Sie sind auf Seite 1von 29

Mech 422 - Stress and

Strain Analysis
D.L. DuQuesnay
(adapted from original text
by R.J. Ferguson)

z
σ zz
σ xz σ yz
R R
1 σ zx
σ zy
2
σ yy
σ yx σ
σ xx xy
y

September 2002
Chapter 1

Introduction

1. What is Stress Analysis?


The aim of stress analysis is to take the geometry of a component or structure and the
externally applied “loads” and determine the state of stress in the material. Subjects that
encompass stress analysis include “strength of materials”, “solid mechanics”, mechanics
of deformable solids” , etc. When the stresses in the body are known, the material
properties are used in a failure theory to decide whether the body can withstand the
design loads. If this is the case, then further analysis may be undertaken to determine the
service life of the structure. In general, these analyses are accomplished by computation
with a calculator or a computer. However, stress analysis may also be performed
experimentally. Since the computer has invaded the design office, the importance of
experimental stress analysis has waned in recent times. It is now mainly used in the
determination of loads and stresses for in-service components or systems.
The following definitions are important:
• Stress analysis: used to find the stresses in a loaded body.
• Exact solution: is one that satisfies the conditions of equilibrium, compatibility
and meets the boundary conditions of the body and its loads. There are
comparatively few exact solutions in stress analysis.
• Closed-form solution: involves mathematical relationships that can be
manipulated by the ordinary operations of algebra and calculus to determine the
stresses. Many closed-form solutions involve simplifying assumptions that may
make the results invalid for some situations or for portions of a loaded body.
• Numerical solution: uses algorithms for equations that cannot be solved easily
(or at all) by conventional algebraic means.
• Finite element method: is a computer-based method wherein the body is divided
into discrete elements and equations solved on an element-by-element basis.
• Failure theories: are used with the state of stress, the geometry of the body and
candidate materials to determine the potential for failure within the desired
service life. Apart from comparing the stresses to the strength of the material,
such theories consider fracture, impact, fatigue, creep, etc.

2. Computer Programs
A multitude of computer programs exist for general purpose computation, stress analysis,
and for solving specific problems. In this course, the basic principles of stress analysis

1
will be introduced and illustrated with some worked examples. The method of solution
may include computer software to assist with computation. However, the emphasis is on
the development of the analytical models and analogues, the assumptions and the
limitations, and the stress analysis solutions for some typical structural systems that may
be encountered in the field.

3. Basic Concepts

PAUSE FOR THOUGHT


A force is a somewhat curious thing, as it cannot be measured directly. That is, we can
observe the effect of a force on its surroundings, but have no way of measuring the force
itself. The forces that we are likely to encounter, such as the load of a vehicle, or the
recoil of a gun, are typically measured using weigh-scales or accelerometers,
respectively. A weigh-scale measures force typically through the deflection of a spring
which is calibrated to load. Hence, when we use such a device to measure a force, we are
actually measuring a deflection or a distance. This is true of many quantities that are used
in stress analysis. Stress itself is not a measurable quantity, although the deformation of a
body may be measurable and may be calibrated back to stress. Hence, the term stress
analysis may be a misnomer as in most cases what we are doing is performing a
displacement or deformation analysis that is calibrated back to stress or force. In all
cases, there are assumptions in our analysis about the material properties that allow us to
make this calibration. The most useful assumption is that the material is Linear-Elastic
and obeys Hooke’s Law. This is valid within limits for most materials you are likely to
encounter. The second most useful assumption is that the displacements are relatively
small, which is also valid for most of the components and systems you are likely to
encounter. A third assumption is that the material is homogeneous and isotropic, meaning
that its composition and properties are the same throughout the component and in every
direction. This assumption is valid within limits for most polycrystalline materials (which
include engineering materials such as steel, cast iron, and aluminum alloys), but may not
be valid for some typical materials such as plastics and timber. Furthermore, composite
materials, which have gained widespread popularity in recent years, are intentionally
non-homogeneous and exhibit properties that are isotropic or orthotropic. Stress analysis
with these materials is inherently complex and is beyond the scope of this course.

3.1. Forces
A force is a quantity having a magnitude, a direction and a sense. A horizontal force of
300 kN acting to elongate a rod has a magnitude of 300 kN, its direction is horizontal and
its sense is tensile. Forces are vector quantities. For equilibrium, we require that the sums
of the forces in the three co-ordinate directions are zero. By convention tensile forces are
generally taken as positive (+) quantities, leading to stresses, displacements, strains, etc.
that are of like sense.

