Sei sulla pagina 1di 153

MAE 477/577

- EXPERIMENTAL TECHNIQUES
IN SOLID MECHANICS -

CLASS NOTES
Version 5.0
by
John A. Gilbert, Ph.D.
Professor of Mechanical Engineering
University of Alabama in Huntsville

FALL 2013

Table of Contents
Table of Contents

Introducing MAE 477/577

iv

MAE 477/577 Outline

MAE 477 ABET Syllabus

vi

MAE 577 SACS Syllabus

viii

Chapter 1. Stress
1.1 Introduction
1.2 Types of Forces
1.3 Traction Vector, Body Forces/Unit Mass, Surface and Body Couples
1.4 Resolution of the Traction Vector Stress at a Point
1.5 Equilibrium Equations Conservation of Linear Momentum
1.6 Stress Symmetry Conservation of Angular Momentum
1.7 Transformation Equations Mohrs Circle
1.8 General State of Stress
1.9 Stresses in Thin Walled Pressure Vessels
1.10 Homework Problems

1.1
1.1
1.1
1.3
1.5
1.7
1.7
1.16
1.19
1.22

Chapter 2. Strain
2.1 Introduction
2.2 Strain-Displacement Equations
2.3 Strain Equations of Transformation
2.4 Compatibility Equations
2.5 Constitutive Equations
2.6 Homework Problems

2.1
2.1
2.2
2.3
2.3
2.5

Chapter 3. Linear Elasticity


3.1 Overview
3.2 Plane Stress
3.3 Plane Strain
3.4 Homework Problems

3.1
3.1
3.2
3.6

Chapter 4. Light and Electromagnetic Wave Propagation


4.1 Introduction to Light
4.2 Electromagnetic Spectrum
4.3 Light Propagation, Phase, and Retardation
4.4 Polarized Light
4.5 Optical Interference
4.6 Complex Notation
4.7 Intensity
4.8 Superposition of Wavefronts
4.9 Reflection and Refraction
4.10 Double Refraction Birefringence
4.11 Homework Problems

4.1
4.3
4.4
4.5
4.6
4.6
4.7
4.8
4.11
4.12
4.14

Chapter 5. Photoelasticity
5.1 Introduction
5.2 Plane Polariscope
5.3 Circular Polariscope
5.4 Calibration Methods for Determining the Material Fringe Value
5.5 Compensation Methods for Determining Partial Fringe Order Numbers
5.6 Calculation of the 1 Direction
5.7 Birefringent Coatings
5.8 Three-Dimensional Photoelasticity
5.9 Homework Problems
5.10 Classroom Demonstration in Photoelasticity

5.1
5.2
5.6
5.10
5.12
5.14
5.15
5.17
5.18
5.24

Chapter 6. Photography
6.1 Introduction
6.2 Single Lens Reflex Cameras
6.3 Depth of Field
6.4 Photographic Processing
6.5 Polaroid Land Film Cameras
6.6 Digital Cameras

6.1
6.1
6.2
6.3
6.5
6.9

Chapter 7. Brittle Coatings


7.1 Introduction
7.2 Testing Procedures
7.3 Calibration
7.4 Measurements
7.5 Coating Selection
7.6 Application
7.7 Crack Patterns
7.8 Relaxation Techniques

7.1
7.2
7.3
7.8
7.9
7.11
7.11
7.11
ii

7.9 Homework Problems

7.12

Chapter 8. Moir Methods


8.1 Introduction
8.2 Analysis
8.3 Optical Filtering
8.4 Stress Analysis
8.5 Shadow Moir
8.6 Homework Problems

8.1
8.2
8.4
8.5
8.6
8.8

Chapter 9. Strain Gages


9.1 Introduction
9.2 Electrical Resistance Strain Gages
9.3 Length Considerations
9.4 Transverse Sensitivity Corrections
9.5 Temperature Considerations
9.6 Rosettes
9.7 Gage Selection and Series
9.8 Dummy Gages and Transducer Design
9.9 P-3500 Strain Indicator
9.10 Two-Wire and Three-Wire Strain Gage Circuits
9.11 Homework Problems

9.1
9.2
9.4
9.5
9.5
9.6
9.8
9.9
9.11
9.13
9.15

Appendix I Review of Statics and Mechanics of Materials


A.1.1 2-D Forces in Rectangular Coordinates
A.1.2 3-D Rectangular Coordinates
A1.3 Addition of Concurrent Forces in Space/3-D Equilibrium of Particles
A1.4 Moments and Couples
A1.5 Equilibrium of Rigid Bodies
A1.6 Center of Gravity and Centroid
A1.7 Moments of Inertia
A1.8 Shear and Moment Diagrams
A1.9 Deflection of Beams
A1.10 Synopsis of Mechanics of Materials
A1.11 Homework Problems
Appendix II Basic Formulas for Mechanics of Materials

iii

A1.1
A1.1
A1.2
A1.3
A1.4
A1.4
A1.6
A1.8
A1.8
A1.10
A1.13
A2.1

Introducing MAE 477/577


EXPERIMENTAL TECHNIQUES IN SOLID MECHANICS - 3 hrs.
Required Text: Experimental Techniques in Solid Mechanics, Version 4.1 by Gilbert
Reference Text: Experimental Stress Analysis, 4th. Edition by Dally and Riley
Prerequisite: MAE/CE 370; Junior Standing
Overview:
When dealing with the majority of real engineering systems, it is not always sufficient, or
advisable, to rely on analytical results alone. The experimental determination of stress, strain,
and displacement is important in both design and testing applications. MAE 477/577,
Experimental Techniques in Solid Mechanics, presents techniques which are valuable
complements to the design and analysis process; and, in some difficult or complex situations,
provide the only practical approach to a real solution.
The bulk of the course constitutes a detailed treatment of the more conventional methods
currently used for experimental stress analysis (photoelasticity, brittle coatings, moir methods,
strain gages, etc.); however, more recent developments in the field are also introduced (hybrid
methods, speckle metrology, holographic interferometry, moir interferometry, fiber optics,
radial metrology, and STARS). In-class laboratory exercises are included so that students gain
some practical experience. The lecture and laboratory exercises are designed to provide enough
exposure for participants to secure an entry level position in the field of experimental mechanics.
Attendance:
Students are required to be present from the beginning to the end of each semester, attend all
classes, and take all examinations according to their assigned schedule. In case of absence,
students are expected to satisfy the instructor that the absence was for good reason. For
excessive cutting of classes (3 or more class periods), or for dropping the course without
following the official procedure, students may fail the course.
Homework:
Homework assignments must be done on only one side of 8 1/2" x 11" paper. Each problem
shall begin on a separate page and each page must contain the following (in the upper right hand
corner): Your name, the date, and page __ of __. All final answers must be boxed and converted
(SI to US or US to SI). Homework is due at the beginning of the class on the date prescribed.
Work must be legible and should be done in pencil. Problems shall be restated prior to solution
and free-body diagrams (FBDs) shall be drawn for problems requiring such. Loose sheets shall
be stapled together in the upper left hand corner.
Note: For further information on the course and registration procedure, call Professor Gilbert by
telephone at (256) 824-6029, or, contact him by e-mail at jag@eng.uah.edu.

iv

MAE 477/577 Experimental Techniques in Solid Mechanics


Fall 2013
Time/Place:
Instructor:
Office:
Telephone:
Fax/E-mail:
Office Hours:
Grading:

Friday 8:00 a.m. - 10:40 a.m.; TH S117


Dr. John A. Gilbert
OB 301E
(256) 824-6029 (direct); (256) 824-5117/5118 (Beth/Cindy)
(256) 824-6758; jag@eng.uah.edu
To be announced.
10% Attendance (100% for being in class; 0% for not being there); 30%
Homework and Labs; 30% Midterm (10/11/13); 30% Final (11/29/13).

Required Text:

Gilbert, J.A., Experimental Techniques in Solid Mechanics, Volume 5.0,


University of Alabama in Huntsville, 2013.
Dally, J.W., Riley, W.F., Experimental Stress Analysis, 4th Edition,
College House Enterprises, 2005, ISBN 0-9762413-0-7.

Reference Text:

Course Outline
1.

Overview of Solid Mechanics - Stress, Strain, and Displacement; Equilibrium,


Transformation, Strain-Displacement, Compatibility, and Constitutive Equations.

2.

Stress Analysis - Method of Attack; Stress Concentration, Failure, Design, and Examples.

3.

Light and Electromagnetic Wave Propagation Amplitude, Phase, Polarization,


Coherence, Interference, Reflection, Refraction, Birefringence, and Stress Optic Law.

4.

Photoelasticity Plane and Circular Polariscopes; Calibration and Compensation


Methods; Reflection Polariscope, Birefringent Coatings, and 3-D Photoelasticity.

5.

Photography Cameras, Lenses, and Photographic Development; Digital Cameras,


Image Acquisition and Processing.

6.

Brittle Coatings Theory, Calibration, Application, and Measurement.

7.

Moir Methods Geometrical Considerations and In-Plane Displacement Measurement;


Diffraction, Optical Filtering, and Stress Analysis; Out-of-Plane Displacement
Measurement and Shadow Moir.

8.

Electrical Resistance Strain Gages - Parametrical Studies, Transverse Sensitivity,


Rosettes, Circuitry, Installation, and Transducer Design.

9.

Advanced Topics Moir Interferometry, Speckle Metrology, Holographic


Interferometry, Fiber Optics, Digital Image Processing, Hybrid Methods, Panoramic
Imaging Systems, Radial Metrology, and STARS.
v

MAE 477 Experimental Techniques in Solid Mechanics


Responsible Department:
Mechanical and Aerospace Engineering
Catalog Description:
477 Experimental Techniques in Solid Mechanics 3 hrs.
Experimental methods to determine stress, strain, displacement, velocity, and
acceleration in various media. Theory and laboratory applications of electrical
resistance strain gages, brittle coatings, and photoelasticity. Application of
transducers and experimental analysis of engineering systems. (Same as CE 477)
Prerequisites:
MAE/CE 370 and junior standing.
Textbook:
Required Text:

Reference Text:

Gilbert, J.A., Experimental Techniques in Solid


Mechanics, Volume 5.0, University of Alabama in
Huntsville, 2013.
Dally, J.W., Riley, W.F., Experimental Stress Analysis,
4th Edition, College House Enterprises, 2005, ISBN 09762413-0-7.

Course Objectives:
1. To present techniques that are valuable complements to the design and
analysis process; and in some difficult or complex situations, provide the
only practical approach to a real solution.
2. The bulk of the course constitutes a detailed treatment of the more
conventional methods currently used for experimental stress analysis
(photoelasticity, brittle coatings, moir methods, strain gages, etc.);
however, more advanced topics are introduced (hybrid methods, speckle
metrology, holographic interferometry, moir interferometry, fiber optics,
radial metrology, and STARS).
3. A portion of the course is devoted to laboratory work.
Topics Covered:
1. Overview of solid mechanics
2. Stress analysis
3. Light and electromagnetic wave propagation
4. Photoelasticity
5. Photography
6. Brittle coatings
7. Moir methods
8. Electrical resistance strain gages
9. Advanced topics

vi

Class Schedule:
Once per week; class is 2 hours 40 minutes.
Contribution of Course to Meeting the Professional Component:
Basic Mathematics & Science:
0 credits.
Engineering Science:
3 credits.
Engineering Design:
0 credits.
Relationship of Course to Program Outcomes:
In this course each undergraduate student will have to:
a) apply their knowledge of mathematics, science, and engineering;
b) apply a knowledge of calculus-based physics;
c) apply advanced mathematics through differential equations;
d) apply a knowledge of linear algebra;
e) design a system, component or process to meet desired needs;
f) work professionally in mechanical systems;
g) use the techniques, skills, and modern engineering tools necessary for
engineering practice;
h) conduct experiments and to analyze and interpret data;
i) identify, formulate and solve engineering problems;
k) communicate effectively.
Person Preparing this Description:
John A. Gilbert, Ph.D., Course Coordinator, Professor
MAE Program
28 November 2012

vii

MAE 577 Experimental Techniques in Solid Mechanics


Responsible Department:
Mechanical and Aerospace Engineering
Catalog Description:
577 Experimental Techniques in Solid Mechanics 3 hrs.
Experimental methods to determine stress, strain, displacement, velocity, and
acceleration in various media. Theory and laboratory applications of electrical
resistance strain gages, brittle coatings, and photoelasticity. Application of
transducers and experimental analysis of engineering systems. (Same as CE 577)
Prerequisites:
MAE 370 and junior standing
Textbook:
Required Text:

Reference Text:

Gilbert, J.A., Experimental Techniques in Solid


Mechanics, Volume 5.0, University of Alabama in
Huntsville, 2013.
Dally, J.W., Riley, W.F., Experimental Stress Analysis,
4th Edition, College House Enterprises, 2005, ISBN 09762413-0-7.

Course Educational Objectives:


1. To present techniques that are valuable complements to the design and
analysis process; and in some difficult or complex situations, provide the
only practical approach to a real solution.
2. The bulk of the course constitutes a detailed treatment of the more
conventional methods currently used for experimental stress analysis
(photoelasticity, brittle coatings, moir methods, strain gages, etc.);
however, more advanced topics are introduced (hybrid methods, speckle
metrology, holographic interferometry, moir interferometry, fiber optics,
radial metrology, and STARS).
3. A portion of the course is devoted to laboratory work.
Topics Covered:
1. Overview of solid mechanics
2. Stress analysis
3. Light and electromagnetic wave propagation
4. Photoelasticity
5. Photography
6. Brittle coatings
7. Moir methods
8. Electrical resistance strain gages
9. Advanced topics

viii

Class Schedule:
Once per week; class is 2 hours 40 minutes.
Relationship of Course to Program Outcomes:
In this course each graduate student will have to:
a) apply mathematics to solve engineering problems;
b) successfully perform advanced analysis in a technical area;
c) communicate effectively.
Person Preparing this Description:
John A. Gilbert, Course Coordinator, Professor
MAE Program
28 November 2012

ix

CHAPTER 1 - STRESS
1.1 Introduction
When dealing with the majority of real engineering systems, it is not always sufficient, or
advisable, to rely on analytical results alone. The experimental determination of stress, strain,
and displacement is important in both design and testing applications. Experimental Techniques
in Solid Mechanics presents techniques which are valuable complements to the design and
analysis process; and, in some difficult or complex situations, provide the only practical
approach to a real solution.
The bulk of the course constitutes a detailed treatment of the more conventional methods
currently used for experimental stress analysis (photoelasticity, brittle coatings, moir methods,
strain gages, etc.); however, more recent developments in the field are also introduced (hybrid
methods, speckle metrology, holographic interferometry, moir interferometry, fiber optics,
radial metrology, and STARS). In-class laboratory exercises are included so that students gain
some practical experience. The lecture and laboratory exercises are designed to provide enough
exposure for participants to secure an entry level position in the field of experimental mechanics.

1.2 Types of Forces


The forces which act to produce stress in a material are classified according to their manner of
application. Surface forces act on the surface of the stressed body and are usually generated
when two bodies come into contact. Body forces, on the other hand, act on each element of the
body and are generated by force fields such as gravity. In a great many cases, the body forces
are small as compared to the surface forces and can be neglected in the analysis.

1.3 Traction Vector, Body Force/Unit Mass, Surface and Body Couples
When a solid body is subjected to external loads, the material from which the body is made
resists and transmits these loads. The strength of the body is measured in terms of stress.
Stress represents force intensity and is a mathematical quantity that must be computed by
dividing an applied force by the area over which the force acts. For a given force; the smaller the
area, the higher the stress. In general, the stress is different at every point in a loaded body.
The stress distribution at a given point is relatively complex and may be represented by a second
order tensor. Each direction in space (typically defined by a unit vector perpendicular to a plane
passed through the point in question) is associated with a stress vector which can be resolved into
components normal (normal stress) and parallel (shear stress) to the plane.

1.1

Figure 1, for example, shows an arbitrary internal or external surface with a small area A
enclosing a point P. As mentioned above, the orientation of the plane is defined with respect to a
reference axes system by a unit vector, n, drawn perpendicular to the surface. In general, the
distributed forces acting on A can be reduced to a force-couple system at point P where Fn
and Mn are the resultant force and couple, respectively. In most cases, the resultant force, Fn,
is not coincident with n and the couple vector lies at an arbitrary angle with respect to the line of
action of the resultant force.

Figure 1. Forces and moments act on a small area surrounding a point.


The average stress (force intensity) acting over the area is:
(1.3-1)
The traction (or resultant stress) vector, tn, at point P is found by taking the limit of the quantity
in Equation (1.3-1) as A goes to zero:
(1.3-2)

Since the magnitude and direction of the terms on opposite sides of Equation (1.3-2) are equal,
the traction vector is aligned with Fn. Simply put, tn depends upon the point in question and the
plane defined by n.
The surface couple at point P is given by:

(1.3-3)
Figure 2, on the other hand, shows a three-dimensional body with point Q surrounded by a
volume element V having a corresponding mass, m. FB and MB are the resultant body
force and body couples acting on the volume.
1.2

Figure 2. Forces and moments act on a small volume corresponding to a given mass.
Following arguments similar to those taken in obtaining Equations (1.3-2) and (1.3-3), the
resultant body force per unit mass acting at point Q is:
(1.3-4)
whereas, the resultant body couple per unit mass is:
(1.3-5)
While formulating the governing equations, surface and body couples are almost always
neglected. As a consequence, higher order terms disappear thereby simplifying the equations.

1.4 Resolution of the Traction Vector - Stress at a Point


Figure 3 shows the traction vector resolved into components normal and tangent to the plane
under consideration. The normal component, , is called the normal stress, whereas the
tangential component, , is called the shear stress.

Figure 3. The traction vector can be resolved into normal and tangential components.

1.3

A more meaningful description is obtained when one axis of a Cartesian system is aligned with
the normal to the plane under consideration. Figure 4, for example, shows the traction vector
acting at point P on a plane having a normal n acting along the Z axis. In this case, the tangential
component of the traction vector can be broken into two in-plane components. A double
subscript notation is used to define the stress components; the first subscript denotes the normal
to the surface under consideration while the second subscript denotes the direction in which the
stress component acts. For the configuration shown in Figure 4, the shear and normal stress
components are denoted by zx, zy and zz, respectively.

Figure 4. The traction vector is resolved into three Cartesian components.


Two additional sets of components are obtained for planes having their normal along the X and
Y axes. The corresponding stress components are xx,xy,xz and yx,yy,yz , respectively. The
stress distributions on these three mutually perpendicular planes can be summarized in a stress
tensor as follows:

(1.4-1)

In general, the magnitude and direction of the traction vector at a point depend upon the plane
considered. The rows in the stress tensor, on the other hand, correspond to the scalar
components of the traction vectors which act on planes having their normal vector acting along
the X,Y,Z axes, respectively.
Even though an infinite number of traction vectors may be defined by passing different planes
through a single point, each can be characterized in terms of the 9 components of stress defined
on three mutually perpendicular planes. Mathematically, the traction vector, tn, on a plane
defined by its normal, n, may be determined from the stress tensor, , using:
.

1.4

(1.4-2)

The infinitesimal region surrounding a point can be modeled by a rectangular parallelepiped. If


the sides are aligned with a Cartesian axes system, the 9 components of stress can be graphically
represented.
Figures 5, 6, and 7, for example, show the stress components on planes with their outer normals
parallel to the X,Y,Z axes, respectively. In this case, each parallelepiped is infinitesimal and the
stress distribution on opposing faces is equal and opposite, thus maintaining equilibrium.

Figure 5. Stresses on planes


with their normal along X.

Figure 6. Stresses on planes


with their normal along Y.

Figure 7. Stresses on planes


with their normal along Z.

By convention, stress components are drawn along the positive coordinate direction when the
outer normal to the plane on which they act lies along a positive coordinate direction; otherwise,
the component is drawn along the negative coordinate direction.
Each component is labeled using the double subscript notation described above. Normal stresses
acting away from the element correspond to tensile stresses while normal stresses acting toward
the element correspond to compressive stresses.

1.5 Equilibrium Equations - Conservation of Linear Momentum


The overall distribution of forces acting on the body must be such that the stress distribution
satisfies equilibrium. Figure 8, for example, shows a small element of volume taken from a
stressed body.

Figure 8. A finite element taken from a stressed body.

1.5

As opposed to the parallelepipeds depicted in Figures 5-7, the one is Figure 8 has finite
dimensions. To simplify the analysis, only the stress components which act in the X direction
have been included on the figure along with Fx, the body force per unit mass.
Assuming that the body is in equilibrium:
.

(1.5-1)

The stresses can be converted to forces by multiplying each stress by the area of the face over
which it acts and the body force per unit volume can be converted to force by multiplying it by
the volume. When this is done and Equation (1.5-1) is applied:

(1.5-2)

Dividing by the volume,


(1.5-3)
When a similar argument is applied to analyze the stresses along Y:
(1.5-4)
and for those along Z:
(1.5-5)
Summarizing,

(1.5-6)

The expressions in Equation (1.5-6) are referred to as the equilibrium equations.


1.6

1.6 Stress Symmetry - Conservation of Angular Momentum


Neglecting surface and body couples and taking moments about an XYZ axes located at the
centroid of an element subjected to stresses:

(1.6-1)

The expressions in Equation (1.6-1) show that the stress tensor, , is symmetrical. Hence, there
are 6 independent components of stress.

