, leading to reduced resistance to deformation as atoms deviate further from their positions. The , defined as \gamma = -\frac{d \ln \omega}{d \ln V}, quantifies this by relating changes in frequencies \omega to volume variations V, typically taking values between 1 and 2 for most solids. At low temperatures, the temperature dependence of elastic moduli M can be approximated empirically by M(T) = M_0 (1 - \beta T^2), where M_0 is the modulus at T=0 and \beta is a material-specific reflecting contributions to . For example, in , the elastic moduli exhibit a modest softening, decreasing by approximately 8% from to 1600 , demonstrating the material's exceptional thermal stability despite the general trend. In contrast, polymers show a dramatic drop in modulus at the temperature T_g, where the material shifts from a rigid glassy state (moduli ~GPa) to a compliant rubbery state (moduli ~), driven by increased molecular mobility. Under hydrostatic pressure, elastic moduli typically increase as the atomic lattice is compressed, enhancing interatomic interactions and stiffness. The pressure derivative of the , dK/dP, exceeds 1 for most solids, often ranging from 3 to 7, indicating that decreases nonlinearly with applied pressure. This links directly to the K, which measures resistance to uniform compression and governs overall volumetric response. At high pressures, the Mie-Grüneisen provides a framework for describing this behavior, expressing pressure P as P = P(V) + \frac{\gamma}{V} E, where E is and \gamma the , allowing prediction of modulus stiffening through volume-dependent thermal contributions. Ab initio predictions of these dependencies often employ the quasi-harmonic approximation, which accounts for volume changes with temperature while assuming harmonic vibrations at fixed volumes, enabling computation of moduli evolution without full anharmonic treatments. This approach accurately reproduces experimental trends in materials like metals and oxides across wide temperature and pressure ranges.

Microstructural Effects

The elastic modulus of a material is profoundly influenced by its crystal structure, particularly in anisotropic crystals where the arrangement of atoms leads to direction-dependent stiffness. In single-crystal materials, the elastic properties vary significantly with orientation due to the underlying lattice symmetry. For instance, in graphite, a hexagonal crystal with layered structure, the Young's modulus in the basal plane reaches approximately 1 TPa, reflecting the strong in-plane covalent bonding, while perpendicular to the basal plane, it drops to about 36 GPa owing to weak van der Waals interlayer forces. This extreme anisotropy arises from the elastic stiffness tensor components, such as c_{11} \approx 1060 GPa for in-plane deformation and c_{33} \approx 36.5 GPa along the c-axis, as determined through nanoindentation and theoretical modeling. Such directional variations are captured by the full elastic tensor in anisotropic materials, linking microstructural symmetry to macroscopic mechanical response. Defects within the crystal lattice, such as vacancies, voids, and dislocations, generally soften the material by disrupting the coherent atomic bonding and introducing compliance. Vacancies and voids act as stress concentrators, reducing the effective load-bearing cross-section and thereby lowering the ; for example, a 10% increase in can decrease the by 5-10% in low- crystalline materials, following a near-linear relationship for small void fractions. Dislocations, as line defects, further contribute to this softening by enabling local anelastic deformation and altering the apparent constants, with their and configuration influencing the measured in single crystals through interactions that promote internal friction and reduced rigidity. In metals and semiconductors, these defects collectively diminish the modulus from its ideal single-crystal value, with the extent depending on defect concentration and type. In composite materials, the microstructure—comprising distinct phases like embedded in a matrix—determines the overall through volume fraction-weighted contributions. For unidirectional -reinforced composites under longitudinal loading, the Voigt provides an upper-bound