2
F2

F1
F3

In the above sketch, equilibrium is only achieved in the vertical direction if


F1 + F2 + F3 = 0
Note that forces can be added algebraically as long as they have the same direction. A
force that is neither entirely horizontal or vertical will have components in both of these
directions. The components of forces can, therefore, be added algebraically in any
direction.
Forces are often shown as distributed over a surface. Many closed-form solutions allow
for uniformly distributed loads or for those that are distributed in accordance with some
simple function.

Linearly distributed
Uniformly distributed

Distributed Loads

In checking equilibrium in such cases, an equivalent point load is used. The equivalent
point load has the same overall magnitude as the distributed load and acts at the centroid
of the distributed load, and hence, produces the same moment.

3.2. Moments
Equilibrium of forces is not sufficient to ensure equilibrium of a body. For example, in
the following figure if F1 = 300 kN then F2 and F3 can have any magnitudes that add to
300 kN and equilibrium of forces in the vertical direction will be satisfied. But, it is
obvious that F1 should be twice as large as F3 by considering the points of application of
the forces.

3
F2
1m

F1 3m
F3

We deal with the location of forces by using the principle that the sums of the moments
about the three co-ordinate axes must be zero. The moment of a force about a particular
axis is defined as the product of the force and the perpendicular distance between the axis
and the line of action of the force. If we take moments about an axis perpendicular to the
line of action of F1 (and the plane of the paper), we obtain
1 +
∑M F1 = 0 = −1 × F2 + 3 × F3 , hence F3 = F
3 2
Then, by equilibrium of forces in the vertical direction, F2 = 200 kN and F3 = 100 kN.
You can use the right-hand rule in categorising moments. To use the rule, think of your
thumb as the axis and your fingers as representing the force. Thus if your fingers point in
the same direction as F3, your thumb points out of the paper and represents a positive
moment (about F1). For F2, your thumb points into the paper representing a negative
moment. Moments can be added algebraically if they have the same direction.
Note also that moments are vectors. In the current example, the moment vectors for the
moments associated with if F1, F2 and F3 are perpendicular to the paper.

Try It!
Use the right-hand rule to help you decide what moments are created by vectors AB, BC,
CD and DE in the following sketch. Assume the cube has sides of unit length. State the
sign of each of the moments.

z
A
E C

y
D
x

4
3.3. St. Venant's principle
It is common to show loads applied at a point as has been done in previous sections. But,
loads must be distributed over a finite area. In practice it is quite difficult to determine
accurate stresses near the point of application of a load. Engineers deal with this problem
by applying St. Venant's principle which states that statically equivalent systems of
forces produce the same stresses and strains within a body except in the immediate region
where the loads are applied. Thus the stresses calculated in the middle of a beam are not
influenced by the way the ends are supported as long as the supporting forces and
moments are statically equivalent to those in the mathematical model. Local problems are
dealt with by separate methods.

3.4. Free Body Diagrams


An important aid in thinking clearly about problems in mechanics is the free body
diagram. In such a diagram, the body is considered by itself and the effect of the
surroundings on the body is shown by forces and moments. Free body diagrams are also
used to show internal forces and moments by cutting away the unwanted portion of a
body.

F2
1m M

V F3

F1 x

In this figure, the right-hand portion of the beam has been cut away and the aim is to find
the internal moment and vertical shear force at a distance x from the left-hand end.
Taking moments about the cross-section and applying equilibrium of forces and
moments, we obtain.
M + F2 ( x − 1) − F1 x = 0
V − F2 + F1 = 0
F3 = 0

With known values for F1 and F2, M and V can be found for a given x. Note F3 = 0 for
equilibrium of forces in the horizontal direction.

Try It!
If F1 = 200 kN and F2 = 100 kN, find M and V for x = 1.75 m. Can you also find them at x
= 0.75 m?

5
Boundary and End Conditions

It was assumed in the above example and discussion that there were no horizontal forces
acting or developing in our beam analysis. This would not be true in a real structure as
forces and moments would develop at the beam supports. The supports would offer some
moment resistance to the beam bending, and the change in length of the beam would pull
the supports closer together developing horizontal forces. To get around these
complexities, we may assume that certain releases exist at the boundaries (supports) of
structures wherever possible. It is then our responsibility to assure that these assumptions
are valid. Typical boundary conditions (releases) are:
1. Pin (no moment resistance)
2. Slider (no force resistance in one direction)
3. Free end (no resistance)
4. Fixed end (all degrees of freedom fixed)

PIN SLIDER FIXED PIN and SLIDER

4. Basic Load Cases


The formulas given in this section are applicable only to certain special cases as noted.
The material must be homogeneous and linearly elastic. The stresses are only due to
mechanical load and do not take residual stress or temperature into account.