1.7 Transformation Equations - Mohr's Circle


In Section 1.4, it was shown that for every point in a loaded member, the components of shear
and normal stress change as different planes are passed through the point. Even though an
infinite number of planes can be passed through the point, it is sufficient to determine only the
stresses on three mutually perpendicular planes to completely characterize the state of stress on
an arbitrary plane. There are, however, certain critical planes, the analysis of which help to
determine the structural integrity. This section considers the transformation equations required
to pinpoint the orientations of these critical planes and the corresponding stresses which act on
them.
This exercise is important in experimental mechanics where the first step is usually to identify
critical points on the loaded structure. For an interior point, the measurements must be made on
three mutually perpendicular planes. In general, the stresses obtained on these planes are not the
maximum stresses at the point under consideration and a coordinate rotation is required.
When the stress tensor is formulated for a Cartesian axes system and the axes system is rotated, a
new stress tensor, ', is obtained. This is illustrated in Figures 9 and 10.
The stress tensor is modified because the shear and normal stress components on the rotated
planes are different from those on the original planes. However, the stresses on the rotated
planes may be obtained from those on the original planes by applying the law of transformation
for a tensor of the second rank. Mathematically, the transformation takes the form,
(1.7-1)
where ij' and kl are the components of the rotated and original tensors, respectively; ik and jl
correspond to the terms contained in the matrix of direction cosines and in its transpose,
respectively.

1.7

Figure 10. The stress distribution referred to


a rotated X'Y'Z' axes system.

Figure 9. The stress distribution referred to


an XYZ axes system.

Fortunately, most experimental measurements are taken on a free surface where the components
of traction on the plane defined by the outer normal vanish. This condition is often referred to as
a plane, 2-D, or, biaxial stress state. In this case, for a point contained in the XY plane, the stress
tensor becomes:

(1.7-2)

From Equation (1.7-2) it is apparent that for plane stress, there are 3 independent stress
components. Thus, from an experimental mechanics standpoint, the complete determination of
the stress distribution at the point requires that 3 independent measurements be made.
The stress transformation can be studied analytically using Equation (1.7-1) or graphically by
manipulating these expressions and using Mohr's circle. Consider, for example, the stress
element shown in Figure 11.
Equation (1.7-1) can be applied to determine the stresses on the rotated element shown in Figure
12 as follows:

(1.7-3)

1.8

Figure 12. The stress distribution on a plane


stress element referred to a rotated X'Y' axes.

Figure 11. The stress distribution on a plane


stress element referred to an XY axes system.

Applying the expressions in Equation (1.7-3) to the configuration shown in Figure 12, noting
that the X and Y axes correspond to = 0o and 90o, respectively,

(1.7-4)

The expressions in Equation (1.7-4) are referred to as the transformation equations for a point in
plane stress.
An alternate approach is to manipulate each of the expressions included in Equation (1.7-3) as
follows:
(1.7-5)
and
(1.7-6)
Adding Equations (1.7-5) and (1.7-6):

1.9

(1.7-7)
Equation (1.7-7) is the equation of a circle drawn in - space. This stress circle is commonly
referred to as Mohr's circle.
In X-Y space, the expression
(1.7-8)
corresponds to a circle centered at x = a and y = b with radius c.
Comparing the terms in Equation (1.7-7) to those in Equation (1.7-8), the Mohr's circle in -
space is centered at:
(1.7-9)
with radius
(1.7-10)

As illustrated in Figure 13, the circle is drawn by plotting points X with coordinates xx and xy
and Y with coordinates yy and -xy. Normal stresses are assumed positive when they act away
from the element; positive shear stresses produce a counterclockwise rotation of the element.

Figure 13. A typical Mohr's circle for graphical transformation of stress.

1.10

Each radial line in the circle represents the orientation of the unit vector normal to one of the
infinite number of planes which can be passed through the point. The intersection of the line
with the circle defines the shear and normal stresses acting on that plane. The orientation of the
planes are usually measured with respect to the X axis and are determined by computing angles
on the circle and dividing these values in half. The resulting angle is measured in the opposite
sense (clockwise or counterclockwise) on the element.
Points A and B are of special interest, since they correspond to the principal planes on which the
principal stresses act. Note that the shear stress is zero at these points. Other points of interest
lie at D and E where the maximum in-plane shearing stress occurs. As discussed later in Section
1.8, this value is the absolute maximum shearing stress when the principal normal stresses are of
opposite sign (i.e., the Mohrs circle encompasses the origin). Note that the normal stress for
both points corresponds to the location of the center of the circle.
The values of normal and shear stress at these critical locations can be computed graphically, as
demonstrated in the example problem outlined at the end of this section, or calculated on the
basis of Equation (1.7-3) as described below.
The orientation of the principal planes with respect to the X axis, p, are obtained by
differentiating the first expression in Equation (1.7-3) with respect to and setting the result
equal to zero. This operation produces,
(1.7-11)

The magnitudes of the principal stresses are found by substituting the values found from
Equation (1.7-11) back into the first expression in Equation (1.7-3). This operation produces:

(1.7-12)

A similar argument can be applied to the second expression in Equation (1.7-3). That is, the
planes of maximum in-plane shear stress are oriented at s with respect to the X axis, where
(1.7-13)

Substituting the values in Equation (1.7-13) into the expressions in Equation (1.7-3) produces:

(1.7-14)

It should be noted, as illustrated in the following example, that the principal planes are oriented
at an angle of 45o with respect to the planes of maximum shear.

1.11

Example:
(a)
(b)
(c)
(d)
(e)

For the stress state shown in Figure 14, determine the:

orientation of the principal planes.


magnitude of the principal stresses.
orientation of the planes of maximum in-plane shearing stress.
magnitude of the maximum in-plane shearing stress.
magnitude of the normal stress acting on the maximum in-plane shear planes.

Figure 14. A point in a state of plane stress.


Solution:

The problem can be solved analytically. From Equation (1.7-11):


(1)

.
Thus, (a) the principal planes are oriented at:
.

(2)

From the first expression in Equation (1.7-12):

(3)

1.12

.
Thus, (b) the principal stresses are:
.

(4)

It is not apparent, however which of the stresses in Equation (4) correspond to the orientations
given in Equation (2). To determine the correspondence, it is necessary to substitute the values
in Equation (2) into the first expression in Equation (1.7-3). This operation shows that max lies
at p = - 13.3o while min corresponds to p = 76.7o. It should be noted that, from the second
expression in Equation (1.7-12), the shear stresses on the principal planes are zero.
From Equation (1.7-13):

(5)
.
Thus, (c) the orientations of the maximum in-plane shearing stresses are:
.

(6)

From the first expression in Equation (1.7-14), (d) the maximum in-plane shearing stress is:

(7)
.
Again, it is not apparent which value of s corresponds to max. This can only be determined by
substituting the values in Equation (6) into the second expression of Equation (1.7-3). This
process reveals that max corresponds to s = 31.7o; min = - max lies at s = 121.7o.
From the second expression in Equation (1.7-14), (e) the corresponding normal stress is:
(8)
The graphical solution is obtained by drawing the Mohr's circle shown in Figure 15. This is
accomplished by plotting the two points corresponding to the planes labeled as 1 and 2 on
Figure 14 in - space. The points are connected and a circle is drawn centered at C with radius
r. The location of the center of the circle is computed from Equation (1.7-9) as:
(9)

1.13

Figure 15. A Mohrs circle can be drawn for the point in question.
The radius is computed from Equation (1.7-10) as:
(10)

The maximum and minimum values of stress are determined by adding and subtracting the
magnitude of the radius to the value of at point C, respectively. This operation produces:
(11)
.
The orientations of the principal planes are obtained by determining the orientation of the point
labeled 3 on Figure 15 with respect to the point labeled 1. Referring to the figure:
(12)
Note that on the circle, 2 is drawn counterclockwise from point 1 to point 3. Also note that
point 3 corresponds to min. Therefore, the normal to the plane corresponding to min lies at =
13.3o measured in the clockwise direction from the normal to the plane labeled as 1 in Figure 14.
These results are summarized in Figure 16 which shows the rotated element corresponding to
the principal planes.

1.14

Figure 16. The principal stresses and their


orientations.

Figure 17. The maximum and minimum inplane shear stresses and their orientations.

Referring to Figure 15, the value of at the point labeled 5 is equal to the radius while the
corresponding value of is equal to the normal stress of point C. At the point labeled 5:
(13)
.
The orientation of the maximum in-plane shearing stresses may be determined by finding the
angle labeled 2 on Figure 15 as:
.

(14)

Note that 2 is drawn from point 1 to point 5 in a clockwise direction. Thus, on the element, the
plane corresponding to the maximum in-plane shearing stress lies at = 31.7o measured
counterclockwise from the plane labeled as 1 on Figure 14. Since the shear is positive, it is
drawn such that it creates a counterclockwise rotation on the element. These results are
summarized in Figure 17.
It is apparent by comparing Figures 16 and 17 that the planes of maximum in-plane shearing
stress are oriented at 45o with respect to those corresponding to the principal stresses. It should
also be noted from Figures 14-17 that the sum of the normal stresses acting on two
perpendicular planes is constant. The sum corresponds to the trace of the stress tensor (xx + yy
+ zz) which is one of three quantities which remain invariant during coordinate rotation (in this
case, about the Z axis).

1.15

1.8 General State of Stress


As discussed in Sections 1.4 and 1.6, the stress tensor consists of 9 components of stress, 6 of
which are independent. The 9 values correspond to 3 normal and 6 shear stress components
acting on three mutually perpendicular planes.
As illustrated in Figure 18, the coordinate axes can be rotated, in much the same manner as in
plane stress, to determine principal planes on which the shear stress is zero. In mathematical
terms, this is called an eigenvalue problem. One obtains a characteristic equation which has
three roots called eigenvalues; each eigenvalue is equal in magnitude to one of the principal
stresses. Eigenvectors are determined for each eigenvalue, and these vectors define the
orientation of the normal to each of the principal planes on which the principal stresses act.
Figure 19 shows one possible configuration for graphically depicting the stress distribution at a
point. In this case, the three eigenvalues are all positive. Coordinate rotations around the three
eigenvectors are pictorially represented by the three circles, and points on their circumferences
define the stress distribution on different planes on a rotated element. A more general coordinate
transformation leads to a point inside of the area between the inner circles and the outer circle.
In all cases,

(1.8-1)

Figure 19. The Mohrs circle for an arbitrary


point at which the three eigenvalues are all
positive.

Figure 18. A rotated element showing the


three principal stresses.

1.16

Two cases are of interest when evaluating Equation (1.8-1) for the case of plane stress (when one
of the eigenvalues is zero).
Figure 20 shows that when the principal stresses obtained from the bi-axial transformation (the
2-D Mohr's circle shown as the solid line in the figure) are of opposite sign, the in-plane
maximum shear stress is the absolute maximum shear stress at the point under study. As
illustrated in Figures 21 and 22, the corresponding planes are oriented at 45 degrees with respect
to the principal directions.

Figure 20. A Mohr's circle for a point at which the principal stresses are of opposite sign.

Figure 21. Orientation of the maximum


shear stress.

Figure 22. Orientation of the maximum


shear stress.

1.17

Figure 23, on the other hand, illustrates that when the in-plane principal stresses are of the same
sign (see the solid circle), the maximum shear stress (shown at D') must be calculated based on
the knowledge that the third eigenvalue is zero; i.e. equal to one-half the stress at A. The
orientations of the maximum shear stress planes are shown in Figures 24 and 25.

Figure 23. A Mohr's circle for a point at which the principal stresses are of the same sign.

Figure 24. Orientation of the maximum


shear stress.

Figure 25. Orientation of the maximum


shear stress.

1.18

1.9 Stresses in Thin Walled Pressure Vessels


A thin walled pressure vessel provides an important application of the analysis of plane stress.
The following discussion will be confined to cylindrical and spherical vessels.
The cylindrical vessel shown in Figure 26 has inside radius r and wall thickness t. The vessel
contains a fluid under pressure, p.

Figure 26. A cylindrical pressure vessel.


The hoop stress, 1, and the longitudinal stress, 2, are given by:

(1.9-1)

The Mohr's circle corresponding to the cylindrical pressure vessel is shown in Figure 27.

Figure 27. Mohrs circle for a cylindrical pressure vessel.


1.19

Note,
(1.9-2)
and
(1.9-3)

Figure 28, on the other hand shows a spherical pressure vessel of inside radius r and wall
thickness t, containing a fluid of gage pressure p.

Figure 28. A spherical pressure vessel.


In this case,
(1.9-4)

The Mohr's circle for the spherical pressure vessel is shown in Figure 29.
Note,
(1.9-5)

1.20

Figure 29. Mohrs circle for a spherical pressure vessel.

1.21

1.10 Homework Problems


1. Draw a rectangular parallelepiped whose sides are aligned with a Cartesian axis system
passing through the point in question. Assign stresses on each of the six faces consistent
with convention.
Answer: Combine the stress states shown in Chapter 1, Figures 5 through 7.
2. Consider a small element of volume in a continuum. Assign the stresses and the body
force/unit volume which act in the Y direction. Use this figure as a basis and derive the
corresponding equilibrium equation (1.5-4) by summing forces along Y.
Answer: See Section 1.5.
3. Consider a small element of volume in a continuum. Assign the stresses and the body
force/unit volume which act in the Z direction. Use this figure as a basis and derive the
corresponding equilibrium equation (1.5-5) by summing forces along Z.
Answer: See Section 1.5.
4. The following stress distribution has been determined for a machine component:

Is equilibrium satisfied in the absence of body forces?


Answer: Expand expressions in Equation (1.5-6); yes.
5. The following stress distribution has been determined for a machine component:

Is equilibrium satisfied in the absence of body forces?


Answer: Expand expressions in Equation (1.5-6); yes.
6. If the state of stress at any point in a body is given by the equations:

1.22

What equations must the body-force intensities Fx, Fy, and Fz satisfy?
Answer: Apply expressions in Equation (1.5-6); Fx = - ( a + 2 p z ), Fy = - ( m + 2 e y ),
Fz = - ( l + 2 n x + 3 i z2 ).
7. If the state of stress at any point in a body is given by the equations:

What equations must the body-force intensities Fx, Fy, and Fz satisfy?
Answer: Apply expressions in Equation (1.5-6); Fx = - 4y (3y2+x), Fy = x, Fz = 0.
8. A two-dimensional state of stress (zz = zx = zy = 0) exists at a point on the surface of a
loaded member. The remaining Cartesian components of stress are:

Using Mohrs circle:


(a) Determine the principal stresses and sketch the element on which they act.
(b) What is the maximum shear stress at the point in question?
Answer: (a) 1 = 95.14 MPa, 2 = - 85.14 MPa, p = - 9.72o (b) maximum = 90.14 MPa.
9. A two-dimensional state of stress (zz = zx = zy = 0) exists at a point on the surface of a
loaded member. The remaining Cartesian components of stress are:
xx = 90 MPa

yy = 40 MPa

xy = 60 MPa .

Using Mohrs circle:


(a) Determine the principal stresses and sketch the element on which they act.
(b) What is the maximum shear stress at the point in question?
Answer: (a) 1 = 130 MPa, 2 = 0 MPa, p = 33.7o (b) maximum = 65 MPa.
10. A two-dimensional state of stress (zz = zx = zy = 0) exists at a point on the surface of a
loaded member. The remaining Cartesian components of stress are:

Using Mohrs circle:

1.23

(a) Determine the principal stresses and sketch the element on which they act.
(b) What is the maximum shear stress at the point in question?
Answer: (a) 1 = 107 MPa, 2 = 43 MPa, p = - 19.3o (b) maximum = 53.5 MPa.
11. A two-dimensional state of stress (zz = zx = zy = 0) exists at a point on the surface of a
loaded member. The remaining Cartesian components of stress are:

Using Mohrs circle:


(a) Determine the principal stresses and sketch the element on which they act.
(b) What is the maximum shear stress at the point in question?
Answer: (a) 1 = 11 ksi, 2 = 1 ksi, p = - 63.4o (b) maximum = 5.5 ksi.
12. A pressure tank is supported by two cradles as shown; one of the cradles is designed so that
it does not exert any longitudinal force on the tank. The cylindrical body of the tank is
fabricated from a 3/8" steel plate by butt welding along a helix which forms an angle of 25 o
with respect to a transverse plane (the vertical direction). The end caps are spherical and
have a uniform wall thickness of 5/16". The internal gage pressure is 180 psi.

8'

30"

25

Figure 30. A pressure vessel with helical welds.


(a) Determine the normal stress and the maximum shearing stress in the spherical
caps.
(b) Determine the normal and shear stresses perpendicular and parallel to the
weld, respectively.
Answer: (a) = 4230 psi, maximum = 2115 psi, (b) weld = 4137 psi, weld = 1345 psi.
13. A pressure vessel of 10-in. inside diameter and 0.25-in. wall thickness is fabricated from a
4-ft section of spirally welded pipe. The vessel is pressurized to 300 psi and a centric axial

1.24

compressive force of 10-kip is applied to the upper end through a rigid end plate.

10 kip

4 ft
o
35
A
Figure 31. A compressed pressure vessel with helical welds.
(a) Draw a stress element for a point located along the weld with horizontal and
vertical faces.
(b) Determine and in directions, respectively, normal and tangent to the weld.
(c) What is the maximum shear stress in the wall?
Answer: (a) weld = 3.15 ksi, (b) weld = - 1.993 ksi, (c) max = 3 ksi.
14. Square plates, each of 5/8 thickness, may be welded together in either of the two ways
shown to form the cylindrical portion of a compressed air tank. Knowing that the
allowable stress normal to the weld is 9 ksi, determine the largest allowable gage pressure
for each case.

Figure 32. Two different pressure vessels constructed from square plates.

Answer: Case No. 1: 62.9 psi, Case No. 2: 83.9 ksi.

1.25

CHAPTER 2 - STRAIN
2.1 Introduction
In Chapter 1, stress relationships were based on equilibrium. No restrictions were made
regarding the deformation of the body or the physical properties of the material from which it
was made. Strain is a purely geometric quantity and no restrictions need be placed on the
material of the body; however, restrictions must be placed on the allowable deformations in
order to linearize the strain-displacement relations. The type of material and/or material
constraints are introduced in the constitutive equations which relate stress to strain.

2.2 Strain-Displacement Equations


Displacement is a vector quantity which characterizes the movement of an arbitrary point. As
illustrated in Figure 1, each point has a displacement vector, d, which can be resolved along
Cartesian axes X,Y,Z such that the corresponding scalar components of displacement are u,v,w,
respectively.
Y

d
_
v_
u
_
_
w

Z
Figure 1. The movement of a point is described by a displacement vector, d, which has scalar
components u,v,w.
The motion of the points in the body consists of two parts. Rigid body motion includes
translations and rotations during which there is no relative motion between neighboring points.
The movement of points relative to one another is called deformation.
Strain is a geometric quantity which describes the deformation which occurs in the body. There
are two types of strain. Shear strain is a measure of the angular change which takes place
between two line segments which were originally perpendicular. Since shear strain is expressed
in non-dimensional terms, the angle is measured in radians. Normal strain, on the other hand, is
defined as the change in length of a line segment divided by the original length of the segment.

2.1

The displacement components shown in Figure 1 are related to the strain components by the
strain-displacement equations which are given as follows:

(2.2-1)

2.3 Strain Equations of Transformation


The general transformation for a second order symmetric tensor illustrated by Equation (1.7-1)
can also be applied to transform strain provided that a slight modification is made when defining
the shear strain. To this end, the following definition is introduced:
(2.3-1)
In terms of Cartesian components, Equation (2.3-1) is written:
(2.3-2)

Then,
(2.3-3)
where ij' and kl are the components of the rotated and original tensors, respectively; ik and jl
correspond to the components in the matrix of direction cosines and its transpose, respectively.
From a rudimentary standpoint, the same relations developed in Section 1.7 for stress apply to
strain when the following substitutions are made:

(2.3-4)

Mohr's circle, for example, is plotted with as the abscissa and /2 as the ordinate.

2.2

2.4 Compatibility Equations


Equation (2.2-1) demonstrates that the strain field can be obtained from a given displacement
field by differentiating the displacement. However, if the strain field is measured, it becomes
necessary to integrate to obtain the displacement field. In this case, a unique displacement field
is obtained if and only if the body is simply connected (two points on the boundary can be
connected by never leaving the body or traversing it) and the strains satisfy compatibility
equations derived from the strain-displacement relations. The compatibility equations are:

(2.4-1)

2.5 Constitutive Equations


In general, stress is related to strain through 81 material constants. The linearity of stress versus
strain reduces this number to 36 while strain energy considerations further reduce the number to
21. If the body is linearly elastic, homogeneous and isotropic, the number reduces to 2.
A linearly elastic body is one which is made of a material which has a linear stress versus strain
curve. The body is homogeneous if it is made of the same material throughout. Isotropic refers
to the fact that a material property at a given point is independent of the direction in which it is
measured. For such a body, the equations which relate the stress to the strain are written in terms
of two independent material constants and are referred to as Hooke's laws. These constitutive
equations may be expressed for strain as a function of stress as:

2.3

(2.5-1)

or expressed for stress in terms of strain as:

(2.5-2)

2.4

2.6 Homework Problems


1. The central portion of the tensile specimen shown in Figure 2 represents a uniaxial stress
condition; P is the applied load and A is the cross-sectional area. For a point located on the
surface and in the central portion of the specimen:

Figure 2.
(a) Draw a two-dimensional stress element having horizontal and vertical faces. Show the
stress distribution on this element. Are these the principal stress planes?
(b) Construct the Mohrs circle for the point and draw a rotated element showing the
maximum in-plane shear stress. Be sure to label all stresses and specify the angle of
orientation with respect to, say, the horizontal direction.
(c) How does the maximum in-plane shear stress compare with the absolute maximum
shear stress at the point?
(d) Develop the constitutive equation(s) for this uniaxial stress condition from the three
dimensional Hookes Law (see Eqn. 2.8-2).
(e) Develop the mathematical relationships for Young's Modulus and Poisson's Ratio and
describe how these quantities could be calculated by making physical measurements.
(f) Specify the six components of strain if the specimen is made of steel (E = 30 x 106 psi,
= 0.3) and loaded to yield (y = 45 x 103 psi).
(g) What would be the angular orientation of the fracture surface with respect to a
horizontal axis if the material were ductile and failed along the planes of maximum shear?
What would happen if the material were brittle?
Answer: (a) xx = 0, yy = P/A, xy = 0, yes, (b) element oriented at 45o with = = P/2A, (c)
equal, (d) yy = Eyy ..., (e) E = yy/yy, = -xx/yy = - zz/yy, measure strain and
compute stress, (f) xx = zz = - 450 , yy = 1500 , xy = xz = yz = 0 , (g) ductile
@ 45o, brittle @ 0o.

