You are on page 1of 7

Strength of materials - Wikipedia 31/01/18, 6)11 PM

Strength of materials
Strength of materials, also called mechanics of materials, is a subject which deals with the behavior of solid
objects subject to stresses and strains. The complete theory began with the consideration of the behavior of one and
two dimensional members of structures, whose states of stress can be approximated as two dimensional, and was then
generalized to three dimensions to develop a more complete theory of the elastic and plastic behavior of materials. An
important founding pioneer in mechanics of materials was Stephen Timoshenko.

The study of strength of materials often refers to various methods of calculating the stresses and strains in structural
members, such as beams, columns, and shafts. The methods employed to predict the response of a structure under
loading and its susceptibility to various failure modes takes into account the properties of the materials such as its
yield strength, ultimate strength, Young's modulus, and Poisson's ratio; in addition the mechanical element's
macroscopic properties (geometric properties), such as its length, width, thickness, boundary constraints and abrupt
changes in geometry such as holes are considered.

Contents
Definition
Types of loadings
Stress terms
Strength terms
Strain (deformation) terms
Stress–strain relations
Design terms
Failure theories
See also
References
Further reading
External links

Definition
In mechanics of materials, the strength of a material is its ability to withstand an applied load without failure or
plastic deformation. The field of strength of materials deals with forces and deformations that result from their acting
on a material. A load applied to a mechanical member will induce internal forces within the member called stresses
when those forces are expressed on a unit basis. The stresses acting on the material cause deformation of the material
in various manners including breaking them completely. Deformation of the material is called strain when those
deformations too are placed on a unit basis. The applied loads may be axial (tensile or compressive), or rotational
(strength shear). The stresses and strains that develop within a mechanical member must be calculated in order to
assess the load capacity of that member. This requires a complete description of the geometry of the member, its

https://en.wikipedia.org/wiki/Strength_of_materials Page 1 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

constraints, the loads applied to the member and the properties of the material of which the member is composed.
With a complete description of the loading and the geometry of the member, the state of stress and state of strain at
any point within the member can be calculated. Once the state of stress and strain within the member is known, the
strength (load carrying capacity) of that member, its deformations (stiffness qualities), and its stability (ability to
maintain its original configuration) can be calculated. The calculated stresses may then be compared to some measure
of the strength of the member such as its material yield or ultimate strength. The calculated deflection of the member
may be compared to a deflection criteria that is based on the member's use. The calculated buckling load of the
member may be compared to the applied load. The calculated stiffness and mass distribution of the member may be
used to calculate the member's dynamic response and then compared to the acoustic environment in which it will be
used.

Material strength refers to the point on the engineering stress–strain curve (yield stress) beyond which the material
experiences deformations that will not be completely reversed upon removal of the loading and as a result the member
will have a permanent deflection. The ultimate strength refers to the point on the engineering stress–strain curve
corresponding to the stress that produces fracture.

Types of loadings
Transverse loadings — Forces applied perpendicular to the longitudinal axis of a member. Transverse loading
causes the member to bend and deflect from its original position, with internal tensile and compressive strains
accompanying the change in curvature of the member.[1] Transverse loading also induces shear forces that
cause shear deformation of the material and increase the transverse deflection of the member.
Axial loading — The applied forces are collinear with the longitudinal axis of the member. The forces cause the
member to either stretch or shorten.[2]
Torsional loading — Twisting action caused by a pair of externally applied equal and oppositely directed force
couples acting on parallel planes or by a single external couple applied to a member that has one end fixed
against rotation.

