You are on page 1of 28

Marius. N. Baba.

, (2019) Synopsis of lectures in Strength of


Materials for Undergraduates. Part. 2, Transilvania University Press,
Brașov, România, ISBN 978-606-19-1160-8.

11 Theories of elastic failure

11.1 Introduction

When the loads are applied to a machine component or mechanical structure, let us say
statically for the sake of simplicity, and whether their magnitude is kept increasing, at some
point, the material will fail. Thus, the questions that may arise are how failure can be predicted
and what level the stresses or strains in the component or structure need to reach for it to fail?
To answer these questions, it is first necessary to define what precisely failure means in the
context of mechanics of materials.

Figure 11-1

It is well known that for ductile materials, failure is usually considered to occur at the
onset of plastic deformation, while for brittle materials, it is taking place by fracture1. These two
distinct failure modes are relatively easy to establish for a uniaxial stress state typically carried
out in a standard tensile test where, as the load increases, the normal stress in the specimen
is reaching the yield strength in case of ductile materials or the ultimate strength for brittle
materials. On the contrary, predicting failure is much less straightforward for a more complex

1
In what follows, the term failure may mean either yielding or fracture, whichever is occurring first. Failure due to variable
loading (i.e., mechanical fatigue), elastic instability, or local buckling shall not be regarded in this chapter.

5
Section 11. Theories of elastic failure

case of multiaxial loading, and this difficulty comes out because, so far, a universally applicable
method has not been evolved. Thus, to accurately predict the failure of machine components
or structural elements under three or two-dimensional stress states, one needs to select the
most suitable criterion from a range of different theories. Each of them is experimentally
proven to provide reliable results under certain well-defined circumstances. As illustrated in
Figure 11-1, since the ductile materials fail in fundamentally different ways than the brittle
ones, failure theories that are adequate for the former do not apply to the latter and vice-
versa2.
With all the above considerations in mind, one may wonder what does the failure theories
do? The answer ought to be reasonable both from a theoretical but also from experimental
perspective, i.e., they allow predicting the failure by comparing the stress state at the critical
points of a component or structure carrying a complex multiaxial loading with a limit stress
value that can be obtained through simple standard uniaxial tests, such as the yield or the
ultimate strength of the material in question. A convenient way to put in practice this definition
is to introduce the concept of equivalent stress:

σeq = σeq (σx ; σy ; σz ; τxy ; σyz ; σzx ) , (11.1)

which represents a fictitious uniaxial stress function that causes the same effect on the
material as done by an actual multiaxial loading when several or even all components of stress
tensor are involved. Since the materials react differently to the applied loads, such functions
are established phenomenologically and may take various expressions. However, as long as
the behavior of isotropic materials is the same in any direction, the principal stresses can
always be used as a reference, which simplifies the handling of different functions3 defining the
failure theories. From this point of view, the equivalent stress may be thought of as a function
of principal stress components that must reach the value of limit stress 4 in simple tension for

2
Whereas the ductile materials typically experience yielding about a plane inclined by approximatively 45 degrees to the loading
direction when undergoing tensile, the brittle materials fracture about a fairly cross-section relative to the their longitudinal
axis. The reverse is found to be in the case of pure shear loading. Moreover, under uniaxial compression, the yield plane is
almost never reported for ductile materials as the specimens exhibit a higher resistance after the onset of yielding, and thus
no ultimate compression strength is to be obtained. On the other hand, in the case of brittle materials undergoing compression,
the fracture angle is generally lower than 45 degrees relative to the loading axis [13].
3
To be physically meaningful in practical applications, such functions must be invariant relative to any coordinate system.
4
For the purposes of this text, in the case of ductile materials, the limit stress is taken to be equal with the yield strength
(which is generally the same in tension and compression). This is because ductile materials are considered to have failed when
they start to yield. Brittle materials, do not yield, but they fracture instead, and thus, the limit stress is assumed to be equal
with their ultimate strength in tension or compression (which are typically more or less different).

6
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

failure to occur:
SY ;
σeq = σeq (σ1 ; σ2 ; σ3 ) = { (11.2)
Sut or Suc ,

where the limit stress is expressed by the yield strength (SY ) in case of ductile materials, and
the ultimate strength in tension (Sut ) or compression (Suc ) for brittle materials.
One needs to be aware of the possible uncertainties when calculating the stresses at
specific points of mechanical components or structures under investigation. These are
commonly related to the nature and magnitude of the applied loads, material properties,
manufacturing processes, environmental conditions, and not the last, their geometry and
dimensions. Therefore, based on the expression (11.2), the strength condition (i.e., non-failure
condition) may be written by accounting for the safety factor nsf :

SY
with nsf >1 ;
nsf
σeq ≤ (11.3)
Sut Suc
or with nsf >1 ,
{ nsf nsf

which can be interpreted as a rebate in strength by a factor of nsf to account for uncertainties.
This chapter aims to present only the classical theories applicable to bodies made of
ductile or brittle materials, assumed to be isotropic and homogeneous, and whose failure is
considered to have taken place just at, if not slightly over, their elastic limits or when they begin
to deform plastically once the yield strength is reached.

11.2 Maximum principal stress failure theory5

This theory establishes that failure of a material under a state of complex loading occurs
when at any of its points, the maximum principal normal stress reaches a limit stress
determined through a simple standard uniaxial tensile or compression test. Data reported in
the literature show that it may be successfully applied to predict the occurrence of failure by
fracture for some materials subjected to static multiaxial loading. Namely, for brittle materials
having nearly the same ultimate strength under tension and compression, the maximum
principal stress theory leads to results that agree quite well with the experimental data (e.g.,
some glasses or some wrought materials as the fully hardened tool steels). Nevertheless, the
maximum principal stress theory provides a lower prediction accuracy of yield onset, and hence,
it is not recommended to be used for materials that typically fail by yielding.