estimate of the effective , given by E_c = f_f E_f + f_m E_m, where f_f and f_m = 1 - f_f are the volume fractions of the and matrix, respectively, and E_f and E_m are their individual moduli. This model assumes uniform strain across phases and is particularly applicable when align with the loading direction, as in carbon- composites where high-modulus (e.g., E_f \approx 200-500 GPa) dominate the response for fractions above 50%. Deviations occur in real microstructures due to interfacial effects, but the captures the scaling of with phase distribution. At the nanoscale, microstructural features like surface-to-volume ratio introduce size-dependent effects on elastic modulus, primarily through surface and elasticity. In nanowires, as dimensions shrink below 100 nm, the heightened surface contribution often increases the beyond the bulk value; for example, surface tensile in metallic nanowires stiffens the by inducing compressive residual strains that enhance overall to deformation. This phenomenon, observed in and nanowires, arises because surface atoms experience altered bonding, leading to an effective modulus that scales inversely with , with increases up to 20-50% reported for diameters under 50 nm. Such effects underscore the role of free surfaces in modulating stiffness at mesoscales. A illustrative example of microstructural influence is seen in , where the transition from crystalline to amorphous form alters the elastic modulus due to atomic disorder. Single-crystal silicon exhibits a of approximately 160 GPa (averaged over orientations), reflecting its lattice with strong directional bonds. In contrast, , lacking long-range order, has a lower modulus of approximately 70-80 GPa, as the irregular network reduces the average bond stiffness and introduces softer local configurations, though both remain close due to similar short-range Si-Si bonding.

References

  1. [1]
    Elastic Moduli - an overview | ScienceDirect Topics
    Elastic modulus, also known as Young's modulus, is defined as a measurement of a material's stiffness, calculated as the ratio of elastic stress to elastic ...
  2. [2]
    Understanding Elastic Moduli | Dewesoft
    Oct 27, 2025 · An elastic modulus is a fundamental property of an object or material that describes its ability to deform non-permanently under stress.What are elastic moduli? · Young's modulus (E) · Shear modulus (G) · Bulk modulus
  3. [3]
    12.3 Stress, Strain, and Elastic Modulus - University Physics Volume 1
    Sep 19, 2016 · Young's modulus Y is the elastic modulus when deformation is caused by either tensile or compressive stress, and is defined by Equation 12.33.<|control11|><|separator|>
  4. [4]
    Elastic Modulus: Definition, Values, and Examples - Xometry
    Apr 15, 2023 · Elastic modulus, often referred to as Young's modulus, is a measure of a material's stiffness or resistance to elastic (non-permanent) deformation under stress.
  5. [5]
    Modulus of Elasticity - Instron
    Modulus is the measurement of a material's elasticity. Elastic modulus quantifies a material's resistance to non-permanent, or elastic, deformation.
  6. [6]
    12.4: Stress, Strain, and Elastic Modulus (Part 1) - Physics LibreTexts
    Mar 16, 2025 · Young's modulus Y is the elastic modulus when deformation is caused by either tensile or compressive stress, and is defined by Equation 12.4.
  7. [7]
    Introduction to Stress and Equations of Motion - COMSOL
    Apr 19, 2018 · When solid bodies are deformed, internal forces get distributed in the material. These are called stresses. Stress has the unit of force per area.
  8. [8]
    Cauchy stress tensor – Knowledge and References - Taylor & Francis
    The Cauchy stress tensor is a mathematical representation of the relationship between forces and areas in the present (deformed) configuration.
  9. [9]
    [PDF] 3.3 The Cauchy Stress Tensor
    Here, the stress is defined more formally, Cauchy's Law. Cauchy's Law states that there exists a Cauchy stress tensor σ which maps the normal. to a surface to ...
  10. [10]
    Stress and Stress Types Explained - saVRee
    Stress intensity within the body of a component is expressed as one of three basic types of internal load. They are known as tensile, compressive, and shear.
