Download as docx, pdf, or txt
Download as docx, pdf, or txt
You are on page 1of 17

In continuum mechanics, stress is a physical quantity that expresses the internal forces that

neighboring particles of a continuous materialexert on each other. For example, when a solid vertical
bar is supporting a weight, each particle in the bar pulls on the particles immediately above and
below it. When a liquid is under pressure, each particle gets pushed inwards by all the surrounding
particles, and, in reaction, pushes them outwards. These macroscopic forces are actually the
average of a very large number of intermolecular forces and collisionsbetween the particles in
those molecules.
Stress inside a material may arise by various mechanisms, such as reaction to external forces
applied to the bulk material (like gravity) or to its surface (like contact forces, external pressure,
or friction). Any strain (deformation) of a solid material generates an internal elastic stress,
analogous to the reaction force of a spring, that tends to restore the material to its original non-
deformed state. In liquids and gases, only deformations that change the volume generate persistent
elastic stress. However, if the deformation is gradually changing with time, even in fluids there will
usually be some viscous stress, opposing that change. Elastic and viscous stresses are usually
combined under the namemechanical stress.
Significant stress may exist even when deformation is negligible or non-existent (a common
assumption when modeling the flow of water). Stress may exist in the absence of external forces;
such built-in stress is important, for example, in prestressed concrete and tempered glass. Stress
may also be imposed on a material without the application of net forces, for example by changes in
temperature or chemical composition, or by external electromagnetic fields (as
in piezoelectric and magnetostrictive materials).
The relation between mechanical stress, deformation, and the rate of change of deformation can be
quite complicated, although a linear approximation may be adequate in practice if the quantities are
small enough. Stress that exceeds certain strength limits of the material will result in permanent
deformation (such as plastic flow, fracture, cavitation) or even change its crystal
structure and chemical composition.
In some branches of engineering, the term stress is occasionally used in a looser sense as a
synonym of "internal force". For example, in the analysis of trusses, it may refer to the total traction
or compression force acting on a beam, rather than the force divided by the area of its cross-section.
Contents
[hide]
1 History
2 Overview
3 Simple stresses
4 General stress
5 Stress analysis
6 Alternative measures of stress
7 See also
8 Further reading
9 References
History[edit]

Roman-era bridge in Switzerland

Inca bridge on the Apurimac River
Since ancient times humans have been consciously aware of stress inside materials. Until the 17th
century the understanding of stress was largely intuitive and empirical; and yet it resulted in some
surprisingly sophisticated technology, like the composite bow and glass blowing.
[citation needed]

Over several millennia, architects and builders, in particular, learned how to put together carefully
shaped wood beams and stone blocks to withstand, transmit, and distribute stress in the most
effective manner, with ingenious devices such as the capitals, arches, cupolas,trusses and the flying
buttresses of Gothic cathedrals.
Ancient and medieval architects did develop some geometrical methods and simple formulas to
compute the proper sizes of pillars and beams, but the scientific understanding of stress became
possible only after the necessary tools were invented in the 17th and 18th centuries: Galileo's
rigorous experimental method, Descartes's coordinates and analytic geometry, and Newton's laws of
motion and equilibrium and calculus of infinitesimals.
[1]
With those tools, Cauchy was able to give the
first rigorous and general mathematical model for stress in a homogeneous medium.
[citation
needed]
Cauchy observed that the force across an imaginary surface was a linear function of its normal
vector; and, moreover, that it must be a symmetric function (with zero total momentum).
[citation needed]

The understanding of stress in liquids started with Newton, who provided a differential formula for
friction forces (shear stress) in parallellaminar flow.
Overview[edit]
Definition[edit]
Stress is defined as the average force per unit area that some particle of a body exerts on an
adjacent particle, across an imaginary surface that separates them.
[2]

Being derived from a fundamental physical quantity (force) and a purely geometrical quantity (area),
stress is also a fundamental quantity, like velocity, torque or energy, that can be quantified and
analyzed without explicit consideration of the nature of the material or of its physical causes.
Following the basic premises of continuum mechanics, stress is a macroscopic concept. Namely, the
particles considered in its definition and analysis should be just small enough to be treated as
homogeneous in composition and state, but still large enough to ignore quantum effects and the
detailed motions of molecules. Thus, the force between two particles is actually the average of a
very large number of atomic forces between their molecules; and physical quantities like mass,
velocity, and forces that act through the bulk of three-dimensional bodies, like gravity, are assumed
to be smoothly distributed over them.
[3]:p.90106
Depending on the context, one may also assume that
the particles are large enough to allow the averaging out of other microscopic features, like the
grains of a metal rod or the fibers of a piece of wood.