5
In general it is linked to the names of Saint Venant – Rankine – Navier.

7
Section 11. Theories of elastic failure

Let us consider, for instance, a general stress state at a certain point of a generic body,
given by all the principal stress components σ1 , σ2 and σ3 , arranged in descending order, such
that σ1 ≥ σ2 ≥ σ3 . In line with the maximum principal stress theory, the condition of failure can
be mathematically formulated as:

σ1 = Sut ; (11.4)

or6

σ3 = Suc if σ3 ≤ 0 and |σ3 | ≥ σ1 , (11.5)

With the understanding that this theory mainly applies to materials exhibiting a brittle fracture
behavior, in the above two conditions, the limit stresses refer to the ultimate tensile strength
Sut , and the ultimate compression strength Suc , respectively.

Figure 11-2

Suppose the limit conditions expressed through the equations (11.4) and (11.5) are
represented in a plane coordinate system with the maximum principal stress taken on the
abscissa axis and the minimum principal stress on the ordinate. In that case, a rectangular
failure locus is obtained, as illustrated by the shaded surface area in Figure 11-2. That is, the
maximum principal stress theory predicts failure for any combination of σ1 and σ3 that lies
outside of this surface or upon its boundary lines. Instead, any state of stress lying inside the
surface represents a safely elastic stress state.

6
This is due to the fact that many cast brittle materials typically have a more or less uneven behaviour in tensile and
compression (i.e. the absolute value of ultimate compression strength is higher than their ultimate tensile strength).

8
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

If the particular case of a pure state of shear stress is considered7, as graphically


represented by the dotted line in Figure 11-2, it can be described by the equation:

σ1 = - σ3 = τmax . (11.6)

As stated in its definition, the maximum principal stress theory predicts failure when the
magnitude of first principal stress reaches the value of limit stress, in the case of brittle
materials typically expressed by their ultimate tensile strength, Sut . For this purpose, in what
concerns the validity of condition (11.6), the experimental data reported in the literature shows
that for some brittle materials, the ultimate shear strength is almost equal or even greater than
the value of ultimate tensile strength. Thus, for such materials, the use of maximum principal
stress theory can be considered reasonable. On the contrary, it no longer applies to ductile
materials whose shear strength is considerably lower than the value of tensile strength8.

Figure 11-3

The concept of Mohr's circle can be used to help illustrate the nature of maximum normal
stress theory. Thus, Figure 11-3 displays the three Mohr's circles, which may result due to a
given combination of principal stresses σ1 , σ2 and σ3 , customarily arranged so that σ1 is
maximum, and σ3 is minimum. The largest one (i.e., shaded) corresponds to the critical plane
where the highest stress (in absolute value) acts. On the same plot, with dashed-lines are

7
Such a stress state is specific to circular or annular shafts subjected to torsion.
8
The experimental evidence shows that for such materials, failure takes place much earlier; e.g., when τmax equals about 0,55
to 0,6 of SY , with an average value of 0,57SY .

9
Section 11. Theories of elastic failure

drawn the Mohr's circles associated with simple uniaxial tension and compression, while with
a dotted-line is sketched the Mohr's circle which defines the limit stress states, as it is tangent
to the vertical boundaries identified by the equations σ = Sut and σ = Suc , respectively.
Within the context introduced above, one needs to recall that a general state of stress
can be decomposed into a part that causes the volume change (i.e., hydrostatic part) and a part
causing the shape distortion (i.e., deviatoric part). Moreover, it is worth noting that for metals,
the hydrostatic component has no influence on yielding, and only the deviatoric component
governs their plastic deformation. Shifting the location of Mohr's circle along the abscissa axis
represents an increase of the hydrostatic component. On the contrary, increasing the radius of
Mohr's circle without changing the average stress value represents an expansion of the
deviatoric component. As long as the failure of ductile materials depends only on the deviatoric
component, a failure theory intended to be used for ductile materials should produce the same
result regardless of where the Mohr's circle is located along the horizontal axis. This explains
why the maximum principal stress theory is not adequate for ductile materials, i.e., since it is
not consistent with the observation that yielding does not depends on the hydrostatic stress.
Having in mind what was mentioned so far, the following limitations of maximum
principal stress theory could be briefly summarized:
• It only considers the maximum principal stress (or minimum principal stress if
appropriate), disregarding the effects of the other principal stress components as well as
the effect of shear stress;
• In the case of hydrostatic pressure loading, the maximum principal stress theory fails in
predicting the real response behavior in the sense that it indicates failure when the value
of ultimate compression strength is reached, even though the experiments show that
materials can withstand much higher hydrostatic compression without failure.
In light of the practical use of maximum principal stress theory, it is more comfortable to
combine the critical stress components at a point, under an actual multiaxial loading state, into
a single quantity represented through the equivalent stress and then, by comparing it with the
ultimate stress, in some appropriate form, one may predict the failure occurrence. Therefore,
based on the equations (11.4) and (11.5), the strength condition can be written as:

Sut Sut
σeq = max {σ1 ; ∙ σ3 } ≤ with nsf >1 , (11.7)
Suc nsf

where the ratio Sut /Suc always takes a negative value, and nsf is the factor of safety.

10
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

A graphical representation of failure locus, similar to the one illustrated in Figure 11-2,
could also be obtained when the particular case of biaxial stress state (i.e., plane stress state
with∙σ2 = 0) is involved.

Example 11.2-1
A cast pipe made of an aluminum alloy with the outer diameter D = 90 mm, and the inner
diameter d = 50 mm is submitted to static torsion by an applied torque moment Mt = 15 kNm.
Consider the maximum principal stress failure theory to determine the factor of safety, supposing
that the values of ultimate strengths in tension and compression are Sut = 290 MPa and
Suc = -330 MPa , respectively.

