Jump to content

Young's modulus: Difference between revisions

From Wikipedia, the free encyclopedia
Content deleted Content added
Luckas-bot (talk | contribs)
m r2.7.1) (robot Adding: hr:Youngov modul elastičnosti
No edit summary
Line 28: Line 28:
where
where
:<var>E</var> is the Young's modulus (modulus of elasticity)
:<var>E</var> is the Young's modulus (modulus of elasticity)
:<var>F</var> is the force applied to the object;
:<var>F</var> is the force exerted by the object;
:<var>A<sub>0</sub></var> is the original cross-sectional area through which the force is applied;
:<var>A<sub>0</sub></var> is the original cross-sectional area through which the force is applied;
:<var>ΔL</var> is the amount by which the length of the object changes;
:<var>ΔL</var> is the amount by which the length of the object changes;

Revision as of 21:40, 9 September 2011

Young's modulus is a measure of the stiffness of an elastic material and is a quantity used to characterize materials. It is defined as the ratio of the uniaxial stress over the uniaxial strain in the range of stress in which Hooke's Law holds.[1] In solid mechanics, the slope of the stress-strain curve at any point is called the tangent modulus. The tangent modulus of the initial, linear portion of a stress-strain curve is called Young's modulus, also known as the tensile 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. In anisotropic materials, Young's modulus may have different values depending on the direction of the applied force with respect to the material's structure.

It is also commonly, but incorrectly, called the elastic modulus or modulus of elasticity, because Young's modulus is the most common elastic modulus used, but there are other elastic moduli measured, too, such as the bulk modulus and the shear modulus.

Young's modulus is named after Thomas Young, the 19th century British scientist. However, 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, predating Young's work by 25 years.[2]

Units

Young's modulus is the ratio of stress, which has units of pressure, to strain, which is dimensionless; therefore, Young's modulus has units of pressure.

The SI unit of modulus of elasticity (E, or less commonly Y) is the pascal (Pa or N/m² or m−1·kg·s−2). The practical units used are megapascals (MPa or N/mm²) or gigapascals (GPa or kN/mm²). In United States customary units, it is expressed as pounds (force) per square inch (psi).

Usage

The Young's modulus calculates 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. Some calculations also require the use of other material properties, such as the shear modulus, density, or Poisson's ratio. It also helps in selection of materials for particular structural applications.

Linear versus non-linear

For many materials, Young's modulus is essentially constant over a range of strains. Such materials are called linear, and are said to obey Hooke's law. Examples of linear materials are steel, carbon fiber and glass. Non-linear materials include rubber and soils, except under very small strains.

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 from which the force is applied. Anisotropy can be seen in many composites as well. For example, carbon fiber has much higher Young's modulus (is much stiffer) when force is loaded parallel to the fibers (along the grain). 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 tensile strain in the elastic (initial, linear) portion of the stress-strain curve:

where

E is the Young's modulus (modulus of elasticity)
F is the force exerted by the object;
A0 is the original cross-sectional area through which the force is applied;
ΔL is the amount by which the length of the object changes;
L0 is the original length of the object.

Force exerted by stretched or compressed 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 compressed or stretched by ΔL.

Hooke's law can be derived from this formula, which describes the stiffness of an ideal spring:

where

Elastic potential energy

The elastic potential energy stored is given by the integral of this expression with respect to L:

where Ue is the elastic potential energy.

The elastic potential energy per unit volume is given by:

, where is the strain in the material.

This formula can also be expressed as the integral of Hooke's law:

Relation among elastic constants

For homogeneous isotropic materials simple relations exist between elastic constants (Young's modulus E, shear modulus G, bulk modulus K, and Poisson's ratio ν) that allow calculating them all as long as two are known:

Approximate values

Influences of selected glass component additions on Young's modulus of a specific base glass

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.

Approximate Young's modulus for various materials[3]
Material GPa lbf/in² (psi)
Rubber (small strain) 0.01–0.1 1,500–15,000
PTFE (Teflon)[citation needed] 0.5 75,000
Low density polyethylene[citation needed] 0.2 30,000
HDPE 0.8
Polypropylene 1.5-2 217,000–290,000
Bacteriophage capsids[4] 1–3 150,000–435,000
Polyethylene terephthalate 2-2.7
Polystyrene 3-3.5 435,000–505,000
Nylon 2–4 290,000–580,000
Diatom frustules (largely silicic acid)[5] 0.35–2.77 50,000–400,000
Medium-density fiberboard[6] 4 580,000
Pine wood (along grain)[citation needed] 8.963 1,300,000
Oak wood (along grain) 11 1,600,000
High-strength concrete (under compression) 30 4,350,000
Hemp fiber [7] 35
Magnesium metal (Mg) 45 6,500,000
Flax fiber [8] 58
Aluminium 69 10,000,000
Stinging nettle fiber [9] 87
Glass (see chart) 50–90
Aramid[10] 70.5–112.4
Mother-of-pearl (nacre, largely calcium carbonate) [11] 70 10,000,000
Tooth enamel (largely calcium phosphate)[12] 83 12,000,000
Brass and bronze 100–125 17,000,000
Titanium (Ti) 16,000,000
Titanium alloys 105–120 15,000,000–17,500,000
Copper (Cu) 117 17,000,000
Glass-reinforced plastic (70/30 by weight fibre/matrix, unidirectional, along grain)[citation needed] 40–45 5,800,000–6,500,000
Glass-reinforced polyester matrix [13] 17.2 2,500,000
Carbon fiber reinforced plastic (50/50 fibre/matrix, unidirectional, along grain)[citation needed] 125–150 18,000,000–22,000,000
Carbon fiber reinforced plastic (70/30 fibre/matrix, unidirectional, along grain)[14] 181 26,300,000
Silicon[15] 130-185
Wrought iron 190–210
Steel 200 29,000,000
polycrystalline Yttrium iron garnet (YIG)[16] 193 28,000,000
single-crystal Yttrium iron garnet (YIG)[17] 200 30,000,000
Beryllium (Be) 287 42,000,000
Molybdenum (Mo) 329
Tungsten (W) 400–410 58,000,000–59,500,000
Sapphire (Al2O3) along C-axis[citation needed] 435 63,000,000
Silicon carbide (SiC) 450 65,000,000
Osmium (Os) 550 79,800,000
Tungsten carbide (WC) 450–650 65,000,000–94,000,000
Single-walled carbon nanotube[18][19] 1,000+ 145,000,000+
Graphene 1000
Diamond (C)[20] 1220 150,000,000–175,000,000