The stress across a surface element (yellow disk) is the force that the material on one side (top ball) exerts on the
material on the other side (bottom ball), divided by the area of the surface.
Quantitatively, the stress is expressed by the Cauchy traction vector T defined as the traction
force F between adjacent parts of the material across an imaginary separating surface S, divided by
the area of S.
[4]:p.4150
In a fluid at rest the force is perpendicular to the surface, and is the
familiar pressure. In a solid, or in a flow of viscous liquid, the force F may not be perpendicular to S;
hence the stress across a surface must be regarded a vector quantity, not a scalar. Moreover, the
direction and magnitude generally depend on the orientation of S. Thus the stress state of the
material must be described by a tensor, called the (Cauchy) stress tensor; which is a linear
function that relates the normal vector n of a surface S to the stress T across S. With respect to any
chosen coordinate system, the Cauchy stress tensor can be represented as a symmetric matrix of
3x3 real numbers. Even within a homogeneous body, the stress tensor may vary from place to
place, and may change over time; therefore, the stress within a material is, in general, a time-
varying tensor field.
Normal and shear stress[edit]
Further information: compression (physical) and Shear stress
In general, the stress T that a particle P applies on another particle Q across a surface S can have
any direction relative to S. The vector Tmay be regarded as the sum of two components: the normal
stress (compression or tension) perpendicular to the surface, and the shear stress that is parallel to
the surface.
If the normal unit vector n of the surface (pointing from Q towards P) is assumed fixed, the normal
component can be expressed by a single number, the dot product T n. This number will be positive
if P is "pulling" on Q (tensile stress), and negative if P is "pushing" against Q(compressive stress)
The shear component is then the vector T (T n)n.
Units[edit]
The dimension of stress is that of pressure, and therefore its coordinates are commonly measured in
the same units as pressure: namely, pascals (Pa, that is, newtons persquare metre) in
the International System, or pounds per square inch (psi) in the Imperial system.
Causes and effects[edit]

Glass vase with thecraquel effect. The cracks are the result of brief but intense stress created when the semi-molten
piece is briefly dipped in water.
[5]

Stress in a material body may be due to multiple physical causes, including external influences and
internal physical processes. Some of these agents (like gravity, changes in temperature and phase,
and electromagnetic fields) act on the bulk of the material, varying continuously with position and
time. Other agents (like external loads and friction, ambient pressure, and contact forces) may
create stresses and forces that are concentrated on certain surfaces, lines, or points; and possibly
also on very short time intervals (as in the impulses due to collisions). In general, the stress
distribution in the body is expressed as a piecewise continuous function of space and time.
Conversely, stress is usually correlated with various effects on the material, possibly including
changes in physical properties like birefringence,polarization, and permeability. The imposition of
stress by an external agent usually creates some strain (deformation) in the material, even if it is too
small to be detected. In a solid material, such strain will in turn generate an internal elastic stress,
analogous to the reaction force of a stretched spring, tending to restore the material to its original
undeformed state. Fluid materials (liquids, gases and plasmas) by definition can only oppose
deformations that would change their volume. However, if the deformation is changing with time,
even in fluids there will usually be some viscous stress, opposing that change.
The relation between stress and its effects and causes, including deformation and rate of change of
deformation, can be quite complicated (although a linear approximation may be adequate in practice
if the quantities are small enough). Stress that exceeds certain strength limits of the material will
result in permanent deformation (such as plastic flow, fracture, cavitation) or even change its crystal
structure and chemical composition.
Simple stresses[edit]
In some situations, the stress within a body may adequately be described by a single number, or by
a single vector (a number and a direction). Three such simple stresssituations, that are often
encountered in engineering design, are the uniaxial normal stress, the simple shear stress, and
the isotropic normal stress.
[6]

Uniaxial normal stress[edit]