2.5

2. Figures 3 and 4 show that when a plane is passed through a circular shaft perpendicular to its
longitudinal axis, a state of pure shear exists.

T
_

T_

Axis of shaft

A
Axis of shaft

Figure 3.

Figure 4.

(a) Specify the formula used to determine the shear stress and define all terms. Draw a
diagram showing how the stress varies over the circular cross section.
(b) Express the magnitude of the maximum shear stress at the surface in terms of the applied
torque and the radius.
(c) Assuming that the shaft is oriented so that its longitudinal axis is horizontal, draw a twodimensional stress element having horizontal and vertical faces, and show the stress
distribution on this element. Are these the principal stress planes?
(d) Construct the Mohrs circle for the point and draw a rotated element showing the
principal stresses. Be sure to label all stresses and specify the angle of orientation with
respect to, say, the horizontal direction.
(e) What would be the angular orientation of the fracture surface with respect to a
horizontal axis if the material were ductile and failed along the planes of maximum shear?
What would happen if the material were brittle?
Answer: (a) = T/J, (b) = 2T/c3, (c) xx = yy = 0, xy = 2T/c3, no, (d) element oriented at
45o with = 2T/c3, (e) ductile @ 90o, brittle @ 45o.

2.6

CHAPTER 3 LINEAR ELASTICITY


3.1 Overview
At each point in a loaded body, there are 15 unknowns: 3 displacements (u,v,w), 6 strains
(xx,yy,zz,xy,xz,yz) and 6 stresses (xx,yy,zz,xy,xz,yz). A complete solution to the general
elasticity problem must satisfy 15 equations (3 equilibrium, 6 compatibility and 6 constitutive)
plus boundary conditions formulated in terms of displacement and/or traction.
In general, it is possible to obtain a closed-form solution only for problems having relatively
simple loading and geometry. In more complex situations, it is necessary to resort to an
analytical formulation using a technique such as finite element analysis. In this method, the
body is mathematically modeled by partitioning it into a finite number of elements connected
together at points called nodes. Displacements and/or tractions are specified at the boundary and
a mathematical wave front is passed over the mesh. Stresses are predicted based on either a
flexibility or stiffness approach.
An alternative solution to the problem is to experimentally measure the displacements, strains or
stresses. When displacements are measured, the strain-displacement equations are used to
determine the strains. Stresses are obtained using the constitutive equations, after which the
transformation equations may be applied to obtain the principal stresses and the maximum shear
stress. When strains are measured, the compatibility equations must be satisfied to obtain a
unique displacement field.

3.2 Plane Stress


The geometry of the body and the nature of the loading allow two important types of problems to
be defined. Figure 1 illustrates an example of plane stress in which the geometry is essentially a
flat plate with a thickness much small than the other dimensions. The loads applied to the plate
act in the plane of the plate and are uniform over the thickness.
Y

Figure 1. An example of plane stress.

3.1

Mathematically, for a plate contained in the XY plane,

(3.2-1)
.
Although the normal stress perpendicular to the surface vanishes, the normal strain, zz, does not.
The relationship between the latter and the other two in-plane strain components is found by
setting zz = 0 in the expression given for it in Equation (2.5-2). This results in:
(3.2-2)

An alternate form of Hookes law for plane stress is found by substituting Equation (3.2-2) into
the remaining expressions contained in Equation (2.5-2) as

(3.2-3)

3.3 Plane Strain


Figure 2, on the other hand, illustrates the condition of plane strain. In this case, the body is
considered a prismatic cylinder with a length much larger than the other dimensions. The loads
applied to the cylinder are distributed uniformly with respect to the large dimension and act
perpendicular to it.
Mathematically, for a point in the XY plane,

3.2

(3.3-1)

Figure 2. An example of plane strain.


Example:

A flat plate (located in the XY plane of a Cartesian coordinate system, with


normal along Z) was designed and built as part of a compressor system using steel
(Elastic Modulus = 30 x 106 psi; Poisson's Ratio = 0.3). The allowable shear
stress of the material is 12.5 ksi but the manufacturer wants a factor of safety of
2.0 to be incorporated into the design.

Preliminary tests conducted by the manufacturer indicated that the most critical location on the
structure was located at x = 1.0, y = 2.0, z = 0.0. This point was on a traction free surface where
the in-plane displacements were found to be expressed as follows:
(1)
(a) Determine the magnitudes and directions of the principal stresses at the critical location.
Sketch your results on a rotated element and label the stress distribution. Specify the angle
of orientation of the element with respect to the x direction.
(b) Does the part meet the required specifications?
Solution:

The point is in plane stress and the in-plane strain components may be obtained
using the strain-displacement equations given by Equation (2.2-1) as

(2)

For plane stress, zz = 0, and


(3)

3.3

The in-plane stresses are computed using Equation (2.5-2) [or Equation (3.2-3)] as

(4)

Figure 3 shows the corresponding stress element; and, Figure 4 shows the Mohrs circle for the
stress distribution at the critical point.

Figure 4. Mohrs circle.

Figure 3. Stress element.

The circle is drawn by plotting the two points corresponding to the planes labeled as 1 and 2 on
Figure 3 in - space. The points are connected and a circle is drawn centered at C with radius r.
The location of the center of the circle is computed from Equation (1.7-9) as:
(5)
The radius is computed from Equation (1.7-10) as:

(6)

The maximum and minimum values of stress are determined by adding and subtracting the
magnitude of the radius to the value of at point C, respectively. This operation produces:

3.4


(7)

The orientations of the principal planes are obtained by determining the orientation of the point
labeled 3 on Figure 4 with respect to the point labeled 1. Referring to the figure:
(8)

Note that on the circle, 2 is drawn clockwise from point 1 to point 3. Also note that point 3
corresponds to max. Therefore, the normal to the plane corresponding to max lies at = 10.9o
measured in the counterclockwise direction from the normal to the plane labeled as 1 in Figure
3. These results are summarized in Figure 5 which shows the rotated element corresponding to
the principal planes.

Figure 5. Principal planes.


The in-plane shear stress is equal in magnitude to the radius. However, since the circle does not
encompass the origin, the absolute maximum shear stress is equal to one-half of the maximum
value,
(9)
The allowable stress, on the other hand, is found by dividing the maximum shear stress that the
material can take by the factor of safety,
(10)

3.5

Since the maximum stress is greater than the allowable stress, the part does not meet the design
specifications.
3.4 Homework Problems
1. Figure 6 shows a flowchart for the solution to the general elasticity problem. Add an
appropriate description to the items numbered 1 through 8.

u, v, w
1

6
xx, yy, zz, xy, xz, yz
2

xx, yy, zz, xy, xz, yz


3

1, 2, 3
4

Figure 6. The general elasticity problem.

Answer: 1. displacements; 2. strain, 3. stress, 4. principal stress, 5. compatibility equations, 6.


strain-displacement equations, 7. constitutive equations, and 8. transformation
equations.
2. Assume that a state of plane stress exists such that zz = yz = zx = 0.
(c) Reduce the equilibrium equations given in Equation (1.5-6) for plane stress.
(d) Reduce the constitutive equations given in Equation (2.5-1) for plane stress.
(e) Knowing zz = 0, develop an expression for zz = f (xx, yy) for plane stress and show by
substituting this expression into Equation (2.5-2) that the constitutive equations become:

Answer: (a) xx/x + yx/y + Fx = 0 ..., (b) xx = 1/E [xx - yy] ..., (c) zz = - /(1-) [xx +
yy] ...

3.6

3. A critical point on the surface of a part fabricated using steel (E = 30 x 10 6 psi; = 0.3) is
under plane stress conditions with zz = yz = zx = 0. Knowing that xx = 505 x 10-6 in/in [or,
505 ], yy = 495 x 10-6 in/in [or, 495 ] and xy = 100 x 10-6 in/in [or, 100 ], determine
the
(a) strain zz.
(b) principal stresses and their orientation with respect to the XYZ axes system.
(c) maximum shear stress at the point.
Answer: (a) zz = - 428.6 , (b) 22.59 ksi, 20.27 ksi, 0 ksi, x1 = 42o (c) max = 11.3 ksi.

3.7

CHAPTER 4 LIGHT AND ELECTROMAGNETIC WAVE


PROPAGATION
4.1 Introduction to Light
Electromagnetic radiation is predicted by Maxwell's theory to be a transverse wave motion
which propagates in free space with a velocity of approximately c = 3 x 108 m/s (186,000 mi/s).
The wave consists of oscillating electric and magnetic fields which are described by electric and
magnetic vectors E and H, respectively. These vectors are in phase, perpendicular to each other,
and at right angles to the direction of propagation, z.
A simple representation of the electric and magnetic vectors associated with an electro-magnetic
wave at a given instant of time is illustrated in Figure 1. As described below, the wave is
assumed to have a sinusoidal form and may be characterized by either the electric or magnetic
vector. In the arguments that follow, light propagation is described in terms of the electric
vector, E.

H
z
Figure 1. An electromagnetic wave.
A mathematical representation for the magnitude of the electric vector as a function of position,
z, and time, t, can be determined by considering the one-dimensional wave equation

(4.1-1)
and setting E = u. The DeLambert solution to this partial differential equation is
(4.1-2)
where z is the position along the axis of propagation and t is time.
In Equation (4.1-2), the quantity f (z - ct) represents a wave moving at velocity c in the positive z
direction while g (z + ct) represents a wave moving at velocity c in the negative z direction.

4.1

Many of the optical effects of interest in experimental mechanics and optical metrology can be
described by considering a harmonic vibration propagating along positive z such that

(4.1-3)

where a is the instantaneous amplitude, is the wavelength, and 2/ is the wave number.
Equation (4.1-3) neglects the attenuation associated with the expanding spherical wavefront and
is rigorously valid only for a plane wave.
Figure 2 shows a graphical representation of the magnitude of the light vector as a function of
position along the positive z axis, at two different times. The wavelength, , is defined as the
distance between successive peaks. The time required for the wave to propagate through a
distance equal to the wavelength is called the period, T.

ct

t>0

z
t=0
E = a cos 2/ (z - ct)
Figure 2. A wave, originally at time t = 0, propagates through space.
Since the wave is propagating with velocity, c,
(4.1-4)
The angular frequency of the wave is defined as
(4.1-5)
where the frequency,
(4.1-6)

represents the number of oscillations which take place per second.


4.2

The above definitions make it possible to express E(z,t) in several ways. For example,

(4.1-7)

4.2 Electromagnetic Spectrum


Although the electromagnetic spectrum has no upper or lower limits, the radiations commonly
observed have been classified in the table shown in Figure 3.

Figure 3. The electromagnetic spectrum.


The visible range of the spectrum is centered about a wavelength of 550 nm and extends from
approximately 400 to 700 nm. Figure 4 lists the different colors observed by the human eye as a
function of wavelength.

Figure 4. The visible spectrum.

4.3

The characteristics of an electromagnetic wave such as color depend upon the frequency; since,
the frequency is independent of the medium through which propagation occurs. The wavelength,
, on the other hand, depends upon the medium; since the velocity of propagation changes from
medium to medium (see Section 4.9).
When the light vectors corresponding to E(z,t) are all of one frequency, the light is referred to as
monochromatic. Light consisting of several frequencies is recorded by the eye as white light.

4.3 Light Propagation, Phase, and Retardation


In Equation (4.1-3) [the first form of E(z,t) shown in Equation (4.1-7)], the term in the argument
of the cosine function represents the angular phase, , while the term in the parenthesis
corresponds to the linear phase, . It is apparent that these quantities are related through the
wave number as follows:
(4.3-1)

Figure 5, for example, shows two waves having the same wavelength, , and equal amplitudes.
In this case, the waves are plotted as a function of position, along the axis of propagation.

ct

E = a cos 2/ (z - ct)

a
z

E1 = a cos 2/ (z - ct + 1)
E2 = a cos 2/ (z - ct + 2)
Figure 5. Two waves out of phase.
Assuming that each wave has different amplitude given by a1 and a2, respectively,

(4.3-2)

4.4

where 1 (1) and 2 (2) are the initial linear (angular) phases of E1 and E2, respectively. The
linear phase difference, , is given by

(4.3-3)
Since one wave trails the other, the linear phase difference is often referred to as the retardation.

4.4 Polarized Light


Most light sources consist of a large number of randomly oriented atomic or molecular emitters.
The rays emitted in any direction from such sources will have electric fields that have no
preferred orientation. In this case, the light beam is said to be unpolarized.
If, however, a light beam is made up of rays with electric fields that show a preferred direction of
vibration, the beam is said to be polarized. Figure 6, for example, shows the most general case
of elliptical polarized light where the tip of the electric vector describes an elliptical helix as it
propagates along z.

Light vector
b
z

a
y

Light vector

Ellipical helix

Figure 6. Elliptically polarized light.

z
a
Circular helix
Figure 7. Circularly polarized light.

Figure 7 shows a more restrictive case where the tip of the light vector describes a circular helix
as it propagates along z. This condition, referred to as circularly polarized light, is studied in
Section 4.8. It is important because it will be encountered in Chapter 5 when studying the
circular polariscope in conjunction with the method of photoelasticity.
The most restrictive case, illustrated in Figure 8, occurs when E(z,t) vibrates in a single known
direction. This condition, referred to as linearly polarized light, is perhaps the most important
case as far as experimental mechanics is concerned. The figures used to characterize light in the
beginning of this chapter, for example, were drawn by making the assumption that the light
vector was linearly polarized. This state of polarization can be produced by using sheetpolarizers, a Nicol prism, or, by reflection at the Brewster angle.

4.5

Tip sweeps out a sine curve

Plane of polarization
Figure 8. Linearly polarized light.

4.5 Optical Interference


The light emitted by a conventional light source, such as a tungsten-filament light bulb, consists
of numerous short pulses originating from a large number of different atoms. Each pulse
consists of a finite number of oscillations known as a wave train. Each wave train is thought to
be a few meters long with duration of approximately 10-8 s. Since the light emissions occur in
individual atoms which do not act together in a cooperative manner, the wave trains differ from
each other in the plane of vibration, frequency, amplitude, and phase. Radiation produced in this
manner is referred to as incoherent light and the wavefronts simply add as scalar quantities.
For other light sources, such as a laser, the atoms act cooperatively in emitting light and produce
coherent light. In this case, the wave trains are monochromatic, in phase, linearly polarized, and
extremely intense. When the light coming from these sources is split and later recombined, the
wavefronts superimpose as vectors to produce optical interference.
Although coherent light must be monochromatic (of a single frequency) to interfere, light
coming from two different monochromatic sources is not, in general, coherent. In most cases,
interference will not take place unless the wavefronts in question originate from the same source.

4.6 Complex Notation


A convenient way to represent both the amplitude and phase of the light wave described by
Equation (4.1-3) for calculations involving a number of optical elements is through the use of
complex or exponential notation. To this end, the Euler identity is

(4.6-1)
where i2 = -1. The first term on the right hand side is called the real part of the complex function
whereas the term associated with i is the imaginary part.
Using this notation, the sinusoidal wave characterized by Equation (4.1-3) can be represented by

4.6

(4.6-2)
When the operations performed are linear, the symbol Re is dropped and calculations are
performed using the complex function

(4.6-3)
where and are the angular and linear phase, respectively.
If the waves have an initial phase, such as those described by Equation (4.3-2), they can be
expressed as

(4.6-4)

4.7 Intensity
In many practical applications, the light vector is related to a measurable quantity called the
intensity where
(4.7-1)

Assuming that k = 1, and using the form of E(z,t) given in Equation (4.6-3),
(4.7-2)
where E(z,t)* is the complex conjugate of E(z,t) and the operation is a dot product.
E(z,t)* is obtained from E(z,t) by changing the sign of the imaginary part of the latter which
amounts to changing the sign in the exponent of the exponential function. When referring to a
single wavefront, the dot product operation becomes a simple multiplication and
(4.7-3)
Equation (4.7-3) shows that the intensity of light is proportional to the square of the
instantaneous amplitude of the light vector.

4.7

4.8 Superposition of Wavefronts


As mentioned in Section 4.5, incoherent wavefronts superimpose as if they were scalars.
Assuming that two waves are represented by vectors E1 and E2, the resulting intensity for
incoherent waves is simply
(4.8-1)
If the wavefronts are coherent, the light vectors add as vectors. In this case, E1 and E2 can be
combined into a resultant ER and the intensity is
(4.8-2)
Consider, for example the superposition of two coherent (monochromatic; same frequency),
rectilinear harmonic vibrations (plane polarized wavefronts) that are of unequal amplitude and
out of phase. In this case, the expressions in Equation (4.6-4) can be used to characterize the
wavefronts; and, Equation (4.8-2) represents the intensity.
Since the two wavefronts are co-linear, the dot product becomes a simple multiplication, and
Equation (4.8-2) can be expanded as follows:

(4.8-3)

From complex variables, the last term in brackets on the right hand side of Equation (4.8-3) is
proportional to the cosine of the argument in the exponential function and the intensity becomes
(4.8-4)
where the term in the brackets of Equation (4.8-4) represents the difference in angular phase of
the wavefronts.
Two cases are of special interest. In the case in where the angular phase difference is a multiple
of , Equation (4.8-4) becomes,
(4.8-5)
If a1 = a2, then I = 0. In this case, there is no measurable intensity; implying that light plus light
results in darkness. Destructive interference is said to take place.

4.8

If, on the other hand, the angular phase difference is either zero or a multiple of 2, then
Equation (4.8-4) becomes
(4.8-6)
If a1 = a2, then I = 4a2, thereby resulting in constructive interference.
Thus, the intensity observed when two rectilinear coherent wavefronts are superimposed depends
upon the amplitudes of the wavefronts and the relative difference in phase between them. In the
case considered above, the electric vector is restricted to a single plane and the superposition
results in a plane polarized, or, linearly polarized light. Interference effects have important
applications in optical metrology; for example, in photoelasticity, moir, holography, and
speckle metrology.
Two other important forms of polarized light, previously discussed in Section 4.4, arise as a
result of the superposition of two linearly polarized light waves having the same frequency but
mutually perpendicular planes of vibration. This condition is depicted in Figure 9.

Ex

Ex = ax cos 2/ (z - ct + x)
Ey = ay cos 2/ (z - ct + y)
ax

Ey

ay

Figure 9. Superposition of two wavefronts having perpendicular planes of vibration.


In this case, at a fixed position along the axis of propagation, the magnitude of the light vector
can be expressed as
(4.8-7)
where is the angular phase and is the angular velocity at which the electric vector rotates. To
achieve a better understanding of this case, the vector can be resolved into components along the
x- and y-axes, and the wavefronts shown in Figure 9 represented by

(4.8-8)

4.9

where x and y are the phase angles associated with waves in the xz and yz planes, and ax and ay
are the amplitudes of waves in the xz and yz planes.
The magnitude of the resulting light vector is given by vector addition as
(4.8-9)
Considerable insight into the nature of the light resulting from the superposition of two mutually
perpendicular waves is provided by studying the trace of the tip of the resulting electric vector on
a plane perpendicular to the axis of propagation at points along the axis of propagation. An
expression for this trace can be obtained by eliminating time from Equation (4.8-8). Pursuing
this argument in terms of angular phase results in the relation
(4.8-10)
Or, since
(4.8-11)
Equation (4.8-10) becomes
(4.8-12)
Equation (4.8-12) is the equation of an ellipse, and light exhibiting this behavior is known as
elliptically polarized light.
As illustrated in Figure 10, tips of the electric vectors at different positions along the z axis form
an elliptical helix. As mentioned previously, the electric vector rotates with an angular velocity,
. During an interval of time t, the helix will translate a distance z = ct in the positive direction.
As a result, the electric vector at position z will rotate in a counter-clockwise direction as the
translating helix is observed in the positive z direction.

x
Light vector E

z
y

Ellipical helix
Figure 10. Elliptically polarized light.
4.10

A special case of elliptically polarized light occurs when the amplitudes of the two waves are
equal [ax = ay = a] and = /2 [or, = /4]. In this case, Equations (4.8-10) and (4.8-12)
reduce to the equation of a circle given by
(4.8-13)
As mentioned previously in Section 4.4, light exhibiting this behavior is known as circularly
polarized light, and the tips of the light vectors form a circular helix along the z axis. For = /2
[ = /4], the helix is a left circular helix, and the light vector rotates counter-clockwise with time
when viewed from a distant position along the z axis. For = - /2 [ = - /4], the helix is a right
circular helix, and the electric vector rotates clockwise.