  11. [11]
    Stress: Cauchy Sensor Tensor - Engineering at Alberta Courses
    A principal state of stress (no shear stresses) is characterized by a Cauchy matrix with zero shear stresses (zero off diagonal components). A hydrostatic state ...Definitions · The Cauchy Stress Tetrahedron · Normal and Shear Stress
  12. [12]
    Strain - Continuum Mechanics
    Strain, like stress, is a tensor. And like stress, strain is a tensor simply because it obeys the standard coordinate transformation principles of tensors. It ...Pure Shear Strains · 2-D Notation · Tensor Shear Terms
  13. [13]
    Strain - Mechanical Engineering
    The change in length, divided by the original length, is the strain. We use strain to normalize deformations with respect to the size of the geometry.
  14. [14]
    Mechanics of Materials: Strain - Boston University
    A positive value corresponds to a tensile strain, while negative is compressive. Shear strain occurs when the deformation of an object is response to a shear ...
  15. [15]
    Why All These Stresses and Strains? | COMSOL Blog
    Nov 21, 2013 · The engineering strain is defined as the ratio between the elongation and the original length, . For larger stretches, however, it is more ...The Tensile Test · Stress Measures · Conjugate Quantities
  16. [16]
    Infinitesimal Strain Theory - an overview | ScienceDirect Topics
    The infinitesimal strain theory, or small deformation theory, is a mathematical approach to the description of the deformation of a solid body.Development Of Elastoplastic... · 7.2. 1 Infinitesimal... · 5.1. 1 Linear Elastic...
  17. [17]
    Stress | SpringerLink
    Jan 27, 2014 · The concept of stress traces back nearly two centuries to the published works of Cauchy (1827). Cauchy generalized Euler's concept of pressure ...
  18. [18]
    [PDF] Roark's Formulas for Stress and Strain - Jackson Research Group
    ... strain and stress. Hooke's law for a linear, homogeneous, isotropic material is simply that the normal strain is directly proportional to the normal stress ...
  19. [19]
    [PDF] 12.510 Introduction to Seismology - MIT OpenCourseWare
    Strain can be represented by a similar tensor, in the form εkl. The ... We can now write Hooke's Law in a general form. σ ij. = c ijkl ε kl. Later ...
  20. [20]
  21. [21]
    “Historical Introduction” from A Treatise on the Mathematical Theory ...
    Jan 1, 2007 · In the interval between the discovery of Hooke's Law and that of the general differential equations of Elasticity by Navier, the attention of ...<|control11|><|separator|>
  22. [22]
    [PDF] 6 Linear Elasticity
    This assumption turns out to be an excellent predictor of the response of components which undergo small deformations, for example steel and concrete ...
  23. [23]
    [PDF] Hooke's law
    In continuous elastic materials Hooke's law implies that strain is a linear function of stress. Some materials that we usually think of as highly elastic, for ...
  24. [24]
    Elastic deformation: Hooke's law and stiffness - DoITPoMS
    Hooke's law states that the stress is directly proportional to strain. This upper limit of stress that can be applied for Hooke's law to be obeyed is called the ...
  25. [25]
    Hooke's Law - Continuum Mechanics
    Uniaxial Tension Example ; σxx=2 σ x x = 2 ; E=10 E = 10 ; ν=0.5 ν = 0.5 ; ϵxx=0.2 ϵ x x = 0.2 ; ϵyy=ϵzz=−0.1 ϵ y y = ϵ z z = − 0.1 ...
  26. [26]
    nglos324 - Young's Modulus
    The linking equation between stress and strain is known as Hooke's law: s = E e, and governs the linear elastic response of materials. The elastic modulus, E, ...<|control11|><|separator|>
  27. [27]
    [PDF] Chapter 26 Elastic Properties of Materials
    Hooke's Law relationship holds in that the shear stress is proportional to shear strain,. Page 6. 26-5. F tan. A. = S δx h. (Hooke's Law) . (26.3.3) where the ...Missing: uniaxial | Show results with:uniaxial
  28. [28]
    Mechanical Properties of Materials | MechaniCalc
    The elastic modulus of the material affects how much it deflects under a load, and the strength of the material determines the stresses that it can withstand ...