4.1. Axial Loading


The critical parameter in determining the load intensity is the cross-sectional area
perpendicular to the load.
F
σ axial =
A
Here the cross-section is assumed to be constant or varying slowly along the length. The
load is applied along a line joining the centroids of the cross-sections. If the load is
constant along the length, the elongation of a bar under axial load is
FL
∆=
AE

6
Try It!

Consider a bridge truss with a span of 10m and a constant depth of 2m as shown in the
diagram below. If the members are all made of high strength aluminum alloy tubing 100
mm in diameter with a wall thickness of 10 mm, determine the maximum stress in the
truss for the loading condition shown.

500kN

4.2. Bending
In bending it is assumed that the cross-sections of the beam remain plane and the bending
moment is directed along a principal axis of inertia of the beam. The material of the beam
must have the same modulus in tension and compression so that the stress will vary
linearly with distance from the neutral surface.
The formula for bending stress is only exact for pure bending of long slender beams
where the depth is greater than the width. In short beams, shear stress effects are more
significant. In the limit, wide beams become plates which have a separate analysis
procedure

My d 2v M
σ= =
I dx 2 EI

The second moment of area or moment of inertia is the critical section parameter.

Moment of Inertia:

7
A
2
I xx = Σ A y
y
X X h 3
For a rectangle, I xx = bh /12
Neutral Axis

Try it!

Determine the maximum bending stress in the welded girder shown below.

w = 10 kN/m
300 x 25 mm

300x8 mm

300 x 25 mm
10 m

4.3. Torsion
In torsion the shear stress and the angle of twist are given by simple formulas for circular
cross-sections. The section may be hollow but the hole most be round and concentric
with the outside circumference. Plane sections are assumed to remain plane.

Tr TL
τ= θ=
J JG
The polar moment of area or polar moment of inertia is the critical section parameter.

8
Polar Moment of Inertia for typical solid cross-sections:
π D4
Circle: J =
32
where D = Diameter, Do = Outer diameter, and Di = inner diameter.
π ( Do4 − Di4 )
Tube: J =
32

Try it!

Determine the angle of twist for the steel shaft shown below. Calculate the safety factor
against yield. The shear strength of the steel is 200 MPa, and the shear modulus is 87000
MPa.

D=50 mm

500 mm T = 2500 Nm

4.4. Disks and Cylinders


Stress analysis solutions for problems where there is axial symmetry are, generally
speaking, exact from the point of view of the theory of elasticity. You must remember
that local effects such as keyways, rifling, and so forth can cause large stress
concentrations.

Rotating solid disk of outer radius ro

9
3+ν
σ rr = ρω 2 (ro2 − r 2 )
8
ρω 2
σθθ =
8
[ (3 + ν )ro2 − (1 + 3ν )r 2 ] (1)
3+ν
σ max = ρω 2 ro2 (at centre)
8

Rotating disk of outer radius ro and hole radius ri

3+ν ri 2 ro2
σ rr = ρω (ro + ri − 2 − r 2 )
2 2 2
8 r
3+ν 2 2 ri 2 ro2 1 + 3ν 2 
σθθ = ρω ro + ri + 2 −
2
r (2)
8  r 3 + ν 
3+ν  1 − ν  ri 2  
σθθ = ρω 2 ro2 1 +  2  ( max value at ri )
4  3 + ν  ro  

Thick cylinders subjected to external and internal pressure (plane strain or plane
stress), Lamé’s equations:

pi ri 2 − po ro2 ri 2 ro2 ( po − pi ) 1
σ rr = +
ro2 − ri 2 ro2 − ri 2 r2
(3)
pi ri 2 − po ro2 ri 2 ro2 ( po − pi ) 1
σθθ = −
ro2 − ri 2 ro2 − ri 2 r2

The relationship between diametral interference δ and interference pressure p is

 (1 + ν1 )ri 2 + (1 − ν1 )rb2 (1 + ν2 )ro2 + (1 − ν2 )rb2 


δ = 2rb p  +  (4)
 E1 (rb2 − ri 2 ) E 2 (ro2 − rb2 ) 

where 1 refers to the inner cylinder and 2 to the outer cylinder. The radius rb is that at the
interface or boundary between the two cylinders. If both cylinders are of the same
material, then

4rb3 (ro2 − ri 2 ) p
δ= (5)
(rb2 − ri 2 )(ro2 − rb2 ) E

10
To assemble two cylinders, the outer must be warmer than the inner by an amount

δ
∆T = (6)
2rbα

Elastic-perfectly-plastic cylinder subject to a pre-stress pressure ps

r   r 
σ rr = S y ln  − ps and σθθ = S y 1 + ln   − ps (7)
 ri    ri  

The pressure at the elastic-plastic interface can be found in two ways

r  S y  rep2 
pep = S y ln i  + ps or pep = 1 −  (8)
 rep  2  ro 

The stress distribution in the outer or elastic part of the cylinder can be found by Lamé’s
equations.

