Fact-checked by Grok 2 weeks ago

Young's modulus

Young's modulus is a fundamental mechanical property of linearly elastic, isotropic materials that measures their stiffness under tensile or compressive loading, defined as the ratio of applied to the resulting axial within the elastic deformation regime. Named after the English Thomas Young, who introduced the concept in his 1807 lectures on , it extends Robert Hooke's earlier 1660 of elasticity by providing a material-specific constant for the linear relationship between and . The modulus, often denoted as Y or E, is mathematically expressed as Y = \frac{\sigma}{\epsilon}, where \sigma is the normal (force per unit cross-sectional area, in pascals) and \epsilon is the longitudinal (change in divided by original , dimensionless). This yields units of , typically gigapascals (GPa) for solids, with values reflecting inherent atomic bonding strength; for instance, exhibits a Young's modulus of approximately 200 GPa, indicating high , while rubber is around 0.01 GPa, showing greater deformability. It applies only below the material's yield strength, where deformation is reversible, and assumes small strains to maintain linearity. In and , Young's modulus is essential for predicting structural behavior, such as beam deflection or cable elongation, and is one of three primary elastic constants alongside the and , interconnected via for isotropic materials. Tabulated values for common materials guide design in , civil , and , where mismatches in modulus can lead to failure under load. Experimental determination typically involves on universal testing machines, ensuring uniform distribution.

Definition

Mathematical formulation

Young's modulus, denoted E, quantifies the stiffness of a under uniaxial loading and is defined as the of \sigma to the corresponding axial \varepsilon within the linear regime: E = \frac{\sigma}{\varepsilon}.
This relation arises from the empirical observation that, for small deformations, is directly proportional to , as described by .
The normal stress \sigma represents the internal force per unit area acting perpendicular to the cross-section of the material and is expressed as \sigma = F / A, where F is the applied force and A is the original cross-sectional area.
The axial strain \varepsilon measures the relative elongation or contraction along the direction of the applied force and is given by \varepsilon = \Delta L / L_0, where \Delta L is the change in length and L_0 is the original length.
Since strain is dimensionless and stress has dimensions of force per unit area, Young's modulus inherits the units of stress.
In the (SI), the unit of Young's modulus is the pascal (Pa), equivalent to one per square meter (N/m²), though values for materials are often reported in gigapascals (GPa, or 10⁹ Pa) to reflect their typical magnitude. The concept and terminology of Young's modulus originated with the English Thomas Young, who introduced it in his 1807 publication A Course of Lectures on and the Mechanical Arts, where he systematically explored the elastic properties of materials and defined the modulus as the constant ratio characterizing longitudinal elasticity under or .

Physical interpretation

Young's modulus quantifies a material's , representing its resistance to deformation under uniaxial ; higher values indicate that the material undergoes minimal or for a given applied , while lower values signify greater compliance. For instance, exhibits an exceptionally high Young's modulus of approximately 1200 GPa, making it one of the stiffest known materials, whereas has a much lower value around 0.01 GPa, allowing it to stretch easily without permanent damage. At the and molecular scale, Young's modulus arises from the of interatomic bonds, which act analogously to springs governed by functions such as the . These bonds provide restorative forces that oppose deformation, with the modulus directly related to the of the curve near —the steeper the , the stronger the bonds and higher the —and to the density of such bonds per unit volume. In crystalline solids, for example, covalent bonds in contribute to its high modulus due to their strong, directional nature and high packing density. This interpretation applies exclusively to the elastic regime, where deformations are reversible and the fully recovers its original upon removal of the , distinguishing it from plastic deformation that involves irreversible rearrangements or bond breaking. Young's modulus thus characterizes only the initial, linear portion of the -strain response, beyond which permanent changes occur. The value of Young's modulus exhibits temperature dependence, generally decreasing as temperature rises because increased enhances vibrations, which effectively soften the interatomic bonds through anharmonic effects that alter the average and reduce the restoring force . This qualitative trend holds for most , with the modulus softening more pronounced in materials with weaker bonds.