4.9 Reflection and Refraction


The discussions so far have been limited to light propagating in free space. Most optical effects
of interest, however, occur as a result of the interaction between a beam of light and some
physical material.
As mentioned in the beginning of the chapter, light propagates in free space with a velocity c
equal to approximately 3 x 108 m/s (186,000 mi/s). In any other medium, the velocity is less
than the velocity in free space.
The ratio of the velocity in free space to the velocity in a medium, v, is called the index of
refraction, n, of the medium. The latter is defined as
(4.9-1)
The index of refraction for most gases is only slightly greater than unity (for air, n = 1.0003).
Values for liquids range from 1.3 to 1.5 (for water, n = 1.33) and for solids range from 1.4 to 1.8
(for glass, n = 1.5). The index of refraction for a material is not constant but varies slightly with
wavelength of the light being transmitted. This dependence of index of refraction on wavelength
is referred to as dispersion.
Figure 11 illustrates that when a beam of light strikes a surface between two transparent
materials with different indices of refraction, it is divided into a reflected ray and a refracted ray.
The reflected and refracted rays lie in the plane formed by the incident ray and the normal to the
surface, known as the plane of incidence. The angle of incidence, , the angle of reflection , and
the angle of refraction are related. For reflection, light bounces off the surface with
(4.9-2)
Refraction is governed by Snells law and can be quantified in terms of the velocity of
propagation and/or index of refraction of the media as follows:

4.11

(4.9-3)

Incident ray

Reflected ray

Wavefront
Material 1
Material 2

Refracted ray
Figure 10. Reflection and refraction.
4.10 Double Refraction - Birefringence
Light traveling through an isotropic medium propagates with the same velocity in all directions;
however, when light enters an anisotropic or crystalline medium along an optical axis it is
divided or refracted into two plane polarized components which are orthogonal and are
propagating at different velocities. This phenomenon is known as double refraction or
birefringence, and occurs in photoelastic models subjected to stress.
'The angular phase difference that occurs between these two wavefronts when light passes
through a plate of thickness d is
(4.10-1)
where is the wavelength and n is the index of refraction.
The corresponding linear phase difference, , is given by Equation (4.3-1); and,
(4.10-2)
The fringe order, R, to be defined later in conjunction with the method of photoelasticity,
provides a measure of the relative retardation as follows:
(4.10-3)

4.12

The difference in indices of refraction is proportional to the difference in principal stress; in


particular,
(4.10-4)
where C is the stress optic coefficient.
A more useful description of this phenomenon is given by the relation
(4.10-5)
where c is the velocity of propagation outside the model (usually air, n = l), d is the thickness of
the model, R is the fringe order, F = f/c is the model fringe value, and f = /c is the material
fringe value.
Equation (4.10-5) is used to determine the stress state in a loaded photoelastic model from the
analysis of the isochromatic fringe pattern. As illustrated in Chapter 5, the most commonly used
form of this equation is
(4.10-6)

where d is found by direct measurement, R is determined from the photoelastic fringe pattern,
and f is determined by performing a calibration test on a sample of the photoelastic material in
question.

4.13

4.11 Homework Problems


1. Figure 12 shows two coherent (monochromatic; same frequency), rectilinear harmonic
vibrations (plane polarized wavefronts) which have unequal amplitude and are out of phase.

ct

E = a cos 2/ (z - ct)

a
z

E1 = a cos 2/ (z - ct + 1)
E2 = a cos 2/ (z - ct + 2)
Figure 12.
(a) Express the magnitudes of these light vectors in complex notation using the linear phase
descriptions labeled in the figure.
(b) Using complex notation, derive the intensity distribution resulting from their
superposition.
Hint: See Equation (4.6-4); apply Equation (4.8-2), and follow the arguments used to derive
Equation (4.8-4).
2. Analyze the intensity distribution derived in Problem 1 and specify the linear phase and any
other conditions that must occur to produce destructive and constructive interference.
Quantify the magnitude of the intensity for each of these conditions.
Hint: See the arguments made in Section 4.8. The results can be checked by applying Equation
(4.3-1).

4.14

CHAPTER 5 - PHOTOELASTICITY
5.1 Introduction
Photoelasticity has long been a widely used experimental technique for determining stresses and
strains. The technique is applicable to any state of stress, but historically it has been for the most
part limited to the determination of static two-dimensional stress distributions. Later
developments led to its application to three dimensional stress problems, two dimensional
dynamic problems including both cyclic and transient, and some plasticity and viscoelastic
problems.
The technique is limited to a class of materials which are birefringent; however, the technique
can be applied to other materials and actual prototypes by cementing films or thin sheets of
photo-elastic material directly to the surface of the specimen. These techniques are covered in
this chapter.
The photoelastic technique is based upon the stress-optic law (see Section 4.10) that applies to
many transparent materials when they are stressed. Depending on the material, the phenomena
may be due to stress or strain, or both. If the material is linear and elastic the effect can be
referred to either stress or strain. If the material is viscoelastic both the strain-optic and the
stress-optic relationships must be known.
The materials in question exhibit a temporary birefringence when they are under a temporary
load. The principal optical axes coincide when the directions of the principal stresses and strains
are linearly related to the refractive indices along their two respective principal axes. The result
is that if a plane-polarized light wave enters a birefringent material under load, it will be resolved
into two plane-polarized waves that propagate through the stressed material at different
velocities. When the two waves emerge from the material they are out of phase, where the phase
difference is proportional to the difference between the principal stresses and the thickness of the
material. A detailed analysis is included herein.
Photoelastic materials that exhibit birefringence are commonly analyzed with respect to their
stress distributions in a device called a polariscope. The simplest type of polariscope is a plane
polariscope which makes use of two polarizers whose axes are crossed with respect to each
other. Light is passed through the first polarizer and then through the loaded specimen where it
is resolved into two plane-polarized waves which are analyzed by the second polarizer. As
discussed in Section 5.2, the resulting pattern is an isochromatic fringe pattern superimposed
upon an isoclinic fringe pattern. Isoclinic fringes are the loci of points where the principal stress
directions coincide with the axis of the polarizer and analyzer; isochromatic fringes are the loci
of constant principal stress difference.
A widely used polariscope is the circular polariscope. It has the advantage that the isoclinic
fringe pattern is eliminated leaving only the isochromatic fringes and is discussed in Section 5.3.

5.1

5.2 Plane Polariscope


As mentioned above, many transparent noncrystalline materials which are optically isotropic
when free of stress become optically anisotropic and display characteristics similar to crystals
when they are stressed. These characteristics persist while loads on the material are maintained
but disappear when the loads are removed. This behavior is known as temporary double
refraction or birefringence.
The state of stress in birefringent materials can be analyzed by using a polariscope. Figure 1, for
example, shows a plane polariscope which consists of a light source (white or monochromatic)
and two linear polarizers called the polarizer and the analyzer. The configuration shown in the
figure is referred to as a dark field polariscope because the crossed polarizers result in a dark
background.

Axis of
polarization
Principal stress
direction 1

Light
source

Polarizer

Analyzer

Axis of
polarization

Principal stress
direction 2
Model

Figure 1. A dark field plane polariscope.


Light coming from the source passes through the first linear polarizer (called the polarizer)
shown in the figure with its plane of vibration in the vertical direction.
In accordance with the arguments presented in Chapter 4, the amplitude of the light leaving the
polarizer can be expressed represented by its electric vector, E1. The magnitude of the latter can
be expressed as a rectilinear harmonic vibration of the form
(5.2-1)
It should be noted that the ensuing argument could just as easily have been pursued using the
cosine function as illustrated in Experimental Stress Analysis by Dally and Riley, 4th Edition,
College House Enterprises, 2005, ISBN 0-9762431-0-7.
A doubly refracting plate, labeled as the model in Figure 1, is located between the polarizer and
the analyzer with one of its principal axes making an angle with respect to the horizontal.

5.2

Figure 2 illustrates that when E1 enters the model, it is resolved into two vector components, E2
and E3 with magnitudes
(5.2-2)

1
E1

E2

E3

Figure 2. Linearly polarized light coming from the polarizer is resolved along the principal
stress directions in the model.
The orientation of these components remains constant as these plane waves pass through the
plate; since, the model is assumed to be in a state of plane stress. But, because the model is
birefringent and each of the waves propagate with a different velocity as they travel through the
thickness, d, the components emerge with an angular phase difference, , given by
(5.2-3)

where is the wavelength.


Assuming that E4 and E5 correspond to the wavefronts which emerge as a result of E2 and E3,
respectively,
(5.2-4)

The wavefronts continue to propagate until they reach the analyzer; another linear polarizer
oriented with its plane of vibration in the horizontal direction.
As illustrated in Figure 3, the analyzer allows only the horizontal components of the vectors E4
and E5 to pass.

5.3

E4

E5

Figure 3. The two components of light emerging from the model are resolved by the analyzer.
The resulting wavefront is characterized by a resultant vector, say E6, with magnitude
(5.2-5)

Equation (5.2-5) can be simplified somewhat by using the following trigonometric identities:

(5.2-6)

Applying the identities in Equation (5.2-6) to the relation in Equation (5.2-5),

(5.2-7)

In addition to the identities stipulated in Equation (5.2-6), the following relations hold true:

(5.2-8)

Applying the latter to the expression contained in Equation (5.2-7),

5.4

(5.2-9)
By recognizing that
(5.2-10)
Equation (5.2-9) can be written as

(5.2-11)

where
(5.2-12)
The term in Equation (5.2-12) is the instantaneous amplitude of the emerging beam; and, a
measure of the intensity, observed by looking through the analyzer, is given by the square of its
amplitude:

(5.2-13)
There are two conditions of extinction. The first occurs when sin2 2 = 0. This happens when 2
= n, where n is an integer; or, when = n/2. For n = 0 and n = 1, = 0 and = /2,
respectively.
For the configuration studied in Figure 1, represents the orientation of the principal stress
measured counterclockwise from the horizontal. Thus, the critical values found for coincide
with the directions of the polarizer and analyzer. The implication is that all points which have
their principal stress direction aligned with the polarizing elements appear dark. The loci of all
such points create families of fringes called isoclinics.
The second condition of extinction occurs when sin2 /2 = 0. This occurs when /2 = n, where
n is an integer; or, when = 2n. The integer multiples of 2 correspond to linear phase
differences which are integer multiples of the wavelength, . Thus, extinction also occurs when
linear phase changes by an integral number of wavelengths. An isochromatic fringe pattern

5.5

results in which the fringe order number, R, is related to the principal stress difference by the
equation developed in Section 4.10. In particular,

(5.2-14)
In Equation (5.2-14), R is the fringe order number (measured at a point in the isochromatic fringe
pattern), d is the thickness of the model (found by direct measurement), and f is the material
fringe value (determined by performing a calibration test on a sample of the photoelastic material
used to construct the model). Integer order fringes correspond to the color of the background.
In conclusion, two families of fringes result when a loaded model is studied with a plane
polariscope: isoclinic fringes represent the loci of points which have their principal stress
directions aligned with the polarizer and analyzer; isochromatic fringes represent the loci of
points which have a constant principal stress difference. Since the difference between the two
principal stresses is twice the in-plane shear stress, the isochromatic pattern also represents the
loci of constant shear stress.

5.3 Circular Polariscope


The isoclinics are not always needed in photoelastic analysis and often obscure the isochromatics
to the point where a stress analysis becomes impossible. Figure 4 shows a device called a
circular polariscope which is used to eliminate the isoclinics.

P
/4
1 F
2
S

Analyzer
A

Light
source

S
Polarizer
1st /4 plate

Model
2nd /4 plate

z
Figure 4. A dark field circular polariscope.
The dark field polariscope depicted in the figure represents one of many possible configurations
for a circular polariscope. In this case, light comes from the source and passes through the first

5.6

polarizer with its plane of vibration in the vertical direction. The amplitude of the light vector
leaving the polarizer can be represented by
(5.3 -1)
As illustrated in Figure 4, the light hits a quarter wave plate having its fast and slow axes
oriented at 450 with respect to the axis of the polarizer.
The plate is made of a birefringent material and is produced so that every point has the same
principal stress directions. Thus, the incoming wave is split into two equal components which
are plane polarized at 900. The components are represented by E2 and E3 in Figure 5.

E1

0
E2 45

E3

Figure 5. Linearly polarized light coming from the polarizer is resolved along the fast and slow
axes of the first quarter wave plate.
The magnitudes of these vectors are

(5.3-2)

where b = a/ 2.
The thickness of the quarter wave plate is adjusted so that after the waves travel through the plate
and emerge, the one along the fast axis leads the other by /4; a linear phase shift corresponding
to an angular phase shift of /2. Since E2 is assumed to lie along the fast axis, the waves coming
out of the plate are represented by

(5.3-3)

where the out-coming waves E4 and E5 correspond to the incoming waves E2 and E3,
respectively.
5.7

Although the above argument was made by adding the angular phase into the argument of the
expression for E4, this could have been done by including -/2 into the argument of E5.
The two waves continue to propagate along their respective directions until they encounter the
model. Referring to the argument made in Section 4.8, the model is illuminated with circularly
polarized light; that is, the components in Equation (5.3-3) are plane polarized in perpendicular
directions, in phase, of equal amplitude, and have a relative phase difference of /2.
As illustrated in Figure 6, it is assumed that, for the point in question, one of the principal axes
makes an angle of with respect to the fast axis of the first quarter wave plate.

x
F
y

E4
450

E5

Figure 6. The two light vectors exiting the first quarter wave plate are resolved along the
principal directions in the model.
Light is resolved into vector components along directions x and y; and, if these components are
denoted by E7 and E6, respectively,

(5.3-4)

The orientation of these components remains the same as these plane waves pass through the
plate; since, the model is assumed to be in a state of plane stress. But, because the model is
birefringent and each of the waves propagate with a different velocity as they travel through the
thickness, d, the components emerge with an angular phase difference, , given by Equation
(5.2-3). Assuming that E8 and E9 correspond to the wavefronts which emerge as a result of E6
and E7, respectively,
(5.3-5)

5.8

The resulting vectors are now aligned along the principal stress directions and propagate until
they reach the second quarter wave plate. As illustrated in Figure 7, the second quarter wave
plate is oriented having its fast and slow axes crossed with respect to the first quarter wave plate.

x
S
y

E9

450
E8

Figure 7. The two components of light emerging from the model are resolved along the fast and
slow axes of the second quarter wave plate.
When the vectors E8 and E9 reach the plate, they are resolved into vector components E10 and E11
along the slow and fast axis, respectively, where

(5.3-6)

These beams emerge from the quarter wave plate with an additional phase difference equal to /2
which can be added into the argument of E11. Assuming that vectors E12 and E13 correspond to
the out-coming wavefronts associated with the incoming wavefronts E10 and E11, respectively,

(5.3-7)
The analyzer allows only the horizontal components of these vectors to pass. Since both vectors
make an angle of 450 with respect to the analyzer, the resulting vector, E14, is

(5.3-8)

5.9

where
(5.3-9)
The term in Equation (5.3-9) is the instantaneous amplitude of the emerging beam; and, a
measure of the intensity (observed by looking through the analyzer) is given by the square of its
amplitude:

(5.3-10)
There is only one condition of extinction corresponding to sin2 /2 = 0. This occurs when /2 =
n, where n is an integer; or, when = 2n. The integer multiples of 2 correspond to linear
phase differences which are multiples of the wavelength, . Thus, extinction occurs when the
linear phase changes by an integral number of wavelengths. The isochromatic fringe pattern
which results is governed by Equation (5.2-14). As mentioned previously, integer order fringes
correspond to the color of the background.
In conclusion, isochromatic fringes are observed when a loaded model is studied with a
circular polariscope. The isoclinic pattern is eliminated, since the model is illuminated with
circularly polarized light.

5.4 Calibration Methods for Determining the Material Fringe Value


It is evident from Equation (5.2-14) that the material fringe value, f, must be known before an
isochromatic fringe pattern can be interpreted. Since this quantity may vary for a given material
because of its age, supplier, batch of resin, temperature, etc., a test is usually conducted to
determine f even if a value is provided by the manufacturer. The associated calibration
procedure is conducted by constructing a model and testing it in a known stress state which is
easily applied.
A tensile test is one of the most popular methods for the calibration of the material fringe value.
In this case, the principal stresses are given by

(5.4-1)
where the load, P, is applied along the 1" direction; w and d are the width and thickness of
the cross section, respectively. Substituting the quantities in Equation (5.4-1) into Equation (5.214) yields

5.10

(5.4-2)
where m is the slope of the curve obtained by plotting the load versus fringe order number.
Typical tensile specimens are shown in Figures 8 and 9. A reference point must be selected for
the tapered specimen, since the stress varies over the shank.
34 4
4.5

0 12 4
3
P

Reference point

Figure 8. A standard tensile specimen.

Figure 9. A tapered tensile specimen.

Another popular calibration specimen is a beam in pure bending. In this case, the principal
stresses at the free surface on the tensile side of the beam are
(5.4-3)

where M is the moment, c is the distance between the neutral axis and the free surface (equal to
half the height, w, for a rectangular cross section), and I is the centroidal moment of inertia about
the axis around which the moment is applied (equal to 1/12 dw3, for a rectangular cross section
of thickness, d). Substituting the quantities in Equation (5.4-3) into Equation (5.2-14) gives

(5.4-4)

where Rbdry is the fringe order number at the boundary on the tensile side of the specimen.
Figure 10 shows a typical calibration specimen recorded with a dark field circular polariscope.
The integer order fringes are black and correspond to the background. The principal stresses are
parallel and perpendicular to the longitudinal axis of the beam.

5.11

0 1 2 3 4 5
P/2

P/2

a
P/2

Tensile side

Neutral axis

P/2

Figure 10. The central portion of the beam is in pure bending.


At each point in the central section of the span, the stress in the vertical direction is zero while
the stress in the horizontal direction increases linearly with thickness. The neutral axis is located
at the center of the beam. For the case shown, the material above the neutral axis is in uniaxial
compression while the material below the axis is in uniaxial tension. For the beam shown, the
applied moment is equal to Pa/2; and, Rbdry is approximately 5.0.
Note that at the corners, there are two perpendicular free surfaces. Since both principal stresses
are zero there, the fringe order number is also zero. These isotropic points (where 1 = 2 = 0)
are very good places to begin labeling the fringe order numbers.

5.5 Compensation Methods for Determining Partial Fringe Order Numbers


In photoelasticity, it is necessary to accurately determine the fringe orders in the isochromatic
fringe pattern. There are three basic techniques used to obtain partial fringe order numbers.
The color sequence method is accurate to approximately 1/10th of a fringe. It involves observing
the fringe orders of the isochromatic fringes surrounding the point in question, as well as the
color at that point. The color is compared with those listed on the chart shown in Table 1.
As an example, consider a point which lies between fringes having R = 2 and R= 3 where the
color is yellow. From the above chart, the retardation is 57 x 10-6 in. The partial fringe order at
the point in question, RP, is computed based on the retardation corresponding to R = 1.0 (22.7 x
10-6 in.) as follows:
(5.5-1)

5.12

RETARDATION (10-6 in)

COLOR OBSERVED

Black

12

Yellow

18

Red

22.7

1st Fringe

25

Blue-Green

35

Yellow

40

Red

45.4

2nd Fringe

50

Green

57

Yellow

63

Red

68.1

3rd. Fringe

73

Green

Table 1. Color compensation chart.


Once a fringe is located (usually 0 or 1st) the 2nd then 3rd fringe order can be found by tracking
the color sequence yellow-red-green toward areas of higher strain. The sequence green-redyellow corresponds to decreasing strain, leading to progressively lower fringe orders.
The second method is called the Tardy compensation method. The procedure is accurate to
approximately 1/100th of a fringe and is executed by first using a dark field plane polariscope to
position the polarizer and analyzer in the directions of principal stress. This is accomplished by
rotating the crossed wave plates until an isoclinic fringe passes through the point in question.
The quarter wave plates are then inserted at 450 relative to the polarizer and analyzer directions
to form a circular polariscope. Once this has been accomplished, the analyzer is rotated until one
of the adjacent isochromatic fringes moves to the point. The rotation angle is measured and the
fringe order computed based on
(5.5-2)
where Radj is the fringe order number of the fringe which moved to the point; the plus sign is
used if this is the lower fringe order number while the negative sign is used if it is the higher
fringe order number. In Equation (5.5-2), the rotation angle, , is measured in radians.

5.13

As illustrated in Figure 11, digital compensation may also be performed by placing a small
variable wave plate between the model and one of the quarter wave plates. The Babinet-Soleil
compensator shown in the figure, for example, consists of two adjustable wedges. The unit is
placed along one of the principal stress directions with the help of a plane polariscope. Then the
quarter wave plates are inserted to form a circular polariscope and the compensator is adjusted
until its relative retardation is equal and opposite to that of the point in question. The partial
fringe order number is based on a digital read out and determined from a calibration chart/curve
provided with the compensator by the manufacturer. Units typically have a linear response and
can determine fringe orders to within 1/100th of a fringe.

Compensator

P /4 Model

/4 A

Figure 11. A Babinet-Soeil compensator.