Try It!
(1) The tube used in the M109 A2 self-propelled howitzer has an outside diameter at the
breach of 11.2 inches. The inner diameter is 6.1 inches in the smooth part of the bore and
6.2 inches in the rifled part. If the steel has a yield strength of 1100 MPa, what is the
factor of safety for the gun barrel if the design pressure1 is 4400 bar (1 bar = 14.5038 psi)
and stress concentration effects are ignored?
(2) If you redesign the barrel as a compound cylinder with a radial interference of 0.010
inches at an intermediate diameter of 8.6 inches, what will the factor of safety be then?
(Use E = 205 GPa, Sy = 1100 MPa, ν = 0.3). If α = 11×10−6 /°C What temperature
difference is required between the two parts of the barrel in order to permit assembly?
(3) Optimize the design from part (2) so that the demands on the material in the two parts
of the barrel are the same.

1
As a matter of interest, the rated maximum pressure for the cannon is 3475 bar. The design pressure thus provides a
margin of safety in itself.

11
(4) Plot the stress distributions in the compound gun barrel for the barrel subjected to the
design firing pressure and also for the case when the assembly or interference stresses are
acting alone.
(5) Redesign the barrel assuming an autofrettage manufacturing process such that the
elastic-plastic boundary is in the middle of the wall thickness. Plot all the stresses.

5. Stress and Strain


A clear understanding of stress and strain is vital to any work in stress analysis.

5.1. Definition of stress

S = δF/ δA
δF

δA

Stress is the force per unit area where the area is infinitesimally small so that variations
of stress from point to point can be monitored. The force δF is the portion of the total
force applied to the infinitesimal area δA. The force does not have to be perpendicular to
the area. For the general case, the force is inclined at some angle ψ to the normal to the
plane of area δA. Thus we write,
δF
S = lim , S n = S cosψ and S s = S sinψ
δA→0 δA

where the subscripts s and n refer to the normal and shear stress components of the stress.

12
Try It!
In a beam it's not unusual for the stress to go from 400 MPa to 0 in a distance of 75 mm.
How big should δA be so that the stress does not vary by more than 1% across the face?
Because the stress at a point depends on the magnitude of the force and the orientation of
the plane containing δA, it is obviously impossible to specify stress with a single number.
It can be shown that stress is defined unambiguously when the stresses are known on
three orthogonal planes. These are usually shown in the form of a cube with stresses
given as the normal and shear components on the cube faces. In the following diagram,
the stresses are all positive.

z
σ zz
σ xz σ yz
σ zy
σ zx
σ yy
σ yx σ
σ xx xy
y

Although the cube is large when viewed on the printed page, don't forget that it
represents the stress state at a point. For visualisation purposes, you can think of it as the
size of the point on a sharp pencil.
There are nine separate components of stress on the cube as shown in the following
equation,
σ σ xy σ xz 
 xx 
σ ij =  σ yx σ yy σ yz 

 σ zx σ zy σ zz 

Fortunately, only six of them are independent because σxy = σyx, etc.

13
5.2. Stresses on an oblique plane.
5.2.1. One-dimensional case
Consider a square bar loaded in uniaxial tension where a section has been cut out of the
bar in the central region where the stress is uniform.

Sn
Ss
é
Fx Fx åx
é S = Sx

If A is the area of the inclined face on which S acts, then we can use equilibrium of forces
in the x-direction to write,
σ x A cosθ = S x A
Now, because the normal and shear components of S are given by
S n = S x cosθ and S s = S x sin θ ,

we can see that these components are related to the stress created by the externally
applied load according to
S n = σ x cos2 θ and S s = σ x sin θ cosθ

These expressions reflect the significance of the orientation of the plane on the
magnitude of the stress components.

5.2.2. Three-dimensional case


If you pass a cutting plane through a point in a body, its orientation is determined by the
angles made by the normal to the plane with the co-ordinate axes. These are commonly
specified as direction cosines, that is, the cosines of these angles.

