Download as pdf or txt
Download as pdf or txt
You are on page 1of 99

Failure prevention

Chapter 5 Failure resulting from static loading

Chapter 6 Fatigue failure resulting from variable loading

© McGraw Hill 1
Failure Examples 1
Fig. 5–1

Failure of truck driveshaft spline due to corrosion fatigue


腐蝕疲勞:材料疲勞伴隨腐蝕加重而引起材料在
它的一般疲勞極限之下就可能發生斷裂。
For permission to reprint Figures 5–1 through 5–5, the authors are grateful for the personal photographs of
Larry D. Mitchell, coauthor of Mechanical Engineering Design, 4th ed., McGraw-Hill, New York, 1983.

© McGraw Hill 2
Failure Examples 2

Fig. 5–2

Impact failure of a lawn-mower blade driver hub.


The blade impacted a surveying pipe marker.

© McGraw Hill 3
Failure Examples 3

Fig. 5–3

Failure of an overhead-pulley retaining bolt on a weightlifting


machine.
A manufacturing error caused a gap that forced the bolt to take the
entire moment load.
© McGraw Hill 4
Failure Examples 4

Fig. 5–4

Chain test fixture that failed in one cycle.


To alleviate complaints of excessive wear, the manufacturer decided to
case-harden the material.
(a) Two halves showing brittle fracture initiated by stress concentration.
(b) Enlarged view showing cracks induced by stress concentration at the
support-pin holes.
© McGraw Hill 5
Failure Examples 5

Fig. 5–5

Valve-spring failure caused by spring surge in an overspeed engine.


The fractures exhibit the classic 45 degree shear failure

© McGraw Hill 6
Stress Concentration

Localized increase of stress near discontinuities.


Kt is Theoretical (Geometric) Stress Concentration Factor.
 max = K t nom (a )
 max = K ts nom (b)

Nominal Stress
Stress calculated on the basis of the net
cross section of a specimen without taking
into account the effect of geometric
discontinuities such as holes, grooves,
fillets, etc.

Access the text alternative for slide images.

© McGraw Hill 7
Theoretical Stress Concentration Factor

Graphs available for


standard configurations.
See Appendix A–15 and
A–16 for common
examples.
Many more in Peterson’s Fig. A–15–1
Stress-Concentration
Factors (may check PDF
file from website).
Note the trend for higher
Kt at sharper discontinuity
radius, and at greater
Fig. A–15–9
disruption. Access the text alternative for slide images.

© McGraw Hill 8
Stress Concentration for Ductile and Brittle material

• Ductile material with static loading, we can consider the strain


strengthening of the material and an increasing in yield strength
at critical notch location (locally).
• This local part can carry the load properly with no general
yielding. Thus, Stress-Concentration Factor Kt was set to one.
• In other words, in a ductile metal plastic deformation enables the
stress to redistribute and blunt the stress concentration.
Note that:
• As the loading stress exceeds yield strength of the material,
plastic deformation occurs for ductile materials.
• Brittle materials, a fracture criterion based upon the maximum
normal stress is generally used.

© McGraw Hill 9
© McGraw Hill 10
Ductile material Brittle material

© McGraw Hill 11
ductile materials w/quasi-static loads

© McGraw Hill 12
Why SCFs are not applied to ductile/damaging materials w/quasi-static loads ?

Source:
https://www.youtube.com/watch?v=8NbIZXLfPu4

© McGraw Hill 13
© McGraw Hill 14
Static Failure Theories 1
• Ductile materials are normally classified 𝜀𝑓 >= 0.05 and have an
identifiable yield strength that is often the same in compression as in
tension 𝑆𝑦𝑡 = 𝑆𝑦𝑐 = 𝑆𝑦 .
• Brittle materials (𝜀𝑓 < 0.05) do not exhibit an identifiable yield strength,
and are typically classified by ultimate tensile 𝑺𝒖𝒕 and compressive
strength 𝑺𝒖𝒄 .

The generally accepted Failure Theories:

Ductile materials (yield criteria): Brittle materials (fracture criteria):

• Maximum shear stress • Maximum normal stress


• Distortion energy • Brittle Coulomb-Mohr
• Ductile Coulomb-Mohr • Modified Mohr

* In materials science and engineering, the yield point is the point on a stress-strain
curve that indicates the limit of elastic behavior and the beginning of plastic behavior.
Access the text alternative for slide images.

© McGraw Hill 15
Maximum Shear Stress Theory ~ for ductile materials (MSS) 1

Theory: Yielding begins when the maximum shear stress in a stress


element exceeds the maximum shear stress in a tension test
specimen of the same material when that specimen begins to yield.
For a tension test specimen, the maximum shear stress is 1 /2.
At yielding, when 1 = Sy, the maximum shear stress is Sy /2 .
The theory can be restated as follows:
• Theory: Yielding begins when the maximum shear stress in a
stress element exceeds Sy/2.

© McGraw Hill 16
Maximum Shear Stress Theory (MSS) 2
For any stress element, use Mohr’s circle (see next page in detail)
to find the maximum shear stress. Compare the maximum shear
stress to Sy/2.
Ordering the principal stresses such that σ1 ≥ σ2 ≥ σ3,
1 −  3 S y
 max =  or  1 −  3  S y (5 - 1)
2 2
Incorporating a factor of safety n.
Sy Sy
 max = or  1 −  3 = (5 - 3)
2n n
Or solving for factor of safety.
Sy 2
n=
 max
© McGraw Hill 17
Mohr’s Circle Diagram 2

Fig. 3−10

Access the text alternative for slide images.