5.6 Calculation of the 1 Direction
In photoelasticity, the isoclinics pinpoint the directions of the principal stresses but they do not
specify which one of them corresponds to the algebraically largest eigenvalue, 1. However, a
beam in pure bending can be used to establish a rule that can later be applied to any isochromatic
pattern observed in the polariscope for which the rule was formulated.
The procedure is to align the polarizer with the maximum principal stresses found on the tensile
side of the beam. After the quarter waves plates are inserted to form a circular polariscope, the
analyzer is rotated in a known direction, say counterclockwise. It is observed whether the higher
order isochromatics move toward the lower orders on the tensile side of the beam, or, visa versa.
If, for example, the higher order isochromatics move toward the lower orders (fringes on the
boundary move toward the neutral axis) the following rule would be established:
The polarizer is in the direction of 1 if for a counterclockwise
rotation of the analyzer, higher order isochromatics move
toward lower orders.
The rule is tested at points on an arbitrary specimen. To do this, a plane polariscope is used to
position the polarizer and analyzer in the principal stress directions by passing an isoclinic

5.14

through the point. Then, the quarter wave plates are inserted to form a circular polariscope. The
rotation is executed and the movement of the isochromatics noted. If the rule holds true at the
point, the polarizer is along 1. If it does not hold true, then the analyzer is oriented along 1
(i.e., the polarizer is not).

5.7 Birefringent Coatings


In general, the method of photoelasticity relies on a birefringent model which must be fabricated
and placed in a state of plane stress. However, tests can be conducted on a prototype made from
another engineering material, such as steel or aluminum, by using a reflective adhesive to bond a
thin layer of photoelastic plastic to the surface of the specimen. During this procedure, the
assumption is made that the photoelastic coating does not reinforce the structure to the point
where the overall structural response is changed.
As illustrated in Figure 12, the photoelastic coating is studied with a reflection polariscope.
Note that, since the light reflects off the cement, it passes through the coating twice.
Coating

1st /4
Polarizer
Light source

Observer
Analyzer
2nd /4

Loads

Thickness, d

Part

Figure 12. A reflection polariscope.

The stresses in the specimen can be related to those in the coating by making the assumption that
the strains in the coating are the same as those in the specimen. This assumption can be
expressed mathematically as
(5.7-1)
where the superscripts c and s represent the coating and specimen, respectively.
For convenience, Equation (5.7-1) is expressed in terms of the principal directions. Applying
Hookes laws separately to the specimen and the coating [Equation (2.5-1) with xx = 1, yy =
2, and zz = 3 = 0] gives

5.15

(5.7-2)

where E is the elastic modulus and is the Poissons ratio. When the expressions contained in
Equation (5.7-2) are used to satisfy Equation (5.7-1) and the resulting equations solved
simultaneously for the stresses in the coating,

(5.7-3)

Subtracting the second expression in Equation (5.7-3) from the first yields
(5.7-4)
The term on the left hand side of Equation (5.7-4) is governed by a modified from of Equation
(5.2-14), by taking into account that light passes through the thickness twice, as
(5.7-5)
Substituting Equation (5.7-5) into Equation (5.7-4) and solving for the stress difference in the
specimen gives

(5.7-6)

where R is the fringe order observed in the coating, d is the thickness, and f is the material
fringe value obtained by calibrating the coating on a simple cantilever beam. Figure 13, for
example, shows a typical calibration specimen. In this case, the stress on the upper surface of the
beam is uniaxial with 1s = Mc/I and 2s = 0. The latter are principal stresses directed parallel
and perpendicular to the longitudinal axis of the beam, respectively.
As illustrated in the figure, a reference point is selected at a convenient location along the span.
The moment is computed at this location based on statics and the material fringe value is
obtained from the isochromatic fringe order at the point in question by applying Equation (5.76).

5.16

Cantilever beam

W
Reference point
Color
3

Increasing stress/strain
Figure 13. A cantilever beam is used for calibration.
Problems may occur when using this technique when the coating is applied and used under
different environmental conditions. Edge effects may also be present at free boundaries, since it
takes a finite distance for the strains to transfer from the specimen to the coating. It is also
important to keep the thickness constant to obtain a stress plot; otherwise, Equation (5.7-6) is
only valid on a point-wise basis.

5.8 Three-Dimensional Photoelasticity


Most of the studies done in photoelasticity involve a model in plane stress. In a more general
case, the method of stress freezing can be applied. The procedure consists of thermally locking
in stresses, slicing the model into pieces sufficiently thin for them to be assumed to be in a state
of plane stress, and analyzing the slices using conventional methods.
The method works based on the di-phase behavior exhibited by most photoelastic materials. The
material is heated beyond its transition temperature to break its secondary bonds. After loads are
applied, the material is cooled slowly to reform the bonds and lock in the stresses. The model is
then sliced and analyzed.
A few factors need to be taken into account. First, since the material fringe value is a function of
temperature, a calibration specimen must be placed in the oven along with the specimen.
Second, since the slices are usually thin and not too birefringent, compensation methods that
work well at low fringe order are usually required. The Babinet-Soleil compensator is an
excellent tool for this task.

5.17

5.9 Homework Problems


1. During the 1930's, Max Frocht developed the stress concentration factor chart included below
in Figure 14 by conducting a series of tests on photoelastic specimens having different
geometries.

Figure 14. Stress concentration factor charts for bars containing a hole and fillets.
The isochromatic pattern for one of these tests is illustrated in the photographs shown in
Figure 15. A blow-up of the region adjacent to the hole is shown in the photo to the left.

Figure 15. Isochromatic fringes form around the hole.

5.18

Dr. Frocht specified that the bar had a width of 0.789 in. and a thickness of 0.136 in. The
hole was 0.258 in. in diameter and the material fringe value was 86 psi/fringe/in. For the
case shown, the applied load was 146.6 lb. What should the fringe order number be
adjacent to the hole if his chart is accurate? Does this appear to be the case?
Answer: R ~ 7.2; yes.
2. Figure 16 shows the isochromatic pattern created when a disk is subjected to a diametrical
compressive load, P = 275 lb. The disk has a diameter, D = 1.25 in., and a thickness, t =
0.2 in.

P
Figure 16. Isochromatic fringes for a disk under diametrical compression.
From symmetry, it is known that the principal stress directions at the center point of the
disk are aligned with the horizontal and vertical directions; assumed as x and y,
respectively. The theory of elasticity predicts that at the center point

where P is the applied load, D is the diameter, and t is the thickness.


(a) Sketch the configuration for the polariscope used to conduct this investigation.
(b) Calculate the material fringe value. Be sure to include units in your answer.
Answer: (a) light field circular polariscope, (b) f = 86.2 psi/fringe/in.

5.19

3. Figure 17 shows the stress pattern in a beam with fillets subjected to pure bending. The
depths of the beam are 1.222 in. and 0.622 in., respectively; and, the radii of the fillets are
0.0675 in.
A stress concentration factor chart for this case is included below as Figure 18.
(a) What are the fringe orders at the points labeled as A, B and C?
(b) Assuming that the fringe order at the boundary of the shank is R = 4.0, use the
chart to predict the fringe order at point D.

A
B

R = 4.0

D
C

Figure 17. Isochromatic fringes in filleted bar.

Figure 18. Stress concentration factor chart.

Answer: (a) RA = 0, RB = 1.5, RC = 0; (b) RD = 7.6.


4. Figure 19 shows the isochromatic pattern for a portion of a beam with shallow grooves
subjected to a bending moment equal to 26.5 in-lb. A stress concentration factor chart is
shown in Figure 20. The pattern was recorded using a light field circular polariscope. The
overall depth of the beam is 0.655 in.; the depth of the minimum section located between
the grooves is 0.432 in.; the radii of the grooves are 0.222 in.; and the thickness of the
specimen is 0.246 in.
(a) Determine the material fringe value of the specimen based on the pattern
observed in the region farthest away from the grooves.
(b) Compute the stress concentration factor at the root of the notch based on the
nominal bending stress in the critical section containing the discontinuity.
(c) Determine the stress concentration factor from the chart and then compare the
result with that obtained from part (b). Present the answer in terms of
percentage error.
(d) What might the error computed in part (c) be due to?

5.20

2
12

Figure 19. Bar with shallow grooves.

Figure 20. Stress concentration factor chart.

Answer: (a) f = 90 psi/fringe/in.; b) K = 1.27; c) Kchart = 1.22, error = 4%; d) chart


formulated based on deep as opposed to shallow grooves.
5. Figure 21 shows the bending calibration specimen used to compute the material fringe
value for a material used to study the isochromatic fringe pattern for a ring subjected to
diametrical compression (see Figure 22). The four loads applied to the calibration
specimen are each 96 lb.

Figure 21. Isochromatic pattern in a beam in bending.


The height (vertical dimension) of the calibration beam is 0.763" and the thickness is
0.180". The distance between the upper loads is 3.0" while the distance between the lower
loads is 4.0". Tardy compensation showed that the fringe order number at the top/bottom
of the beam was 5.5.
The isochromatic pattern in the ring was created by applying loads equal to 202.5 lb along
the vertical diameter. The ring was cut from the same sheet of photoelastic material as the
calibration specimen; it has an inner diameter equal to 0.766" and an outer diameter of
1.532". The fringe order number at the critical point located on the inner radius and along
the horizontal diameter is 16.75.

5.21

202.5 lb

202.5 lb

(a) Determine the material fringe value of the calibration specimen.


(b) Determine the stress concentration factor at the critical location, assuming that
the area along the horizontal diameter is subjected to pure compression.
(c) What is the maximum in-plane shear stress at the critical location.
Answer: (a) f = 90 psi/fringe/in.; b) K = 5.7; c) max = 4185 psi.
6. Figure 23 shows a 6 in. long aluminum calibration specimen being used to compute the
material fringe value of a photoelastic coating. A 50-lb load, labeled as W, is applied to the
beam. The beam is 1.0 in. wide and 0.25 in. thick. It has an elastic modulus of 10 x 106 psi
and a Poissons ratio of 0.3.
The coating is 0.10 in. thick. It has an elastic modulus of 450 ksi and a Poissons ratio of
0.42. The color at the reference point is yellow; a color compensation chart is included as
Figure 24.
Compute the material fringe value based on the stress at the reference point using the

5.22

relation

Be sure to include units in your answer.

3"

3"

Figure 23. A cantilever beam is coated with a birefringent coating.

Figure 24. Color compensation chart.

Answer: f = 77 psi/fringe/in.

5.23

5.10 Classroom Demonstration in Photoelasticity


1. Photoelasticity requires a basic knowledge of plane and circular polariscopes.
conjunction with a classroom demonstration of these units:

In

a) Sketch the configuration for a dark field plane polariscope. Explain how the device
was constructed in class. After observing the isochromatics and isoclinics in a beam
subjected to pure bending, discuss the meaning of each system of fringes and
comment on any observations made during the demonstration.
b) Sketch four possible configurations for a circular polariscope. Explain how these
light and dark field configurations were constructed in class.
c) What is the main advantage of using a circular polariscope over a plane polariscope?
2. The isoclinics characterize the principal directions; however, it is impossible to determine
which direction corresponds to the algebraically largest component of the stress unless a
rule is established with a well-known stress state.
a) Establish a rule for the polariscope demonstrated in class using a beam in bending.
b) Discuss the procedure for applying the rule to an arbitrary specimen.
3. To evaluate the stress distribution at a point, it is necessary to know the material fringe
value, f; and, accurately determine the fringe order number. Calibration and
compensation methods are used to obtain f and interpolate between fringes, respectively.
a) Outline three basic methods that can be used for compensation.
b) After having placed a beam with a circular hole in uniaxial tension, calibrate the
specimen using the fringe order observed in the uniformly stressed portion of the
specimen removed from the hole. Specify the apply load and the fringe order
number.
c) Assume that the beam is loaded along the vertical direction and consider a point
adjacent to the hole along the horizontal centerline. Determine the sign of the
boundary stresses adjacent to the hole using the rule established in part 2 of the
demonstration.

5.24

CHAPTER 6 - PHOTOGRAPHY
6.1 Introduction
The use of a camera is an effective means of recording experimental data for storage or later
analysis. Conventional photography relies on recording the image of an object by focusing it
onto the emulsion side of a silver-based photographic film and/or paper. Any silver salt particle
in the emulsion which is exposed to light undergoes a physical change.
During development in a darkroom, the physical change becomes a chemical change leading to
the formation of metallic silver atoms. This transition takes place in a solution called the
developer. The action of the developer is stopped after a specified period using a second
solution called a stop bath. A third solution, called the fixer, dissolves any unexposed silver
salts remaining in the emulsion.
An alternate approach is to capture the image using a digital camera. In this case, the image is
captured on a focal plane array and stored in a computer as a three-dimensional array consisting
of the pixel locations (x,y) of points in the image and their intensities (z).

6.2 Single Lens Reflex Cameras


Figure 1 illustrates that the basic components of a camera are the lens, the diaphragm, the
shutter, and the film plane. The lens (usually compound) is used to focus the image of the object
onto a photographic film. The specifications of the lens are given in terms of focal length and
maximum f-number (f-numbers are defined as the focal length divided by the aperture). The
diaphragm adjusts the lens aperture while the shutter controls the amount of time that the film is
exposed to the admitted light.

Diaphragm
Viewer

Lenses
Film
Shutter
Figure 1. Basic components of a camera.

6.1

The amount of light allowed to react with the film can be controlled by varying the shutter speed
(in seconds or fractions of a second) or the lens aperture (the larger the aperture, the more light
admitted). When the f-number setting is increased, the aperture decreases. On most cameras,
the shutter speeds and apertures are adjusted so that a change in one f-stop setting corresponds to
one step change in shutter speed.
Figure 2, for example, shows a single lens reflex camera in more detail. With the SLR the
image is viewed through the lens (TTL). Depressing the shutter release button first causes a
hinged mirror to swing upwards clear of the light path before the shutter opens, and the film
records the image exactly as it appeared in the viewfinder. SLR cameras using 35 mm film are
most popular for general, good-quality photography. Large-format SLR cameras are used mostly
by professionals. Many have optional prism-finder attachments, with built-in exposure meters.
Film speed dial
Focusing screen
Rewind knob
Pentaprism
Shutter speeed ring
Focusing ringg
Aperture ring

Filter thread
Reflex mirror
Self timer
Shutter release
Film advance lever
Frame counter

Figure 2. A single lens reflex camera.


As illustrated in Figure 3, the controls found on most lens barrels are the aperture control ring
and the focus control ring. Other information to be found on lenses is a depth of field scale, in
both feet and meters, and an infra-red focusing mark. Apertures are aligned against a fixed
reference mark and the focusing ring should be turned until the viewfinder shows a-sharp image.
As mentioned above, the lower the f-number, the larger the aperture.

6.3 Depth of Field


The depth of field is the zone extending in front of and behind the plane of sharpest focus. In
this zone, blur is negligible, and everything may be considered as being in focus.

6.2

Focus reference mark


Depth of field scale
Focusing distances
Focus control ring
Aperture control ring

f2.8

f4

f5.8

f8

f11

f16

Figure 3. A typical compound lens.


The depth of field depends on the focal
length, lens to subject distance, and
aperture. The shorter the focal length, the
greater the depth of field. The closer the
subject, the smaller the depth of field.
And, the smaller the aperture, the greater
the depth of field.
In most engineering applications, focusing
is performed with an open aperture to
obtain the best focus and then the lens is
stopped down to take the picture with a
larger depth of field.
Figure 4 illustrates that the greater a lens's
focal length, the less its angle of view. As
the focal length increases the minimum
focusing distance also increases.

Figure 4. Lens characteristics.

6.4 Photographic Processing


As explained in data sheet included as Figure 6, there are three basic steps in processing black
and white photographic film. First, the exposed film is placed in a developer solution which
reacts with and sensitizes the emulsion (this reaction varies with developer temperature,
concentration, and with the amount of time the film remains in the bath). Second, the exposed
film is placed in a "stop bath" which stops the action of the developer. Third, the film is placed

6.3

in a "fixer" solution, which removes the backing and excess emulsion from the negative. The
negative is now ready for printing.

Figure 5. Data sheet for a 35mm photographic film.

6.4

Figure 6 shows a typical configuration for the darkroom facility used to process the film. Once
this is done, an enlarger is used to expose a piece of photographic paper by passing light through
the negative. The same development process used to process the film is used to process the
paper. Figure 7 includes the data sheet for the latter.

Figure 6. A typical darkroom.

6.5 Polaroid Land Film Cameras


Some cameras are designed to allow a user to develop the film immediately after it has been
exposed. A Polaroid Land camera relies on a packet filled with developer to do this. It takes
approximately 30 seconds to develop the image. Most of the film packets for land cameras do
not provide the user with a negative making it difficult to make copies of the print. For details,
see Figures 8 and 9.

6.5

Figure 7. Data sheet for a photographic paper.

6.6

Figure 8. Data sheet for a Polaroid Land Camera.

6.7

Figure 9. Data sheet for a Polaroid Land Camera (continued).

6.8

6.6 Digital Cameras


An alternate approach to conventional photography is to capture the image using a digital
camera, such as the ones shown in Figure 6.10. In this case, the image is captured on a focal
plane array and may be stored in a computer as a three-dimensional array consisting of the pixel
locations (x,y) of points in the image and their intensities (z).

Figure 10. Digital cameras provide an alternative to conventional photographic processing.


Digital photography has many advantages over traditional film photography. Digital photos are
convenient, allow users to see the results instantly, don't require the costs of film and developing,
and are suitable for software editing and uploading to the Internet.
The basic attribute of a digital camera that determines image quality is its megapixel rating. This
number refers to the amount of information that the camera sensor can capture in a single
photograph. Cameras with high megapixel ratings take larger pictures with more detail. Those
photos will also look better when printed, especially in bigger sizes. At five megapixels, image
quality gets in the neighborhood of film.
It should be noted that the focal plane of most cameras contains rectangular pixels which have a
length to height ratio of 4:3. In this case, care must be taken while making measurements, since
the different scale factors must be taken into consideration. For this reason, some manufacturers
provide cameras having arrays with square pixels.
There are also many additional features available on digital cameras, including image
stabilization, on-board image editing, color correction functions, auto-bracketing and burst
modes. Many of these can be handled by image editing software, and so they can be
unnecessary (and often inferior) when built into a camera.

6.9

CHAPTER 7 BRITTLE COATINGS


7.1 Introduction
Brittle coatings are lacquers which are sprayed onto a test part. When dried, they become brittle
and crack well below the yield stress of the material that they are deposited on. As illustrated in
Figure 1, cracks occur in tension, initially perpendicular to the maximum principal stress. The
cracks form isostatics; lines tangent to the principal stress directions.

Figure 1. Cracks form along isostatics in a coating applied to a tubular structure.


The brittle coating technique is widely employed in experimental stress analysis because it is
inexpensive and easy to use; and, quickly provides an engineer with vital information about the
distribution and severity of stresses in a test part. The coating provides a visual overall picture of
the strain distribution and can be applied to almost any part or structure, irrespective of the size,
shape, or material to highlight areas of stress concentration. The technique locates the critically
stressed areas, and gives the principal strain directions, as well as information on the strain
gradients; all of which are needed for the proper selection, location, and orientation of strain
gages for accurate measurement of peak stresses. Perhaps more importantly, the technique is
inexpensive and simple to use, and requires no instrumentation. In most cases, the coating has
negligible reinforcing effect on the test part. Most coatings dry to vibrant colors, offering good
contrast for observation of the crack patterns.

7.1

7.2 Testing Procedures


In most practical applications, the brittle coating is sprayed onto the surface of the part or
structure to be tested, and allowed to air-dry. After drying, the part is subjected to test loads
usually applied in incremental fashion. As the incremental loading proceeds, the lacquer first
cracks in the region where the tensile strain is greatest; the cracks are always oriented along
directions perpendicular to the maximum tensile strain.
With each additional increment of load, the cracks multiply and grow outward from the area of
peak stress, terminating at points where the strain on the surface of the test part falls just below
the lacquers threshold strain. The loci of the crack ends is typically outlined with a grease
pencil after each load increment, and identified with the corresponding load magnitude.
Figure 2 illustrates this testing procedure and shows a line called an isoentatic being drawn on a
loaded ring. The isoentatic, P, encloses an area where the strain equals or exceeds the threshold
strain. As the load is increased incrementally, several isoentatics (Pl, P2, P3 ... ) can be traced.

Figure 2. Isoentatics are drawn as the specimen is loaded incrementally.


For field testing under service loads (as in the case of aircraft or spacecraft, for example), it is
common practice to apply the brittle coating over all areas of interest and, after the lacquer has
dried thoroughly, subject the structure to the prescribed working conditions. The coating is then
carefully examined to locate the regions where the strains induced by service loads were high
enough to produce cracks in the lacquer. Strain gages can subsequently be installed in these
critical areas for accurate strain measurements. Figure 3, for example, shows the crack pattern
developed on a nuclear heat exchanger as a result of hydrostatic pressure.

7.2

Figure 3. Cracks on a heat exchanger.

7.3 Calibration
Quantitative measurement of the strains in the test part requires that the coating be calibrated to
obtain the strain level at which cracking occurs. This is accomplished by spraying a number of
calibration beams with the lacquer at the same time the test part is sprayed. The part and the
calibration beams are then allowed to dry together so that the two coatings experience identical
curing conditions. Coincident with loading the test part, the beams are placed in a cantilever
calibration fixture and subjected to a predetermined deflection at the free end. This imposes a
known linear strain gradient along the length of the beam, from which the threshold strain level
is read on the calibrator measuring scale. Figure 4 shows a typical calibration test fixture.
As the deflection is applied, cracks begin to form perpendicular to the longitudinal axis of the
beam at the fixed end where the stress is the highest. The threshold strain is determined by
pinpointing the location of the last full crack and calculating the strain at this point.
Figure 5, for example, shows a cantilever beam of length, L, fixed into a wall at the left end and
subjected to a concentrated load, P, applied at the right end.

7.3

Figure 4. A cantilever calibration fixture.


Y
_P

R
_ Ax

_P

R
_ Ay
A

Figure 5. An end-loaded cantilever beam.

M
_A

Figure 6. FBD of the cantilever beam.