Sn Ss
σ xy σ xx
σ yy
S
σ zy σ zx
y
σ zz
x

In the general case, the resultant stress S is not perpendicular to the


plane and so both the normal stress Sn and the shear stress Ss are non-
zero.

14
Try It!
Use Stresses on an oblique plane to find the stresses and their directions on a plane
oriented at 45° to the load axis in a rod under a tensile stress of 250 MPa (see figure,
Section 4.1).

5.3. Definition of strain


Strains are classified as either direct or shear. Direct or normal strain is the change of
length of a linear element divided by the original length. Shear strain is a change of shape
conventionally defined as the tangent of α. But, because the angles are assumed to be
small, tan α = α when α is measured in radians.

α
L δL
Direct strain Shear strain

δL
εn = and γ = tan α ≈ α
L
We have seen that six stresses are needed to define the stress at a point. Hence, it is not
surprising that six strains are associated with them. We can write (following the example
of Section 4.2) the strains at a point as:

 ε 1
γ xy 1
γ xz 
 xx 2 2

εij =  21 γ yx ε yy 2 γ yz 
1

1
 2 γ zx 2 γ zy
1
εzz 

Note that the six strains are not all independent because they are defined in terms of only
three displacements. Also, the factor of 1/2 multiplying the shear strains is present
because of a conflict between the traditional definition of shear strain and the tensor
definition.

5.4. Converting from stress to strain


If a material is isotropic, homogeneous and linearly-elastic in its stress-strain behaviour,
Hooke's law can be used to convert from stress to strain and vice-versa.

15
ε xx =
1
[ ]
σ − ν (σ yy + σ zz + α∆T
E xx
1
[ ]
ε yy = σ yy − ν (σ zz + σ xx + α∆T
E
1
[ ]
εzz = σ zz − ν (σ xx + σ yy + α∆T
E
2(1 + ν )
γ xy = σ xy
E
2(1 + ν )
γ yz = σ yz
E
2(1 + ν )
γ zx = σ zx
E
In these expressions, E is the elastic or Young’s modulus, ν is Poisson's ratio and α is the
coefficient of linear expansion. In an unrestrained body, an increase in temperature leads
to strains but no stress.
In relating shear strain and shear stress, it is common to write γ xy = σ xy G , where G is
the shear modulus. G is not another independent property because,
E
G=
2(1 + ν )

Example
Two circular rods, one made of aluminum and the other of copper, are fixed to rigid walls
such that there is a gap of 0.2 mm between them when the temperature is 20°C. Each rod
has a diameter of 30 mm. Given αAl = 13×10−6 /°F, αCu = 9.4×10−6 /°F, EAl = 10,000 ksi,
ECu = 18,000 ksi and ν = 0.33 for both materials, answer the following questions:

(a) At what temperature will the gap between the ends of the rods close?
(b) What is the complete state of stress in each rod when the temperature is 200°F?
(c) What is the complete state of strain in each rod at 200°F?

0.2mm

Copper Aluminum

100 mm 200 mm

16
5.5. Manipulating stress and strain tensors
You can only manipulate quantities like stress and strain, which have six components in
their definition, with special procedures. We have seen evidence of the complexity that
this produces in Section 4.2. We do not need to plunge into the details of the mathematics
of tensors, however, you should remember that techniques learned from manipulating
vectors do not work with tensors. You must use the methods to be described next.

5.5.1. Principal stresses


The six stress components constitute a symmetric matrix with three principal values and
directions also known as eigenvalues and directions. In stress analysis, these values are
called principal stresses and directions. They are significant because they represent
extreme values of the normal stresses which are seen as controlling the onset of failure
according to many failure theories. When we refer to a principal element, we mean an
element that has been rotated into a position where the faces are free of shear stress and
hence the normal stresses on the faces are the principal values.

5.5.2. Mohr's circle for stress


The graphical technique called Mohr’s circle was used by earlier generations of engineers
to make calculations. Thanks to the computer, graphical methods are no longer needed
for this purpose. However, Mohr’s circle survives because it helps us to visualise the
relationship between normal and shear stress and the co-ordinate system for a loaded
body.
The Mohr’s circle method is based on the observation that the normal and shear stresses
on an oblique plane trace out a circle when plotted against each other. To verify this,
recall the one-dimensional case in Section 4.2.1 where we established that
S n = σ x cos2 θ and S s = σ x sin θ cosθ

for a uniaxial stress field. If you set σx = 200 MPa and plot Ss as a function of Sn , you’ll
obtain a semicircle where a 15° rotation of the plane produces a 30° angular displacement
in the semicircle. So, if we are going to use the semicircle as a guide to stress variation,
we must remember that angles are doubled.