Solution:
The maximum shear stress develops on the outside surface of the shaft, and it is related to
the applied torque through the well-known elastic torsion formula:

Mt 15 ∙10 6
τmax = = = 115,8 MPa .
Wp π ∙90 3 50 4
16 [1 - (90) ]

Provided that a plane state of pure shear stress develops when the circular or annular cross-section
shafts are subjected to torsion, the principal stresses can be obtained graphically from the Mohr's
circle:

σ1 = -σ3 = τmax = 115,8 MPa and σ2 = 0 .

Then, applying the strength condition (11.7), conforming to maximum principal stress failure theory,
we obtain:

290 290 MPa


σeq = max {115,8 ; ∙115,8 } ≤ ,
-330 nsf

which leads to the following minimum value of the safety factor:

290 MPa
nsf = = 2,5 .
115,8 MPa

Thus, it can be observed that, according to the maximum principal stress failure theory, the elastic
state of stress due to the applied static torque is far enough from the material's ultimate tensile
strength.
11
Section 11. Theories of elastic failure

11.3 Maximum shear stress failure theory9

This theory states that the material yields when the maximum shear stress at any of its
points reaches the critical value of shear stress corresponding to yielding in a standard uniaxial
tensile test. It is generally suited to predict the yielding of ductile materials, as it is consistent
with the mechanisms of yielding occurrence at micromechanical level; i.e., involving shear slips
along with the directions of maximum shear stress, which are inclined to an angle of 45 degrees
relative to the planes of maximum and minimum principal stresses.
As already shown in the chapter related to basics in the theory of elasticity, for a general
state of stress, the maximum shear stress may be expressed as:

|σ1 - σ3 | |σ2 - σ3 | |σ1 - σ2 |


τmax = max { ; ; }, (11.8)
2 2 2

where σ1 , σ2 and σ3 are the principal stresses customarily arranged in descending order, such
that σ1 ≥ σ2 ≥ σ3 .
In the case of simple standard specimens submitted to uniaxial static tests, the yield
stress at failure is reached by the first principal stress component so as the expression of
maximum shear stress at yielding (11.8) becomes:

σ1 - 0 SY
τmax = = , (11.9)
2 2

where SY is the yield strength of the material.


Hence, conforming to Tresca's theory, yielding under a combined loading state occurs
when:

|σ1 - σ3 | |σ2 - σ3 | |σ1 - σ2 | SY


max { ; ; }= , (11.10)
2 2 2 2

which simplifies to:

max {|σ1 - σ3 |; |σ2 - σ3 |; |σ1 - σ2 |} = S Y . (11.11)

As far as the computation relationships intended for safety design purposes are
concerned, it is often convenient to involve the equivalent stress, which is nothing but a

9
Commonly linked to the names of Tresca – Rankine – Guest.

12
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

function of the stress components that must reach the yield stress value in simple tension for
yielding to commence. Thus, the strength condition in line with the maximum shear stress
failure theory may be written as:

Sut
σeq = max {|σ1 - σ3 |; |σ2 - σ3 |; |σ1 - σ2 |} ≤ with nsf >1 , (11.12)
nsf

where the safety factor is denoted with nsf .

Figure 11-4

For a plane stress problem, taking for instance σ3 = 0, the equation (11.11) becomes:

σeq = max {|σ1 - σ2 |; |σ1 |; | σ2 |} = S Y , (11.13)

which plotted on a coordinate system whose axes are identified with the principal stresses σ1
and σ2 , gives a hexagon representing the yield locus10, associated with maximum shear stress
failure theory (see Figure 11-4). Yielding of material is considered to occur when the actual
biaxial stress state corresponds to a point lying upon its boundary line, while the combination
of principal stresses that fall inside the hexagon is thought to be safe.
Regarding the condition (11.13), two different cases may arise. In the first case, the sign
of principal stresses σ1 and σ2 , are assumed to be the same, and this corresponds to the first

10
Sometimes termed as the failure surface.

13
Section 11. Theories of elastic failure

and third quadrant (see Figure 11-4). For instance, taking the first quadrant, so as both principal
stresses are positive, the resulting Mohr’s circles are represented in Figure 11-5.

Figure 11-5

Although σ3 = 0, the highest shear stress results from the semi-difference between σ1 and σ3 ,
and thus it would be equal to half of σ1 . Therefore, yielding commences when either:

σ1 = S Y or σ2 = S Y , (11.14)

representing the equations of two straight lines, the 1st is vertical while 2nd is horizontal.

Figure 11-6

In the second case, the sign of principal stresses σ1 and σ2 , are assumed to be the opposite,
and this applies in the second and fourth quadrant (see Figure 11-4). For example, in the fourth

14
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

quadrant, with σ1 positive and σ2 negative, the resulting Mohr’s circles are illustrated in Figure
11-6. The largest Mohr’s circle passes through σ1 and σ2 , and thus, the maximum shear stress
is given by their semi-difference. Hence, the corresponding yield condition would be:

σ1 - σ2 = SY , (11.15)

and represents the equation of a straight line with an inclination of 45 degrees and the intercept
equal to SY . The analyses related to the second and third quadrant should follow the same
reasoning in the development as described in Figure 11-5 and Figure 11-6, respectively.
According to the maximum shear stress theory, by adding hydrostatic compressive or
tensile stresses, no change in the material response is expected. In terms of Mohr’s circles, only
their sizes (and not their locations) are essential in assessing whether yielding commences.
In what concern the limitations in the applicability of maximum shear stress failure
theory, the following remarks may be emphasized:
• It gives conservative results for states of stress where a high amount of shear is involved
(i.e., in case of torsion tests, an average overestimation error of 15% is reported relative
to experimental data);
• Its application is limited to materials having similar strength in tension and compression,
and thus, it is not to be used for brittle materials whose elastic limit stress in tension and
compression differ significantly;
• It is not applicable to materials loaded under hydrostatic pressure, as it predicts almost
zero shear stress and consequently no failure, however high the applied pressure can be,
but this is unrealistic.