Idealized stress in a straight bar with uniform cross-section.
A common situation with a simple stress pattern is when a straight rod, with uniform material and
cross section, is subjected to tensionby opposite forces of magnitude along its axis. If the system
is in equilibrium and not changing with time, and the weight of the bar can be neglected, then
through each transversal section of the bar the top part must pull on the bottom part with the same
force F. Therefore the stress throughout the bar, across any horizontal surface, can be described by
the number = F/A, where A is the area of the cross-section.
On the other hand, if one imagines the bar being cut along its length, parallel to the axis, there will
be no force (hence no stress) between the two halves across the cut.
This type of stress may be called (simple) normal stress or uniaxial stress; specifically,
(uniaxial, simple, etc.) tensile stress.
[6]
If the load is compression on the bar, rather than stretching
it, the analysis is the same except that the force F and the stress change sign, and the stress is
called compressive stress.

The ratio may be only an average stress. The stress may be unevenly distributed over the cross
section (mm), especially near the attachment points (nn).
This analysis assumes the stress is evenly distributed over the entire cross-section. In practice,
depending on how the bar is attached at the ends and how it was manufactured, this assumption
may not be valid. In that case, the value = F/A will be only the average stress, called engineering
stress ornominal stress. However, if the bar's length L is many times its diameter D, and it has no
gross defects or built-in stress, then the stress can be assumed to be uniformly distributed over any
cross-section that is more than a few times D from both ends. (This observation is known as
the Saint-Venant's principle).
Normal stress occurs in many other situations besides axial tension and compression. If an elastic
bar with uniform and symmetric cross-section is bent in one of its planes of symmetry, the
resulting bending stress will still be normal (perpendicular to the cross-section), but will vary over
the cross section: the outer part will be under tensile stress, while the inner part will be compressed.
Another variant of normal stress is the hoop stress that occurs on the walls of a
cylindrical pipe or vessel filled with pressurized fluid.
Simple shear stress[edit]

Shear stress in a horizontal bar loaded by two offset blocks.
Another simple type of stress occurs when a uniformly thick layer of elastic material like glue or
rubber is firmly attached to two stiff bodies that are pulled in opposite directions by forces parallel to
the layer; or a section of a soft metal bar that is being cut by the jaws of ascissors-like tool. Let F be
the magnitude of those forces, and M be the midplane of that layer. Just as in the normal stress
case, the part of the layer on one side of M must pull the other part with the same force F. Assuming
that the direction of the forces is known, the stress across M can be expressed by the single
number = F/A, where F is the magnitude of those forces and A is the area of the layer.
However, unlike normal stress, this simple shear stress is directed parallel to the cross-section
considered, rather than perpendicular to it.
[6]
For any plane S that is perpendicular to the layer, the
net internal force across S, and hence the stress, will be zero.
As in the case of an axially loaded bar, in practice the shear stress may not be uniformly distributed
over the layer; so, as before, the ratio F/A will only be an average ("nominal", "engineering") stress.
However, that average is often sufficient for practical purposes.
[7]:p.292
Shear stress is observed also
when a cylindrical bar such as a shaft is subjected to opposite torques at its ends. In that case, the
shear stress on each cross-section is parallel to the cross-section, but oriented tangentially relative
to the axis, and increases with distance from the axis. Significant shear stress occurs in the middle
plate (the "web") of I-beams under bending loads, due to the web constraining the end plates
("flanges").
Isotropic stress[edit]

Isotropic tensile stress. Top left: Each face of a cube of homogeneous material is pulled by a force with magnitude F,
applied evenly over the entire face whose area is A. The force across any section S of the cube must balance the
forces applied below the section. In the three sections shown, the forces are F (top right), F (bottom left), and F
(bottom right); and the area of S is A, A and A , respectively. So the stress across S is F/Ain all
three cases.
Another simple type of stress occurs when the material body is under equal compression or tension
in all directions. This is the case, for example, in a portion of liquid or gas at rest, whether enclosed
in some container or as part of a larger mass of fluid; or inside a cube of elastic material that is being
pressed or pulled on all six faces by equal perpendicular forces provided, in both cases, that the
material is homogeneous, without built-in stress, and that the effect of gravity and other external
forces can be neglected.
In these situations, the stress across any imaginary internal surface turns out to be equal in
magnitude and always directed perpendicularly to the surface independently of the surface's
orientation. This type of stress may be called isotropic normal or justisotropic; if it is compressive,
it is called hydrostatic pressure or just pressure. Gases by definition cannot withstand tensile
stresses, but liquids may withstand very small amounts of isotropic tensile stress.
Cylinder stresses[edit]
Parts with rotational symmetry, such as wheels, axles, pipes, and pillars, are very common in
engineering. Often the stress patterns that occur in such parts have rotational or even cylindrical
symmetry. The analysis of such cylinder stresses can take advantage of the symmetry to reduce the
dimension of the domain and/or of the stress tensor.
General stress[edit]
Often, mechanical bodies experience more than one type of stress at the same time; this is
called combined stress. In normal and shear stress, the magnitude of the stress is maximum for
surfaces that are perpendicular to a certain direction , and zero across any surfaces that are
parallel to . When the stress is zero only across surfaces that are perpendicular to one particular
direction, the stress is called biaxial, and can be viewed as the sum of two normal or shear stresses.
In the most general case, called triaxial stress, the stress is nonzero across every surface element.
The Cauchy stress tensor[edit]
Main article: Cauchy stress tensor