© McGraw Hill 18
Principal Stresses for Plane Stress
Differentiating Eq. (3–8) with respect to f and setting equal to zero
maximizes  and gives.(check next page)
2 xy
tan 2f p = (3 - 10)
x −y
The two values of 2fp are the principal directions.
The stresses in the principal directions are the principal stresses.
The principal direction surfaces have zero shear stresses.
Substituting Eq. (3–10) into Eq. (3–8) gives expression for the non-zero
principal stresses.
x +y x −y 
2

1 ,  2 =   +  2
(3 - 13)
 2 
xy
2
(check previous slide figure)

Note that there is a third principal stress, equal to zero for plane stress.
© McGraw Hill 19
Plane-Stress Transformation Equations

Cutting plane stress element at an arbitrary angle and balancing


stresses gives plane-stress transformation equations.
x +y x −y
= + cos 2f +  xy sin 2f (3 - 8)
2 2
x −y
 =− sin 2f +  xy cos 2f (3 - 9)
2

Access the text alternative for slide images. Fig. 3−9


© McGraw Hill 20
© McGraw Hill 21
General Three-Dimensional Stress 1
All stress elements are actually 3-D.
Plane stress elements simply have one surface with zero stresses.
For cases where there is no stress-free surface, the principal stresses are found
from the roots of the cubic equation.
 3 − ( x +  y +  z ) 2 + ( x y +  x z +  y z −  xy2 −  2yz −  zx2 )
(3 - 15)
−( x y z + 2 xy yz zx −  x yz −  y zx −  z xy ) = 0
2 2 2

Fig. 3−12
Access the text alternative for slide images.

© McGraw Hill 22
Maximum Shear Stress Theory (MSS) 3

To compare to experimental data, express τmax in terms of principal


stresses and plot.
To simplify, consider a plane stress state.
Let σA and σB represent the two non-zero principal stresses, then
order them with the zero principal stress such that σ1 ≥ σ2 ≥ σ3.
Assuming σA ≥ σB there are three cases to consider
• Case 1: σA ≥ σB ≥ 0
• Case 2: σA ≥ 0 ≥ σB
• Case 3: 0 ≥ σA ≥ σB

© McGraw Hill 23
Maximum Shear Stress Theory (MSS) 4
1 −  3 Sy
 max =  or  1 −  3  S y (5 - 1)
Case 1: σA ≥ σB ≥ 0 2 2

• For this case, σ1 = σA and σ3 = 0


• Eq. (5–1) reduces to σA ≥ Sy
Case 2: σA ≥ 0 ≥ σB
• For this case, σ1 = σA and σ3 = σB
• Eq. (5–1) reduces to σA − σB ≥ Sy
Case 3: 0 ≥ σA ≥ σB
• For this case, σ1 = 0 and σ3 = σB
• Eq. (5–1) reduces to σB ≤ −Sy

© McGraw Hill 24
Maximum Shear Stress Theory (MSS) 5

Plot three cases on


principal stress axes.
Case 1: σA ≥ σB ≥ 0
• σA ≥ Sy
Case 2: σA ≥ 0 ≥ σB
• σA − σB ≥ Sy
Case 3: 0 ≥ σA ≥ σB
• σB ≤ −Sy
Other lines are
symmetric cases
Inside envelope is
Fig. 5–7
predicted safe zone.
Access the text alternative for slide images.

© McGraw Hill 25
Maximum Shear Stress Theory (MSS) 6

Comparison to
experimental data.
Conservative in all
quadrants.
Commonly used for
design situations.

Access the text alternative for slide images.

© McGraw Hill 26
Distortion Energy (DE) Failure Theory 1 ~ for ductile materials

Also known as:


• Octahedral (八面體(八個面均為相同大小的三角形的立體))Shear Stress.
• Shear Energy.
• Von Mises.
• Von Mises – Hencky.

© McGraw Hill 27
Distortion Energy (DE) Failure Theory 2

Originated from observation that ductile materials stressed hydrostatically


(equal principal stresses) exhibited yield strengths greatly in excess of
expected values.
Theorizes that if strain energy is divided into hydrostatic volume changing
energy and angular distortion energy, the yielding is primarily affected by
the distortion energy.

Fig. 5–8
Access the text alternative for slide images.

© McGraw Hill 28
Distortion Energy (DE) Failure Theory 3

Theory: Yielding occurs when the distortion strain energy per unit
volume reaches the distortion strain energy per unit volume for
yield in simple tension or compression of the same material.

Fig. 5–8

© McGraw Hill 29
Deriving the Distortion Energy 1

Hydrostatic stress is average of principal stresses


1 +  2 +  3
 av = (a )
3
Strain energy per unit volume, u = 12 [1 1 +  2 2 +  3 3 ]
Substituting Eq. (3–19) for principal strains into strain energy
equation,
1
( )
 x =  x − v  y +  z 
E
 y =  y − v ( x +  z ) 
1
(3 - 19)
E
1
( )
 z =  z − v  x +  y 
E
 21 +  22 +  23 − 2v ( 1 2 +  2 3 +  3 1 ) 
1
u= (b)
2E
© McGraw Hill 30
Deriving the Distortion Energy 2

 12 +  22 +  32 − 2v ( 1 2 +  2 3 +  3 1 ) 
1
u= (b)
2E
Strain energy for producing only volume change is obtained by
substituting σav for σ1, σ2, and σ3 in to the above equation
3 av2
uv = (1 − 2v) (c )
2E
Substituting σav from Eq. (a),
1 − 2v 2
uv =
6E
(  1 +  22 +  32 + 2 1 2 + 2 2 3 + 2 3 1 )
(5 - 7)

Obtain distortion energy by subtracting volume changing energy,


Eq. (5–7), from total strain energy, Eq. (b)
1 + v  ( 1 −  2 ) 2 + ( 2 −  3 ) 2 + ( 3 −  1 ) 2 
u d = u − uv =
3E   (5 - 8)
2 
© McGraw Hill 31
Deriving the Distortion Energy 3