Example 11.3-1
Given a structural element made of a ductile material with the yield strength equal to 250
MPa, and assuming at one of its points a critical stress state with σ1 = 105 MPa, σ2 = 0 and
σ3 = -130 MPa, calculate the safety factor using the maximum shear stress failure theory.

Solution:
Applying the strength condition (11.13), one may write:

250 MPa
σeq = max {|105 + 130 |; |0 + 130 |; |105 - 0 |} ≤ ⇒ nsf = 1,06 .
nsf

It follows that conforming to maximum shear stress theory, the safety factor falls just at the limit
but on the safe side as it is slightly greater than unity.

15
Section 11. Theories of elastic failure

11.4 Mohr's theory of failure

This theory is generally applied to predict the failure of brittle materials exhibiting a
pronounced uneven behavior in tension and compression11. Before getting into its features, it
is to be recalled that the Mohr's circle can be used to illustrate principal stresses and stress
transformations via a graphical format in a planar coordinate system with the normal stresses
plotted along the abscissa and the shear stresses along the ordinate axis. All that being said,
one may state that according to this theory, it is assumed that of all the planes undergoing the
same magnitude of normal stress, the critical one on which failure is likely to occur is that where
the maximum shear stress develops12. In other words, this theory predicts shear failure about
a specific fracture plane, but the involved critical shearing stress is supposed to be governed by
the effective normal stress acting on that plane.

Figure 11-7

Several experimental tests have to be conducted on the same material to obtain its
graphical representation, as the envelope of all Mohr's circles corresponding with the particular
states of stress that cause failure. Instead, suppose at least the results of experiments in

11
Gray cast irons typically have the ultimate compression strength about three to four times greater than the value of ultimate
tensile strength. For ceramics, even larger ratios are possible. Certain magnesium alloys are ductile but are stronger in tension
than in compression [22].
12
Here, it should be understood that the failure plane does not strictly refer to the particular plane where the shear stress
takes its maximum value. Rather, it is a combined state of shear and normal stress developing at a certain point within the
material, which becomes critical.

16
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

simple tension, compression, and pure shear are available in terms of principal stresses, in
addition to a hydrostatic tension test intended to determine the finite value of the tension limit.
In that case, they can be plotted on the same graphic through the associated stress circles, as
portrayed in Figure 11-7. A curve linking the circles for uniaxial tension, compression, and pure
shear establishes the failure envelope for Mohr's theory. Its main characteristics shall be briefly
summarized in what follows:
• It is symmetrical to the abscissa since the failure does not depends on the sign of shear
stress(quite often in practical applications, the use of the upper half is enough because
of symmetry; however, the entire curve allows to easier identify those stress
combinations that correspond to the enclosed points and do not cause failure);
• Its tangent slope is directly related to the critical combination of shear and normal stress
(if it is horizontal, the maximum shear stress indicated by the Mohr's circle coincides with
the shear stress acting on the failure plane, if it is inclined with a specific angle relative to
the horizontal axis, it means that the shear stress on the failure plane is lower than the
maximum shear stress determined with the help of Mohr's circle representation);
• It is generally open on the negative side of the abscissa axis since, under hydrostatic
compression loading, no elastic limit exists, and therefore, the envelope is supposed to
rise indefinitely;
• It is closed on the positive side of abscissa into an extreme point representing the finite
limit for hydrostatic tensile loading (at this point, which had marked by D in Figure 11-7,
the tangent to the envelope curve takes a vertical orientation indicating a fracture plane
relatively straight and perpendicular to the direction of the 1st principal stress).
The largest Mohr's circle for a given stress state turns out to be tangent at the envelope
curve when the failure is expected to occur. Once the contact point, denoted by L in Figure 11-
7, is located, not only the shear and normal stress components acting on the failure plane can
be calculated, but also the effective failure angle (i.e.,  ). As indicated in the same figure, 2
refers to the angle by which the tangent radius of the largest circle to the failure envelope is
rotated counterclockwise relative to the abscissa axis. It may also be observed that the same
as for the other failure criteria described so far, this theory disregards the effect of the
intermediate principal stress component.
One frequently assumes the failure envelope curves to be straight lines for practical
applications, as shown in Figure 11-8. The most common straight-line assumption relies on
the well-known Coulomb's friction hypothesis. Combined with Mohr's failure theory, it gives

17
Section 11. Theories of elastic failure

the so-called Mohr-Coulomb failure criterion, sometimes known as the internal friction theory.
Accordingly, the failure is supposed to coincide with the contact point between the Mohr's circle
for a given state of stress and the envelope straight-lines linking only the circles for uniaxial
tension and compression. A state of stress whose Mohr's circle lies inside these lines does not
determine failure. Mohr's circles that extend outside the envelope straight-lines have no
meaning in this context since the stresses are assumed to slowly increase from a particular
stress state (say, initially safe), and thus, failure is supposed to occur when the Mohr's circle
touches the failure envelope lines.

Figure 11-8

The above considerations lead to a mathematical condition, which defines the failure limit
if the following equation is satisfied13:

τ = c - μσ , (11.16)

where τ is the intercept of failure envelope line, c represents its slope (commonly called the
cohesive strength of the material), while the product μσ is termed as the friction stress, where:

μ = tan (α ) , (11.17)

is the coefficient of internal friction, with the friction angle denoted by α in Figure 11-8.

13
If all principal stresses are positive, the Mohr-Coulomb failure theory might lead to results which are not allways proven by
experimentas (i.e., the two straight-lines aproximation yields to an intersection point, denoted by V in Figure 11-8, which does
not appear on the experimental Mohr’s failure envelope curve [10], [13], [16]).