  29. [29]
    12.4 Elasticity and Plasticity – University Physics Volume 1
    When stress is gradually increased beyond the elastic limit, the material undergoes plastic deformation. Rubber-like materials show an increase in stress with ...
  30. [30]
    [PDF] 10-1 CHAPTER 10 DEFORMATION 10.1 Stress-Strain Diagrams ...
    The constant of proportionality is called the Elastic Modulus or Young's Modulus (E). The relationship between stress and strain in this region is given by ...
  31. [31]
    materials - elasticconstants
    In the elastic response range for the material being tested stress and strain are related through a material property known as an "Elastic Constant." For the ...
  32. [32]
    12.3 Stress, Strain, and Elastic Modulus - UCF Pressbooks
    From Figure, Young's modulus for steel is Y = 2.0 × 10 11 Pa . Solution. Substituting numerical values into the equations gives us.
  33. [33]
    Material - TAM 2XX
    ... elastic. Typical values: #youngs‑modulus. Material, Young's Modulus [GPa]. Mild Steel, 210. Copper, 120. Bone, 18. Plastic, 2. Rubber, 0.02. 2. Yielding.
  34. [34]
    Elastic Constant - an overview | ScienceDirect Topics
    It can be shown quite generally that, if no body torques act on the material, the maximum number of independent elastic constants is 21. As a result, pairs of ...
  35. [35]
    Isotropic Linear Elastic Material - an overview | ScienceDirect Topics
    Here λ and μ are elastic parameters called Lamé constants. Due to their ... An isotropic elastic material has two independent material constants.
  36. [36]
    Thomas Young and the Theory of Structures 1807-2007
    In 1807 in the space of a single paper, Thomas Young introduced the concept of an elastic modulus to define material stiffness and presented the first full and ...
  37. [37]
    Calculations of single-crystal elastic constants made simple
    The lower the symmetry, the more the independent elastic constants: Cubic crystals have three but triclinic crystals have 21 independent elastic constants, for ...
  38. [38]
    [PDF] Materials and Elasticity Lecture M17: Engineering Elastic Constants
    Let us examine how we can estimate the Young's modulus of the resulting composite material. Initially we will consider a two dimensional case. 2-D fibers ...
  39. [39]
    Wrought Iron - Microstructure and Properties - Nuclear Power
    Young's modulus of elasticity of wrought iron is 190 GPa. Young's modulus of elasticity is the elastic modulus for tensile and compressive stress in the linear ...
  40. [40]
    E111 Standard Test Method for Young's Modulus, Tangent ... - ASTM
    Sep 8, 2017 · This test method covers the determination of Young's modulus, tangent modulus, and chord modulus of structural materials.
  41. [41]
    Thermal expansion and temperature-dependent Young's modulus of ...
    Apr 28, 2022 · Up to a temperature of approximately 240 °C, Young's modulus steadily increased, after which it showed a maximum between 260 °C and 270 °C, ...
  42. [42]
    What is Poisson's ratio?
    Poisson's ratio is related to elastic moduli K (also called B), the bulk modulus; G as the shear modulus; and E, Young's modulus, by the following (for ...
  43. [43]
    [PDF] Torsion (Additional Topics)
    Shear Stress. T. J ρ τ = (2). Angle of Twist. TL. GJ θ = (3). Where: θ = Angle of twist (radians). T = Applied torque (moment). L = length of member. G = Shear ...
  44. [44]
    Torsion Test | Mechanical Testing Instructional Laboratory | Illinois
    The tester measures the ... A property obtained from this test is the shear modulus of rupture in torsion. Mechanical Testing Instructional Laboratory.
  45. [45]
    Bulk Modulus: Definition, How it Works, Formula, Examples, and ...
    Apr 15, 2023 · Bulk modulus is a measure of a material's resistance to uniform compression. It describes how much a material decreases in volume when pressure ...