Figure 6, on the other hand, shows the free body diagram corresponding to the beam illustrated
in Figure 5. Since all of the forces are contained in a single plane, the reactions at the fixed
support consist of two force reactions and one moment reaction.
Referring to Figure 6, and taking into account overall equilibrium,

(7.3-1)

Expressions for the shear force, V, and bending moment, M, can be obtained for all points along
the span by passing a section through a point, O, located at an arbitrary distance, x, from the
fixed support.

7.4

Figure 7 shows this procedure and illustrates the standard convention for assigning the shear and
moment distribution on the cut section.

M
_

O
x

PL

V_

Figure 7. A cut at position x.


Applying the equilibrium equations to the section:

(7.3-2)
The expressions in Equation (7.3-2) can be used to plot the shear and moment diagrams shown in
Figure 8. The shear force is constant over the span and equals the applied load. The bending
moment is maximum at the fixed support and progressively decreases with increasing x.

P
V
A

-PL

Figure 8. Shear and moment diagrams.


The diagrams are consistent with the fact that the shear force at each section is equal to the
derivative of the bending moment (the slope of the moment diagram). Conversely, the difference
in the bending moment between two sections is equal to the area under the shear curve. That is,
for small deflections,
(7.3-3)

The shear force creates a shear stress at every section that can be approximated by dividing the

7.5

force by the cross-sectional area. However, this approximation only gives an average value for
the stress and does not adequately define the actual distribution through the thickness. A better
description is given by,
(7.3-4)
where V is the shear force, Q is the first moment of the area measured about the neutral axis of
the portion of the cross section located either above or below the point under consideration, I is
the centroidal moment of inertia of the entire cross section, and t is the width of a cut made
through the point in question perpendicular to the applied load.
When Equation (7.3-4) is applied to a beam of rectangular cross section, it is found that the shear
stress distribution is parabolic through the thickness. The shear stress is zero at the upper and
lower surfaces of the beam and maximizes at the neutral axis.
Shear stresses always come in pairs, and it should be noted that the vertical shear force creates a
horizontal shear flow. That is, horizontal shear stresses develop between the longitudinal fibers
in the beam. This effect is zero on the free surface and maximum at the neutral axis.
The bending moment creates normal stresses parallel to the longitudinal axis of the beam. The
magnitude of these stresses is given by the elastic flexure formula,
(7.3-5)
where M is the moment, y is the distance measured from the neutral axis to the point under
consideration, and I is the centroidal moment of inertia corresponding to the axis about which the
moment is applied.
Equation (7.3-5) shows that the normal stress is linearly distributed through the thickness. It
becomes a maximum at the free surfaces and reduces to zero at the neutral axis.
Equation (7.3-5) can be applied to determine the stress on the upper surface of the cantilever
beam shown in Figure 5. Assume that the beam has width, b, and thickness, t. Since the neutral
axis is located at the geometric center of the beam, y = t/2. The centroidal moment of inertia
measured around the bending axis (z) is I = 1/12 bt3; the moment is given as a function of x by
Equation (7.3-2). Substituting these values into Equation (7.3-5),
(7.3-6)

Equation (7.3-6) shows that the normal stress at the top of the section is tensile. Since the fibers
on this surface do no experience any shear stress, the points located there are in a state of
uniaxial stress. In this simple case, the constitutive relations (Hookes Laws) reduce to

7.6

(7.3-7)
where E is the elastic modulus and xx is the normal strain measured along the longitudinal axis
of the beam.
Substituting Equation (7.3-7) into (7.3-6) and solving for the strain,
(7.3-8)
Equation (7.3-8) implies that the strain along the beam varies linearly with x.
The load can also be obtained by knowing the deflection under the load. The governing
expression can be found by applying the equation,
(7.3-9)
where M(x) is the expression for the applied moment given by Equation (7.3-2). Thus,
(7.3-10)
Integrating once,
(7.3-11)
Integrating again,
(7.3-12)
Applying the boundary conditions [@ x = 0, dy/dx = 0] and [@ x = 0, y = 0] to the expressions
in Equations (7.3-11) and (7.3-12), respectively, reveals that C1 = C2 = 0. Therefore,
(7.3-13)
Under the load at x = L,
(7.3-14)
Denoting the absolute value of this deflection by end, and solving for P,

7.7

(7.3-15)

7.4 Measurements
Although considered primarily a qualitative tool, the method can provide an approximate
measure of surface strain magnitudes. If care is taken in the preparation, application, and testing
procedure, a coating that is properly calibrated can yield quantitative results accurate to 100 .
Under optimum conditions, accuracy on the order of 50 can be achieved.
The nominal sensitivity, or threshold strain, t*, of most brittle coatings is around 500 u. The
stress in the specimen is calculated based on its elastic modulus, Es, from the simple relation

(7.4-1)
Equation (7.4-1) assumes that the stress at the target point is uniaxial; whereas, in general, the
part is in a state of biaxial, or plane, stress. This, in itself, may give rise to substantial error. But
other errors exist due to a mismatch in the Poissons ratio between the specimen and coating
materials.
In general, the stress in the specimen is not equal to that in the coating. This can be
demonstrated by making the assumption that the strains in the coating are the same as those in
the specimen. This assumption can be expressed mathematically as
(7.4-2)
where the superscripts c and s represent the coating and specimen, respectively.
By assuming plane stress (zz = 0) and applying the Hookes laws, previously given in Equation
(2.5-1), separately to the specimen and the coating:

(7.4-3)

where E is the elastic modulus and is the Poissons ratio. When the expressions contained in
Equation (7.4-3) are used to satisfy Equation (7.4-2) and the resulting equations solved
simultaneously for the stresses in the coating,

7.8

(7.4-4)

During calibration, the stress on the upper surface of the cantilever beam is in uniaxial tension
and, when the x axis is directed along the span, characterized by
(7.4-5)
Since the strain in the specimen is equal to the strain in the coating,
(7.4-6)
Substituting these stresses into Equation (7.4-4) reveals

(7.4-7)

Equation (7.4-7) shows that, in general, a biaxial stress state is produced in the coating even
when the stress state is uniaxial in the specimen. The only condition under which Equation (7.47) predicts a uniaxial stress state occurs when c = s. But this is far from the case for most
engineering materials that have a Possons ratio of approximately 0.3; since, the Possions ratio
of brittle coatings are typically 0.42.

7.5 Coating Selection


As mentioned previously, the threshold strain of most brittle coatings is around 500 u.
However, the actual threshold for a particular test depends upon the application and the test
environment. Because its cracking behavior is affected by both temperature and humidity, many
different coatings are manufactured to provide optimum performance over a wide range of
environmental conditions.
The proper coating for a given test condition is obtained from a selection chart, such as the one
shown in Figure 9 (for Tens-Lac manufactured by Photolastic Inc.).
If the anticipated test temperature were 700 F (210 C) and the relative humidity 50%, for
example, the appropriate coating indicated by the selection chart would be TL-500-75. At the
same temperature, but with 30% relative humidity, either TL-500-75 or TL-500-70 would be

7.9

suitable. The lacquer with the higher number (TL-500-75) would have greater sensitivity (that
is, a lower threshold), while the TL-500-70 would begin cracking at higher strains.

Figure 9. Coating selection chart.


The lacquer normally begins to crack when a nominal strain level of 500 is reached.
However, an increase in temperature or humidity will raise the threshold of the brittle coating
and cause a reduction in sensitivity. At thresholds greater than 1200 u, the cracks will not
always remain open after removing the load, and it may be necessary to observe the crack
patterns on the test part while under load. The threshold of most brittle lacquers can be lowered
by decreasing the temperature. However, there is a practical lower threshold limit of about 300
.

7.10

7.6 Application
A brittle coating can be applied to almost any test material. Application of the lacquer must be
preceded, however, by proper preparation of the surface to render it completely free of dirt, oils,
rust, loose paint, etc. After cleaning, an undercoating is sprayed on the part and allowed to air
dry (15 to 30 minutes). The undercoating is an aluminized covering agent which provides a
uniformly reflective surface for promoting crack visibility in the overlying lacquer.
When the undercoating has dried, the brittle lacquer is sprayed over it in a succession of thin, wet
layers, allowing about two minutes between coats. For most test situations a uniform thickness
of. about 0.15 mm (0.006 in.) is optimum. The thickness is monitored by comparing the color to
a test strip provided by the manufacturer. Thicker coatings appear as darker colors whereas
thinner coatings appear lighter. After spraying the final layer, the lacquer must then be dried at
3-60 C (5-100 F) above the expected test temperature. Best performance of the lacquer is
achieved after a drying period of 18-24 hours.

7.7 Crack Patterns


Cracks only occur in areas where the threshold strain is reached. Since the surface is in plane
stress, zz = 0. In the case where the in-plane principal stresses are of opposite sign, cracks will
occur perpendicular to the maximum tensile stress provided, of course, that the threshold is
reached. When both principal stresses are positive, two families of cracks may result. Cracks
will initially form perpendicular to the largest stress and then, if the strain is high enough,
perpendicular to the lower tensile stress. For the special case where both principal stresses are
positive and equal, the cracks will appear random but will always intersect at 900.

7.8 Relaxation Techniques


A brittle coating cracks only when placed in tension. To test for compressive loads, the
specimen is first loaded and then coated. As the load is relaxed, the coating cracks.

7.11

7.9 Homework Problems


1. Figure 10 shows the isostatic fringe pattern observed in a brittle coating that was applied to
the top surface of a 9 long end-loaded cantilever beam (Elastic Modulus = 30 x 106 psi;
Poisson's Ratio = 0.3). When a load of P = 100 lb was applied, it was observed that the
crack pattern extended 3" from the fixed end.
(a) Determine the reactions at the fixed support.
(b) Sketch the shear and moment diagrams for the beam.
(c) Determine the threshold strain for the coating.

6 in.
3 in.
P

2 in.

0.25 in.

Figure 10. Isostatic pattern in a brittle coating applied to an end-loaded cantilever.


Answer: (a) Rx = 0, Ry = 100 lb , M = 900 in.-lb clw, b) shear is constant; moment is linear, c)
t* = 480 .

7.12

CHAPTER 8 MOIR METHODS


8.1 Introduction
Moir patterns result from interference caused by superposing two geometrical patterns. These
interactions are very common in everyday life and may be observed if one looks closely for
them. Figure 1, for example, shows a moir formed by two wire mesh bowls used in food
preparation. The fringes correspond to the superposition of two amplitude gratings, comprised
of opaque bars and clear spaces. The pattern is indicative of both the spacing between the bowls
and their differences in shapes.

Figure 1. A moir pattern is formed by two wire bowls.


The arrays used to produce moir fringes may be a series of straight parallel lines, a series of
radial lines emanating from a point, a series of concentric circles, or a pattern of dots. In stressanalysis, arrays typically consist of straight parallel lines, ideally formed with opaque bars
having clear spaces of equal width.
The principles underlying the geometrical moir methods described in this chapter are relatively
straightforward. However, these methods suffer from low sensitivity and are mainly used to
measure displacement, as opposed to strain.
Recent advancements in grating replication techniques, such as interferometry and microlithography, have led to the development of extremely sensitive methods such as moir
interferometry and diffractive-optic interferometry. But these techniques require studying the

8.1

complex superposition of coherent wavefronts diffracted by high frequency gratings deposited or


projected onto test surfaces.

8.2 Analysis
In conventional moir analysis, the procedure used to measure in-plane displacement of a test
specimen is to fasten or deposit a master grating onto its surface. The distance between the lines
in the master when it is undeformed is referred to as the pitch, p. The frequency, f, of the grating
is the reciprocal of the pitch. Hence, the relation between pitch and frequency (density) is
(8.2-1)
Displacement is always measured perpendicular to the lines in the master grating. When the
frequency of the grating is low [0.4 lines per mm (10 lines per inch) or less], image analysis
techniques are normally employed to measure the changes in spacing at each point before and
after loading. Displacements and strains are extracted from these data. For example, the strain
in the direction perpendicular to the lines in the master grating is simply
(8.2-2)
where p is the change in pitch.
For gratings with higher frequency and smaller pitches [2 to 40 lines per millimeter (50 to 1000
lines per inch)], the specimen grating is typically superimposed with a master grating to form a
moir fringe pattern. The master grating can be another grating placed in front of the specimen
or a stored copy (photographically or digitally) of the undeformed specimen grating.
Assuming that the master grating is placed in the xy plane with its lines parallel to the y
direction, the displacement component along the x direction is given by

(8.2-3)
where nx is the fringe order number.
Figure 2, for example, shows the interference produced by a uniform elongation of the
specimen. In this example, it is assumed that p is the pitch of the master grating while p1 is the
pitch of the specimen grating.
A dark fringe appears at the points where an opaque strips of the master grating fall over the
transparent strips of the specimen grating. When the opaque strips of the gratings coincide, there
is a maximum of light intensity and a light fringe appears. Light fringes represent integer orders.

8.2

Center of a light fringe


Center of a
dark fringe

Center of a Center of a
light fringe dark fringe
p1
Specimen
grating

Master
grating
p

Figure 2. A moir pattern is formed due to specimen elongation.


Figure 3, on the other hand, shows a more general case where the specimen deforms differently
at each point. In this case, light fringes appear along the shorter diagonals of the intersecting
grating lines. Even though the superposition is more complex than that illustrated in Figure 2,
Equation (8.2-3) still holds.

-2

-1

2 3

Master grating
Figure 3. A moir pattern for the case of general deformation.
The fringe orders are typically established with respect to a reference point. The black dot on
Figure 3, for example, can be selected as a starting point and the fringe passing through it
labeled as zero (implying that the specimen grating did not move relative to the master grating at
this point. As one moves along fringe order 0" progressively to the right, all intersections
coincide to points where the lines in the master did not move relative to those on the specimen.
At the intersections of the gratings along fringe number one, all points in the specimen move one
pitch to the right relative to the master, etc.
In general, two in-plane displacement maps are required to completely determine the overall

8.3

strain state; and, a cross grating is usually deposited on the specimen.


Assuming that the gratings have equal pitch, the lines parallel to the x axis produce an additional
moir fringe pattern which corresponds to the displacement component in the y direction, v,
governed by

(8.2-4)
where ny is the fringe order number.

8.3 Optical Filtering


When a cross grating is used the two moir patterns are superimposed but they can be separated
by sampling the diffraction spectrum in an optical filtering arrangement. To understand this,
Figure 4 shows that when light emanates from a point source, spherical wavefronts are
produced.

Point
source
2

Collimating lens

Grating 1

4
3
2
1
0
-1
-2
-3
-4

Transform lens

Figure 4. Light is diffracted by a grating.


Plane waves are created when a lens is placed at a focal distance from the source; and, the light
beam is referred to as collimated. When the collimated beam is used to illuminate a grating, the
light is bent due to diffraction. Each clear space in the grating produces a separate source of
light and when the light is collected by another lens, the wavefronts from the different sources
combine to produce interfere in the lens transform plane. The dots represent points where
constructive interference occurs. The dots are oriented perpendicular to the lines in the grating
and spaced at a distance, d, apart. The spacing is given by

8.4

(8.3-1)
where is the wavelength, p is the pitch, and f is the focal distance of the transform lens.
Equation (8.3-1) shows that the smaller the pitch, the larger the spacing between the diffraction
orders.
Figure 5 shows an optical filtering arrangement that can be used to separate the moir pattern
from the gratings that form it. Coherent light from a laser is expanded and collimated by placing
a lens at a focal distance away from a spatial filter. The light is used to illuminate a photographic
film containing the master and specimen gratings, as well as the moir pattern that they produce.

Input

Output
3

2
1

Transform plane

+1
0
-1

Moir
and
gratings

Moir
less
gratings

Figure 5. An optical filtering arrangement.


The gratings coincide at the center of a light moir fringe and produce diffraction that contributes
to the dot pattern observed in the transform plane. A filter is positioned in this plane that allows
only the +1 diffraction order to pass through the system. When the imaging lens creates the
filtered image, moir fringes are observed.
A similar procedure is followed when a crossed grating is used to study the displacement of the
specimen. In this case, the diffraction pattern is relatively complex consisting of a rectangular
array of dots. However, the moir patterns are separated by allowing the corresponding +1
diffraction order to pass through the system.

8.4 Stress Analysis


Once the in-plane displacements have been determined, the strains can be computed based on a
reduced form of the general strain-displacement relations given in Equation (2.2-1). For small
strains measured on a traction free surface in plane stress, these relations become

8.5

(8.4-1)

The terms in Equation (8.4-1) can be evaluated by plotting u versus x, u versus y, v versus x, and
v versus y over the full field and then computing the slopes of the curves at the point(s) in
question, Once the strain components are established, the stresses can be computed by applying
the constitutive equations for plane stress. Principal values are obtained using the transformation
equations.
If the strains are large, higher order terms must be taken into account. In the Lagrangian
description, for example,

(8.4-2)

The derivatives of the in-plane displacement components are determined from the slopes of the
plots produced by analyzing the moir patterns produced by a cross grating. The derivatives of
the out-of-plane displacement can be determined by plotting the information obtained from the
shadow moir technique described below.

8.5 Shadow Moir


As illustrated in Figure 6, the out-of-plane displacement, w, can be measured by projecting a
grating onto the surface of a specimen. In formulating the argument, the assumption is made that
the light source and camera are located at equal distances from the surface.
The displacement is typically measured relative to a flat reference surface. In Figure 6, for
example, a moir pattern is generated by superimposing the image of the grating projected onto
the reference surface (AB) with that projected onto the specimen (AB). The displacement
normal to the surface is given by

(8.5-1)

8.6

where p is the pitch and n is the fringe order number. Although the angle of illumination, , and
the angle of observation, , vary from point to point, the sum of their tangents is constant, and
Equation (8.5-1) holds over the full field.

x
x
A
A

B
B

dz


b c
d

e
x

Light
source

O
D

Camera

z
Figure 6. Set up for the shadow moir method.
The perspective effect of the coordinates can be corrected for a large specimen by using the
relations
(8.5-2)
where x,y and x,y are the apparent and real coordinates, respectively.
For the special case where the angle of illumination is set equal to 450, and the surface is viewed
normal to the reference surface at = 0, Equation (8.5-1) reduces to
(8.5-3)
The shadow moir method can be used to obtain the thickness change in a specimen. If a
specimen of thickness, t, is in a state of plane stress, the strain in the direction normal to the
surface of a specimen is given by

(8.5-4)

8.7

8.6 Homework Problems


1. Figure 7 shows the moir pattern created by printing a grating of equispaced circles with a
pitch of 1/10,000" on a 1.5" wide, 1/4" thick, flat uniaxial tensile specimen. The moir
pattern shown in the figure was observed when a circular master of the original size was
placed in front of the deformed specimen. A load, P = 4,500 lb, was applied to create the
fringe pattern.
(a) What is the Young's modulus of the specimen?
(b) What is the Poisson's ratio of the specimen?
(c) What kind of material do you think this is?

1.5 in.
P

1.5 in.

P
Figure 7. A moir pattern created by concentric circles.
Answer: (a) E = 10 x 106 psi, (b) = 0.3, (c) aluminum.
2.

A moir analysis was performed on the tension specimen shown in Figure 8. Gratings
having a spatial frequency of 1000 lines/in. were used in the analysis. The moir fringe
patterns are shown superimposed on the specimen with the orientation of the grating used
to generate the pattern depicted to the right of each figure. Knowing,
t = 0.5"

w = 3"

d = 6.0"

(a) What is the Young's modulus of the specimen?


(b) What is the Poisson's ratio of the specimen?
Answer: (a) E = 500 x 103 psi, (b) = 0.4.

8.8

P = 1,875 lb.

Figure 8. Moir patterns created by linear gratings.

8.9

CHAPTER 9 STRAIN GAGES


9.1 Introduction
As explained in Chapter 2, there are six components of strain which may be expressed in terms
of the displacements via Equation (2.2-1) as

(9.1-1)

When measurements are made on a free surface contained in the x,y plane, the plane stress
assumption,

(9.1-2)

holds, and the relations in Equation (9.1-1) reduce to

(9.1-3)

Since there are three independent non-zero relations in Equation (9.1-3), it is sufficient to make
three independent strain measurements at the point in question to completely define the strain
state there.
One approach is to measure the normal strain along three different directions and use the strain
transformation relations

(9.1-4)

9.1

to extract the required information. Once this is done, the constitutive equations [see Equation
(3.2-3)] can be applied to obtain the stress; transformation equations can be used to obtain the
principal values and/or maximum shear stress.
There are many different types of strain gages and they are nominally classified into five basic
groups: mechanical, electrical, optical, acoustical, and pneumatic. Most of these devices are
designed to obtain normal strain in one direction by measuring the change in length which occurs
over a relatively large gage length.
The most important characteristics of a strain gage are its length, sensitivity, accuracy, and
range. Since most units measure the average strain over the length, it is advantageous to make
the latter as small as possible. The sensitivity is the smallest reading that can be measured on the
scale associated with the gage. But, sensitivity does not necessarily mean accuracy. The latter
has to do with the smallest quantity that can be extracted when measurements are repeated.
Range, on the other hand, is the extent to which measurements can be made without having to
replace and/or realign the gage.
It is also advantageous to have an output which is linearly proportional to strain, so that
calibration curves and/or look-up tables are not required.

9.2 Electrical Resistance Strain Gages


The electrical resistance strain gage is perhaps the most widely employed strain measurement
device. As illustrated in Figure 1, a typical strain gage consists of a metal foil alloy etched into
a grid pattern. The gage is bonded to the test beam so that the foil experiences the same normal
strain (L/L) as the specimen.