Elasticity Context

Linear elasticity principles

Linear elasticity forms the foundational framework for understanding material deformation under load, predicated on the core assumption that is directly proportional to for small deformations. This relationship, encapsulated in , implies that the material's response is linear and reversible, meaning it returns to its original shape upon unloading without residual deformation or dependence on prior loading history. The theory applies to deformations where changes in dimensions are minute fractions of the original size, ensuring that geometric nonlinearities are negligible. In the context of elastic theory, materials are distinguished by their homogeneity in response: isotropic materials exhibit properties independent of direction, leading to symmetric deformation under uniaxial or , whereas anisotropic materials display direction-dependent and due to internal microstructures like orientations or alignments. For uniaxial loading, isotropic behavior simplifies , as the proportionality between axial stress and remains consistent across orientations, while anisotropy introduces coupling between normal and shear strains. Complementing this framework is , ν, a dimensionless parameter defined as the negative ratio of to axial in a subjected to uniaxial , quantifying the 's tendency to expand or contract transversely. In isotropic , ν ranges theoretically from just above -1 (for auxetic materials that expand laterally under tension) to 0.5 (indicating incompressibility), with most common materials falling between 0.2 and 0.5. These principles hold validity only within the regime of small strains, typically less than 0.1% to 0.2%, where the linear stress-strain relationship persists; exceeding this threshold introduces nonlinearity, plasticity, or failure, rendering the assumptions inapplicable. In uniaxial cases, Young's modulus emerges as the specific proportionality constant linking axial stress to strain under these conditions.

Relation to other elastic moduli

In isotropic materials, Young's modulus E is interconnected with the G and the K through relations that depend on \nu. The G represents a material's resistance to shear deformation, quantified as the ratio of applied to the resulting shear strain. The K measures resistance to volumetric change under uniform hydrostatic pressure, defined as the ratio of that pressure to the relative volume decrease. These moduli are linked by the formulas E = 2G(1 + \nu) and E = 3K(1 - 2\nu), which allow conversion between them for materials exhibiting isotropic linear elasticity. These interrelations stem from the generalized Hooke's law in three dimensions, expressed via the stress-strain tensor for isotropic solids: \sigma_{ij} = \lambda \delta_{ij} \epsilon_{kk} + 2\mu \epsilon_{ij}, where \lambda and \mu are the Lamé constants, with \mu = G. The full set of Lamé relations includes \lambda = \frac{E \nu}{(1 + \nu)(1 - 2\nu)} and K = \lambda + \frac{2}{3} G, enabling a complete description of elastic behavior from any two independent moduli. To illustrate the differences in these moduli across material classes, the following table provides representative values, highlighting how stiff materials like metals and ceramics exhibit high values across all moduli, while compliant ones like polymers and rubber show stark contrasts, particularly in K due to near-incompressibility.
Material ClassExampleYoung's Modulus E (GPa)Shear Modulus G (GPa)Bulk Modulus K (GPa)
Metals20080160
CeramicsAlumina400150250
Polymers10.43
ElastomersRubber0.010.0032
These values underscore the scale: metals and ceramics have moduli in the hundreds of GPa, reflecting rigidity, whereas polymers and rubber are orders of magnitude softer in [E](/page/E!) and [G](/page/G), but rubber's high [K](/page/K) approximates incompressibility.

Calculation Methods

From stress and strain

Young's modulus is determined experimentally by applying a uniaxial tensile F to a prismatic specimen with original L and cross-sectional area A, then measuring the resulting longitudinal extension \Delta L within the linear elastic regime, typically following standards such as ASTM E111. The axial \epsilon is calculated as \epsilon = \frac{\Delta L}{L}, representing the relative deformation. The corresponding engineering \sigma is \sigma = \frac{F}{A}, the force per unit area. Young's modulus E is then obtained directly as the ratio E = \frac{\sigma}{\epsilon}, assuming small deformations where the material behaves linearly. This approach integrates with , which posits that the restorative force in an body is proportional to the displacement from equilibrium. For a tensile specimen, yields F = \left( \frac{E A}{L} \right) \Delta L, where \frac{E A}{L} acts as an effective spring constant. Experimentally, E is derived from the slope of the linear portion of the force-displacement curve, or equivalently, the stress-strain curve in the region below the proportional limit. This slope method ensures E captures the material's intrinsic stiffness under controlled loading. Measurements must account for potential errors arising from specimen geometry and loading conditions. In ductile materials during , necking—a localized reduction in cross-section—can initiate near the yield point, invalidating the uniform assumption if data beyond the elastic limit are included. For slender samples with high length-to-diameter ratios, compressive components from misalignment or unintended bending may induce , prematurely deviating from linear behavior and underestimating E. Proper fixturing, precise extensometry, and restriction to low strains mitigate these issues. Theoretical estimation of Young's modulus employs methods like (DFT) and (MD) simulations to predict E for novel materials without empirical fitting. In DFT, elastic constants are computed from energy-strain relations under , enabling E derivation for crystalline structures. MD simulations extend this by incorporating atomic vibrations and finite-temperature effects, as seen in 2020s advancements modeling 2D materials and alloys. These computational approaches have validated E values for systems like compounds, guiding materials design.