Stress terms
Uniaxial stress is expressed by

where F is the force [N] acting on an area A [m2].[3] The area can be the
undeformed area or the deformed area, depending on whether
engineering stress or true stress is of interest.
A material being loaded in a)
Compressive stress (or compression) is the stress state caused by compression, b) tension, c) shear.
an applied load that acts to reduce the length of the material
(compression member) along the axis of the applied load, it is in
other words a stress state that causes a squeezing of the material. A simple case of compression is the uniaxial
compression induced by the action of opposite, pushing forces. Compressive strength for materials is generally
higher than their tensile strength. However, structures loaded in compression are subject to additional failure
modes, such as buckling, that are dependent on the member's geometry.
Tensile stress is the stress state caused by an applied load that tends to elongate the material along the axis of
the applied load, in other words the stress caused by pulling the material. The strength of structures of equal
cross sectional area loaded in tension is independent of shape of the cross section. Materials loaded in tension

https://en.wikipedia.org/wiki/Strength_of_materials Page 2 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

are susceptible to stress concentrations such as material defects or abrupt changes in geometry. However,
materials exhibiting ductile behavior (most metals for example) can tolerate some defects while brittle materials
(such as ceramics) can fail well below their ultimate material strength.
Shear stress is the stress state caused by the combined energy of a pair of opposing forces acting along parallel
lines of action through the material, in other words the stress caused by faces of the material sliding relative to
one another. An example is cutting paper with scissors[4] or stresses due to torsional loading.

Strength terms
Mechanical properties of materials include the yield strength, tensile strength, fatigue strength, crack resistance, and
other characteristics.[5]

Yield strength is the lowest stress that produces a permanent deformation in a material. In some materials, like
aluminium alloys, the point of yielding is difficult to identify, thus it is usually defined as the stress required to
cause 0.2% plastic strain. This is called a 0.2% proof stress.[6]

Compressive strength is a limit state of compressive stress that leads to failure in a material in the manner of
ductile failure (infinite theoretical yield) or brittle failure (rupture as the result of crack propagation, or sliding along
a weak plane — see shear strength).
Tensile strength or ultimate tensile strength is a limit state of tensile stress that leads to tensile failure in the
manner of ductile failure (yield as the first stage of that failure, some hardening in the second stage and breakage
after a possible "neck" formation) or brittle failure (sudden breaking in two or more pieces at a low stress state).
Tensile strength can be quoted as either true stress or engineering stress, but engineering stress is the most
commonly used.
Fatigue strength is a measure of the strength of a material or a component under cyclic loading,[7] and is usually
more difficult to assess than the static strength measures. Fatigue strength is quoted as stress amplitude or
stress range ( ), usually at zero mean stress, along with the number of cycles to failure under
that condition of stress.

Impact strength is the capability of the material to withstand a suddenly applied load and is expressed in terms of
energy. Often measured with the Izod impact strength test or Charpy impact test, both of which measure the
impact energy required to fracture a sample. Volume, modulus of elasticity, distribution of forces, and yield
strength affect the impact strength of a material. In order for a material or object to have a high impact strength
the stresses must be distributed evenly throughout the object. It also must have a large volume with a low
modulus of elasticity and a high material yield strength.[8]

Strain (deformation) terms


Deformation of the material is the change in geometry created when stress is applied (as a result of applied
forces, gravitational fields, accelerations, thermal expansion, etc.). Deformation is expressed by the displacement
field of the material.[9]
Strain or reduced deformation is a mathematical term that expresses the trend of the deformation change among
the material field. Strain is the deformation per unit length.[10] In the case of uniaxial loading the displacements of
a specimen (for example a bar element) lead to a calculation of strain expressed as the quotient of the
displacement and the original length of the specimen. For 3D displacement fields it is expressed as derivatives of
displacement functions in terms of a second order tensor (with 6 independent elements).
Deflection is a term to describe the magnitude to which a structural element is displaced when subject to an
applied load.[11]

Stress–strain relations
Elasticity is the ability of a material to return to its previous shape after stress is released. In many materials, the
relation between applied stress is directly proportional to the resulting strain (up to a certain limit), and a graph

https://en.wikipedia.org/wiki/Strength_of_materials Page 3 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

representing those two quantities is a straight line.