Backing
Encapsulation

Copper-coated tabs
Figure 1. An electrical resistance strain gage.
The resistance of the gage, R, is measured in Ohms () and may be expressed as
(9.2-1)

9.2

where is the resistivity [measured in -m (-in.)], L is the length [m (in.)], and A is the crosssectional area [m2 (in.2)].
When the strain parallel to the grid is positive, the length of the foil increases while its crosssectional area decreases. Assuming that the resistivity remains constant, Equation (9.2-1) shows
that the resistance of the gage will increase. When a constant current is passed through the gage,
this change in resistance, R, produces a change in voltage (i.e., Ohms law; V = IR). The
voltage change is converted to strain by using a strain indicator.
The manufacturer of the gage provides a gage factor, Sg, so that the change in resistance can be
correlated with the strain. Sg is a dimensionless quantity typically having a magnitude of 2.0.
The gage factor equation is given by,

(9.2-2)

where R is the resistance and is the strain. Rearranging terms,


(9.2-3)
Typically, R = 120 and Sg = 2.0; strain ranges from max = 10,000 to min = 5 .
Substitution of these values into Equation (9.2-2) shows that the resistance changes to be
measured range from 2.4 to 0.012 . The most practical means of accurately measuring these
relatively small resistance changes is by using a Wheatstone bridge. The bridge circuit, shown in
Figure 2, has an input voltage source, V, four resistors, R, and a galvanometer, G.

R1

R2

V
R4

R3
G

Figure 2. A Wheatstone bridge circuit.

9.3

In most cases, R1 is replaced by an active strain gage; R2 is a variable resistor used to balance or
zero the bridge; and, R3 and R4 remain constant.
The galvanometer is balanced and displays a zero reading when the resistance ratio is satisfied:
(9.2-4)
As the gage in R1 is strained, the bridge becomes unbalanced and
(9.2-5)
However, the bridge can be rebalanced by adjusting R2 through a Rm such that
(9.2-6)
If R = R1 = R2 = R3 = R4, the circuit is balanced when R1 = Rm. In this case,
(9.2-7)
Rm is measured by using a commercially available instrument called a strain indicator. The
input must include specifying the gage factor and the resistance of the gage. Detailed operating
instructions for a commercially available unit are included in Section 9.9.

9.3 Length Considerations


The gage length of a strain gage is the active or strain-sensitive length of the grid. The endloops
and solder tabs are considered insensitive to strain because of their relatively large crosssectional area and low electrical resistance.
As mentioned previously, a strain gage measures the average strain over its length;
consequently, it is desirable to make this dimension as small as possible. Figure 3 illustrates that
the average of any nonuniform strain distribution is always less than the maximum.
Consequently, a strain gage that is noticeably larger than the maximum strain region will indicate
a strain magnitude which is too low.

9.4

Peak strain
Indicated strain

x
Figure 3. Strain is averaged over the length of the gage.

9.4 Transverse Sensitivity Corrections


In order to produce a small gage length and achieve the necessary resistance for the Wheatstone
Bridge circuit, the manufacturer is forced to etch the foil into loops. The ends of each loop (see
Figure 1) result in a sensitivity to transverse strain.
This effect, referred to as transverse or cross sensitivity, can be eliminated by introducing a
correction factor found based on the transverse sensitivity factor, Kt, and the ratio of transverse
to axial strain. Kt is provided by the manufacturer, and can be used along with charts (also
provided by the manufacturer) to obtain the true strain readings.
An alternate approach is to read strain from two gages oriented perpendicular to one another.
The true strain measurements are obtained from the measured readings by applying

(9.4-1)

where Kt is the transverse sensitivity factor, and o is the Poissons ratio of the beam used to
calibrate Kt. Most strain gage manufacturers use a calibration beam having o = 0.285.

9.5 Temperature Compensation


Of all modem strain gage alloys, constantan is the oldest, and still the most widely used. This
alloy has the best overall combination of properties needed for many strain gage applications. It

9.5

has, for example, an adequately high strain sensitivity, or gage factor, which is relatively
insensitive to strain level and temperature. Its resistivity is high enough to achieve suitable
resistance values in even very small grids, and its temperature coefficient of resistance is not
excessive. In addition, constantan is characterized by good fatigue life and relatively high
elongation capability. Very importantly, constantan can be processed for self-temperature
compensation to match a wide range of test material expansion coefficients.
Self-temperature-compensated strain gages are designed to produce minimum thermal output
(temperature-induced apparent strain) over the temperature range from about -450 to +2000 C (500 to +4000 F). When selecting either constantan (A-alloy) or modified Karma (K-alloy) strain
gages, the self-temperature-compensation (S-T-C) number must be specified. The S-T-C number
is the approximate thermal expansion coefficient in ppm/0C (ppm/0F) of the structural material
on which the strain gage will display minimum thermal output. The strategy of the designer is to
balance apparent strains caused by differences in the coefficients of expansion of the gage and
backing materials with the fluctuations in resistivity with temperature to obtain linear response
over an extended temperature range.

9.6 Rosettes
At every point in a solid body there are three mutually perpendicular planes that are called the
principal planes. When a Cartesian axes system is assigned, the unit vectors aligned normal to
these planes are called eigenvectors. By definition, there are no shear stresses on the principal
planes; consequently, the normal stresses are aligned with the eigenvectors. The magnitudes of
the normal stresses are referred to as eigenvalues. The three principal stresses (eigenvalues)
include the maximum and minimum values of the normal stress at the point in question. As
discussed in Chapter 1, the eigenvectors and eigenvalues are found by using transformation
equations which depend on rotations of the coordinate axes.
When a point is located on a free surface (plane stress), one of the eigenvectors is normal to the
surface. The corresponding eigenvalue is equal to zero. The other two eigenvectors (principal
stress directions) and eigenvalues (principal stresses) are found by considering a rotation of
coordinates around this eigenvector. This can be done analytically by using the transformation
equations or graphically by using Mohrs circle.
Since the in-plane principal stresses are always perpendicular to one another, it is sufficient to
determine the angle from a reference axis to only one of them. Since the magnitudes of the
stresses are also unknown, three independent equations are required to fully establish the stress
state. In the case of a linearly elastic, homogeneous, and isotropic material, the directions of
principal stress and strain coincide, and three independent strain measurements can be made to
determine the two in-plane principal stresses and their directions. In most cases, this is done by
measuring the normal strains along three different directions.
The three axes along which strains are measured can be arbitrarily oriented about the point of
interest. For computational convenience, however, it is preferable to space the measurement

9.6

axes apart by submultiples of , such as /3 (60o) or /4 (45o). Integral arrays of strain gages
intended for simultaneous multiple strain measurements about a point are known as rosettes.
Figure 4 shows two of the most common commercially available strain rosettes. The rosette at
the left is called a "delta" or equiangular rosette while the rosette at the right is a 45 o rectangular
rosette. The configurations are analyzed by the manufacturer who provides the governing
equations required to extract critical information regarding the strain and the stress.

Figure 4. The delta (left) and rectangular (right) rosette.


To perform this analysis, an axis is typically aligned with one of the gages. The general strain
transformation
(9.6-1)
is applied to each of the gages with measured counterclockwise from the x axis. The three
equations are manipulated to obtain xx, yy, and xy in terms of the three normal strain readings
taken from the gages (1, 2, 3). Then, a final transformation is made to obtain the principal
values.
For the delta rosette, the principal strains (p and q) and the orientation of the principal axes (p
and q) are given in terms of the three measured strains as,
(9.6-2)
and
(9.6-3)

9.7

The rectangular rosette, on the other hand, is characterized by,


(9.6-4)
and
(9.6-5)

The algebraically maximum (p) and minimum (q) principal strains correspond to the plus and
minus alternatives, respectively. As discussed in Chapter 3, even though there is no stress
normal to the surface (i.e., zz = 0), the surface deforms in this direction and the corresponding
strain (zz) is
(9.6-6)
where is the Poissons ratio.
The information in Equation (9.6-6) can be introduced into the generalized Hookes Laws to
obtain the simplified expressions for the biaxial case at hand. For plane stress, the principal
stresses are given by
(9.6-7)
where E is the elastic modulus.
Since the principal directions of stress and strain coincide for the material in question, Equations
such as (9.6-3) and (9.6-5) can be used to obtain the principal stress directions.
Several other configurations are commercially available. Some of these have two pair of gages
with the gages in each pair oriented perpendicular to one another. The configuration provides an
extra reading for redundancy. The main advantage is that the strains in each gage pair can be
corrected for transverse sensitivity using Equation (9.4-8).

9.7 Gage Selection and Series


Most strain gages are characterized by a gage designation series which is provided by the
manufacturer. Figure 5, for example shows how Micro-Measurements, Inc. designates their
gages. It is evident that there are but five parameters to select, not counting options. These are:
the gage series, the S-T-C number, the gage length and pattern, and the resistance.

9.8

Figure 5. Gage designation.


Of the preceding parameters, the gage length and pattern are normally the first and second
selections to be made, based on the space available for gage mounting and the nature of the stress
field in terms of biaxiality and expected strain gradient. The next step is to select the gage series,
thus determining the foil and backing combination, and any other features common to the series.
After selecting the gage series (and option, if any), reference is made to a catalog to record the
gage designation of the desired gage size and pattern in the recommended series. The resistance
and S-T-C number are then specified.

9.8 Dummy Gages and Transducer Design


More than one strain gage can be inserted into the Wheatstone bridge circuit previously shown in
Figure 2. In general, the reading from the bridge is proportional to the net strain given by

(9.8-1)
where the constant K is typically set to one by specifying the resistance and gage factor on the
instrument used to measure the strain.
Equation (9.7-1) shows that the strains in gages placed in opposite arms add while the strains in
gages placed in adjacent arms subtract. This property can be used to compensate for
temperature, as well as, for transducer design.
Figure 6, for example, shows a cantilever beam which is part of a transducer used to measure the
load, P. The beam is equipped with a single strain gage attached to its upper surface.
For argument sake, assume that the strain reading consists of two parts: one corresponding to the
actual strain produced by the load, P, the other an apparent strain due to temperature, T. If the

9.9

gage is placed in Arm No. 1 of the bridge and the bridge adjusted so that K = 1.0, Equation (9.81) predicts
(9.8-2)

L
P

Figure 6. A load transducer.


Since the net strain includes the apparent strain due to temperature, the system is not
compensated for environmental effects.
One approach which can be taken to compensate for environmental effects is to place a dummy
gage in Arm No. 2 of the bridge. The dummy gage is placed on a similar material that is not
loaded but subjected to a similar environment so that it experiences T only. In this case,
Equation (9.8-1) predicts
(9.8-3)
The dummy gage is not the only strategy that can be applied. Figure 7, on the other hand,
depicts a beam equipped with two gages, one on the upper surface and one on the lower surface.
If the beam has a rectangular cross section, the mechanical strain in the gage on the lower surface
is of opposite sign from that on the upper surface; both gages experience the same apparent strain
due to temperature. By placing the second gage into Arm No. 2 with the original gage still in
Arm No. 1, Equation (9.8-1) predicts
(9.8-4)
Equation (9.8-3) shows that not only is the transducer temperature compensated, it is twice as
sensitive to load.

9.10

L
2
P

Figure 7. A temperature compensated unit.


The exercise demonstrates how environmental effects can be compensated using the fact that
similar readings in adjacent arms of a Wheatstone bridge subtract from one another. The
transducer could have been designed with four active gages, two on the upper surface and two on
the lower surface of the cantilever. If they are placed in Arms No. 1 and 3, and 2 and 4,
respectively, Equation (9.8-1) predicts
(9.8-5)
Other transducers can be designed to measure bending, torsion, pressure, etc. using the unique
properties of the Wheatstone bridge and possibly incorporating dummy gages.

9.9 P-3500 Strain Indicator


The P-3500, shown in Figure 8, is a portable, battery-powered precision instrument which can
be used to monitor strain in a surface mounted electrical resistance strain gage. As illustrated in
Figures 9, the unit features an LED output. The push-button controls provide an easy-to-follow,
logical sequence of setup and operational steps.
Connect Strain Gage
Resistive strain gages are normally connected at the binding posts located on the right of the
front panel. These binding posts are color-coded in accordance with conventional practice, and
are clearly labeled. Input connections for full-, half-, and quarter-bridge configurations are
shown on the inside cover of the instrument.
To use this unit, connect the strain gage leadwires to the binding posts. Select the desired bridge
configu1ration using the BRIDGE push button. Set the MULT push button to X1 position.

9.11

Figure 8. P-3500 strain indicator.

Figure 9. P-3500 push-button control panel.

Initially adjust using AMP ZERO - Orange


Depress the AMP ZERO push button. Allow the instrument to warm up for at least two minutes.
Rotate the AMP ZERO finger-tip control for a reading of 0000. [To save time, the instrument
may be left in the AMP ZERO position while the gage(s) is being connected.]
Set the GAGE FACTOR - Orange
Set the GAGE FACTOR switch to the desired gage factor range and depress the GAGE
FACTOR push button. The gage factor will be displayed on the readout. Rotate the GAGE
FACTOR potentiometer for the exact desired gage factor, and lock the control in place. The
knob utilizes a lever which must be rotated clockwise to lock it in place. The knob can be
unlocked simply by rotating the lever back to the counterclockwise stop.
Select RUN - Green
Select the X1 or X10 MULT position as required and depress the RUN push button. In this
position all internal circuitry is configured to make an actual strain measurement. Set the
BALANCE range switch and rotate the BALANCE potentiometer to obtain a reading of 0000
with no load on the test structure, and lock the control in place. The knob utilizes a lever which
must be rotated clockwise to lock it in place. The knob can be unlocked simply by rotating the
lever back to the counterclockwise stop.
Note: A reading of 0000 with flashing colons indicates an off-scale condition that is usually
caused by improper input wiring or a defective strain gage installation.
The test structure may now be loaded and the reading recorded.

9.12

9.10 Two-Wire and Three-Wire Strain Gage Circuits


All commercial static strain indicators employ some form of the Wheatstone bridge circuit to
detect the resistance change in the gage with strain. When a single active gage is connected to
the Wheatstone bridge with only two wires, as illustrated in Figure 10, both of the wires must be
in the same leg of the bridge circuit with the gage. One of the effects of this arrangement is that
temperature induced resistance changes in the lead wires are manifested as apparent strain by the
strain indicator. The errors due to lead wire resistance changes in single-gage installations with
two-wire circuits can be minimized by minimizing the total lead wire resistance; that is, by using
short lead wires of the largest practicable cross-section.

Figure 10. Two-wire circuit for connecting a single active strain gage (quarter-bridge circuit).
When two matched gages are connected as adjacent legs of the bridge circuit (with the same
length lead wires, maintained at the same temperature), the temperature effects cancel since they
are the same in each leg, and like resistance changes in adjacent legs of the bridge circuit are self
-nullifying. The two-wire circuit for connecting strain gages is shown in Figure 11.

Figure 11. Two-wire circuit for connecting two strain gages (half-bridge circuit).
Lead wire effects can be virtually eliminated in single active gage installations by use of the
"three- wire" circuit. In this case, a third lead (representing the center point connection of the

9.13

bridge circuit), is brought out to one of the gage terminals. Resistance changes in the bridge
center point lead do not affect bridge balance. For this method of lead wire compensation to be
effective, the two lead wires in the adjacent bridge arms should be the same length, and should
be maintained at the same temperature. The three-wire circuit is the standard method of
connection for a single active temperature-compensated strain gage in a quarter-bridge
arrangement. This approach is illustrated in Figure 12.

Figure 12. Three-wire circuit for connecting a single active gage (quarter-bridge circuit).
Contact resistance at mechanical connections within the Wheatstone bridge circuit can lead to
errors in the measurement of strain. Connections should be snugly made. Following bridge
balance, a "wiggle" test should be made on wires leading to mechanical connections. No change
in balance should occur if good connections have been made.
If necessary, contact surfaces may be cleaned of oils with a low residue solvent such as isopropyl
alcohol. If long periods of disuse have caused contact surfaces to tarnish, clean them by scraping
lightly with a knife blade.

9.14

9.11 Homework Problems


1. Figure 13 shows how a strain gage is attached to measure the strain in the longitudinal
direction on the cylindrical surface of a pressure vessel of 23.5 inch outside diameter and
0.3 inch wall thickness. The strain gage reads 120 x 10-6 in/in and the material properties
are E = 30 x 106 psi and = 0.25.
(a) Determine the three principal strains on the cylindrical surface of the vessel.
(b) Compute the principal stresses in the wall.
(c) What is the gage pressure inside the vessel?

Figure 13. A strain gage mounted on a cylindrical pressure vessel.


Note: The biaxial stress distribution in a pressure vessel is given in terms of the gage pressure, p,
the inner radius, ri, and the thickness, t, by

Answer: (a) xx = 120 , yy = 420 , zz = -180 , (b) xx = 7200 psi, yy = 14400 psi, (c) p =
377.3 psi.
2. Figure 14 shows how a strain gage is attached to measure the strain in the longitudinal
direction on the cylindrical surface of a pressure vessel of 30 inch outside diameter and 0.5
inch wall thickness. The strain gage reads 630 x 10-6 in/in and the material properties are E
= 10 x 106 psi and = 0.3.
(a) Determine the three principal strains on the cylindrical surface of the vessel.
(b) Compute the principal stresses in the wall.
(c) What is the gage pressure inside the vessel?

Figure 14. A strain gage mounted on a cylindrical pressure vessel.


Note: The biaxial stress distribution in a pressure vessel is given in terms of the gage pressure, p,
the inner radius, ri, and the thickness, t, by
9.15

Answer: (a) xx = 148 , yy = 630 , zz = -334 , (b) xx = 3706 psi, yy = 7412 psi, (c) p =
255.6 psi.

3. Normal strains are measured along three directions in the


rosette shown in Figure 15. Express the principal
strains (1 and 2) and the angle of orientation of the
principal direction, , measured with respect to the x
direction in terms of the readings taken from the gages
(A, B, and C). Be sure to show all work.
Answer: 1,2 = (A + C)/2 {(A - C)2 + [(4B - A 3C)/(3) ]2}, = tan-1 {(4B - A - 3C)/[(3) (A - C)]}.

Figure 15. A strain gage rosette.

4. Normal strains are measured along three directions in the


rosette shown in Figure 16. Express the principal
strains (1 and 2) and the angle of orientation of the
principal direction, , measured with respect to the x
direction in terms of the readings taken from the gages
(A, B, and C). Be sure to show all work.
Answer: 1,2 = (A + C)/2 1/2 [(A - B)2 + (B - C)}2]1/2;
= tan-1 [2B - A - C)] / (A - C).

Figure 16. A strain gage rosette.

5. Four electrical-resistance strain gages have been mounted on a torque wrench and wired
into a Wheatstone bridge as shown in Figure 17. The designer of the system claims that
the output signal from the bridge is proportional to the torque applied by the wrench
irrespective of the point of application of the load P so long as it remains to the right
of the gages. Analyze the system and prove or disprove this claim.

9.16

Figure 17. A torque wrench.


Hint: The applied moment at given point on the wrench is given by multiplying the
magnitude of the applied load, P, by the distance between the point of application of the
load and the point in question.

Answer: The claim is valid.

9.17

APPENDIX I REVIEW OF STATICS


AND MECHANICS OF MATERIALS
A1.1 2-D Forces in Rectangular Coordinates
A force may be represented by a directed line segment. Since it is a vector quantity, the length
represents the magnitude. The direction must be specified relative to a given axes system.
As illustrated in Figure 1, one of the most effective methods for working with concurrent,
coplanar forces is to resolve the forces into components along orthogonal coordinate axes.

Figure 1. A force may be resolved into components along rectangular coordinates.


In this case,

(A1.1-1)

When using rectangular coordinates, a scalar component is assumed to be positive when its
corresponding vector is in a positive coordinate direction; otherwise, it is negative.

A1.2 3-D Rectangular Coordinates


When adding three or more forces, there is no practical trigonometric solution from the polygon
which describes the resultant. The best approach for finding the resultant is to resolve each force
into rectangular components. The components are then composed into their resultant.

A1.1

Figure 2 shows the case in which the line of action of a force is defined by two points. The
force may be resolved into rectangular components using a vector approach.

Figure 2. A force may be directed along a line segment.


Referring to Figure 2,
(A1.2-1)
where
(A1.2-2)

A1.3 Addition of Concurrent Forces in Space/3-D Equilibrium of Particles


When two or more forces act on the same particle, the resultant can be determined from their
components; since,

R x Fx

R y Fy
R

cos x

Rx
R

R z Fz

R x2 R y2 R z2

cos y

Ry
R

cos z

(A1.3-1)

Rz
R

Equilibrium exists when a particle remains at rest or moves with a constant velocity. In this
case,

A1.2

Fx 0

Fz 0 .

(A1.3-2)

The method of attack is to resolve all forces into their components, sum the components and then
apply the appropriate equations.

A1.4 Moments and Couples


Referring to Figure 3, the moment of a force F acting at point A about point O may be expressed
as either
) M F d

(A1.4 -1)

MO r xF

(A1.4-2)

or

where d is the perpendicular distance from the reference point O to the line of action of the force
F; and, r is a position vector drawn from O to any point along the line of action of F.
These equations can also be applied to compute the moment produced by the couple shown in
Figure 4.
Varigon's theorem states that the moment of a force about any axis is equal to the moments of its
components about that axis. Conversely, the moment of several concurrent forces about any axis
is equal to the sum of the moment produced by their resultant about that axis.
It should be noted that the maximum moment for a given force, or the least force necessary to
produce a given moment, occurs when the force is perpendicular to the moment arm.