18
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

Given that for a biaxial stress state, the stress components acting on a plane inclined with
angle  , are:

σ1 + σ3 σ1 - σ3
σ= + cos (2ϕ) ,
2 2
(11.18)
σ1 - σ3
τ= sin (2ϕ) ,
2

from Figure 11-8, one may change the argument for both of the above relations, thus:

π
2ϕ = -α. (11.19)
2

Then, by replacing σ and τ from (11.18) together with μ from (11.17), into the Mohr-Coulomb
equation (11.16), we obtain:

σ1 - σ3 + (σ1 + σ3 ) sin (α ) = 2c ∙cos (α ) , (11.20)

or in a more general form, if the principal stresses σ1 , σ2 and σ3 are no longer ordered, it can
be expressed as:

σmax - σmin + (σmax + σmin ) sin (α ) = 2c ∙cos (α ) . (11.21)

Concerning the form of the above equation, if a particular case of pure uniaxial tension
(σmax = Sut ; σmin = 0) is imposed, the ultimate tensile stress can be derived:

2c ∙cos (α )
Sut = , (11.22)
1 + sin (α )

Similarly, assuming a pure stress state of uniaxial compression (σmin = Suc ; σmax = 0), the
expression for ultimate compression stress is obtained:

2c ∙cos (α )
Suc = . (11.23)
1 - sin (α )

Solving the equations (11.22) and (11.23) for c and α , leads to:

1
c= √Suc ∙Sut and α = sin-1 [(Suc - Sut )/(Suc + Sut )] . (11.24)
2

In other words, given the ultimate limit stresses Suc and Sut for a material, the failure envelope
parameters of Mohr-Coulomb theory can be calculated with the expression (11.24).

19
Section 11. Theories of elastic failure

The data from Figure 11-8 may be replotted in a planar coordinate frame having the 1st
principal stress on the abscissa and the 3rd principal stress on the ordinate axis (see Figure 11-
9). One may observe that in the first and third quadrant where σ1 and σ3 are of the same signs,
the failure loci of Mohr-Coulomb's failure theory and maximum principal stress theory coincide.
However, the area of biaxial tension is made smaller than it is for biaxial compression, and thus,
the limit condition can be expressed in the form:

max{|σ1 |; |σ3 |} = Sut and min{|σ1 |; |σ3 |} = Suc . (11.25)

In the second and fourth quadrant, where σ1 and σ3 have opposite signs, the two theories
differ. In particular, the limit condition can be written through the following relationship:

max{|σ1 |; |σ3 |} min{|σ1 |; |σ3 |}


+ =1, (11.26)
Sut Suc

which describes a linear change between the failure locus for biaxial tension and the one for
biaxial compression.

Figure 11-9

If the material strengths in tension and compression are the same, a similar hexagon with the
one shown in Figure 11-4 is obtained. Nevertheless, whereas the hexagon in Figure 11-4
defines a yield condition for ductile materials, it gives a fracture criterion for brittle materials in
the current framework.
20
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

In what concerns the limit condition (11.26), if the principal stresses are ordered such that
max{|σ1 |; |σ3 |} = σ1 and min{|σ1 |; |σ3 |} = σ3 , by imposing the equality Sut = σeq , it is possible
to obtain the expression of the equivalent stress associated with Mohr-Coulomb's failure
theory. In effect, the strength condition, including the safety factor nsf , can be rewritten as:

Sut Sut Sut


σeq = max {σ1 ; σ3 ; σ1 + σ3 }≤ with nsf >1 . (11.27)
Suc Suc nsf

Although Mohr-Coulomb failure theory has an analytical basis, a modified Mohr theory
arose subsequently by fitting the experimental data. Figure 11-9 depicts the modified Mohr's
theory along with the Mohr-Coulomb theory and maximum principal stress theory, intending
to highlight their differences. All three theories coincide in quadrants one and three, while this
is no longer valid in quadrants two and four. Consider, for instance, a state of pure torsion
described by the dash-dotted line of negative unit slope, lying within the second and fourth
quadrant, as illustrated in Figure 11-9. The intersection between this line and the boundary of
Mohr-Coulomb's failure locus implies that the shear strength of the material under analysis is
less than its tensile strength. On the other hand, in the case of brittle materials, it is generally
accepted that the shear and tensile strength are almost equal, and this is just what the
modified Mohr's theory achieves.
From the mathematical point of view, if the principal stresses are still kept ordered, so
that σ1 is maximum and σ3 is minimum, the modified Mohr's theory can be expressed as:

Suc + Sut 1
σ1 + σ3 = 1 if σ1 > 0 and σ3 < -Sut ; (11.28)
Suc ∙Sut Suc
or

σ1 = Sut if σ3 > -Sut ; (11.29)

otherwise

σ3 = Suc if σ1 < 0 . (11.30)

Then, by following the same approach as in the case of Mohr-Coulomb's theory, the following
strength condition written in terms of equivalent stress and safety factor, is obtained for the
modified Mohr's theory:

Sut Sut Sut Sut


σeq = max {σ1 ; σ3 ; σ1 (1+ ) + σ3 }≤ . (11.31)
Suc Suc Suc nsf

21
Section 11. Theories of elastic failure

Example 11.4-1
Recalculate the safety factor for the case of annular shaft subjected to torsion in example
11.2-1. Follow the criteria of Mohr-Coulomb's theory and the modified Mohr's theory as well.