The slope of this line is known as Young's modulus, or the
"modulus of elasticity." The modulus of elasticity can be
used to determine the stress–strain relationship in the
linear-elastic portion of the stress–strain curve. The linear-
elastic region is either below the yield point, or if a yield
point is not easily identified on the stress–strain plot it is
defined to be between 0 and 0.2% strain, and is defined as
the region of strain in which no yielding (permanent
deformation) occurs.[12]

Plasticity or plastic deformation is the opposite of


elastic deformation and is defined as unrecoverable
strain. Plastic deformation is retained after the release
of the applied stress. Most materials in the linear-
elastic category are usually capable of plastic Basic static response of a specimen under tension
deformation. Brittle materials, like ceramics, do not
experience any plastic deformation and will fracture
under relatively low strain, while ductile materials such
as metallics, lead, or polymers will plasticly deform much more before a fracture initiation.
Consider the difference between a carrot and chewed bubble gum. The carrot will stretch very little before breaking.
The chewed bubble gum, on the other hand, will plastically deform enormously before finally breaking.

Design terms
Ultimate strength is an attribute related to a material, rather than just a specific specimen made of the material, and
as such it is quoted as the force per unit of cross section area (N/m2). The ultimate strength is the maximum stress
that a material can withstand before it breaks or weakens.[13] For example, the ultimate tensile strength (UTS) of AISI
1018 Steel is 440 MN/m2. In general, the SI unit of stress is the pascal, where 1 Pa = 1 N/m2. In Imperial units, the
unit of stress is given as lbf/in² or pounds-force per square inch. This unit is often abbreviated as psi. One thousand
psi is abbreviated ksi.

A factor of safety is a design criteria that an engineered component or structure must achieve. , where
FS: the factor of safety, R: The applied stress, and UTS: ultimate stress (psi or N/m2)[14]

Margin of Safety is also sometimes used to as design criteria. It is defined MS = Failure Load/(Factor of Safety ×
Predicted Load) − 1.

For example, to achieve a factor of safety of 4, the allowable stress in an AISI 1018 steel component can be calculated
to be = 440/4 = 110 MPa, or = 110×106 N/m2. Such allowable stresses are also known as "design
stresses" or "working stresses."

Design stresses that have been determined from the ultimate or yield point values of the materials give safe and
reliable results only for the case of static loading. Many machine parts fail when subjected to a non steady and
continuously varying loads even though the developed stresses are below the yield point. Such failures are called
fatigue failure. The failure is by a fracture that appears to be brittle with little or no visible evidence of yielding.
However, when the stress is kept below "fatigue stress" or "endurance limit stress", the part will endure indefinitely. A

https://en.wikipedia.org/wiki/Strength_of_materials Page 4 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

purely reversing or cyclic stress is one that alternates between equal positive and negative peak stresses during each
cycle of operation. In a purely cyclic stress, the average stress is zero. When a part is subjected to a cyclic stress, also
known as stress range (Sr), it has been observed that the failure of the part occurs after a number of stress reversals
(N) even if the magnitude of the stress range is below the material’s yield strength. Generally, higher the range stress,
the fewer the number of reversals needed for failure.

Failure theories
There are four failure theories: maximum shear stress theory, maximum normal stress theory, maximum strain
energy theory, and maximum distortion energy theory. Out of these four theories of failure, the maximum normal
stress theory is only applicable for brittle materials, and the remaining three theories are applicable for ductile
materials. Of the latter three, the distortion energy theory provides most accurate results in majority of the stress
conditions. The strain energy theory needs the value of Poisson's ratio of the part material, which is often not readily
available. The maximum shear stress theory is conservative. For simple unidirectional normal stresses all theories are
equivalent, which means all theories will give the same result.