1 + v  ( 1 −  2 ) 2 + ( 2 −  3 ) 2 + ( 3 −  1 ) 2 
u d = u − uv =
3E   (5 - 8)
2 
Tension test specimen at yield has σ1 = Sy and σ2 = σ3 = 0
Applying to Eq. (5–8), distortion energy for tension test specimen is
1+ v 2
ud = Sy (5 - 9)
3E
DE theory predicts failure when the considered object’s distortion
energy, Eq. (5–8), exceeds distortion energy of tension test specimen,
Eq. (5–9)
12
 ( 1 −  2 ) + ( 2 −  3 ) + ( 3 −  1 ) 
2 2 2

   Sy (5 - 10)
 2 
© McGraw Hill 32
Von Mises Stress
12
 ( 1 −  2 ) + ( 2 −  3 ) + ( 3 −  1 ) 
2 2 2

   Sy (5 - 10)
 2 
Left hand side is defined as von Mises stress
12
 ( 1 −  2 ) + ( 2 −  3 ) + ( 3 −  1 ) 
2 2 2
 =   (5 - 12)
 2 
For plane stress, simplifies to
  = ( −  A B + 
2
A B)
2 12
(5 - 13)
In terms of xyz components, in three dimensions
1 12
 = ( x −  y ) + ( y −  z ) + ( z −  x ) + 6( xy +  yz +  zx )  (5 - 14)
2 2 2 2 2 2

2
For plane stress
( )
12
  =  −  x y +  + 3
2
x
2
y
2
xy (5 - 15)
© McGraw Hill 33
Distortion Energy Theory With Von Mises Stress

Von Mises Stress can be thought of as a single, equivalent, or


effective stress for the entire general state of stress in a stress
element.
Distortion Energy failure theory simply compares von Mises stress
to yield strength.
   Sy (5 - 11)
Introducing a design factor,
Sy
 = (5 - 19)
n
Expressing as factor of safety,
Sy
n=

© McGraw Hill 34
Octahedral Stresses

Same results obtained by evaluating octahedral stresses.


Octahedral stresses are identical on 8 surfaces symmetric to the
principal stress directions.
Octahedral stresses allow representation of any stress situation with
a set of normal and shear stresses.

Principal stress element with All 8 octahedral planes showing


single octahedral plane showing
Access the text alternative for slide images.

© McGraw Hill 35
Octahedral Shear Stress

Octahedral normal stresses are normal to the octahedral surfaces,


and are equal to the average of the principal stresses.
Octahedral shear stresses lie on the octahedral surfaces.
1 2 12
 oct = ( 1 −  2 ) + ( 2 −  3 ) + ( 3 −  1 ) 
2 2
(5 - 16)
3 Derivation see Arthur Boresi, op. cit., pp. 36-37

Fig. 5–10
Access the text alternative for slide images.

© McGraw Hill 36
Octahedral Shear Stress Failure Theory

Theory: Yielding begins when the octahedral shear stress in a


stress element exceeds the octahedral shear stress in a tension test
specimen at yielding.
The octahedral shear stress is
1 12
 oct = ( 1 −  2 ) 2 + ( 2 −  3 ) 2 + ( 3 −  1 ) 2  (5 - 16)
3
For a tension test specimen at yielding, σ1 = Sy , σ2 = σ3 = 0.
Substituting into Eq. (5–16),
2
 oct = Sy (5 - 17)
3
The theory predicts failure when Eq. (5–16) exceeds
Eq. (5–17). This condition reduces to
12
 ( 1 −  2 ) 2 + ( 2 −  3 ) 2 + ( 3 −  1 ) 2 
   Sy (5 - 18)
 2 
© McGraw Hill 37
Failure Theory in Terms of von Mises Stress

Equation is identical to Eq. (5–10) from Distortion Energy


approach.
Identical conclusion for:
• Distortion Energy.
• Octahedral Shear Stress.
• Shear Energy.
• Von Mises.
• Von Mises – Hencky.
Sy
n=

© McGraw Hill 38
DE Theory Compared to Experimental Data

Plot von Mises stress on


principal stress axes to compare
to experimental data (and to
other failure theories).
DE curve is typical of data.
Note that typical equates to a
50% reliability from a design
perspective.
Commonly used for analysis
situations.
MSS theory useful for design
situations where higher
reliability is desired. Fig. 5–15

Access the text alternative for slide images.

© McGraw Hill 39
Shear Strength Predictions 1

For pure shear loading, Mohr’s circle shows that σA = − σB = τ


Plotting this equation on principal stress axes gives load line for pure
shear case.
Intersection of pure shear load line with failure curve indicates shear
strength has been reached.
Each failure theory predicts shear strength to be some fraction of normal
strength.

Fig. 5–9
Access the text alternative for slide images.

© McGraw Hill 40
Shear Strength Predictions 2

For MSS theory, intersecting pure shear load line with failure
line [Eq. (5–5)] results in.
S sy = 0.5S y (5 - 2)

Fig. 5–9

© McGraw Hill 41
Shear Strength Predictions 3

For DE theory, intersection pure shear load line with failure


curve [Eq. (5–11)] gives Eq. 5-15 with x, y =0

(3 )
12 Sy
2
xy = S y or  xy = = 0.577 S y (5 - 20)
3
Therefore, DE theory predicts shear strength as
S xy = 0.577 S y (5 - 21)

Fig. 5–9

© McGraw Hill 42
Example 5–1 (1)

A hot-rolled steel has a yield strength of Syt = Syc = 100 kpsi and a true strain at
fracture of εf = 0.55. Estimate the factor of safety for the following principal
stress states:
(a) σx = 70 kpsi, σy = 70 kpsi, τxy = 0 kpsi
(b) σx = 60 kpsi, σy = 40 kpsi, τxy = −15 kpsi
(c) σx = 0 kpsi, σy = 40 kpsi, τxy = 45 kpsi
(d) σx = −40 kpsi, σy = −60 kpsi, τxy = 15 kpsi
(e) σ1 = 30 kpsi, σ2 = 30 kpsi, σ3 = 30 kpsi

Solution
Since εf > 0.05 and Syt and Syc are equal, the material is ductile and both the
distortion-energy (DE) theory and maximum-shear-stress (MSS) theory apply.
Both will be used for comparison. Note that cases a to d are plane stress states.