From elastic potential energy

In the context of uniaxial deformation, the elastic potential energy U stored in a volume V is given by the integral of the work done by the applied , which for linear behavior equals U = \frac{1}{2} \sigma \varepsilon V, where \sigma is the axial and \varepsilon is the corresponding . Substituting \sigma = E \varepsilon, this simplifies to U = \frac{1}{2} E \varepsilon^2 V, revealing Young's modulus E as the constant that scales the quadratic dependence of stored energy on . The derivation of E from energy density follows from equating the strain energy density u = U / V = \frac{1}{2} \sigma \varepsilon to the work per unit volume under gradually increasing load, which integrates to u = \frac{1}{2} [E](/page/E!) \varepsilon^2 for infinitesimal deformations within the elastic limit. This quadratic form confirms E as the proportionality factor in the energy-strain relation, distinct from the linear stress-strain definition, and underscores its role in quantifying the material's resistance to deformation through recoverable energy storage. In finite element analysis, Young's modulus E is incorporated into the strain energy component of the total functional, which is minimized to solve for displacements in structural simulations. Specifically, the strain \frac{1}{2} \sigma \varepsilon or \frac{1}{2} E \varepsilon^2 contributes to the , enabling numerical minimization of the potential energy for complex geometries under various loads. Unlike gravitational potential energy, which arises from positional configuration in a field and depends on mass and height, is configuration-dependent through material deformation and is fully recoverable upon unloading, restoring the original shape without dissipation in ideal linear cases.

Practical Usage

In and isotropic materials

In applications for isotropic materials, where elastic properties are uniform in , Young's modulus serves as a fundamental for assessing and designing structural under linear elastic conditions. A primary use is in predicting beam deflection to ensure structural integrity and compliance. For a cantilever beam subjected to a point load F at its free end, the maximum deflection \delta is calculated as \delta = \frac{F L^3}{3 E I} where L is the beam length and I is the second moment of area of the cross-section; this equation allows engineers to quantify bending deformation and optimize designs for applications like support structures. Young's modulus also features prominently in compliance matrices for basic structures such as rods and wires under axial loading. In isotropic materials, the compliance matrix relates \epsilon to \sigma via elements like S_{11} = 1/E for uniaxial cases, enabling straightforward computation of extensions or contractions in or scenarios. Engineers select materials based on Young's modulus to align with functional needs; high-modulus materials like (E \approx 200 GPa) are preferred for bridges to limit deflection under load, whereas lower-modulus options such as elastomers are chosen for springs to permit substantial elastic recovery and deformation. Standardized measurement of Young's modulus in metals and alloys follows ASTM E8, which prescribes at ambient temperatures to derive E as the of the initial linear portion of the - , ensuring consistent and comparable for design validation.