Components of stress in three dimensions

Illustration of typical stresses (arrows) across various surface elements on the boundary of a particle (sphere), in a
homogeneous material under uniform (but not isotropic) triaxial stress. The normal stresses on the principal axes are
+5, +2, and 3 units.
Combined stresses cannot be described by a single vector. Even if the material is stressed in the
same way throughout the volume of the body, the stress across any imaginary surface will depend
on the orientation of that surface, in a non-trivial way.
However, Cauchy observed that the stress vector across a surface will always be alinear
function of the surface's normal vector , the unit-length vector that is perpendicular to it. That
is, , where the function satisfies

for any vectors and any real numbers . The function , now called the(Cauchy)
stress tensor, completely describes the stress state of a uniformly stressed body. (Today, any
linear connection between two physical vector quantities is called atensor, reflecting Cauchy's
original use to describe the "tensions" (stresses) in a material.) In tensor calculus, is classified
as second-order tensor of type (0,2).
Like any linear map between vectors, the stress tensor can be represented in any
chosen Cartesian coordinate system by a 33 matrix of real numbers. Depending on whether
the coordinates are numbered or named , the matrix may be written as
or
The stress vector across a surface with normal vector with
coordinates is then a matrix product , that is

The linear relation between and follows from the fundamental laws of conservation
of linear momentum and static equilibrium of forces, and is therefore mathematically
exact, for any material and any stress situation. The components of the Cauchy stress
tensor at every point in a material satisfy the equilibrium equations (Cauchys equations
of motion for zero acceleration). Moreover, the principle of conservation of angular
momentum implies that the stress tensor is symmetric, that is ,
, and . Therefore, the stress state of the medium at any point and instant
can be specified by only six independent parameters, rather than nine. These may be
written

where the elements are called the orthogonal normal
stresses (relative to the chosen coordinate system),
and the orthogonal shear stresses.
Change of coordinates[edit]
The Cauchy stress tensor obeys the tensor transformation law under a change in
the system of coordinates. A graphical representation of this transformation law is
the Mohr's circle of stress distribution.
As a symmetric 33 real matrix, the stress tensor has three mutually orthogonal
unit-length eigenvectors and three real eigenvalues , such
that . Therefore, in a coordinate system with axes , the
stress tensor is a diagonal matrix, and has only the three normal
components the principal stresses. If the three eigenvalues are equal,
the stress is an isotropic compression or tension, always perpendicular to any
surface; there is no shear stress, and the tensor is a diagonal matrix in any
coordinate frame.
Stress as a tensor field[edit]
In general, stress is not uniformly distributed over a material body, and may vary
with time. Therefore the stress tensor must be defined for each point and each
moment, by considering an infinitesimal particle of the medium surrounding that
point, and taking the average stresses in that particle as being the stresses at the
point.
Stress in thin plates[edit]