© McGraw Hill 43
Example 5–1 (2)

(a) Since there is no shear stress on this stress element, the normal stresses are
equal to the principal stresses. The ordered principal stresses are σA = σ1 = 70, σB =
σ2 = 70, σ3 = 0 kpsi.
DE From Equation (5–13),
  = 702 − 70 ( 70) + 702 
12
= 70 kpsi
From Equation (5–19),
Sy 100
Answer n= = = 1.43
  70
MSS Noting that the two nonzero principal stresses are equal, τmax will be
from the largest Mohr’s circle, which will incorporate the third principal stress at
zero. From Equation (3–16),
 −  3 70 − 0
 max = 1 = = 35 kpsi
2 2
From Equation (5–3),
S y 2 100 2
Answer n = = = 1.43
 max 35
© McGraw Hill 44
Example 5–1 (3)

(b) From Equation (3–13), the nonzero principal stresses are


60 + 40  60 − 40 
2

 A, B =  
  + ( − 15) 2
= 68.0, 32.0 kpsi
2 2
The ordered principal stresses are σA = σ1 = 68.0, σB = σ2 = 32.0, σ3 = 0 kpsi.

DE   = 68 − 68 ( 32) + 32 
2 2 12
= 59.0 kpsi
Sy 100
Answer n= = = 1.70
  59.0
MSS Noting that the two nonzero principal stresses are both positive, τmax
will be from the largest Mohr’s circle which will incorporate the third principal
stress at zero. From Equation (3–16),
 1 −  3 68.0 − 0
 max = = = 34.0 kpsi
2 2
S y 2 100 2
Answer n= = = 1.47
 max 34.0
© McGraw Hill 45
Example 5–1 (4)

(c) This time, we shall obtain the factors of safety directly from the xy components of stress.
DE From Equation (5–15),

( ) ( )
12
=  402 + 3( 45) 
12
  =  −  x y +  + 3 = 87.6 kpsi
2 2 2 2
x y
 xy

S y 100
Answer n= = = 1.14
  87.6
MSS Taking care to note from a quick sketch of Mohr’s circle that one nonzero
principal stress will be positive while the other one will be negative, τmax can be obtained
from the extreme-value shear stress given by Equation (3–14) without finding the principal
stresses.
x −y 
2
 0 − 40 
2

 max =   +  2
=   + 45 2
= 49.2 kpsi
 2 
xy
 2 
S y 2 100 2
n= = = 1.02
Answer  max 49.2
For graphical comparison purposes later in this problem, the nonzero principal stresses
can be obtained from Equation (3–13) to be 69.2 kpsi and −29.2 kpsi.

© McGraw Hill 46
Example 5–1 (5)

(d) From Equation (3–13), the nonzero principal stresses are

−40 + ( −60)  −40 − ( −60) 


2

 A, B = +   + (15) 2
= −32.0, − 68.0 kpsi
2  2 
The ordered principal stresses are σ1 = 0, σA = σ2 = −32.0, σB = σ3 = −68.0 kpsi.

  = ( −32) − ( −32)( −68) + ( −68) 


12
= 59.0 kpsi
2 2
DE

Sy 100
Answer n= = = 1.70
 59.0
MSS From Equation (3–16),
1 −  3 0 − ( −68.0)
 max = = = 34.0 kpsi
2 2
S y 2 100 2
Answer n= = = 1.47
 max 34.0
© McGraw Hill 47
Example 5–1 (6)

(e) The ordered principal stresses are σ1 = 30, σ2 = 30, σ3 = 30 kpsi.


DE From Equation (5–12),

 ( 30 − 30) + ( 30 − 30) + ( 30 − 30)


12
2 2 2

 =   = 0 kpsi
 2 
Sy 100
Answer n= = →
 0

MSS From Equation (5–3),

Answer Sy 100
n= = →
1 −  3 30 − 30

© McGraw Hill 48
Example 5–1 (7)

A tabular summary of the factors of safety is included for comparisons.


(a) (b) (c) (d) (e)
DE 1.43 1.70 1.14 1.70 ∞
MSS 1.43 1.47 1.02 1.47 ∞
Since the MSS theory is on or within the boundary of the DE theory, it will always predict a
factor of safety equal to or less than the DE theory, as can be seen in the table. For each
case, except case (e), the coordinates and load lines in the σA, σB plane are shown in Figure
5–11. Case (e) is not plane stress. Note that the load line for case (a) is the only plane stress
case given in which the two theories agree, thus giving the same factor of safety.

Fig. 5−11
Access the text alternative for slide images.

© McGraw Hill 49
Mohr Theory

Some materials have compressive strengths different from tensile


strengths.
Mohr theory is based on three simple tests: tension, compression, and
shear.
Plotting Mohr’s circle for each, bounding curve defines failure envelope.

Fig. 5–12
Access the text alternative for slide images.

© McGraw Hill 50
Coulomb-Mohr Theory 1

Curved failure curve is difficult to determine analytically.


Coulomb-Mohr theory simplifies to linear failure envelope using
only tension and compression tests (dashed circles).

Fig. 5–13
Access the text alternative for slide images.