See also

References

  1. ^ IUPAC, Compendium of Chemical Terminology, 2nd ed. (the "Gold Book") (1997). Online corrected version: (2006–) "modulus of elasticity (Young's modulus), E". doi:10.1351/goldbook.M03966
  2. ^ The Rational Mechanics of Flexible or Elastic Bodies, 1638–1788: Introduction to Leonhardi Euleri Opera Omnia, vol. X and XI, Seriei Secundae. Orell Fussli.
  3. ^ http://www.engineeringtoolbox.com/young-modulus-d_417.html
  4. ^ Ivanovska IL, de Pablo PJ, Sgalari G, MacKintosh FC, Carrascosa JL, Schmidt CF, Wuite GJL (2004). "Bacteriophage capsids: Tough nanoshells with complex elastic properties". Proc Nat Acad Sci USA. 101 (20): 7600–5. Bibcode:2004PNAS..101.7600I. doi:10.1073/pnas.0308198101. PMC 419652. PMID 15133147.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  5. ^ Subhash G, Yao S, Bellinger B, Gretz MR. (2005). "Investigation of mechanical properties of diatom frustules using nanoindentation". J Nanosci Nanotechnol. 5 (1): 50–6. doi:10.1166/jnn.2005.006. PMID 15762160.{{cite journal}}: CS1 maint: multiple names: authors list (link)
  6. ^ Material Properties Data: Medium Density Fiberboard (MDF)
  7. ^ Nabi Saheb, D.; Jog, JP. (1999). "Natural fibre polymer composites: a review". Advances in Polymer Technology. 18 (4): 351–363. doi:10.1002/(SICI)1098-2329(199924)18:4<351::AID-ADV6>3.0.CO;2-X. {{cite journal}}: line feed character in |journal= at position 12 (help)
  8. ^ Bodros, E. (2002). "Analysis of the flax fibres tensile behaviour and analysis of the tensile stiffness increase". Composite Part A. 33 (7): 939–948. doi:10.1016/S1359-835X(02)00040-4.
  9. ^ Bodros, E.; Baley, C. (15 May 2008). "Study of the tensile properties of stinging nettle fibres (Urtica dioica)". Materials Letters. 62 (14): 2143–2145. doi:10.1016/j.matlet.2007.11.034.
  10. ^ DuPont (2001). "Kevlar Technical Guide": 9. {{cite journal}}: Cite journal requires |journal= (help)
  11. ^ A. P. Jackson,J. F. V. Vincent and R. M. Turner (1988). "The Mechanical Design of Nacre". Proc. R. Soc. Lond. B. 234 (1277): 415–440. Bibcode:1988RSPSB.234..415J. doi:10.1098/rspb.1988.0056.
  12. ^ M. Staines, W. H. Robinson and J. A. A. Hood (1981). "Spherical indentation of tooth enamel". Journal of Materials Science.
  13. ^ Polyester Matrix Composite reinforced by glass fibers (Fiberglass). [SubsTech] (2008-05-17). Retrieved on 2011-03-30.
  14. ^ Epoxy Matrix Composite reinforced by 70% carbon fibers [SubsTech]. Substech.com (2006-11-06). Retrieved on 2011-03-30.
  15. ^ Physical properties of Silicon (Si). Ioffe Institute Database. Retrieved on 2011-05-27.
  16. ^ Chou, H. M.; Case, E. D. (November, 1988). "Characterization of some mechanical properties of polycrystalline yttrium iron garnet (YIG) by non-destructive methods". Journal of Materials Science Letters. 7 (11): 1217–1220. doi:10.1007/BF00722341. {{cite journal}}: Check date values in: |date= (help)
  17. ^ YIG properties
  18. ^ L. Forro; et al. "Electronic and mechanical properties of carbon nanotubes" (PDF). {{cite web}}: Explicit use of et al. in: |author= (help)
  19. ^ Y.H.Yang; et al. (2011). "Radial elasticity of single-walled carbon nanotube measured by atomic force microscopy". Applied Physics Letters. 98: 041901. Bibcode:2011ApPhL..98d1901Y. doi:10.1063/1.3546170. {{cite journal}}: Explicit use of et al. in: |author= (help)
  20. ^ Spear and Dismukes (1994). Synthetic Diamond – Emerging CVD Science and Technology. Wiley, NY. ISBN 9780471535898.

External links

Conversion formulae
Homogeneous isotropic linear elastic materials have their elastic properties uniquely determined by any two moduli among these; thus, given any two, any other of the elastic moduli can be calculated according to these formulas, provided both for 3D materials (first part of the table) and for 2D materials (second part).
3D formulae Notes

There are two valid solutions.
The plus sign leads to .

The minus sign leads to .

Cannot be used when
2D formulae Notes
Cannot be used when