17
Shear Stress as a Function of Normal Stress
(+ = 15° increments)
100 +
90
+ +
80
70
60
50 + +
40
30
Ss (MPa)
20
10
0+ +
0 20 40 60 80 100 120 140 160 180 200
Sn (MPa)

In the traditional presentation of Mohr’s circle, the rotation of the plane is carried through
180° to generate a complete circle. This requires a special sign convention for shear
stress for the purpose of plotting the circle. Here are the steps to follow:
• Draw the 2-D element representing the state of stress at the point of interest.
• Plot the points representing the stresses on the vertical and horizontal faces of the
element using the convention that if the shear stress arrow tends to rotate the
element CCW, plot the point below the Sn axis. If the arrow tends to rotate the
element CW, plot the point above the Sn axis.

( CW )
Ss

å yy
H face
å xy
H
å xx Sn
å yy V
å xx

V face

( CCW )

With this convention, you will obtain a diameter from which the circle can be drawn as
shown in the figure. The circle intersects the Sn axis at two of the three principal stress
values. Note that the shear stress does not change sign as the element is rotated about the
principal axis.

18
5.5.3. 3-D Mohr’s circles
In the explanation in the previous section, two important issues were skipped:
• The Mohr’s circle method works only when there is a single shear stress acting on
the faces of the cube. The front face in the 2-D view of the element must be free
of shear stress—it must be a principal plane.
• Because there are three principal stresses, there are always three circles.
You must never forget that there are three circles because the 2-D element you choose to
examine may not be the critical one.

5.5.4. Example

For the following state of stress, find the principal and critical values. Also, determine the
stresses on the plane stress element after a rotation of 12°.

 120 50 0
 
σ ij =  50 80 0 MPa
 
 0 0 0

Mohr's Circles (V & H given, v & h rotated)


80

60
H
Shear Stress (MPa)

40 h

20

-20

-40 v
V
-60

-80
-25 0 25 50 75 100 125 150 175
Normal Stress (MPa)

The plot shows that the circle created from the 2-D element is not the largest and hence
the maximum shear stress cannot be found without noting that the third principal stress is
zero. This leads to a true maximum shear stress of 76.9 MPa.

Try It!
(1) A helicopter weighing 20,000 lb is supported by a main rotor drive shaft in which the
axial stress is 25,000 psi. If the torque in the shaft develops a shear stress of 18,000 psi,
what are the values of maximum normal and shear stresses in the shaft (in MPa)?

19
(2) A 150 mm diameter pulley is mounted on the extreme end of a 25 mm diameter shaft.
Belt tensile forces of 1500 lb and 500 lb act parallel to each other at the top and bottom
of the pulley. Find the critical stress values (and sketch the corresponding elements) for
the stresses in the shaft at a point 12 mm from the plane of the belt forces. The stresses
are: bending (max) = 10.4 ksi, torsion (max) = 15.3 ksi and shear (max) = 3.4 ksi.

5.6. Strain gauge rosettes


Strain gauges are frequently used to measure strains on components under load. In cases
where a component is failing unexpectedly, engineers will often install strain gauges to
verify that the load conforms with the designers’ assumptions.

Strain gauges are made from fine wire or foil and are bonded to a free surface on the
component. Measurement of the change in resistance of a gauge from the nominal 120 Ω
is converted to strain using constants supplied by the manufacturer. If the strain is non-
uniform, as is often the case in grooves and fillets, then the gauges should be as small as
possible so that the strain can be measured accurately over a short distance. Typical strain
gauges are 6 mm to 25 mm in length for most metal structures although gauge lengths of
a half a millimeter or less are possible. One other thing to note is that strain gauges are
sensitive to temperature changes, and different materials are used in the manufacture of
resistance foil strain gauges either for aluminum or steel structures.
Many factors influence the quality of strain measurement that cannot be dealt with here.
We will assume that normal precautions have been taken to ensure maximum accuracy.
Gauges bonded to a free surface can only measure direct strain. If t6he strain field is
uniaxial. Then a single gauge may be used to measure the strain field, provided the
material properties are known. However, if the strain field is 2-dimensional, then we
require two direct strains and the shear strain on two orthogonal planes to define the
strain. To measure these, the strain gauges are normally applied in the form of a three
member rosette. By solving three simultaneous equations, the three direct strains are
converted into two direct strains and one shear strain. The corresponding stresses are
found with Hooke’s law.