Figure 4. The two diametrically opposed


forces form a couple that causes a rigid body
to rotate about an axis perpendicular to the
plane formed by the forces.

Figure 3. The moment of a force may be


computed as a scalar according to convention
or expressed in terms of a vector cross
product.

A1.3

A1.5 Equilibrium of Rigid Bodies


In general, distributed forces and applied couples cause a rigid body to translate and rotate. For
the body to remain at rest or to move with a constant velocity,

M 0

F 0

(A1.5-1)

The expressions in Equation (A1.5-1) are vector equations which correspond to six scalar
equations; namely,

Fx 0
Fy 0
Fz 0

Mx 0
My 0

(A1.5-2)

Mz 0

When dealing with a 2-D force and moment system, the equilibrium equations become:

Fx 0

Fy 0

MA 0 .

(A1.5-3)

where A is an arbitrary point in space, usually taken in the X-Y plane. The other three
equilibrium equations are automatically satisfied. Alternate equations may be applied to
determine these unknowns provided that they result in independent equations. One could use,
for example,

or

Fx 0

MA 0

MB 0

(A1.5-4)

MA 0

MB 0

MC 0 .

(A1.5-5)

To eliminate unknowns and thus simplify the analytical work, one could select the moment
center so that the line of action of one or more unknowns passes through the point and/or align
an axis with one of the unknowns and take the summation of forces in the perpendicular
direction.

A1.6 Center of Gravity and Centroid


For the purposes of computing the reactions at the points of constraint, the distributed weight of
a body can be considered to be a concentrated force acting through a single point called the
center of gravity. For the flat plate shown in Figure 5,

A1.4

i 1

x i wi

i 1

x i wi

w
i 1

i 1

y i wi

y i wi

i 1

i 1

(A1.6-1)

Figure 5. Finding the centroid of a flat plate located in the XY plane.


For a homogeneous plate of constant thickness, the center of gravity is called the centroid. In
this case,
n

i 1

x i Ai
A

x i Ai

i 1

i 1

i 1

y i Ai
A

Ai

y i Ai

i 1

i 1

(A1.6-2)

Ai

As the number of plates increases to infinity, the plates have differential areas and the finite sum
becomes an integral. In this case,

x el dA
A

x el dA

dA

y el dA
A

y el dA

(A1.6-3)

dA

The expressions xeldA and yeldA are called the first moment of the area about the Y and X axis,
respectively. If the centroid is located on a coordinate axis, the first moment of the area with
respect to that axis is zero. The converse also holds true.
The center of gravity or centroid is determined using either the method of integration or the
method of composites. When the method of composites is used, a rectangular or polar coordinate
axes system is established. The coordinates of the centroid of each individual portion of the area
are assigned as positive or negative depending upon their position with respect to the coordinate
axes system. Areas and/or lengths are considered as positive or negative depending upon
whether the shape must be added or subtracted, respectively, to form the overall shape desired.

A1.5

A1.7 Moments of Inertia


Inertia has significance when used in conjunction with other quantities (equations of motion in
dynamics, flexure formula for beam stress in mechanics of materials, etc.). Basically, the term
inertia is used to describe the tendency of matter to show resistance to change. Referring to the
X and Y axes shown in Figure 6, the area moments of inertia are,
Ix y2 dA

I y x2 dA .

(A1.7-1)

Figure 6. The moment of inertia is defined with respect to a rectangular coordinate system.
The units associated with the area moment of inertia are in4, cm4, etc. Since 2 > 0 (y2 and x2 in
the above equation), the moment of inertia is always positive.
The polar moment of inertia is used to characterize torsional behavior in cylindrical shafts, the
rotation of slabs, etc. It is computed for an axis passing through point O perpendicular to the x,y
plane (i.e., around a z axis). By definition,
JO r2 dA Ix I y

(A1.7-2)

where r is the distance from the element of area to the point O.


Imagine that the area shown in Figure 6 is squeezed either into a strip oriented parallel to one of
the coordinate axes as shown in Figures 7 and 8, or, into a thin ring centered at O as shown in
Figure 9.

Figure 7. The area is


assumed to be a strip
parallel to the X axis.

Figure 8. The area is


assumed to be a strip
parallel to the Y axis.
A1.6

Figure 9. The area is assumed to


be a thin ring centered at the
origin.

For the area to have the same Ix, Iy and Jo, the respective strips must lie at distances kx, ky and ko
respectively, such that,

I x k x2 A

I y k y2 A

J O k O2 A

(A1.7-3)

The distances kx, ky and ko are called the radii of gyration. Note that kx and ky are measured
perpendicular to the x and y axes, respectively, not along them.
As illustrated by Figure 10, it is often convenient to transfer the moment of inertia between a
centroidal axis and an arbitrary parallel axis. This transformation can be accomplished using a
simple formula known as the parallel axis theorem:
(A1.7-4)

I I Ad2

where I is the area moment of inertia about the arbitrary axis, is the moment of inertia about a
parallel axis passing through the centroid of the area, and d is the perpendicular distance between
the two parallel axes.

Figure 10. The moment of inertia may be transferred between two parallel axes.
Utilizing the expressions for the radii of gyration, the parallel axis theorem may also be
expressed as follows,
2

k2 k d2

(A1.7-5)

This approach can also be taken to transfer the polar moment of inertia and the polar radius of
gyration, respectively, as follows,

JO J O Ad 2

k O2 k O d 2 .

(A1.7-6)

If the area under consideration is made up of several parts, its moment of inertia about an axis is
found by adding the moments of inertia of the composite areas about that axis. In general, it is

A1.7

necessary to transfer each moment of inertia to the appropriate axis by the parallel axis theorem
before adding.
Note: The radius of gyration is not the sum of the component areas' radius of gyration. It is first
necessary to perform the superposition of the moments of inertia and then use the
definition the k2 = Itotal/Atotal.

A1.8 Shear and Moment Diagrams


A beam is a structural member that is designed to support loads at various points along its span.
In most cases, loads are perpendicular to the beam. A portion of mechanics of materials deals
with selecting a cross section to resist these shearing forces and bending moments. This often
requires drawing shear and moment diagrams for the beam.
The method of attack is to determine the reactions at points of constraint, then cut the beam at
various sections along the span, indicating on which portion of the beam internal forces are
acting. A convention must be established; the shear V and the bending moment M at a given
point of a beam are assumed to be positive when the internal forces and couples acting on each
element of the beam are directed as shown in Figure 11.
The following relationships are often helpful:
dV
w
d x

dM
V
d x

(A1.8-1)

where w is the distributed load per unit length.

Figure 11. The sign convention used for constructing shear and bending moment diagrams.

A1.9 Deflection of Beams


Recall that a prismatic beam in pure bending bends into a circular arc with curvature
1

M (x)
EI

A1.8

(A1.9-1)

For pure bending, M(x) is constant over the span. However, when the beam is subjected to a
transverse loading, Equation (A1.9-1) remains valid but both the bending moment and curvature
of the neutral surface vary from section to section.
The elastic curve of the beam is the relation between the deflection y measured at a given point
Q on the axis of the beam and the distance x of that point from some fixed origin. The elastic
curve is illustrated for a beam in Figure 12.

Figure 12. The elastic curve represents the deflection of the beam.
It can be shown from calculus that

d2y

dx 2
1

(A1.9-2)

and Eqn. (A1.9-1) becomes

d 2 y M ( x)

EI
dx 2

(A1.9-3)

For a prismatic beam in which E and I are constant, this second order linear differential equation
can be integrated to obtain the parametrical representation of the elastic curve as follows
EI

x
dy
E I ( x) M ( x) dx C 1
0
dx

(A1.9-4)

where dy/dx is the slope of the curve and


x
EI y
0

x
x
M ( x) dx C 1 dx C 2 dx M ( x) dx C 1 x C 2 . (A1.9-5)

0
0

A1.9

The constants are determined from the boundary conditions imposed on the deflection and slope
by considering the beam and the position of its supports. The three types of supports shown in
Figure 13 need be considered for complete analysis of statically determinate beams.

Figure 13. The three basic supports required to determine the equations of the elastic curve.
Concentrated loads, reactions at supports, or discontinuities in a distributed load make it
necessary to divide the beam into several portions, and to represent the moment by a different
function M(x) in each portion of the beam. Additional equations are generated at the interface
between different sections, since the slope and deflection must be continuous at these locations
(the shear and bending moment may be discontinuous).
Recall that in statically indeterminate problems, the reactions may be obtained by considering the
deformations of the structure involved. In these cases, the equilibrium equations may be
combined with the expressions for slope and deflection to evaluate both the constants in the
parametrical equations and the reactions at points of constraint.
A1.10 Synopsis of Mechanics of Materials
When loads are applied to a deformable body they produce stresses. The stresses represent the
force intensity and are computed by dividing the force by the area over which it acts. A normal
stress is produced when the force is perpendicular to the surface under consideration. A tensile
stress results when the force is directed along the outer normal to the surface; a compressive
stress results when the force is directed toward the element. Shear stress results when the force
is tangent to the surface.
The stresses produce changes in shape (deformations) characterized by a quantity called strain.
Normal stresses produce normal strains defined as the change in length of a line segment
divided by the original length of the segment. Shear stresses produce shear strains defined as
the change in angle between two line segments that were originally perpendicular to one another.

A1.10

There are three basic loading conditions that can be superimposed to produce a general state of
deformation. Axial loading produces either tensile or compressive stresses assumed to be
uniformly distributed across a plane cut normal to the longitudinal axis of the member.
Combinations of normal and shear stresses occur on oblique planes. The stresses are computed
by dividing the load (P) by the area (A) considered; thus,

P
A

P
A

(A1.10-1)

Bending produces a uniaxial stress condition in which normal stresses occur parallel to the
longitudinal axis of the member. For a prismatic member possessing a plane of symmetry,
subjected at its ends to equal and opposite couples acting in a plane of symmetry, the stress
distribution is linear through the thickness; compressive stresses occur on one side of the neutral
axis and tensile stress occur on the other side. The stress is computed using the flexure formula

Mc
I

(A1.10-2)

where M is the moment, c is the distance measured from the neutral axis to the point under
consideration, and I is the centroidal moment of inertia measured around the axis about which
the moment is applied.
Torsion produces shear stresses. In a prismatic member of circular cross section subjected to
couples (torques) of magnitude T, the shear stress acts in the direction of the applied torque. It is
zero at the center of the shaft and maximizes at the outer surface. The stress is computed using
the elastic torsion formula,

T
J

(A1.10-3)

where is the distance measured from the center of the shaft to the point under consideration and
J is the polar moment of inertia.
Shear stresses also result in prismatic members subjected to transverse loads. In this case, the
stress is given by

max

It

(A1.10-4)

where V is the shear force, Q is the first moment of the area measured about the neutral axis of
the portion of the cross section located either above or below the point under consideration, I is
the centroidal moment of inertia of the entire cross section, and t is the width of a cut made
through the point perpendicular to the applied load.

A1.11

The stress is related to the strain through constitutive equations (Hookes Laws) that depend
upon material properties. In general, these are very complex relations especially when dealing
with composite materials; however, only two independent material constants [typically Youngs
modulus (E) and Poissons ratio ()] are required when dealing with a linearly elastic (stressstrain curve is linear), homogeneous (mechanical properties are independent of the point
considered), and isotropic (material properties are independent of direction at a given point)
material. Other constants, such as the shear modulus [G = E/2(1 + )], are sometimes used in
these equations. Material constants are evaluated under simple loading conditions (tension test,
bending test, torsion tests) on specimens having simple geometry (typically rods, bars, and
beams).
It is often necessary to find the weakest link in a structure and this can be done by studying the
stress distribution at a critical location. This is not as easy as it may sound, since stress is a
tensor and must be studied using transformation equations. That is, the stress distribution
changes from point to point in a loaded member. However, there is usually a critical point or
region in the structure where the stress is the highest. The stress distribution at this critical
location depends upon the plane considered. There is a set of planes on which the normal stress
becomes maximum; shear stress is zero on these planes. The planes of maximum shear typically
have a non-zero normal stress component. Analytical equations (transformation equations) or
graphical techniques (Mohrs circle) can be employed to pinpoint the maximum stresses and the
planes on which they act. These quantities are used in conjunction with failure criteria to assess
structural integrity.
In some cases, a member may fail due to its unstable geometry as opposed to the imposed stress
state. Column buckling is one of the areas where such considerations come into play. Design
codes have been developed to prevent these cataclysmic failures.

A1.12

A1.11 Homework Problems


1. As illustrated in Figure 14, a prismatic beam is subjected to a uniformly distributed load over
portion AB of its span. The dimensions of the cross section are shown in Figure 15.
15 mm

300 kN/m
60 mm

A B

0.20 m

C
0.15 m

15 mm

40 mm

Figure 14.

Figure 15.

(a) Draw a complete free body diagram showing an equivalent system constructed by
replacing the distributed load, and determine the reactions at the roller (B) and pin (C).
(b) Establish the analytical relationships for the shear force and bending moment distribution
for all points along the span of the beam illustrated in the figure.
(c) Draw the shear and bending-moment diagrams for the beam illustrated in the figure.
(d) Compute the centroid and the centroidal moment of inertia of the cross section.
(e) Describe the exact location (along the span and on the cross section) where the maximum
normal stress due to bending occurs and compute this quantity.
(f) Describe the exact location (along the span and on the cross section) where the maximum
transverse shearing stress in the beam occurs and compute this quantity.
Answer: (a) RB = 100 kN , RCx = 0, RCy = 40 kN ; (b) AB: V = -300x kN and M = 150x2 kN-m; BC: V = 40 kN and M = (40x-14) kN-m; (c) see part b; (d) centroid
is 30 mm above bottom and IC = 787 x 10-9 m4; (e) max = 343 MPa at top with x
= 0.2 m; (f) max = 77.2 MPa at centroid with x = 0.2 m.
2. A prismatic beam is loaded as shown in Figure 16. The dimensions of the cross section are
shown in Figure 17.
8"

5 kips
A
5'

1 kip/ft

3 kips

2"

D
B

C
6'
15'

6"

4'

2"

Figure 16.

Figure 17.

A1.13

(a) Draw a complete free body diagram showing an equivalent system constructed by
replacing the distributed load, and determine the reactions at the pin (B) and roller (D).
(b) Establish the analytical relationships for the shear force and moment distribution for all
points along the span of the beam illustrated in the figure.
(c) Draw the shear and bending-moment diagrams for the beam illustrated in the figure.
(d) Compute the centroid and the centroidal moment of inertia of the cross section.
(e) Describe the exact location (along the span and on the cross section) where the maximum
normal stress due to bending occurs and compute this quantity.
(f) Describe the exact location (along the span and on the cross section) where the maximum
transverse shearing stress in the beam occurs and compute this quantity.
Answer: (a) RBx = 0, RBy = 13.7 k , RD = 4.3 k ; (b) AB: V = -5 k and M = -5x ft-k; BC: V =
13.7-x k and M = -81+13.7x-x2/2 ft-k; CD: V = 10.7-x k and M = -48+10.7x-x2/2 ft-k;
(c) see part b; (d) centroid is 5.29 in. above bottom and IC = 151 in4; (e) max = - 10.5
ksi at bottom with x = 5 ft; (f) max = .81 ksi at centroid with x = 5 ft.
3. A prismatic beam is loaded as shown in Figure 18. The dimensions of the cross section are
shown in Figure 19.
1.5 in.

1600 lb
A

80 lb/ft

9 ft.

11.5 in.

1.5 ft.
Figure 18.

Figure 19.

(a) Draw a complete free body diagram showing an equivalent system constructed by
replacing the distributed load, and determine the reactions at the pin (A) and roller (B).
(b) Establish the analytical relationships for the shear force and moment distribution for all
points along the span of the beam illustrated in the figure.
(c) Draw the shear and bending-moment diagrams for the beam illustrated in the figure.
(d) Determine the maximum transverse shearing stress in the beam assuming that it has the
cross section shown.
(e) Determine the maximum normal stress due to bending.
Answer: ( a) RAx = 0, RAy = 1670 lb , RB = 650 lb ; (b) in the first segment between A and the
1600 lb load, for example, V = 1670 lb () and M = 1670 x ftlb (cclw), etc.; (c) plot
the relationships found in part (c) along the span; (d) max = 145.2 psi; (e) max = 959
psi.

A1.14

4. As shown in Figure 20, a cantilever beam that is fixed into a wall at B is subjected to
uniformly distributed load. The weight of the beam can be neglected. The origin of the
coordinate system is at point A.
Y
w
B
A

Figure 20.
(a) Draw a complete free body diagram of an equivalent beam and determine the reactions
at the fixed support (in terms of w and L).
(b) Draw the shear and bending moment diagrams for the beam illustrated in the figure.
(c) Determine the equation of the elastic curve (in terms of x, L, E, I, w).
(d) Determine the slope at the free end (in terms of L, E, I, w).
Answer: (a) RBx = 0, RBy = wL , MB = wL2/2 clw; (b) V is linear; M parabolic; (c) y = -w/24EI
[x4-4L3x+3L4]; (d) A = wL3/6EI.
5. As illustrated in Figure 21, a cantilever beam that is fixed into a wall at B is subjected to
uniformly distributed load as shown. The weight of the beam can be neglected. The origin
of the coordinate system is at point A.
Y
w0
A

B
L

Figure 21.
(a) Draw a complete free body diagram of the beam and determine the reactions at the fixed
support (in terms of w0 and L).
(b) Draw the shear and bending moment diagrams for the beam illustrated in the figure.
(c) Determine the equation of the elastic curve (in terms of x, L, E, I, w0).
(d) Determine the deflection at the free end (in terms of L, E, I, w0).
Answer: (a) RBx = 0, RBy = w0L/2 , MB = w0L2/6 clw; (b) V is parabolic; M cubic; (c) y = w0/120EIL [x5-5L4x+4L5]; (d) yA = -w0L4/30EI.

A1.15

6. As illustrated in Figure 22, a cantilever beam that is fixed into a wall at B is subjected to a
concentrated load at its free end, A. The weight of the beam can be neglected. The origin of
the coordinate system is at point A.

Y
P
A

B
L

Figure 22.
(a) Draw a complete free body diagram of the beam and determine the reactions at the fixed
support (in terms of P and L).
(b) Draw the shear and bending moment diagrams for the beam illustrated in the figure.
(c) Determine the equation of the elastic curve (in terms of x, L, E, I, P).
(d) Determine the deflection at the free end (in terms of L, E, I, P).
Answer: (a) RBx = 0, RBy = P , MB = PL cclw ; (b) V = P and M = Px; (c) y = P/6EI [x33L2x+2L3] ; (d) yA = PL3/3EI.

A1.16

APPENDIX II BASIC FORMULAS FOR


MECHANICS OF MATERIALS
Axial Loading:

P
A

ave

P
A

Factor of safety F .S .

P
P

A td

ultimate load
allowableload

lateral strain
v

axial strain

PL
AE

Thermal Deformation:

T ( T ) L

T
L

Generalized Hooke's Laws:

xy

v y

xy

v z
E

yz

yz
G

vx y v z

E
E
E

zx

zx

where

v x

v y
E

z
E

E
2 (1 v )

Stress Concentration Factor (Axial Loading):


K

max
ave

Torsion of Circular Shafts:

solid shaft:

1
c4
2

hollow shaft:
A2.1

T
J
J

TL
JG

1
1
1
c24 c14 ( c24 c14 )
2
2
2

Transmission Shafts:

P 2 f T

1 rpm

1 1 1
s
Hz
60
60

1 hp 550 ft lb / s 6600 in lb / s

Stress Concentration Factor (Stepped Shaft in Torsion):


Tc
J

max K
Pure Bending (Homogeneous Beams):
I

M
EI

anticlastic curvature

M y
I

M
S

I
c

Pure Bending (Composite Beams):

E2
E1

nM y
I

M
E1 I

Iz

tan

Stress Concentration Factor (Notched Beam in Bending):

m K

Mc
I

Eccentric Axial Loading:

P Mz y My z

A
Iz
Iy

tan

Iy

Shear and Moment Considerations:

max

max

It

dV
w
d x

Transverse Loading:
q

VQ
I

ave
A2.2

VQ
It

dM
V
dx

Transformation Equations:

xx

xx yy

yy

xx

2
yy
2

xy
max, min
max

xx yy
2

xx yy
xx

2
yy
2

xx yy
2

xx yy

xx yy

cos 2 xy sin 2
cos 2 xy sin 2

sin 2 xy cos 2

2xy

and xy 0 with tan 2 p

xy2

and ave

xx yy
2

with tan 2 s

2 xy

xx yy

xx y y
2 xy

Mohr's Circle:

xx yy
2

c 0

xx yy

xy2

max

1
max min
2

Pressure Vessels:
pr
t

Cylindrical:

pr
2t

Spherical:

1 2

Design of Prismatic Beams:

S min

max

all

Deflection of Beams:

d 2 y M ( x)

EI
dx 2
Buckling of Columns:

Pcr

2 EI
L

2
e

cr

2 E
( Le / r )

A2.3

I A r2

pr
2t

Design of Steel Columns:


I A r2

C c 4.71
y

(
)
e

y
0.658

for Le /r C c :

cr

for L e /r C c :

cr 0.877 e

where

all

all

cr
1.67

cr
1.67

2 E
( Le / r ) 2

Pall all A

Conversions:
g = 32.2 ft/s2 = 9.81 m/s2; 1 slug = 14.59 kg; 1 ft = .3048 m; 1 lb = 4.45 N; 1 psi = 6.895 kPa

A2.4

Potrebbero piacerti anche