© McGraw Hill 51
Coulomb-Mohr Theory 2

From the geometry, derive


the failure criteria.
B2C2 − B1C1 B3C3 − B1C1
=
OC2 − OC1 OC3 − OC1
B2C2 − B1C1 B3C3 − B1C1
=
C1C2 C1C3
B1C1 = St 2, B2C2 = ( 1 −  3 ) 2,
and B3C3 = Sc 2
1 −  3 St S c St
− −
2 2 = 2 2
St  1 +  3 St S c Fig. 5–13
− +
2 2 2 2
1 3
− =1 (5 - 22)
St Sc
© McGraw Hill 52
Coulomb-Mohr Theory 3

Incorporating factor of safety.


1 3 1
− = (5 - 26)
St S c n

For ductile material, use tensile and compressive yield strengths.


For brittle material, use tensile and compressive ultimate strengths.

© McGraw Hill 53
Coulomb-Mohr Theory 4

1  3
− =1 (5 - 22)
St S c
To plot on principal stress axes, consider three cases
Case 1: σA ≥ σB ≥ 0 For this case, σ1 = σA and σ3 = 0
• Eq. (5−22) reduces to
 A  St 1st quadrant (5 - 23)
Case 2: σA ≥ 0 ≥ σB For this case, σ1 = σA and σ3 = σB
• Eq. (5-22) reduces to
A B
− 1 4th quadrant (5 - 24)
St Sc
Case 3: 0 ≥ σA ≥ σB For this case, σ1 = 0 and σ3 = σB
• Eq. (5−22) reduces to
 B  − Sc 3rd quadrant (5 - 25)
© McGraw Hill 54
Coulomb-Mohr Theory 5

Plot three cases on principal stress axes.


Similar to MSS theory, except with different strengths for
compression and tension.

Fig. 5–14
Access the text alternative for slide images.

© McGraw Hill 55
Coulomb-Mohr Theory 6

Intersect the pure shear load line with the failure line to determine
the shear strength. Pure shear 1= − 3 = 
Torsional yield occurs when max = Ssy

Since failure line is a function of tensile and compressive strengths,


shear strength is also a function of these terms.
S yt S yc
S sy = (Set 1= − 3 = Ssy) (5 - 27)
S yt + S yc

© McGraw Hill 56
Example 5–2

A 25-mm-diameter shaft is statically torqued to 230 N · m. It is made of cast 195-T6


aluminium, with a yield strength in tension of 160 MPa and a yield strength in
compression of 170 MPa. It is machined to final diameter. Estimate the factor of safety of
the shaft.
Solution
The maximum shear stress is given by
16 ( 230)
=
16T
= = 75 10 6
N/m( )
2
= 75 MPa
(
 d 3   25 10−3  )
3

 
The two nonzero principal stresses are 75 and −75 MPa, making the ordered principal
stresses σ1 = 75, σ2 = 0, and σ3 = −75 MPa. From Equation (5–26), for yield,
1 1
Answer n= = = 1.10
 1 S yt −  3 S yc 75 160 − ( −75) 170
Alternatively, from Equation (5–27),
S yt S yc 160 (170)
S sy = = = 82.4 MPa
S yt + S yc 160 + 170
and τmax = 75 MPa. Thus, S sy 82.4
n= = = 1.10
Answer  max 75
© McGraw Hill 57
Example 5−3 (1)
A certain force F applied at D near the end of the 15-in lever shown in Figure 5–16,
which is quite similar to a socket wrench, results in certain stresses in the cantilevered bar
OABC. This bar (OABC) is of AISI 1035 steel, forged and heat-treated so that it has a
minimum (ASTM) yield strength of 81 kpsi. We presume that this component would be of
no value after yielding. Thus the force F required to initiate yielding can be regarded as the
strength of the component part. Find this force.

Fig. 5–16
Access the text alternative for slide images.

© McGraw Hill 58
Example 5−3 (2)

We will assume that lever DC is strong enough and hence not a part of the problem. A 1035
steel, heat-treated, will have a reduction in area of 50 percent or more and hence is a ductile
material at normal temperatures. This also means that stress concentration at shoulder A
need not be considered. A stress element at A on the top surface will be subjected to a
tensile bending stress and a torsional stress. This point, on the 1-in-diameter section, is the
weakest section, and governs the strength of the assembly. The two stresses are
M 32 M 32 (14 F )
x = = = = 142.6 F
I c d 3  13
( )
Tr 16T 16 (15 F )
 zx = = 3 = = 76.4 F
J d  13
( )
Employing the distortion-energy theory, we find, from Equation (5–15), that
  = ( + 3 ) = (142.6 F ) + 3( 76.4 F ) 
12
2 12
= 194.5 F
2 2 2
x zx  
Equating the von Mises stress to Sy, we solve for F and get
Sy
81 000
Answer F= = = 416 Ibf
lbf
194.5 194.5
© McGraw Hill 59
Example 5−3 (3)

In this example the strength of the material at point A is Sy = 81 kpsi. The strength of the
assembly or component is F = 416 lbf.
Let us apply the MSS theory for comparison. For a point undergoing plane stress with
only one nonzero normal stress and one shear stress, the two nonzero principal stresses will
have opposite signs, and hence the maximum shear stress is obtained from the Mohr’s
circle between them. From Equation (3–14)

 
2 2
 142.6 F 
 max =  x  +  zx2 = 
 2  2 
+ ( 76.4 F ) 2
= 104.5 F

Setting this equal to Sy/2, from Equation (5–3) with n = 1, and solving for F, we get

81 000 2
F= = 388 Ibf
lbf
104.5
which is about 7 percent less than found for the DE theory. As stated earlier, the MSS
theory is more conservative than the DE theory.

© McGraw Hill 60
Example 5−4 (1)

The cantilevered tube shown in Figure 5–17 is to be made of 2014 aluminum alloy
treated to obtain a specified minimum yield strength of 276 MPa. We wish to select
a stock-size tube from Table A–8 using a design factor nd = 4. The bending load is
F = 1.75 kN, the axial tension is P = 9.0 kN, and the torsion is T = 72 N · m. What
is the realized factor of safety?