20
y y

C B
B A
120° x
120°
45° C
45° A
x

Rectangular rosette Equiangular rosette

The challenge in using strain gauges is to convert their readings into stresses so that an
assessment of the loading conditions can be made. In the slide rule era, Mohr’s circle
graphical techniques were used to solve for the principal strains and then convert them to
stresses.

Example

−0.0005
T
P P
+0.0013
−0.0010 T

A 2 inch diameter shaft made of steel (E = 205 GPa, υ = 0.3) is subjected to a torque T
and axial load P of unknown magnitude and direction. The strain gauge readings are as
shown. What are the stresses, torque and axial load on this shaft?
F Tr
σx = τ xy =
A J
πd2 πd4
A= J=
4 32
The plots for this model are as follows:

21
osette in Relation to Principal Axe Rosette Strains on Mohr's Circle

Tensor Shear Strain (æî)


(Strains are ABSOLUTE values)
2000
c b a c
1500
1000
Principal axis

500
0
-500
-1000
b
-1500
b a c
-2000
c
a
b -2000 0 2000 4000

a Normal Strain (æî)

Principal axis

Mohr's Stress Circles


"a" gauge direction, H 90 deg to "a
300
Shear Stress (MPa)

V
200
100
0
-100
-200
H
-300
-200 0 200 400
Normal Stress (MPa)

The torque and the axial load are 6020 Nm and 525 kN, respectively

22
5.7. Failure and Theories of Failure
A structure or component fails when it can no longer function as intended. Many words
are used in ordinary conversation to describe this phenomenon: Worn out, broken, rusted
out, bent, stretched out of shape, kaput, etc. This is not generally the vocabulary of
engineers who need more precise terms to describe failure. In stress analysis we deal with
modes of failure that can be related to stress, strain and equilibrium considerations2.
1. Failure due to fracture. Fracture is a process in which cracks grow to failure. The
cracks may be pre-existing or they may develop through micro-mechanical processes
such as void coalescence during loading. Fracture can occur under constant load or under
varying load. In the later case, we use the term fatigue failure. In brittle fracture the
failure is accompanied by only a moderate amount of plastic deformation. The fracture
surface is perpendicular to the maximum principal stress.
2. Failure due to yielding. In this mode of failure there is extensive permanent plastic
deformation. This may change the geometry of the object to such an extent that it cannot
fulfill its function. A suspension part bent such that the vehicle will not track properly, is
an example of a failure due to yielding.
3. Failure due to low stiffness. Some components must meet deflection rather than
strength criteria. A frame of a vehicle must have adequate stiffness to maintain alignment
in drivetrain components. Resonant vibration may occur at undesirable frequencies.
4. Buckling. is the loss of stable equilibrium. That is, the structure or part has a sudden
decrease in stiffness. Compressive loading can lead to bucking in columns and shells.
5. Creep. is a high-temperature problem that occurs when the ambient temperature is over
50 percent of the melting temperature of the material (expressed in K). In creep, plastic
strain replaces elastic strain and detracts from the load-carrying capacity of the structure.

Try It!
The melting points of aluminum, iron, lead and polycarbonate are 660°C, 1537°C, 327°C
and 225°C. How do you assess the potential for creep in these materials? Would alloying
make any difference?

5.7.1. Theories for failure by yielding


Of the many theories available in the literature, we will consider only four that apply to
isotropic homogeneous materials. You should note that agreement between theory and
practice is not perfect. Some of the theories work well with certain materials for
particular loading conditions, but not with others. Engineers compensate for this with
factors of safety. Often these are built into codes such as the building code or the codes
for pressure vessel design and certification.

2
Cook, R.D. and Young, W.C. Advanced Mechanics of Materials, Macmillan, 1985.

23
1. The Tresca criterion, which is also known as the maximum shear stress theory is based
on the observation that yielding is begins when the shear stress reaches its maximum
value. In a tensile test this occurs when the diameter of the largest Mohr’s circle is equal
to the tensile yield strength. If the principal stresses are ordered such that σ1 > σ2 > σ3,
then the Tresca criterion is expressed as

σ1 − σ 3 = S y

2. The von Mises criterion. is based on a more complex view of the role of the principal
stress differences. In simple terms, the von Mises criterion considers the diameters of all
three Mohr’s circles as contributing to the characterization of yield onset in isotropic
materials.
When the criterion is applied to the uniaxial tensile test, its relationship to the yield
strength is,
(σ1 − σ 2 ) 2 + (σ 2 − σ 3 ) 2 + (σ 3 − σ1 ) 2 = 2 S y2

3. The maximum principal stress criterion states that a tensile fracture failure will occur
in a previously un-cracked isotropic material when the maximum principal stress reaches
a critical value. Note that this criterion does not characterize fracture in brittle materials
with cracks. In that case, the role of the crack length is highly significant.
4. Mohr’s failure criterion applies to some brittle materials that are much stronger in
compression than in tension. Data from tension and compression tests establish limiting
Mohr’s circles. For any given stress state, failure will not occur if the largest Mohr’s
circle lies within the failure envelope.