  46. [46]
    Bulk Modulus and Fluid Elasticities - The Engineering ToolBox
    The Bulk Modulus Elasticity - or Volume Modulus - is a material property characterizing the compressibility of a fluid - how easy a unit volume of a fluid can ...
  47. [47]
    Mechanics of solids - Stress, Strain, Elasticity - Britannica
    The French mathematician Antoine Parent introduced the concept of shear stress in 1713, but Coulomb was the one who extensively developed the idea, first in ...
  48. [48]
    [PDF] Ultrasonic measurements of elastic properties of small specimens at ...
    The present method for measuring density and elastic constants of solids as a function of pressure uses Bridgman's anvil apparatus and ultrasonic interferometry ...
  49. [49]
    What Is the Shear Modulus? Definition and Examples - ThoughtCo
    Jan 30, 2019 · Table of Shear Modulus Values ; Rubber, 0.0006 ; Polyethylene, 0.117 ; Plywood, 0.62 ; Nylon, 4.1 ; Lead (Pb), 13.1.
  50. [50]
    Bulk Modulus Calculator
    Bulk modulus for lithium, water, and other materials · Water: 2.1 GPa (300,000 psi). · Lithium 11 GPa (1,595,415 psi). · Aluminum: 75 GPa (10,877,828 psi).
  51. [51]
    [PDF] Module 3 Constitutive Equations - MIT
    Write the relationship between stresses and strains for an isotropic elastic material whose Lamé constants are λ and µ and whose coefficient of thermal ...<|separator|>
  52. [52]
    Relation between Young's modulus and shear modulus - DoITPoMS
    We can also relate the bulk modulus K to Young's modulus E . Consider a hydrostatic stress state acting on an isotropic cube shown below. isotropic cube diagram.
  53. [53]
    Extreme Poisson's ratios and related elastic crystal properties
    From a theoretical point of view, it has long been known that for an isotropic elastic medium, Poisson's ratio may assume a value between - 1 and 1 2 and remain ...
  54. [54]
    What Are the Properties of Aluminum? - AZoM
    What Are the Properties of Aluminum? ; Poisson's Ratio, 0.32, 0.36 ; Shear Modulus, 25, 34 ; Tensile Strength, 75, 360 ; Young's Modulus, 68, 88.5 ...
  55. [55]
    [PDF] 2.002 mechanics & materials ii - stress-strain relations
    the number of elastic constants from the number 21 for the most general anisotropic material. Page 11. Recall that if σij ei oe, and of eoe' are the.
  56. [56]
    [PDF] Isotropic linear elastic response
    • Stiffness is a second derivative of energy: Cijkl = d2U/dϵij dϵkl. • Results in 21 unique elastic constants. Better written with Voigt notation: ✏ij = X.
  57. [57]
    [PDF] An In-Depth Tutorial on Constitutive Equations for Elastic Anisotropic ...
    Dec 1, 2011 · ... 21 Independent Constants ... Cijkl = Cjikl εεkl = εεlk. Cijkl = Cijlk. Cijkl = Cjikl. Cijkl = Cijlk. Page 42. 38. GENERAL FORM OF HOOKE'S LAW ...
  58. [58]
    [PDF] Anisotropic Elasticity
    Mar 15, 2020 · ... 21 independent ... How do we show that symmetry reduces the number of independent coefficients in an anisotropic elasticity modulus tensor?
  59. [59]
    Elastic Constants | Materials Project Documentation
    Mar 7, 2023 · The elastic constants from the Materials Project (MP) are calculated from first-principles Density Functional Theory (DFT).
  60. [60]
    [PDF] Charting the complete elastic properties of inorganic crystalline ...
    Mar 17, 2015 · The elastic constant tensor of an inorganic compound provides a complete description of the response of the material to external stresses in ...
  61. [61]
    [PDF] Elastic Modulus Measurement - NPL Publications
    Static methods for measuring modulus include tensile tests and flexure tests. Depending on the application, material availability and data requirement users ...