Fig. 5–17
Access the text alternative for slide images.

© McGraw Hill 61
Example 5−4 (2)

Solution
The critical stress element is at point A on the top surface at the wall, where the bending
moment is the largest, and the bending and torsional stresses are at their maximum values.
The critical stress element is shown in Figure 5–17b. Since the axial stress and bending
stress are both in tension along the x axis, they are additive for the normal stress, giving
P Mc 9 120(1.75)(d o 2) 9 105d o
x = + = + = + (1)
A I A I A I
where, if millimeters are used for the area properties, the stress is in gigapascals.
The torsional stress at the same point is
Tr 72(d o 2) 36d o
 zx = = = (2)
J J J

Fig. 5−17

© McGraw Hill 62
Example 5−4 (3)

For accuracy, we choose the distortion-energy theory as the design basis. The von Mises
stress from Equation (5–15) is
  = ( + 3
2
x )
2 12
zx (3)
On the basis of the given design factor, the goal for σ′ is
Sy 0.276
  = = 0.0690 GPa (4)
nd 4
where we have used gigapascals in this relation to agree with Equations (1) and (2).
Programming Equations (1) to (3) on a spreadsheet and entering metric sizes from
Table A–8 reveals that a 42 × 5-mm tube is satisfactory. The von Mises stress is found to
be σ′ = 0.06043 GPa for this size. Thus the realized factor of safety is
Sy 0.276
Answer n= = = 4.57
  0.06043
For the next size smaller, a 42 × 4-mm tube, σ′ = 0.07105 GPa giving a factor of safety of
Sy 0.276
n= = = 3.88
  0.07105
© McGraw Hill 63
© McGraw Hill 64
Static Failure Theories 1

The generally accepted Failure Theories:

Ductile materials (yield criteria): Brittle materials (fracture criteria):

• Maximum shear stress • Maximum normal stress


• Distortion energy • Brittle Coulomb-Mohr
• Ductile Coulomb-Mohr • Modified Mohr

Done !

Access the text alternative for slide images.

© McGraw Hill 65
Failure Theories for Brittle Materials

Experimental data indicates some differences in failure for brittle


materials.
Failure criteria is generally ultimate fracture rather than yielding.
Compressive strengths are usually larger than tensile strengths.

Fig. 5–19
Access the text alternative for slide images.

© McGraw Hill 66
Maximum Normal Stress Theory 1

Theory: Failure occurs when the maximum principal stress in a


stress element exceeds the strength.
Predicts failure when
 1  Sut or  3  − Suc (5 - 28)

For plane stress,


 A  Sut or  B  − Suc (5 - 29)

Incorporating design factor,


Sut Suc
A = or B = − (5 - 30)
n n

© McGraw Hill 67
Maximum Normal Stress Theory 2

Plot on principal stress axes.


Not recommended for use (is not good as the other theories
mentioned below)

Fig. 5–18
Access the text alternative for slide images.

© McGraw Hill 68
Brittle Coulomb-Mohr

Same as previously derived, using ultimate strengths for failure.


Failure equations dependent on quadrant.

Quadrant condition Failure criteria

Sut
σA ≥ σ B ≥ 0 A = (5 - 31a )
n
A B1 Sut
σA ≥ 0 ≥ σB − = (5 - 31b)
Sut Suc n
Suc
0 ≥ σ A ≥ σB B = − (5 - 31c )
n
-Suc
Fig. 5–14
Access the text alternative for slide images.

© McGraw Hill 69
Brittle Failure Experimental Data

Coulomb-Mohr is conservative in 4th quadrant.


Modified Mohr criteria adjusts to better fit the data in the 4th
quadrant.

Fig. 5–19
Access the text alternative for slide images.

© McGraw Hill 70
Modified-Brittle Coulomb-Mohr
Quadrant condition Failure criteria
σA ≥ σB ≥ 0 Sut
A = (5 - 32a )
B n
A  0  B and 1 S
A  A = ut (5 - 32a )
n
 1 ( uc
B S − Sut )  A  B 1
A  0  B and − = (5 - 32b)
A Suc Sut Suc n
S
 B = − uc (5 - 32c )
0 ≥ σ A ≥ σB n

© McGraw Hill 71
Example 5−5 (1)

Consider the wrench in Example 5–3, Figure 5–16, as made of cast iron, machined to
dimension. The force F required to fracture this part can be regarded as the strength of the
component part. If the material is ASTM grade 30 cast iron, find the force F with
(a) Coulomb-Mohr failure model.
(b) Modified Mohr failure model.

Fig. 5–16

© McGraw Hill 72
Example 5−5 (2)

Solution
We assume that the lever DC is strong enough, and not part of the problem. Since grade
30 cast iron is a brittle material and cast iron, the stress-concentration factors Kt and Kts
are set to unity. From Table A–24 (next page), the tensile ultimate strength is 31 kpsi and
the compressive ultimate strength is 109 kpsi. The stress element at A on the top surface
will be subjected to a tensile bending stress and a torsional stress. This location, on the 1-
in-diameter section fillet, is the weakest location, and it governs the strength of the
assembly. The normal stress σx and the shear stress at A are given by
M 32 M 32 (14 F )
 x = Kt = Kt = (1) = 142.6 F Stress concentration
d 3  (1)
3
I c factors need to be
Tr 16T 16 (15 F )
 xy = K ts = K ts 3 = (1) = 76.4 F considered
d  (1)
3
J

From Equation (3–13) the nonzero principal stresses σA and σB are


142.6 F + 0  142.6 F − 0 
2

 A , B =  
  + ( 76.4 F ) 2
= 175.8F , − 33.2 F
2 2
This puts us in the fourth-quadrant of the σA, σB plane.
© McGraw Hill 73
© McGraw Hill 74
Example 5−5 (3)

(a) For BCM, Equation (5–31b) applies with n = 1 for failure.