In anisotropic and nonlinear materials

In anisotropic materials, such as and composites, Young's modulus is not a scalar but varies with direction, arising from the fourth-rank that governs the - relationship. The tensor \mathbf{S}, the inverse of the stiffness tensor \mathbf{C}, relates \boldsymbol{\epsilon} to \boldsymbol{\sigma} via \epsilon_i = S_{ij} \sigma_j in , where the directional Young's modulus E(\mathbf{n}) in a unit vector direction \mathbf{n} is given by E(\mathbf{n}) = [S'_{11}]^{-1}, with \mathbf{S}' being the tensor transformed to the aligned with \mathbf{n}. This directional dependence requires up to 21 independent components for fully anisotropic triclinic , reducing for higher symmetries like orthorhombic or hexagonal. For example, in , a natural orthotropic composite, the longitudinal Young's modulus along the can be 20–30 times greater than the transverse modulus perpendicular to the , reflecting the aligned microfibrils that provide high axially but low resistance radially. Similarly, in single-crystal , Young's modulus ranges from about 130 GPa along the direction to 188 GPa along the direction, influencing applications in microelectromechanical systems where must be controlled. These variations necessitate tensor-based models for accurate prediction in composites like fiber-reinforced polymers, where the effective modulus aligns with fiber . In nonlinear elastic materials, particularly those undergoing large deformations, Young's modulus is no longer constant but represents a tangent modulus that evolves with applied , captured by hyperelastic constitutive models derived from a W(\mathbf{F}), where \mathbf{F} is the deformation gradient. The Mooney-Rivlin model, a seminal hyperelastic form for nearly incompressible rubbers, expresses W = C_{10}(I_1 - 3) + C_{01}(I_2 - 3), with invariants I_1 and I_2 of the right Cauchy-Green tensor; the initial small-strain Young's modulus is E = 6(C_{10} + C_{01}), but it stiffens or softens nonlinearly under finite s up to 500% in elastomers. This strain-dependent behavior is critical for modeling rubber components like seals or tires, where linear approximations fail. Viscoelastic materials, such as polymers, exhibit a time-dependent apparent Young's modulus due to combined elastic recovery and viscous , making it strain-rate sensitive; under high-speed , the modulus can increase significantly as molecular chains have less time to relax. This is modeled using integral forms like the Boltzmann , where the relaxation E(t) decays from an instantaneous glassy value (e.g., ~3 GPa for many polymers) to a rubbery plateau over seconds to minutes, depending on and . For impact-resistant applications in automotive plastics, reveals rate effects where at 10^3 s^{-1} strain rates exceeds quasi-static values by factors of 2–5. Recent studies on highlight induced by defects; in , pristine sheets exhibit near-isotropic in-plane Young's modulus of approximately 1 TPa, but vacancies or Stone-Wales defects reduce it to 0.15–0.95 TPa and introduce directional variations up to 20% due to altered and formations. Investigations using simulations confirm that defect densities as low as 1% cause pronounced in mechanical response, impacting applications in and composites.

Examples and Applications

Common engineering materials

Young's modulus values for common engineering materials vary significantly across metals, ceramics, and polymers, reflecting their differing stiffness under tensile or compressive loads. Metals like and aluminum exhibit high values, making them suitable for load-bearing structures, while polymers like have much lower moduli, allowing for flexibility in applications such as or . Ceramics such as and fall in between, providing rigidity for and optical uses without the ductility of metals. These values are typically measured under standard conditions and serve as a key indicator of stiffness in isotropic engineering contexts. The following table summarizes representative Young's modulus values for selected common engineering materials:
MaterialYoung's Modulus (GPa)
200
Aluminum70
30
70
1
These values are approximate and can depend on specific compositions and testing conditions. Young's modulus in metals like is largely unaffected by alloying and , as it depends primarily on interatomic bonding; minor variations (typically <2%) may occur with significant compositional changes or in specific alloys, but standard engineering practice treats it as constant around 200 GPa. Early measurements of Young's modulus on metals date back to the , with Thomas Young introducing the concept in 1807 through experiments on materials including metals to quantify elasticity. Subsequent work by researchers like Wertheim in the mid-1800s refined measurements on metallic wires and rods, establishing foundational data for applications. In practical engineering selection, materials with low Young's modulus like are chosen for flexible components such as or bumpers to absorb impacts without permanent deformation, whereas high-modulus ceramics like are preferred for rigid tools and fixtures requiring minimal deflection under load. This selection process prioritizes stiffness to match design requirements for durability and performance in everyday structures and devices.