Solution:
According to the strength criterion associated with Mohr-Coulomb's failure theory (11.27),
we obtain:

290 290 290 MPa


σeq = max {115,8; (-115,8 ); 115,8 + (-115,8 ) } ≤ ⇒ nsf = 1,33 .
-330 -330 nsf

On the other hand, applying the modified Mohr's failure theory (11.7), we have:

290 290 290 290 MPa


σeq = max {115,8; (-115,8 ); 115,8 (1 + )+ (-115,8 ) } ≤ ⇒ nsf = 2,5 .
-330 -330 -330 nsf

Such findings fit the expectations that Mohr-Coulomb's failure theory leads to a conservative
safety factor compared to the modified Mohr's theory.

Example 11.4-2
Given a mechanical part made of brittle material and assuming at one of its points a particular
stress state with σ1 = 105 MPa, σ2 = 0 and σ3 = -550 MPa, calculate the safety factor using
maximum normal stress theory, Mohr-Coulomb theory, and the modified Mohr's theory. Consider
the ultimate tensile strength Sut = 275 MPa, and the compression strength Suc = -1240 MPa.

Solution:
The safety factors calculated according to each of the required failure criteria are reported:
Maximum normal stress theory (11.7):

275 275 MPa


σeq = max {105; ∙ (-550 )} ≤ ⇒ nsf = 2,25 .
-1240 nsf

Mohr-Coulomb’s theory (11.27):

275 275 275 MPa


σeq = max {105; ∙ (-550 ); 105 + ∙ (-550 ) } ≤ ⇒ nsf = 1,21 .
-1240 -1240 nsf

Modified Mohr’s theory (11.7):

275 275 275 290 MPa


σeq = max {105; (-550 ); 105 (1 + )+ (-550 ) } ≤ ⇒ nsf = 1,35 .
-1240 -1240 -1240 nsf

22
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

11.5 The distortion strain energy failure theory14

This theory predicts failure when the distortion strain energy density at any point of
material reaches the same value as in the case of uniaxial tension at incipient yield. Therefore,
according to this theory, material failure occurs through yielding and is assumed to be entirely
governed by the elastic strain energy associated with shear deformation. For this reason, it is
valid only for ductile materials and can be effectively applied to identify the combination of
stresses that cause the yield onset, or in other words, the critical stress states corresponding
to the upper limit of the elastic deformation. However, data reported in the literature show that
it agrees with experiments somehow better than the maximum shear stress theory (i.e.,
Tresca) does. In fact, the distortion strain energy theory may be thought of as a refinement of
Tresca theory in the sense that it accounts for all three principal stresses in a comprehensive
manner.
Since the ductile materials typically withstand higher hydrostatic pressure loads in
comparison with the case of a simple uniaxial loading, the experimental evidence suggested
that yielding should be somehow related to the deviatoric part of the elastic strain energy
density (i.e., distortion strain energy per unit volume), rather than to its hydrostatic component
(i.e., volumetric strain energy). With regard to this, one needs to recall that the total strain
energy of an infinitesimal element of material, denoted by Ut , can be decomposed into two
parts:

Ut = Uv + Us , (11.32)

where Uv is the strain energy related to the volumetric change of the element, while Us refers
to the distortion strain energy associated with changing the shape of the element.
Since the expression of strain energy per unit volume, commonly referred to as the elastic
strain energy density, is:

1
UD = σ ε. (11.33)
2

Given the principal stresses σ1 , σ2 and σ3 , for a generic volume of material subjected to normal
average stresses:

σ1 + σ2 + σ3
̅=
σ (11.34)
3

14
Linked to the names of Maxwell - Huber - Hencky - von Mises.

23
Section 11. Theories of elastic failure

upon each of its six facets, the strain energy density in only one direction would be simply
calculated as:

1
UD (1) = σ
̅ε̅, (11.35)
2

where ε
̅ is the strain due to the applied average stress σ
̅. Then, the volumetric strain energy
per unit volume shall be obtained by multiplying the expression (11.35) with a factor of three:

1
UV = 3∙ σ
̅ε̅. (11.36)
2

Assuming a linear-elastic material behavior, based on Hooke's law, one may write:

σ
̅ 2νσ
̅ σ
̅
̅=
ε - = (1 - 2ν) , (11.37)
E E E

so as the expression (11.36) becomes:

1 σ
̅
UV = 3∙ σ
̅ [ (1 - 2ν)] , (11.38)
2 E

and recalling the expression of σ


̅ given by (11.34), it turns out that:

(1 - 2ν) 2
UV = (σ1 + σ2 + σ3 ) . (11.39)
6E

By means of a similar approach, the total strain energy per unit volume, when the
principal stresses σ1 , σ2 and σ3 are involved, can be written as:

1 1 1
Ut = σ1 ε1 + σ2 ε2 + σ3 ε3 . (11.40)
2 2 2

where:

σ1 ν
ε1 = - (σ2 + σ3 );
E E
σ2 ν
ε2 = - (σ3 + σ1 ); (11.41)
E E
σ3 ν
ε 3 = - (σ + σ2 ).
{ E E 1

Substituting first for ε1 , ε2 and ε3 into (11.40), and then for UV from (11.39) likewise for

24
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

Ut from (11.40), back into the equality (11.32), the expression of distortional strain energy per
unit volume shall be obtained as:

1 (1 - 2ν) 2
US = [σ1 2 + σ2 2 + σ3 2 - 2ν(σ1 σ2 + σ2 σ3 + σ3 σ1 ) - (σ1 + σ2 + σ3 ) ] , (11.42)
2E 6E

which simplifies to:

(1 + ν) 2 2 2
US = [(σ1 - σ2 ) + (σ2 - σ3 ) + (σ3 - σ1 ) ] , (11.43)
6E

or, if the relationship between the elastic constants E = 2G/(1 + ν) is considered15, the
expression (11.43) becomes:

1 2 2 2
US = [(σ1 - σ2 ) + (σ2 - σ3 ) + (σ3 - σ1 ) ] . (11.44)
12G

On the other hand, in case of a simple uniaxial tensile test, we may write:

σ1 = SY , (11.45)

while all the other principal normal stresses, i.e., σ2 and σ3 , are equal to zero. It follows that
the distortion strain energy per unit volume, given by the relation (11.44), transforms into:

SY 2 (11.46)
US = ,
6G

where SY is the yield limit of the material.