A B 175.8 F ( −33.2 F )
− = − =1
Sut Suc 31 103 ( )
109 10 3
( )
Solving for F yields
Answer F = 167 Ibf
(b) For MM, the slope of the load line is ∣σB ∕σA∣ = 33.2 ∕ 175.8 = 0.189 < 1. Obviously,
Equation (5–32a) applies.
A 175.8F
= =1
Sut ( )
31 10 3

Answer F = 176 Ibf


As one would expect from inspection of Figure 5–19, Coulomb-Mohr is more
conservative.

© McGraw Hill 75
Selection of Failure Criteria

First determine ductile vs. brittle.


For ductile.
• MSS is conservative, often used for design where higher reliability is desired.
• DE is typical, often used for analysis where agreement with experimental data
is desired.
• If tensile and compressive strengths differ, use Ductile Coulomb-Mohr.
For brittle.
• Mohr theory is best, but difficult to use.
• Brittle Coulomb-Mohr is very conservative in 4th quadrant.
• Modified Mohr is still slightly conservative in 4th quadrant, but closer to
typical.

© McGraw Hill 76
Selection of Failure Criteria in Flowchart Form

Fig. 5−21
Access the text alternative for slide images.

© McGraw Hill 77
Introduction to Fracture Mechanics

Linear elastic fracture mechanics (LEFM) analyzes crack growth during


service.
Assumes cracks can exist before service begins, e.g. flaw, inclusion, or
defect.
Attempts to model and predict the growth of a crack.
Stress concentration approach is inadequate when notch radius
becomes extremely sharp, as in a crack, since stress concentration factor
approaches infinity. The limitation of this method
Ductile materials often can neglect effect of crack growth, since local
plastic deformation blunts sharp cracks.
Relatively brittle materials, such as glass, hard steels, strong aluminum
alloys, and steel below the ductile-to-brittle transition temperature,
benefit from fracture mechanics analysis.
https://www.nuclear-power.com/nuclear-engineering/materials-science/material-properties/ductility/ductile-brittle-
transition-temperature/
© McGraw Hill 78
© McGraw Hill 79
Quasi-Static Fracture 1

Though brittle fracture seems instantaneous, it actually takes time


to feed the crack energy from the stress field to the crack for
propagation.

A static crack may be stable and not propagate.

Some level of loading can render a crack unstable, causing it to


propagate to fracture.

© McGraw Hill 80
Quasi-Static Fracture 2

Foundation work for fracture mechanics established by Griffith in 1921.


Considered infinite plate with an elliptical flaw.
Maximum stress occurs at (±a, 0).
 a
( )
 y max = 1 + 2  
 b
( 5 - 33)
When a>>b ➔ y becomes inf.

Fig. 5−22
Access the text alternative for slide images.

© McGraw Hill 81
Quasi-Static Fracture 3

Crack growth occurs when energy release rate from applied loading
is greater than rate of energy for crack growth.

Unstable crack growth occurs when rate of change of energy


release rate relative to crack length exceeds rate of change of crack
growth rate of energy.

© McGraw Hill 82
Crack Modes and the Stress Intensity Factor

Three distinct modes of crack propagation.


• Mode I: Opening crack mode, due to tensile stress field.
• Mode II: Sliding mode, due to in-plane shear.
• Mode III: Tearing mode, due to out-of-plane shear.
Combination of modes possible.
Opening crack mode is most common, and is focus of this text.

Fig. 5−23
Access the text alternative for slide images.

© McGraw Hill 83
Derivation may check for details:
Mode I Crack Model Applied Mechanics of Solids
Chapter : Modeling Material Failure

Stress field on dx dy element at crack tip.


a   3 
x =  cos 1 − sin sin  (5 - 34a )
2r 2 2 2

a   3 
y = cos 1 + sin sin  (5 - 34b)
2r 2 2 2

a   3
xy =  sin cos cos (5 - 34c )
2r 2 2 2

0 ( for plane stress)


z = 
( )
(5 - 34d )
v  x +  y ( for plane strain )

Fig. 5−24 Access the text alternative for slide images.

© McGraw Hill 84
Stress Intensity Factor

Common practice to define stress intensity factor.


K1 =   a (5 - 35)
Incorporating KI, stress field equations are
KI   3 
x = cos 1 − sin sin  (5 - 36a )
2 r 2 2 2
KI   3 
y = cos 1 + sin sin  (5 - 36b)
2 r 2 2 2
KI   3
 xy = sin cos cos (5 - 36c )
2 r 2 2 2
0 ( for plane stress)
z = 
( )
(5 - 36d )
v  x +  y ( for plane strain )
© McGraw Hill 85
Stress Intensity Modification Factor 1

Stress intensity factor KI is a function of geometry, size, and shape


of the crack, and type of loading.
For various load and geometric configurations, a stress intensity
modification factor b can be incorporated.

K I = b  a (5 - 37)
Tables for b are available in the literature.
Figures 5−25 to 5−30 present some common configurations.

© McGraw Hill 86
Stress Intensity Modification Factor 2

Off-center crack in plate in


longitudinal tension.
Solid curves are for crack tip
at A.
Dashed curves are for tip at B.

Fig. 5−25
Access the text alternative for slide images.

© McGraw Hill 87
Stress Intensity Modification Factor 3

Plate loaded in longitudinal


tension with crack at edge.
For solid curve there are no
constraints to bending.
Dashed curve obtained with
bending constraints added.

Fig. 5−26
Access the text alternative for slide images.

© McGraw Hill 88
Stress Intensity Modification Factor 4

Beams of rectangular cross


section having an edge crack.