Maximum Shear Stress Theory — This theory postulates that failure will occur if the magnitude of the maximum
shear stress in the part exceeds the shear strength of the material determined from uniaxial testing.
Maximum Normal Stress Theory — This theory postulates that failure will occur if the maximum normal stress
in the part exceeds the ultimate tensile stress of the material as determined from uniaxial testing. This theory
deals with brittle materials only. The maximum tensile stress should be less than or equal to ultimate tensile
stress divided by factor of safety. The magnitude of the maximum compressive stress should be less than
ultimate compressive stress divided by factor of safety.
Maximum Strain Energy Theory — This theory postulates that failure will occur when the strain energy per unit
volume due to the applied stresses in a part equals the strain energy per unit volume at the yield point in uniaxial
testing.
Maximum Distortion Energy Theory — This theory is also known as shear energy theory or von Mises-Hencky
theory. This theory postulates that failure will occur when the distortion energy per unit volume due to the applied
stresses in a part equals the distortion energy per unit volume at the yield point in uniaxial testing. The total
elastic energy due to strain can be divided into two parts: one part causes change in volume, and the other part
causes change in shape. Distortion energy is the amount of energy that is needed to change the shape.
Fracture mechanics was established by Alan Arnold Griffith and George Rankine Irwin. This important theory is
also known as numeric conversion of toughness of material in the case of crack existence.
Fractology was proposed by Takeo Yokobori because each fracture laws including creep rupture criterion must be
combined nonlinearly.
A material's strength is dependent on its microstructure. The engineering processes to which a material is subjected
can alter this microstructure. The variety of strengthening mechanisms that alter the strength of a material includes
work hardening, solid solution strengthening, precipitation hardening, and grain boundary strengthening and can be
quantitatively and qualitatively explained. Strengthening mechanisms are accompanied by the caveat that some other
mechanical properties of the material may degenerate in an attempt to make the material stronger. For example, in
grain boundary strengthening, although yield strength is maximized with decreasing grain size, ultimately, very small
grain sizes make the material brittle. In general, the yield strength of a material is an adequate indicator of the
material's mechanical strength. Considered in tandem with the fact that the yield strength is the parameter that
predicts plastic deformation in the material, one can make informed decisions on how to increase the strength of a
material depending its microstructural properties and the desired end effect. Strength is expressed in terms of the
limiting values of the compressive stress, tensile stress, and shear stresses that would cause failure. The effects of
dynamic loading are probably the most important practical consideration of the strength of materials, especially the

https://en.wikipedia.org/wiki/Strength_of_materials Page 5 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

problem of fatigue. Repeated loading often initiates brittle cracks, which grow until failure occurs. The cracks always
start at stress concentrations, especially changes in cross-section of the product, near holes and corners at nominal
stress levels far lower than those quoted for the strength of the material.

See also
Creep (deformation)
Deformation mechanism map
Dynamics
Fatigue (material)
Forensic engineering
Fracture mechanics
Fracture toughness
List of materials properties#Mechanical properties
Material selection
Molecular diffusion
Specific strength
Statics
Universal testing machine

References
1. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 210. ISBN 978-0-07-352938-7.
2. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 7. ISBN 978-0-07-352938-7.
3. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 5. ISBN 978-0-07-352938-7.
4. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. pp. 9–10. ISBN 978-0-07-352938-7.
5. Kokcharov I. Strength of Structural Materials http://www.kokch.kts.ru/me/t3/SIA_3_Mechanical_Properties.pdf
6. Mechanics of Materials (5th ed.). p. 52. ISBN 978-0-07-352938-7.
7. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 60. ISBN 978-0-07-352938-7.
8. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. pp. 693–696. ISBN 978-0-07-352938-7.
9. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 47. ISBN 978-0-07-352938-7.
10. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 49. ISBN 978-0-07-352938-7.
11. R. C. Hibbeler (2009). Structural Analysis (7 ed.). Pearson Prentice Hall. p. 305. ISBN 978-0-13-602060-8.
12. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. pp. 53–56. ISBN 978-0-07-352938-7.
13. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. pp. 27–28. ISBN 978-0-07-352938-7.
14. Beer & Johnston (2006). Mechanics of Materials (5th ed.). McGraw Hill. p. 28. ISBN 978-0-07-352938-7.