As stated in its definition, according to this theory, failure occurs when the value of US for
a given multiaxial stress state, as described by (11.44), reaches a threshold value corresponding
to a simple standard uniaxial tensile test as indicated by (11.46), so that:

2 2 2
(σ1 - σ2 ) + (σ2 - σ3 ) + (σ3 - σ1 ) = 2SY 2 , (11.47)

or

σ1 2 + σ2 2 + σ3 2 - σ1 σ2 - σ2 σ3 - σ3 σ1 = SY 2 . (11.48)

Based on either of two equations (11.47) or (11.48), the strength condition may be written
more conveniently in terms of equivalent stress, which may be understood as the value of a

15
Here, G stands for the transverse shear modulus of material, and ν is the Poisson’s coefficient.

25
Section 11. Theories of elastic failure

virtual uniaxial tensile stress that would produce the same amount of distortion energy as the
actual stress state:

1 Sut
σeq = √(σ1 - σ2 )2 + (σ2 - σ3 )2 + (σ3 - σ1 )2 ≤ with nsf >1 , (11.49)
√2 nsf

where nsf is the safety factor.


For a plane stress state, assuming σ3 = 0, the equation (11.48) simplifies to:

SY = √σ1 2 - σ1 σ2 + σ2 2 , (11.50)

or, in terms of Cartesian stress components, it can be written as:

SY = √σx 2 - σx σy + σy 2 + 3τxy 2 . (11.51)

The equation (11.50) can be plotted in a planar coordinate system with the 1st principal stress
along the abscissa axis and the 2nd principal stress along the ordinate axis. As illustrated in
Figure 11-10, it describes an ellipse of major and minor axes inclined at 45 degrees relative to
principal axes σ1 and σ2 .

Figure 11-10

The ellipse boundary line indicates the limit state (sometimes termed as the yield locus)
corresponding to the onset of yielding. In contrast, the points lying inside the ellipse define the
biaxial stress state region where the material is considered safe against yielding under static
loading conditions. As the ellipse passes through the six corner points of the Tresca yield locus,

26
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

at these points, both theories are supposed to give identical results. However, it can be
observed that for some other combinations of σ1 and σ2 , Tresca’s theory is slightly
conservative as it predicts failure, while the von Mises theory does not.
Let us consider, for instance, the particular case of a pure shear state, such that σx = σy = 0
and τxy = τ. That is, the principal stresses shall be σ1 = -σ2 = τ and σ3 = 0. It results in a straight
line, as represented in Figure 11-10, passing through the origin and inclined at an angle of -45
degrees to the direction of positive σ1 axis. The line intersects the von Mises yield locus at two
particular points where the magnitudes of principal stresses can easily be found from the
equation (11.48):

SY 2 = σ1 2 - σ1 σ1 + σ1 2 = 3σ1 2 = 3τ2 , (11.52)

which means that, if τ = SY /√ 3 = 0,577SY , the material commences to yields. Thus, according
to the distortion strain energy failure theory, a ductile material is 57,7% stronger in shear as
compared to tension. Experimental evidence shows that this prediction is slightly more
accurate than that provided by the maximum shear stress failure theory.
Concerning the practical use of the distortion strain energy failure theory, the following
remarks may be briefly outlined:
• Its application is limited to materials having similar strength in tension and compression,
and thus, it is not to be used for brittle materials whose elastic limit stress in tension and
compression differ significantly;
• It is suitable to be applied for ductile materials that are pressure independent in the sense
that they are not likely to fail under hydrostatic pressure conditions;
• It accounts for all three principal stress components.

Example 11.5-1
It is required to recalculate the safety factor for the stress state given in example 11.3-1,
using the von Mises failure theory.

Solution:
According to the expression (11.49), we have:

1 250 MPa
σeq = √105 2 + 130 2 + (-130 - 105 )2 ≤ ⇒ nsf = 1,22 .
√2 nsf

which indeed is a little less conservative compared to the value found based on Tresca's failure
theory.
27
Section 11. Theories of elastic failure

11.6 The octahedral shear stress failure theory

According to this theory, yielding commences when the shear stress acting on an
octahedral plane reaches a critical value for the material, determined in a simple uniaxial tensile
test. As this definition states, it applies only to ductile isotropic materials that typically fail by
yielding. Moreover, it leads to the same expression for the failure envelope as the distortion
strain energy theory does, although they essentially use different measures to interpret the
critical stress state at failure within material under complex loading conditions. Whereas the
latter is based on an energy principle and assumes a linear stress-strain behavior, the
octahedral shear stress theory relies on a stress concept, and it does not consider the stress-
strain relationship anymore.