Biological and advanced materials

Biological tissues display a broad spectrum of Young's moduli, reflecting their hierarchical architectures that integrate organic and inorganic components across multiple scales to achieve tailored mechanical responses. For instance, human cortical exhibits a longitudinal Young's modulus of 16-23 GPa, stemming from its organization into -mineral composites at the nanoscale, , lamellae, and larger osteons that distribute loads effectively. Tendons, such as the human , possess a Young's modulus around 1 GPa, enabled by a hierarchical assembly of molecules into , s, and fascicles that provide high tensile strength along the fiber direction. In contrast, human shows a much lower Young's modulus of 0.1-1 MPa, arising from its stratified structure where the dermis's interwoven and networks allow for extensibility and resilience under deformation. Advanced materials push the boundaries of Young's modulus extremes, often with implications for composite performance and functionality. Individual carbon nanotubes demonstrate a remarkable Young's modulus near 1 TPa, the highest among known materials, due to their covalent bonding and tubular geometry; measurements of individual CNTs within polymer composites yield effective moduli of 530-700 GPa for the nanotubes themselves owing to matrix interactions and alignment issues, while the overall composite modulus remains much lower (typically 1-150 GPa depending on CNT ). Similarly, exhibits a Young's modulus of approximately 1 TPa, showcasing atomic-scale for potential use in nanocomposites and electronics. Shape-memory alloys, like NiTi, feature a strain-dependent Young's modulus that varies with martensitic phase transformations, typically ranging from 20-80 GPa in the phase but softening under superelastic loading to enable large recoverable strains up to 10%. Measuring Young's modulus in soft biological materials presents unique challenges due to their low , heterogeneity, and , often requiring nanoscale techniques. (AFM) is widely employed for this purpose, applying controlled indentation to map local elastic properties in cells and tissues with resolutions down to nanometers, revealing moduli as low as tens of kPa in cellular membranes. In emerging applications, Young's modulus plays a pivotal role in for prosthetic design, where materials are engineered to match bone's 10-20 GPa range longitudinally, minimizing stress shielding and promoting . Similarly, in 2020s developments for , stretchable conductors utilize substrates with Young's moduli below 1 MPa—such as elastomers at 7-10 kPa—to accommodate strains exceeding 100% without failure, enabling wearable and conformal devices. Biological materials frequently show , with Young's modulus differing significantly between longitudinal and transverse directions due to oriented hierarchies.