  62. [62]
    Measurement Methods for Materials Properties: Elasticity | NIST
    Feb 19, 2017 · Simple static mechanical tests are used to evaluate the elastic constants. Specimens are either pulled in tension, squeezed in compression, bent in flexure or ...
  63. [63]
    [PDF] Tension Testing of Metallic Materials1
    This standard is issued under the fixed designation E8/E8M; the number immediately following the designation indicates the year of.
  64. [64]
    Modulus of Elasticity, What is it? - Industrial Physics
    It uses the ratio of shear stress to the shear strain to measure elasticity. It is typically carried out via a static torsion test or by a dynamic test that ...Missing: experimental | Show results with:experimental
  65. [65]
    [PDF] Comparison of Dynamic Methods for Determining Elastic Property ...
    These dynamic methods, which are categorized into pulse techniques for measuring transit time of an ultrasonic pulse (UPA), and resonance techniques that set ...
  66. [66]
    Application of resonant ultrasound spectroscopy to determine elastic ...
    RUS is an advanced ultrasonic method based on determination of the elastic moduli from resonant spectra of free elastic vibrations of a small sample of the ...
  67. [67]
    X‐ray diffraction study of the single‐crystal elastic moduli of ε‐Fe up ...
    May 13, 2005 · The accuracy of the calculated elastic moduli has improved compared to previous measurements using similar techniques because of an increase in ...
  68. [68]
    A method for measuring single-crystal elastic moduli using high ...
    This paper presents a method – based on high-energy synchrotron X-ray diffraction data and a crystal-based finite element simulation formulation
  69. [69]
    [PDF] Investigation of anisotropy in elastic modulus of steel
    Aug 30, 2010 · The 'measured' steel modulus easily varies by from 180-220. GPa. 2. The variation in the steel modulus is due to. •. Variations in measurements.
  70. [70]
    Oliver-Pharr Method - an overview | ScienceDirect Topics
    The Oliver-Pharr method is defined as a technique for measuring the elastic modulus of materials based on a purely-elastic contact assumption, although it can ...
  71. [71]
    Density functional theory predictions of the mechanical properties of ...
    Jul 2, 2021 · Here we introduce the concepts behind DFT, and highlight a number of case studies and methodologies for predicting the elastic constants of materials.Missing: seminal | Show results with:seminal
  72. [72]
    Density functional theory for calculation of elastic properties of ...
    Nov 1, 1998 · A theoretical formalism to calculate the single crystal elastic constants for orthorhombic crystals from first principle calculations is described.Missing: seminal | Show results with:seminal
  73. [73]
    Embedded-atom method: Derivation and application to impurities ...
    Jun 15, 1984 · We develop the embedded-atom method [Phys. Rev. Lett. 50, 1285 (1983)], based on density-functional theory, as a new means of calculating ground-state ...
  74. [74]
    Analytic elastic coefficients in molecular calculations: Finite strain ...
    Jul 10, 2023 · Here, we revisit the theory behind zero-temperature, finite-strain elastic moduli and extend it to explicitly consider nonaffine displacements.Abstract · Article Text · FINITE DEFORMATION AND... · RESULTS
  75. [75]
    Assessing Effects of van der Waals Corrections on Elasticity of ... - NIH
    Sep 29, 2023 · Van der Waals corrections affect the absolute values of elasticity, not the trend, and PBE-D3 causes a 1% lattice constant shrinkage. Young's  ...
  76. [76]
    Challenges for Density Functional Theory | Chemical Reviews
    Whether DFT can predict the small energy differences associated with van der Waals interactions or if additional corrections or nonlocal functionals of the ...
  77. [77]
    Temperature dependence of elastic moduli and isobaric volume ...
    The Anderson–Grüneisen parameter corresponding to each elastic modulus is determined and found to be directly proportional to the volume, with the variation in ...