Figure 11-11

One of the octahedral planes is shaded with gray in Figure 11-11. Eight such planes
define a regular octahedron. The vertices of the octahedron lie on the principal stress axes, and
the normal vectors of its facets make equal angles with the directions of principal stresses σ1 ,
σ2 and σ3 . For instance, relative to a given normal vector n, if these angles are denoted with α,
β and γ, we have:

α = β = γ = 54,74° (11.53)

and

1
cos (54,74°) = = 0,577. (11.54)
√3

With the understanding that this theory relies on actual occurring stress rather than the

28
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

fictitious equivalent stress, the normal and shear stresses acting upon the octahedral planes
can be straightforwardly calculated by applying the static equilibrium conditions:

σ1 + σ2 + σ3
σh = (11.55)
3
and
1 2 2 2 1/2
τh = [(σ1 - σ2 ) + (σ2 - σ3 ) + (σ3 - σ1 ) ] . (11.56)
3

Regarding the above two relationships, consider, for instance, a material subjected to
hydrostatic pressure, σ1 = σ2 = σ3 = -p , where p is the applied pressure, then it gives σh = -p
and τh = 0. One may conclude that upon the octahedral plane, the normal stress solely
contributes to the dilatation while the magnitude of octahedral shear stress determines the
yield onset. This theory is fairly well supported by experimental evidence.
At the yield point, in case a uniaxial tensile test (i.e., σ1 = SY , σ2 = σ3 = 0), the expression
(11.56) becomes:

√2 (11.57)
τh = S = 0,471SY
3 Y

so that the introductory statement of octahedral shear stress theory can be simply written in
mathematical terms by comparing the relations (11.56) and (11.57):

1 √2
τh = √(σ1 - σ2 )2 + (σ2 - σ3 )2 + (σ3 - σ1 )2 = S . (11.58)
3 3 Y

and then, by invoking the concept of equivalent stress, it may be considered in another form:

1
σeq = √(σ1 - σ2 )2 + (σ2 - σ3 )2 + (σ3 - σ1 )2 . (11.59)
√2

As mentioned above, this is similar to the expression of equivalent stress derived through the
assumptions of distortion strain energy theory.
The physical nature of equivalent stress becomes a bit more comprehensible when the
reference is made to the octahedral face shaded with grey in Figure 11-11. Strictly speaking,
since the magnitude of the shear stress acting on this face is given by the equation (11.56), it
can be simply related to the expression of equivalent stress (11.59) by a simple algebra; i.e.,

29
Section 11. Theories of elastic failure

σeq = (3/√2)τh . In other words, the octahedral stress criterion can be reformulated as follows:
the yield condition is reached under a complex state of multiaxial loading when the octahedral
shear stress reaches 3/√2 from the value of yield stress in a uniaxial stress state.

30
References

[1] Asaro, R., & Lubarda, V. (2006). Mechanics of solids and materials. Cambridge University
Press.
[2] Bauchau, O. A., & Craig, J. I. (2009). Structural analysis: with applications to aerospace
structures (Vol. 163). Springer Science & Business Media.
[3] Bejan, M., Simion, M., Cherecheș, I. A., Lakatos, D. G., & Vidican, I. (2013). Compendii din
rezistența materialelor, vol. 1 și vol. 2. Editura AGIR, Bucureşti și Editura MEGA, Cluj
Napoca.
[4] Boresi, A. P., & Schmidt, R. J. Advanced mechanics of materials, 2003. John Wiley & Sons,
117, 122.
[5] Bouvet, C. (2017). Mechanics of aeronautical solids, materials and structures. John Wiley
& Sons.
[6] Chiriacescu, S. T. (2008). On the loads limit according to classical failure theories of
materials. The Romanian Journal of Technical Sciences. Applied Mechanics., 53(3), 247-
262.
[7] Constantinescu, I. N., Picu, C., Hadăr, A., & Gheorghiu, H. (2006). Rezistența materialelor
pentru ingineria mecanică. Editura BREN, Bucureşti.
[8] Christensen, R. M. (2013). The theory of materials failure. Oxford University Press.
[9] Curtu, I., Roșca, C., & Crișan, R. (1995). Memorator de rezistența materialelor.
Universitatea" Transilvania" din Brașov, Facultatea de Mecanică.
[10] Da Silva, V. D. (2005). Mechanics and strength of materials. Springer Science & Business
Media.
[11] Dumitru, I., & Faur, N. (1999). Elemente de calcul şi aplicații în rezistența materialelor.,
Editura Politehnică, Timişoara.
[12] Govindjee, S. (2012). Engineering Mechanics of Deformable Solids: A Presentation with
Exercises. Oxford University Press.
[13] Gu, J., & Chen, P. (2018). A failure criterion for isotropic materials based on Mohr’s failure
plane theory. Mechanics Research Communications, 87, 1-6.
[14] Jong, I. C., & Springer, W. (2009). Teaching von Mises stress: from principal axes to non-
principal axes. In American Society for Engineering Education. American Society for
Synopsis of Lectures in Strength of Materials for Undergraduates. Part 2 Marius Nicolae BABA

Engineering Education.
[15] Keil, S. (2017). Technology and practical use of strain gages: with particular consideration
of stress analysis using strain gages. John Wiley & Sons.
[16] Li, S. (2020). A reflection on the Mohr failure criterion. Mechanics of Materials, 103442.
[17] Megson, T. H. G. (2019). Structural and stress analysis. Butterworth-Heinemann.
[18] Munteanu, M. G., Radu, G. N., & Biț, C. S. (1994). Teoria elasticității și rezistența
materialelor. Universitatea" Transilvania", Brașov.
[19] Popov, E. P., & Balan, T. A. (1990). Engineering mechanics of solids (Vol. 2). Englewood
Cliffs, NJ: Prentice Hall.
[20] Radeş, M. (2007). Rezistența materialelor. Editura Printech, București.
[21] Rees, D. W. (2016). Mechanics of solids and structures. World Scientific Publishing
Company.
[22] Schmid, S. R., Hamrock, B. J., & Jacobson, B. O. (2014). Fundamentals of machine
elements: SI version. CRC Press.
[23] Sofonea, G., & Pascu, A. M. (2007). Rezistența materialelor. Editura Universității "Lucian
Blaga", Sibiu.
[24] Srinath, L. S. (2003). Advanced mechanics of solids. Tata McGraw-Hill.
[25] Ugural, A. C. (2018). Mechanical Design of Machine Components: SI Version. Taylor &
Francis.
[26] Yu, M. H., Yu, M., & Yu, M. H. (2004). Unified strength theory and its applications.

You might also like