Ss

Failure envelope

Sc St
Sn

24
Try It!
Tension and compression tests of a brittle material give failure strengths of Sc = 140 MPa
and St = 14 MPa. If the loading on the body produces a state of plane stress with σxx = 0,
σyy = −18 MPa and σxy = 20 MPa, assess the risk of failure according to two failure
theories.

5.7.2. Buckling Failure

Structural elements under compression are susceptible to a failure mode known as


buckling which may occur prior to yielding of the material. Buckling failures are
common in columns and in the compression flanges of girders. The Euler buckling
formula can be used in many cases to predict the critical load that causes buckling for
most column configurations. The buckling failure is at first elastic, provided the
geometric changes do not lead to progressive collapse of the structure. Hence, the column
may return to its original straight configuration when the load is removed. Try it with
your ruler!

π 2 EI
Euler Buckling Load, Pcr =
( k L) 2

where I is the moment of inertia of the cross section about its weakest axis. The factor, k,
depends on the end conditions for the column as follows:

P P P P

k=1 k =2 k = 0.5
k = 0.75

25
6. Finite Element Analysis
It is very difficult to develop equations to model real structures exactly. When such
equations can be developed, their solution is often difficult or impossible. In the finite
element method (FEM), no attempt is made to find a mathematical model valid
everywhere in the structure. Instead, the structure is divided into elements in which a
simple and locally valid mathematical model is used. A good analogy to the FEM is the
common technique of modeling a curve with a series of short straight lines.

6.1. Advantages of the FEM


• Any boundary can be modeled.
• Each element could, in theory, have its own material properties.
• Much cheaper than “build and try” approach

6.2. Disadvantages of the FEM


• Needs considerable computer power
• Good FEM package is costly
• Difficult to use FEM software without proper training
• Results easily invalid if support conditions improperly modeled.
• Needs experienced analyst for modeling and interpretation of results
• Time consuming and expensive

6.3. The steps in the FEM

1. Divide the structure into finite elements


Element

Node

Elements are connected at nodes, which have 1,2, 3,… degrees of freedom (dof).

26
1
Triangular element showing
2 dof per node

3
2

2. Define the properties of each element by obtaining the element stiffness matrices.
3. Assemble the elements to obtain the finite element approximation of the structure.
4. Apply loads to the node points.
5. Specify how the structure is to be supported.
6. Solve simultaneous linear equations to find the nodal dofs (usually displacements) in
terms of the loads and the stiffness matrix of the structure.
7. Use displacements to compute the stresses

6.4. What to watch for in FEM results


The most common element types in commercial software are based on an assumption
about the shape the element will adopt under load. The element shape function defines
the strain within the element and from the strain the stress can be found. Unfortunately,
strains are calculated by differentiating displacements which leads to a magnification of
errors. Most FEM models give more accurate results for displacements than for strains
and stresses.
log|error| vs log of the step size in differentiation of sin x at pi
0

-2

-4

-6
log (error)

-8

-10

-12

-14
-16 -14 -12 -10 -8 -6 -4 -2 0
log(h)
Figure showing how error in numerical differentiation drops as step size decreases, but
that it increases again if the step size is excessively small.

The design of the mesh is critical. More elements are needed in regions of rapidly
varying stress, but elements that are too small lead to numerical errors.

27
In a thorough FEM analysis, different mesh sizes or element types are tried to check for
convergence. Here is an example where a tip-loaded cantilever beam has been modeled
with 1,2,4 and 8 elements. Each element in the model had eight nodes with two DOF
each.

Covergence in a FEM Model


(S = stress, D = Displacement)
100 D
S D
S
90 D
FEM/Exact (percent)

S
80 D
70
60 S
50
40
30
20
10
0
0 10 20 30 40 50 60 70 80
Degrees of freedom in the model

Finally, and it cannot be repeated too often, badly modeled support conditions can lead to
gross errors. The consumer of FEM results must be wary.

*********

28

Das könnte Ihnen auch gefallen