Young's modulus


Young's modulus, or the Young modulus, is a mechanical property that measures the stiffness of a solid material. It defines the relationship between stress and strain in a material in the linear elasticity regime of a uniaxial deformation.
Young's modulus is named after the 19th-century British scientist Thomas Young; but the concept was developed in 1727 by Leonhard Euler, and the first experiments that used the concept of Young's modulus in its current form were performed by the Italian scientist Giordano Riccati in 1782, pre-dating Young's work by 25 years. The term modulus is derived from the Latin root term modus which means measure.

Definition

Linear elasticity

A solid material will undergo elastic deformation when a small load is applied to it in compression or extension. Elastic deformation is reversible.
At near-zero stress and strain, the stress–strain curve is linear, and the relationship between stress and strain is described by Hooke's law that states stress is proportional to strain. The coefficient of proportionality is Young's modulus. The higher the modulus, the more stress is needed to create the same amount of strain; an idealized rigid body would have an infinite Young's modulus. Conversely, a very soft material such as a fluid, would deform without force, and would have zero Young's Modulus.
Not many materials are linear and elastic beyond a small amount of deformation.

Formula used and units

, where
Both and have units of pressure, while is dimensionless. Young's moduli are typically so large that they are expressed not in pascals but in megapascals or gigapascals.

Note

Material stiffness should not be confused with these properties:
Young's modulus enables the calculation of the change in the dimension of a bar made of an isotropic elastic material under tensile or compressive loads. For instance, it predicts how much a material sample extends under tension or shortens under compression. The Young's modulus directly applies to cases of uniaxial stress, that is tensile or compressive stress in one direction and no stress in the other directions. Young's modulus is also used in order to predict the deflection that will occur in a statically determinate beam when a load is applied at a point in between the beam's supports. Other elastic calculations usually require the use of one additional elastic property, such as the shear modulus, bulk modulus or Poisson's ratio. Any two of these parameters are sufficient to fully describe elasticity in an isotropic material.

Linear versus non-linear

Young's modulus represents the factor of proportionality in Hooke's law, which relates the stress and the strain. However, Hooke's law is only valid under the assumption of an elastic and linear response. Any real material will eventually fail and break when stretched over a very large distance or with a very large force; however all solid materials exhibit nearly Hookean behavior for small enough strains or stresses. If the range over which Hooke's law is valid is large enough compared to the typical stress that one expects to apply to the material, the material is said to be linear. Otherwise the material is said to be non-linear.
Steel, carbon fiber and glass among others are usually considered linear materials, while other materials such as rubber and soils are non-linear. However, this is not an absolute classification: if very small stresses or strains are applied to a non-linear material, the response will be linear, but if very high stress or strain is applied to a linear material, the linear theory will not be enough. For example, as the linear theory implies reversibility, it would be absurd to use the linear theory to describe the failure of a steel bridge under a high load; although steel is a linear material for most applications, it is not in such a case of catastrophic failure.
In solid mechanics, the slope of the stress–strain curve at any point is called the tangent modulus. It can be experimentally determined from the slope of a stress–strain curve created during tensile tests conducted on a sample of the material.

Directional materials

Young's modulus is not always the same in all orientations of a material. Most metals and ceramics, along with many other materials, are isotropic, and their mechanical properties are the same in all orientations. However, metals and ceramics can be treated with certain impurities, and metals can be mechanically worked to make their grain structures directional. These materials then become anisotropic, and Young's modulus will change depending on the direction of the force vector. Anisotropy can be seen in many composites as well. For example, carbon fiber has a much higher Young's modulus when force is loaded parallel to the fibers. Other such materials include wood and reinforced concrete. Engineers can use this directional phenomenon to their advantage in creating structures.

Calculation

Young's modulus E, can be calculated by dividing the tensile stress,
, by the engineering extensional strain,, in the elastic portion of the physical stress–strain curve:
where

Force exerted by stretched or contracted material

The Young's modulus of a material can be used to calculate the force it exerts under specific strain.
where F is the force exerted by the material when contracted or stretched by.
Hooke's law for a stretched wire can be derived from this formula:
where it comes in saturation
But note that the elasticity of coiled springs comes from shear modulus, not Young's modulus.

Elastic potential energy

The elastic potential energy stored in a linear elastic material is given by the integral of the Hooke's law:
now by explicating the intensive variables:
This means that the elastic potential energy density is given by:
or, in simple notation, for a linear elastic material:
, since the strain is defined.
In a nonlinear elastic material the Young's modulus is a function of the strain, so the second equivalence no longer holds and the elastic energy is not a quadratic function of the strain:

Relation among elastic constants

For homogeneous isotropic materials simple relations exist between elastic constants that allow calculating them all as long as two are known:

Temperature dependence

The Young's modulus of metals varies with the temperature and can be realized through the change in the interatomic bonding of the atoms and hence its change is found to be dependent on the change in the work function of the metal. Although classically, this change is predicted through fitting and without a clear underlying mechanism, the Rahemi-Li model demonstrates how the change in the electron work function leads to change in the Young's modulus of metals and predicts this variation with calculable parameters, using the generalization of the Lennard-Jones potential to solids. In general, as the temperature increases, the Young's modulus decreases via
Where the electron work function varies with the temperature as and is a calculable material property which is dependent on the crystal structure. is the electron work function at T=0 and is constant throughout the change.

Approximate values

Young's modulus can vary somewhat due to differences in sample composition and test method. The rate of deformation has the greatest impact on the data collected, especially in polymers. The values here are approximate and only meant for relative comparison.
MaterialGPaMpsi
Rubber 0.01–0.11.45–
Low-density polyethylene0.11–0.861.6–
Diatom frustules 0.35–2.770.05–0.4
PTFE 0.50.075
HDPE0.80.116
Bacteriophage capsids1–30.15–0.435
Polypropylene1.5–20.22–0.29
Polycarbonate2–2.40.29-0.36
Polyethylene terephthalate 2–2.70.29–0.39
Nylon2–40.29–0.58
Polystyrene, solid3–3.50.44–0.51
Polystyrene, foam0.0025–0.0070.00036–0.00102
Medium-density fiberboard 40.58
Wood 111.60
Human Cortical Bone142.03
Glass-reinforced polyester matrix17.22.49
Aromatic peptide nanotubes19–272.76–3.92
High-strength concrete304.35
Amino-acid molecular crystals21–443.04–6.38
Carbon fiber reinforced plastic 30–504.35–7.25
Hemp fiber355.08
Magnesium metal 456.53
Glass 50–907.25–13.1
Flax fiber588.41
Aluminum6910
Mother-of-pearl 7010.2
Aramid70.5–112.410.2–16.3
Tooth enamel 8312
Stinging nettle fiber8712.6
Bronze96–12013.9–17.4
Brass100–12514.5–18.1
Titanium 110.316
Titanium alloys105–12015–17.5
Copper 11717
Carbon fiber reinforced plastic 18126.3
Silicon Single crystal, different directions130–18518.9–26.8
Wrought iron190–21027.6–30.5
Steel 20030
polycrystalline Yttrium iron garnet 19328
single-crystal Yttrium iron garnet 20029
Cobalt-chrome 220–25829
Aromatic peptide nanospheres230–27533.4–40
Beryllium 28741.6
Molybdenum 329–33047.7–47.9
Tungsten 400–41058–59
Silicon carbide 45065
Tungsten carbide 450–65065–94
Osmium 525–56276.1–81.5
Single-walled carbon nanotube1,000+150+
Graphene 1050152
Diamond 1050–1210152–175
Carbyne 32100