A tank car made from bent and welded steel plates.
Man-made objects are often made from stock plates of various materials by
operations that do not change their essentially two-dimensional character, like
cutting, drilling, gentle bending and welding along the edges. The description of
stress in such bodies can be simplified by modeling those parts as two-dimensional
surfaces rather than three-dimensional bodies.
In that view, one redefines a "particle" as being an infinitesimal patch of the plate's
surface, so that the boundary between adjacent particles becomes an infinitesimal
line element; both are implicitly extended in the third dimension, straight through the
plate. "Stress" is then redefined as being a measure of the internal forces between
two adjacent "particles" across their common line element, divided by the length of
that line. Some components of the stress tensor can be ignored, but since particles
are not infinitesimal in the third dimension one can no longer ignore the torque that
a particle applies on its neighbors. That torque is modeled as a bending stressthat
tends to change the curvature of the plate. However, these simplifications may not
hold at welds, at sharp bends and creases (where the radius of curvature is
comparable to the thickness of the plate).
Stress in thin beams[edit]

For stress modeling, a fishing pole may be considered one-dimensional.
The analysis of stress can be considerably simplified also for thin bars, beams or
wires of uniform (or smoothly varying) composition and cross-section that are
subjected to moderate bending and twisting. For those bodies may consider only
cross-sections that are perpendicular to the bar's axis, and redefine a "particle" as
being a piece of wire with infinitesimal length between two such cross sections. The
ordinary stress is then reduced to a scalar (tension or compression of the bar), but
one must take into account also a bending stress (that tries to change the bar's
curvature, in some direction perpendicular to the axis) and a torsional stress (that
tries to twist or un-twist it about its axis).
Other descriptions of stress[edit]
The Cauchy stress tensor is used for stress analysis of material bodies
experiencing small deformations where the differences in stress distribution in most
cases can be neglected. For large deformations, also called finite deformations,
other measures of stress, such as the first and second PiolaKirchhoff stress
tensors, the Biot stress tensor, and the Kirchhoff stress tensor, are required.
Solids, liquids, and gases have stress fields. Static fluids support normal stress but
will flow under shear stress. Moving viscous fluids can support shear stress
(dynamic pressure). Solids can support both shear and normal stress,
with ductile materials failing under shear and brittle materials failing under normal
stress. All materials have temperature dependent variations in stress-related
properties, and non-Newtonian materials have rate-dependent variations.
Stress analysis[edit]
Stress analysis is a branch of applied physics that covers the determination of the
internal distribution of stresses in solid objects. It is an essential tool
in engineering for the study and design of structures such as tunnels, dams,
mechanical parts, and structural frames, under prescribed or expected loads. It is
also important in many other disciplines; for example, in geology, to study
phenomena like plate tectonics, vulcanism and avalanches; and in biology, to
understand the anatomy of living beings.
Goals and assumptions[edit]
Stress analysis is generally concerned with objects and structures that can be
assumed to be in macroscopic static equilibrium. By Newton's laws of motion, any
external forces are being applied to such a system must be balanced by internal
reaction forces,
[8]:p.97
which are almost always surface contact forces between
adjacent particles that is, as stress.
[4]
Since every particle needs to be in
equilibrium, this reaction stress will generally propagate from particle, creating a
stress distribution throughout the body.
The typical problem in stress analysis is to determine these internal stresses, given
the external forces that are acting on the system. The latter may be body
forces (such as gravity or magnetic attraction), that act throughout the volume of a
material;
[9]:p.4281
or concentrated loads (such as friction between an axle and
a bearing, or the weight of a train wheel on a rail), that are imagined to act over a
two-dimensional area, or along a line, or at single point.
In stress analysis one normally disregards the physical causes of the forces or the
precise nature of the materials. Instead, one assumes that the stresses are related
to deformation (and, in non-static problems, to the rate of deformation) of the
material by known constitutive equations.
[10]

Methods[edit]
Stress analysis may be carried out experimentally, by applying loads to the actual
artifact or to scale model, and measuring the resulting stresses, by any of several
available methods. This approach is often used for safety certification and
monitoring. However, most stress analysis is done by mathematical methods,
especially during design.
The basic stress analysis problem can be formulated by Euler's equations of
motion for continuous bodies (which are consequences of Newton's laws for
conservation of linear momentum and angular momentum) and the Euler-Cauchy
stress principle, together with the appropriate constitutive equations. Thus one
obtains a system of partial differential equations involving the stress tensor field and
the strain tensor field, as unknown functions to be determined. The external body
forces appear as the independent ("right-hand side") term in the differential
equations, while the concentrated forces appear as boundary conditions. The basic
stress analysis problem is therefore a boundary-value problem.
Stress analysis for elastic structures is based on the theory of
elasticity and infinitesimal strain theory. When the applied loads cause permanent
deformation, one must use more complicated constitutive equations, that can
account for the physical processes involved (plastic flow, fracture, phase change,
etc.).
However, engineered structures are usually designed so that the maximum
expected stresses are well within the range of linear elasticity (the generalization
of Hookes law for continuous media); that is, the deformations caused by internal
stresses are linearly related to them. In this case the differential equations that
define the stress tensor are linear, and the problem becomes much easier. For one
thing, the stress at any point will be a linear function of the loads, too. For small
enough stresses, even non-linear systems can usually be assumed to be linear.