Further reading
Fa-Hwa Cheng, Initials. (1997). Strength of material. Ohio: McGraw-Hill
Mechanics of Materials, E.J. Hearn
Alfirević, Ivo. Strength of Materials I. Tehnička knjiga, 1995. ISBN 953-172-010-X.
Alfirević, Ivo. Strength of Materials II. Tehnička knjiga, 1999. ISBN 953-6168-85-5.
Ashby, M.F. Materials Selection in Design. Pergamon, 1992.

https://en.wikipedia.org/wiki/Strength_of_materials Page 6 of 7
Strength of materials - Wikipedia 31/01/18, 6)11 PM

Beer, F.P., E.R. Johnston, et al. Mechanics of Materials, 3rd edition. McGraw-Hill, 2001. ISBN 0-07-248673-2
Cottrell, A.H. Mechanical Properties of Matter. Wiley, New York, 1964.
Den Hartog, Jacob P. Strength of Materials. Dover Publications, Inc., 1961, ISBN 0-486-60755-0.
Drucker, D.C. Introduction to Mechanics of Deformable Solids. McGraw-Hill, 1967.
Gordon, J.E. The New Science of Strong Materials. Princeton, 1984.
Groover, Mikell P. Fundamentals of Modern Manufacturing, 2nd edition. John Wiley & Sons,Inc., 2002. ISBN 0-
471-40051-3.
Hashemi, Javad and William F. Smith. Foundations of Materials Science and Engineering, 4th edition. McGraw-
Hill, 2006. ISBN 0-07-125690-3.
Hibbeler, R.C. Statics and Mechanics of Materials, SI Edition. Prentice-Hall, 2004. ISBN 0-13-129011-8.
Lebedev, Leonid P. and Michael J. Cloud. Approximating Perfection: A Mathematician's Journey into the World of
Mechanics. Princeton University Press, 2004. ISBN 0-691-11726-8.
Chapter 10 – Strength of Elastomers (https://dx.doi.org/10.1016/B978-0-12-394584-6.00010-8), A.N. Gent, W.V.
Mars, In: James E. Mark, Burak Erman and Mike Roland, Editor(s), The Science and Technology of Rubber
(Fourth Edition), Academic Press, Boston, 2013, Pages 473-516, ISBN 9780123945846, 10.1016/B978-0-12-
394584-6.00010-8
Mott, Robert L. Applied Strength of Materials, 4th edition. Prentice-Hall, 2002. ISBN 0-13-088578-9.
Popov, Egor P. Engineering Mechanics of Solids. Prentice Hall, Englewood Cliffs, N. J., 1990. ISBN 0-13-279258-
3.
Ramamrutham, S. Strength of Materials.
Shames, I.H. and F.A. Cozzarelli. Elastic and inelastic stress analysis. Prentice-Hall, 1991. ISBN 1-56032-686-7.
Timoshenko S. Strength of Materials, 3rd edition. Krieger Publishing Company, 1976, ISBN 0-88275-420-3.
Timoshenko, S.P. and D.H. Young. Elements of Strength of Materials, 5th edition. (MKS System)
Davidge, R.W., Mechanical Behavior of Ceramics, Cambridge Solid State Science Series, (1979)
Lawn, B.R., Fracture of Brittle Solids, Cambridge Solid State Science Series, 2nd Edn. (1993)
Green, D., An Introduction to the Mechanical Properties of Ceramics, Cambridge Solid State Science Series,
Eds. Clarke, D.R., Suresh, S., Ward, I.M.Babu Tom.K (1998)

External links
Failure theories (http://www.mech.uwa.edu.au/DANotes/SSS/failure/theories.html)
Case studies in structural failure (http://materials.open.ac.uk/mem/index.htm)

Retrieved from "https://en.wikipedia.org/w/index.php?title=Strength_of_materials&oldid=821948337"

This page was last edited on 23 January 2018, at 15:17.

Text is available under the Creative Commons Attribution-ShareAlike License; additional terms may apply. By using
this site, you agree to the Terms of Use and Privacy Policy. Wikipedia® is a registered trademark of the Wikimedia
Foundation, Inc., a non-profit organization.

https://en.wikipedia.org/wiki/Strength_of_materials Page 7 of 7

You might also like