Fig. 5−27
Access the text alternative for slide images.

© McGraw Hill 89
Stress Intensity Modification Factor 5

Plate in tension containing circular hole with two cracks.

Fig. 5−28
Access the text alternative for slide images.

© McGraw Hill 90
Stress Intensity Modification Factor 6

Cylinder loaded in axial tension having a radial crack of depth a


extending completely around the circumference.

Fig. 5−29
Access the text alternative for slide images.

© McGraw Hill 91
Stress Intensity Modification Factor 7

Cylinder subjected to internal


pressure p, having a radial
crack in the longitudinal
direction of depth a.

Fig. 5−30
Access the text alternative for slide images.

© McGraw Hill 92
Fracture Toughness

Crack propagation initiates when the stress intensity factor reaches


a critical value, the critical stress intensity factor KIc.
KIc is a material property dependent on material, crack mode,
processing of material, temperature, loading rate, and state of stress
at crack site.
Also know as fracture toughness of material.
Fracture toughness for plane strain is normally lower than for
plain stress.
KIc is typically defined as mode I, plane strain fracture toughness.

© McGraw Hill 93
Typical Values for KIc

Table 5–1 Values of KIc for Some Engineering Materials at


Room Temperature
Material KIc, MPa m Sy, MPa
Aluminum
2024 26 455
7075 24 495
7178 33 490
Titanium
Ti-6AL-4V 115 910
Ti-6AL-4V 55 1035
Steel
4340 99 860
4340 60 1515
52100 14 2070

© McGraw Hill 94
Brittle Fracture Factor of Safety

Brittle fracture should be considered as a failure mode for


• Low-temperature operation, where ductile-to-brittle transition
temperature may be reached.
• Materials with high ratio of Sy /Su , indicating little ability to
absorb energy in plastic region.
A factor of safety for brittle fracture.

K Ic
n= (5 - 38)
KI

© McGraw Hill 95
Example 5−6

A steel ship deck plate is 30 mm thick and 12 m wide. It is loaded with a nominal uniaxial
tensile stress of 50 MPa. It is operated below its ductile-to-brittle transition temperature
with KIc equal to 28.3 MPa. If a 65-mm-long central transverse crack is present, estimate
the tensile stress at which catastrophic failure will occur. Compare this stress with the
yield strength of 240 MPa for this steel.
Solution
For Figure 5–25, with d = b, 2a = 65 mm and 2b = 12 m, so that d ∕ b = 1 and a ∕ d =
65 ∕ 12(103) = 0.00542. Since a ∕ d is so small, β = 1, so that
( )
K I =   a = 50  32.5  10−3 = 16.0 MPa m
From Equation (5–38),
K Ic 28.3
n= = = 1.77
K I 16.0
The stress at which catastrophic failure occurs is
K Ic 28.3
Answer c = = ( 50) = 88.4 MPa
KI 16.0
The yield strength is 240 MPa, and catastrophic failure occurs at 88.4 ∕ 240 = 0.37, or at 37
percent of yield. The factor of safety in this circumstance is KIc ∕ KI = 28.3 ∕ 16 = 1.77 and
not 240 ∕ 50 = 4.8.
© McGraw Hill 96
Example 5−7 (1)

A plate of width 1.4 m and length 2.8 m is required to support a tensile force in the 2.8-m
direction of 4.0 MN. Inspection procedures will detect only through-thickness edge cracks
larger than 2.7 mm. The two Ti-6AL-4V alloys in Table 5–1 are being considered for this
application, for which the safety factor must be 1.3 and minimum weight is important.
Which alloy should be used?
Solution
(a) We elect first to estimate the thickness required to resist yielding. Because σ = P∕wt,
we have t = P∕ wσ. For the weaker alloy, we have, from Table 5–1, Sy = 910 MPa. Thus,
Sy 910
 all = = = 700 MPa
n 1.3
Thus, 4.0 (10)
3
P
t= = = 4.08 mm or greater
w all 1.4 ( 700)
For the stronger alloy, we have, from Table 5–1,
1035
 all = = 796 MPa
1.3
and so the thickness is
4.0 (10)
3
P
Answer t= = = 3.59 mm or greater
w all 1.4 ( 796)

© McGraw Hill 97
Example 5−7 (2)
(b) Now let us find the thickness required to prevent crack growth. Using Figure 5–26, we
have
h 2.8 2 a 2.7
= =1 = = 0.001 93
b 1.4 ( )
b 1.4 103
Corresponding to these ratios we find from Figure 5–26 that β ≈ 1.1, and K I = 1.1  a .
K Ic 115 103 K Ic
n= = , =
K I 1.1  a 1.1n  a
K Ic = 115 MPa m
From Table 5–1, for the weaker of the two alloys.

Solving for σ with n = 1 gives the fracture stress


115
= = 1135 MPa
(
1.1  2.7  10 −3
)
which is greater than the yield strength of 910 MPa, and so
yield strength is the basis for the geometry decision.
Fig. 5−26
© McGraw Hill 98
Example 5−7 (3)

For the stronger alloy Sy = 1035 MPa, with n = 1 the fracture stress is

K Ic 55
= = = 542.9 MPa
(
nK I 1(1.1)  2.7  10−3 )
which is less than the yield strength of 1035 MPa. The thickness t is

t=
P
=
( )
4.0 103
= 6.84 mm or greater
w all 1.4 ( 542.9 1.3)

This example shows that the fracture toughness KIc limits the geometry when the stronger
alloy is used, and so a thickness of 6.84 mm or larger is required. When the weaker alloy
is used the geometry is limited by the yield strength, giving a thickness of only 4.08 mm
or greater. Thus the weaker alloy leads to a thinner and lighter weight choice since the
failure modes differ.

© McGraw Hill 99

You might also like