  78. [78]
    Elastic Moduli: a Tool for Understanding Chemical Bonding and ...
    Dec 14, 2022 · Elastic constants are a measure of a material's resistance to small (i.e., elastic) deformations under an applied stress. Strain in the elastic ...
  79. [79]
    Temperature dependent Grüneisen parameter | Science China ...
    Aug 7, 2019 · A new thermodynamic model of temperature-dependent potential energy is proposed here to investigate the temperature dependent Grüneisen parameter of bulk ...
  80. [80]
    Temperature variation of the isothermal bulk modulus in solids
    Feb 12, 2019 · The quasi-harmonic Debye-Gruneisen model is used to show that upon heating, the solid loses its elasticity and melts, when the thermal pressure ...
  81. [81]
    Temperature dependence of the elastic moduli of diamond
    Feb 1, 1998 · The elastic moduli decrease by only ≈ 8 % up to 1600 K, indicating that diamond retains its reputed hardness even at high temperatures.Missing: above 1000
  82. [82]
    Glass Transition Region - an overview | ScienceDirect Topics
    The glass transition region is characterized by a steep drop in the polymer instantaneous or storage modulus.<|control11|><|separator|>
  83. [83]
    Elastic moduli, pressure derivatives, and temperature derivatives of ...
    Ultrasonic wave velocities in single-crystal forsterite (F) and single-crystal olivine (O) have been measured as a function of pressure and of temperature ...
  84. [84]
    Thermodynamics with the Grüneisen parameter: Fundamentals and ...
    The incomplete and complete Mie-Grüneisen EOSs show a concave surface in the pressure-specific volume-internal energy (P–V-E), pressure-specific volume-shock ...<|control11|><|separator|>
  85. [85]
    Thermo-Elasticity of Materials from Quasi-Harmonic Calculations
    Within the quasi-harmonic approximation, the thermal dependence of the bulk modulus can be determined at a relatively low computational cost by computing the ...
  86. [86]
    [PDF] Nanoindentations in kinking nonlinear elastic solids - Research and ...
    Graphite has one of the highest elastic anisotropies observed, with a basal plane stiffness (perpendicular to the c-axis) c11 = 1060 GPa and (parallel to ...
  87. [87]
    [PDF] A Finite Element Model to Predict the Effect of Porosity on Elastic ...
    May 27, 2018 · The data observed a linear trend of increasing percent reduction in elastic modulus with increasing porosity (R2 = . 77). Therefore, it is ...
  88. [88]
    Influence of dislocations on the apparent elastic constants in single ...
    The present paper takes dislocation character, crystal structure and dislocation network geometry into account and obtains the anisotropic mechanical response.
  89. [89]
    Rule-of-Mixture Equation - an overview | ScienceDirect Topics
    The rule of mixtures equation refers to a set of equations used to predict the properties, such as modulus and strength, of fiber-reinforced composites ...
  90. [90]
    Common Misconceptions on Rules of Mixtures for Predicting Elastic ...
    Mar 25, 2024 · This paper derives the correct formulas for predicting the elastic properties of composites according to the Voigt and Reuss assumptions. The ...Introduction · II. Rules of Mixtures According... · III. Hybrid Rules of Mixtures for...Missing: original | Show results with:original
  91. [91]
    Surface Effect on the Elastic Behavior of Static Bending Nanowires
    Surface stress and surface elasticity have been recognized as important factors that may explain the experimentally measured size dependent elastic modulus of ...
  92. [92]
    Effect of size and shape on the elastic modulus of metal nanowires
    Aug 2, 2021 · Here, we demonstrate a profound influence of shape and size on the elastic properties of materials on the example of gold nanowires.
  93. [93]
    Elastic constants of defected and amorphous silicon with the ...
    Oct 29, 2004 · The elastic constants of a wide range of models of defected crystalline and amorphous silicon are calculated, using the environment-dependent interatomic ...Article Text · METHODS · RESULTS · DISCUSSION