Simplified model of a truss for stress analysis, assuming unidimensional elements under
uniform axial tension or compression.
Stress analysis is simplified when the physical dimensions and the distribution of
loads allow the structure to be treated as one- or two-dimensional. In the analysis of
trusses, for example, the stress field may be assumed to be uniform and uniaxial
over each member. Then the differential equations reduce to a finite set of
equations (usually linear) with finitely many unknowns. In other contexts one may be
able to reduce the three-dimensional problem to a two-dimensional one, and/or
replace the general stress and strain tensors by simpler models like uniaxial
tension/compression, simple shear, etc.
Still, for two- or three-dimensional cases one must solve a partial differential
equation problem. Analytical or closed-form solutions to the differential equations
can be obtained when the geometry, constitutive relations, and boundary conditions
are simple enough. Otherwise one must generally resort to numerical
approximations such as the finite element method, the finite difference method, and
the boundary element method.
Alternative measures of stress[edit]
Main article: Stress measures
Other useful stress measures include the first and second PiolaKirchhoff stress
tensors, the Biot stress tensor, and the Kirchhoff stress tensor.
PiolaKirchhoff stress tensor[edit]
In the case of finite deformations, the PiolaKirchhoff stress tensors express the
stress relative to the reference configuration. This is in contrast to the Cauchy stress
tensorwhich expresses the stress relative to the present configuration. For
infinitesimal deformations and rotations, the Cauchy and PiolaKirchhoff tensors are
identical.
Whereas the Cauchy stress tensor relates stresses in the current configuration,
the deformation gradient and strain tensors are described by relating the motion to
the reference configuration; thus not all tensors describing the state of the material
are in either the reference or current configuration. Describing the stress, strain and
deformation either in the reference or current configuration would make it easier to
define constitutive models (for example, the Cauchy Stress tensor is variant to a
pure rotation, while the deformation strain tensor is invariant; thus creating problems
in defining a constitutive model that relates a varying tensor, in terms of an invariant
one during pure rotation; as by definition constitutive models have to be invariant to
pure rotations). The 1st PiolaKirchhoff stress tensor, is one possible solution to
this problem. It defines a family of tensors, which describe the configuration of the
body in either the current or the reference state.
The 1st PiolaKirchhoff stress tensor, relates forces in the present configuration
with areas in the reference ("material") configuration.

where is the deformation gradient and is
the Jacobian determinant.
In terms of components with respect to an orthonormal basis, the first Piola
Kirchhoff stress is given by

Because it relates different coordinate systems, the 1st PiolaKirchhoff
stress is a two-point tensor. In general, it is not symmetric. The 1st Piola
Kirchhoff stress is the 3D generalization of the 1D concept of engineering
stress.
If the material rotates without a change in stress state (rigid rotation), the
components of the 1st PiolaKirchhoff stress tensor will vary with material
orientation.
The 1st PiolaKirchhoff stress is energy conjugate to the deformation
gradient.
2nd PiolaKirchhoff stress tensor[edit]
Whereas the 1st PiolaKirchhoff stress relates forces in the current
configuration to areas in the reference configuration, the 2nd Piola
Kirchhoff stress tensor relates forces in the reference configuration to
areas in the reference configuration. The force in the reference
configuration is obtained via a mapping that preserves the relative
relationship between the force direction and the area normal in the
reference configuration.

In index notation with respect to an orthonormal basis,

This tensor, a one-point tensor, is symmetric.
If the material rotates without a change in stress state (rigid
rotation), the components of the 2nd PiolaKirchhoff stress tensor
remain constant, irrespective of material orientation.
The 2nd PiolaKirchhoff stress tensor is energy conjugate to
the GreenLagrange finite strain tensor.

You might also like