References

  1. [1]
    12.3 Stress, Strain, and Elastic Modulus - UCF Pressbooks
    Young's modulus Y is the elastic modulus when deformation is caused by either tensile or compressive stress, and is defined by Figure. Dividing this equation by ...
  2. [2]
    [PDF] Hooke's law
    Thomas Young (1773–1829). English physician, physicist and egyptologist. He ... Young's modulus (1807). Given Young's modulus we may calculate the ...
  3. [3]
    Young's Modulus
    The Young's Modulus (or Elastic Modulus) is in essence the stiffness of a material. In other words, it is how easily it is bended or stretched.
  4. [4]
    Strength and Stiffness Characteristics
    Young's Modulus: Also known as the Modulus of Elasticity, is a measure of material resistance to axial deformation. Its value is obtained by measuring the slope ...
  5. [5]
    Young's modulus on UTM (Introduction to MOS) : Mechanics of ...
    When external forces act on a body, internal forces arise within the body to resist the deformation. A measure of this internal resistance is stress (ð Â ...
  6. [6]
    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.
  7. [7]
    Stress, Strain and Young's Modulus - The Engineering ToolBox
    Stress is proportional to load and strain is proportional to deformation as expressed with Hooke's Law. E = stress / strain. = σ / ε. = (Fn / A) / (dl / lo) ...
  8. [8]
    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.Learning Objectives · Tensile or Compressive Stress... · Example 12 . 4 . 1...
  9. [9]
    Elastic Modulus: Definition, Values, and Examples - Xometry
    Apr 15, 2023 · The modulus of elasticity is measured in pascals (Pa), the same SI unit used for stress. MPa and GPa are usually used to express the modulus of ...
  10. [10]
    Thomas Young and the Theory of Structures 1807-2007
    In one remarkable lecture to the Royal Institution, published in 1807, Thomas Young set out most of the modern theory of structures: Young's Modulus, ...
  11. [11]
    Thomas Young (1773 - 1829) - Biography - MacTutor
    Thomas Young was an English mathematician and physicist best known for Young's modulus and his work on interference. ... In 1807 Young published A course of ...
  12. [12]
    sp3: diamond properties
    Young's modulus, 1200, GPa. Poisson's ratio, 0.2, Dimensionless. Atomic density, 1.77 x 1023, atoms/cm3. Thermal expansion coefficient, 1.1–5.0 (300-1300K), ppm ...
  13. [13]
    [PDF] ATOMISTIC BASIS OF ELASTICITY - MIT
    Jan 27, 2000 · K is the Gas Constant. Example 6. The Young's modulus of a rubber is measured at E = 3.5 MPa for a temperature of T = 300.
  14. [14]
    Young's modulus, Hooke's law and material properties - Physclips.
    In this section, we explain the origin of elastic properties and Hooke's law by consideration of the forces and energies between atoms or molecules in a solid.
  15. [15]
    A new mechanism for low and temperature-independent elastic ...
    Jun 25, 2015 · The elastic modulus of most solids decreases when temperature increases as a consequence of thermal expansion, and such a temperature dependence ...
  16. [16]
    The Feynman Lectures on Physics Vol. II Ch. 38: Elasticity
    When you push on a piece of material, it “gives”—the material is deformed. If the force is small enough, the relative displacements of the various points in ...
  17. [17]
    Constitutive laws - 3.2 Linear Elasticity - Applied Mechanics of Solids
    The shear terms are new – in an isotropic material, no shear strain is induced by uniaxial tension. 3.2.10 Strain energy density for anisotropic, linear elastic ...
  18. [18]
    Stress, strain and Poisson's ratio - DoITPoMS
    The Poisson's ratio for most materials is in the range 0.2 to 0.5, with most metals having a value around 0.3. However, there are some materials that have ...
  19. [19]
    Relation between Young's modulus and shear modulus - DoITPoMS
    We obtained the relation between Young's modulus \( E \) and shear modulus \( G \). We can also relate the bulk modulus \( K \) to Young's modulus \( E \).
  20. [20]
    [PDF] Module 3 Constitutive Equations - MIT
    For a linearly elastic, homogeneous, isotropic material, the constitutive laws involve three parameters: Young's modulus, E, Poisson's ratio, ν, and the shear ...
  21. [21]
    Convert Elastic Modulus Constants (Shear, Young's, Bulk)
    E = Young's Modulus, also known as Modulus of Elasticity · G = Shear Modulus, also known as Modulus of Rigidity · K = Bulk Modulus · v = Poisson's Ratio.
  22. [22]
    Lamé Constants -- from Eric Weisstein's World of Physics
    where E is Young's modulus, is the Poisson ratio, G is the shear modulus, K is the bulk modulus, is the density, is P-wave speed, and is the S-wave speed.
  23. [23]
    Young's Modulus of Elasticity – Values for Common Materials
    Young's Modulus (Elastic Modulus) of various materials, including metals, plastics, and composites. How stiffness and elasticity influence material performance.
  24. [24]
    Modulus of Rigidity - The Engineering ToolBox
    Modulus of Rigidity of some Common Materials ; Rubber, 0.0003 ; Structural Steel, 79.3 ; Stainless Steel, 77.2 ; Steel, Cast, 78.
  25. [25]
    Bulk Modulus - My DataBook
    This table contains typical bulk modulus value ranges for different materials. ... Rubber, 257. Silver, 96 – 106. Solder (Tin-Lead), 33 – 58. Steel, 138 – 179.
  26. [26]
    Alumina - Aluminium Oxide - Al2O3 - A Refractory Ceramic ... - AZoM
    Alumina - Aluminium Oxide - Al2O3 - A Refractory Ceramic Oxide. View ... Bulk Modulus, 137, 324, GPa, 19.8702, 46.9922, 106 psi. Compressive Strength, 690 ...
  27. [27]
    Polymers elastic modulus and Poisson ratio | Sonelastic®
    For example, PVC has a modulus of elasticity of 2.41-4.14 GPa and a Poisson's ratio of 0.35-0.60. Nylon 6.6 has 1.59-3.79 GPa and 0.230-0.550.
  28. [28]
    [PDF] Stress – Strain Relationships
    ... stress-strain diagram is called the Modulus of Elasticity or Young's Modulus. E = σ/ε. (normal stress – strain). G = τ/γ. (shear stress – strain). E = Elastic ...
  29. [29]
  30. [30]
    [PDF] 10-1 CHAPTER 10 DEFORMATION 10.1 Stress-Strain Diagrams ...
    The relationship between stress and strain in this region is given by Equation 5-4. where: σ is the stress (psi). E is the Elastic Modulus (psi).
  31. [31]
    [PDF] Experiment 4 - Testing of Materials in Tension - CSUN
    Modulus of elasticity - Young's modulus, or the slope of the stress-strain curve in the elastic region. Necking - Local deformation of a tensile specimen.<|separator|>
  32. [32]
    Young's Modulus, 3-Point Bending & Column Buckling | Webinar
    Dec 14, 2022 · Learn how to perform tensile tests, determine Young's Modulus, and perform column buckling and three-point bending of an I-beam using the ...Missing: errors slender samples
  33. [33]
    Measurement error of Young's modulus considering the gravity and ...
    Apr 8, 2025 · Specimens with larger l o and ρ present more obvious measurement errors for Young's modulus, but the effect of ΔT is not significant. The ...Missing: slender | Show results with:slender
  34. [34]
    DFT simulation to study the mechanical, electronic, thermal ... - Nature
    Aug 16, 2025 · ... Young's modulus, bulk modulus, hardness, Poisson's ratio, mechinable index etc. ... An ab-initio study on structural, elastic, electronic, bonding ...
  35. [35]
    First-principles investigation of elastic and thermodynamic ...
    Sep 25, 2025 · The structural stability and electronic properties of U 1 − x Th x O 2 were investigated using DFT calculations, implemented in the Vienna Ab ...
  36. [36]
    Ab initio analysis of the structural, hydrogen storage, mechanical ...
    Sep 30, 2025 · Density Functional Theory (DFT) studies have been instrumental in assessing their structural stability, hydrogen adsorption, and desorption ...
  37. [37]
    High-speed and low-power molecular dynamics processing unit ...
    Nov 7, 2024 · Ab initio molecular dynamics: Concepts, recent developments, and future trends. Proc. Natl. Acad. Sci. 102, 6654–6659 (2005). Article CAS ...
  38. [38]
    [PDF] 8.2 Elastic Strain Energy
    Feb 7, 2010 · Derive an expression for the vertical displacement at point. A using the direct work-energy method, in terms of L, F, A and the Young's modulus.
  39. [39]
    [PDF] The Principle of Minimum Potential Energy
    The objective of this chapter is to explain the principle of minimum potential energy and its application in the elastic analysis of structures. Two fundamental ...
  40. [40]
    Potential Energy - The Physics Classroom
    Elastic potential energy is the energy stored in elastic materials as the result of their stretching or compressing. Elastic potential energy can be stored in ...
  41. [41]
    Compliance Matrix - an overview | ScienceDirect Topics
    The compliance matrix is defined as a matrix that relates strain to stress in a material, allowing for the calculation of strain from a given stress state.Missing: rods wires
  42. [42]
    Beam Deflection: Definition, Formula, and Examples - SkyCiv
    May 3, 2024 · The tutorial provides beam deflection definition and equations/formulas for simply supported, cantilever, and fixed beams · Beam deflection ...
  43. [43]
    Young's Modulus and Specific Stiffness - Property Information
    Young's modulus measures the resistance of a material to elastic (recoverable) deformation under load. A stiff material has a high Young's modulus and changes ...
  44. [44]
    [PDF] Bridge Steels and Their Mechanical Properties - AISC
    Grade 50 rapidly became the material of choice for primary bridge ... Young's modulus also decreases at higher temperatures leading to an increase in elastic.<|separator|>
  45. [45]
    What is Young's Modulus? Understanding the Basics of Elasticity
    A high Young's Modulus means the material is very stiff, like the metal spring, and a low Young's Modulus means it's stretchy, like the rubber band.
  46. [46]
    Metals Tensile Testing Concepts, Equations, And Theory - ADMET
    Feb 26, 2023 · This blog post goes over metals tensile testing concepts and theory with an emphasis on equations to measure each property.
  47. [47]
    [PDF] Anisotropic Elasticity
    Mar 15, 2020 · Linear and non-linear properties can be anisotropic; a constitutive relation can also be anisotropic. • Both single and poly-crystals can be ...
  48. [48]
    Anisotropic elasticity of silicon and its application to the modelling of ...
    The stiffness and compliance coefficient matrix depend on crystal orientation and, consequently, Young's modulus, the shear modulus and Poisson's ratio as well.
  49. [49]
    Study of the Mechanical Properties of Wood under Transverse ...
    Dec 28, 2021 · The modulus of elasticity of wood perpendicular to the grain is only about 1/30th of what it is parallel to grain [9].
  50. [50]
    Mooney-Rivlin Models - Continuum Mechanics
    Mooney-Rivlin models are popular for modeling the large strain nonlinear behavior of incompressible materials, ie, rubber.
  51. [51]
    Applied Mechanics of Solids (A.F. Bower) Section 3.5: Hyperelastic ...
    Hyperelastic constitutive laws are used to model materials that respond elastically when subjected to very large strains.<|separator|>
  52. [52]
    A viscoelasticity model for polymers: Time, temperature, and ...
    In general, Young's modulus (stiffness) will increase with hydrostatic pressure and strain rate (Lin, 2011) (Cho, 2016). Therefore, linear viscoelasticity can ...
  53. [53]
    [PDF] ENGINEERING VISCOELASTICITY - MIT
    Oct 24, 2001 · This document is intended to outline an important aspect of the mechanical response of polymers and polymer-matrix composites: the field of ...
  54. [54]
    Effect of Strain Rate, Temperature, Vacancy, and Microcracks ... - NIH
    Mar 21, 2024 · The results indicate that wrinkling significantly reduces the ultimate tensile strength and Young's modulus of graphene, while increasing the ...
  55. [55]
    Tuning graphene mechanical anisotropy via defect engineering
    Aug 9, 2025 · For instance, the Young's modulus of defect-containing graphene ranges from 0.15 to 0.95 TPa, compared to pristine graphene's 1 TPa [11,12]. ...Missing: post- | Show results with:post-<|control11|><|separator|>
  56. [56]
    Young's Modulus - an overview | ScienceDirect Topics
    Young's modulus is a measurement of stress over strain. Simply put, Young's modulus is the slope of a line on stress vs strain plot.
  57. [57]
    The variation of Young's modulus and the hardness with tempering ...
    After low-temperature tempering of high-alloy chromium steels Young's modulus begins to increase and then remains almost constant up to tempering temperatures ...
  58. [58]
    Will the Young's Modulus (E) value change with heat treatment?
    Oct 24, 2014 · Adeleke: If it is an steel alloy, the Youngs Modulus changes very only slighty by heat treatment. The ultimate strength and yield can change ...
  59. [59]
    Thomas Young: Young's Modulus | TecQuipment
    Jun 13, 2018 · Young described the characterization of elasticity that came to be known as Young's Modulus, denoted as E, in 1807. This is a ratio of the ...
  60. [60]
    Young's Modulus: Modulus of Elasticity Units & Formula
    Jul 11, 2025 · Young's modulus is the ratio of stress to the strain applied to the material. The force is applied along the longitudinal axis of the specimen tested.
  61. [61]
    [PDF] Recent advances on the measurement and calculation of the elastic ...
    The elastic modulus of human compact bone in the longitudinal direction is reported to be in the range of 16-23. GPa; whereas in the transverse direction is ...
  62. [62]
    [PDF] A Multimodal Exploration of Altered Mechanical and Histological ...
    Young's Modulus. Studies reporting the elastic modulus of the Achilles tendon are fairly consistent. They fall into a range of 1.16 (0.15) GPa21 to 2.0 (0.4) ...<|control11|><|separator|>
  63. [63]
    Age-dependent biomechanical properties of the skin - PMC - NIH
    In the literature, the Young's modulus (E) of the skin fluctuates between 0.42 MPa and 0.85 MPa [1, 20] for the torsion tests, between 4.6 MPa and 20 MPa [21] ...
  64. [64]
  65. [65]
    The effective Young's modulus of carbon nanotubes in composites
    The effective modulus of the SWNTs has been found to be in the range 530-700 GPa which, while being impressive, is lower than the generally accepted value of ...
  66. [66]
    [PDF] Properties of a Ni19.5Pd30Ti50.5 high-temperature shape memory ...
    The elastic modulus of the material was very dependent on strain level; therefore, dynamic Young's Modulus was determined as a function of temperature by an ...
  67. [67]
    Atomic Force Microscopy Methods to Measure Tumor Mechanical ...
    Jun 22, 2023 · Atomic force microscopy (AFM) is a popular tool for evaluating the mechanical properties of biological materials (cells and tissues) at high ...<|control11|><|separator|>
  68. [68]
    Young's modulus variation along the prosthesis length; profile 1 (a),...
    6 . The Young ' s modulus of the prosthesis is a critical design variable that largely determines how load is transferred via the prosthesis to the bone.
  69. [69]
    Elastic and ultra stable ionic conductors for long-life-time soft robots ...
    Jul 1, 2025 · Our EG gels possess high ionic conductivity (1.31 mS cm −1 ), low Young's modulus (7.3 kPa), high elasticity (~100%), and good extreme environmental stability.
  70. [70]
    Hierarchical Structure and Properties of the Bone at Nano Level
    Nov 10, 2022 · Bone is a highly hierarchical complex structure that consists of organic and mineral components represented by collagen molecules (CM) and ...Missing: Young's | Show results with:Young's