Student Version Fall2020
Student Version Fall2020
Student Version Fall2020
Structures
1.1 Introduction
1.2 Structural System
1.3 Load Classification
1.4 Supports And Reactions
1.5 Equations Of Static Equilibrium
1.6 Statically Determinate And Indeterminate Structures
1.7 Applications
Loads also may be categorized as dynamic, static, or thermal. Dynamic loads are
time-dependent, whereas static loads are independent of time. Thermal loads are
created on a restrained structure by a uniform and/or non uniform temperature
change.
F 0
(i x, y, z )
i
Eq. (1.1)
Mi 0
While Eq. (1.1) applies for general space structures. For the case of planar-type
structures it reduces to
F 0i
(i x, y) Eq. (1.2)
M 0 z
Note that only six independent equations exist for any free body in space and
three independent equations exist for a coplanar free body.
A structure is said to be determinate if all its external reactions and the internal
loads on its members can be obtained by utilizing only the static equations of
equilibrium. Otherwise the structure is said to be statically indeterminate.
In the latter, or what is commonly referred to as a redundant structure, there are
more unknown forces than the number of independent equations of statics which
can be utilized.
The additional equations required for the analysis of redundant structures can
be obtained by considering the deformations (displacements) in the structure.
Figure 1.6:
Indeterminate Structures
Chapter 1: Statics Analysis Of Structures 15
1.7 Applications
Example 1.1: Find the internal loads acting on each member of the structure shown
in Fig. 1.7.
Figure 1.7
Figure 1.8
M T 0(...)
1000 2 2T 0
T 1000lb
F y 0( )
FTx 1000 0
FTx 1000lb
F y 0( )
FTy 1000 0
FTy 1000lb
Chapter 1: Statics Analysis Of Structures 18
1.7 Applications
Solution of example 1.1 :
For the members 3-6-7
M 3 0(...)
1000 7 2.4 F4 0
F4 2915lb
F x 0( )
F3x 2915cos36.9o 1000 0
F3x 1335lb
F y 0( )
F3y 2915sin 36.9o 1000 0
F3y 750lb
Chapter 1: Statics Analysis Of Structures 19
1.7 Applications
Solution of example 1.1 :
For the members 2-3-4-5
M 5 0(...)
1335 5 2915 2cos36.9o 4 F1 0
F1 500lb
F x 0( )
F5x 2915cos36.9o 1335 500cos 60o 0
F5x 1250lb
F y 0( )
F5 y 2915sin 36.9o 750 500sin 60o 0
F5 y 1433lb
Chapter 1: Statics Analysis Of Structures 20
1.7 Applications
Example 1.2 Find the internal load in member 5 of the coplanar truss structure
shown in Fig. 1.9.
(a) Method of joints
Figure 1.9
M 4 0(...)
2000 10 4000 10 1000 30 20 R6 y 0
R6 y 4500lb
F x 0( )
2000 R4 x 0
R4 x 2000lb
F 0( )
y
F x 0( )
F2 2000 0
F2 2000lb
F y 0( )
500 F1 0
F1 500lb
F y 0( )
500 F4 sin 45o 0
F4 707lb
F x 0( )
2000 707 cos 45o F3 0
F3 2500lb
Figure 1.10
F y 0( )
707 sin 45o F5 0
F5 500lb
Chapter 1: Statics Analysis Of Structures 25
1.7 Applications
Figure 1.11
F y 0( )
500 F5 0
F5 500lb
Figure 1.12
Figure 1.13
Chapter 1: Statics Analysis Of Structures 30
Problems
Prob. 1.1:
A 5000lb airplane is in a steady glide with the flight path at an angle θ below the
horizontal (Fig. P.1.1) The drag force in the direction of the flight path is 750 lb.
Find the lift force L normal to the flight path and the angle θ.
Figure P1.1
A jet-propelled airplane in steady flight has forces acting as shown in figure P1.2.
Find the jet thrust T, lift L and the tail load P
Figure P1.2
Find the forces at points A and B of the landing gear shown in Fig P1.3.
10
Figure P1.3
2.1 Introduction
2.2 General Considerations
2.3 BASIC FLIGHT LOADING CONDITIONS
2.4 Flight-Vehicle Aerodynamic Loads
2.5 Flight Vehicle Inertia Loads
2.6 Load Factors for Translational Acceleration
2.7 Velocity Load Factor Diagram
2.8 Gust Load Factor
2.9 Examples
To ensure safety, structural integrity, and reliability of flight vehicles along with
the optimality of design, government agencies, both civil and military, have
established definite specifications and requirements in regard to the magnitude
of loads to be used in structural design of the various flight vehicles.
Figure 2.1a
Figure 2.1a
Experiments show that these high angles of attack and high lift coefficients
may be obtained momentarily in a sudden pull-up before the airflow reaches a
steady condition, but it is difficult to obtain accurate lift measurements during
the unsteady conditions.
Figure 2.1a
Figure 2.1a
For a given lift on the wing, the angle of attack decreases as the indicated airspeed
increases, and consequently the PLAA condition corresponds to the maximum
indicated airspeed at which the airplane will dive.
This limit on the permissible diving speed depends on the type of aircraft, but
usually is specified as 1.2 to 1.5 times the maximum indicated speed in level flight,
according to the function of the aircraft.
Even fighter aircraft are seldom designed for a diving speed equal to the terminal
velocity, since the terminal velocity of such airplanes is so great that difficult
aerodynamic and structural problems are encountered. Aircraft are placarded so
that the pilot will not exceed the diving speed limit.
The wing bending moment in this condition produce the maximum compressive
stresses on the upper rear spar flange and adjacent stringers and maximum tensile
stresses on the lower front spar flange and adjacent stringers.
Figure 2.1b
The moment of this force about the center of gravity of the airplane has the
maximum negative (pitching) value; consequently, the download on the horizontal
tail required to balance the moments of other aerodynamic forces will be larger
than for any positive flight condition.
Figure 2.1b
Figure 2.1c
Figure 2.1c
The wing bending moments in the negative high angle of attack condition
produce the highest compressive stresses in the lower forward region of the wing
cross section and the highest tensile stresses in the upper aft region of the wing
cross section. The line of action of the resultant force R is farther aft than for any
other negative flight attitude, and it will probably produce the greatest balancing
upload on the horizontal tail for any negative flight attitude.
Figure 2.1d
Figure 2.1d
Figure 2.2
The wing, of course, must be strong enough to resist loads at medium angles of
attack, but normally the wing will have adequate strength if it meets the
requirements for the four limiting conditions.
Each of the four flight conditions must be investigated for each extreme
position of the center of gravity.
To account for greater balancing tail loads which may occur for another location
of center of gravity, it may be possible to make some conservative assumption and
still compute balancing tail loads for only one location.
It is therefore necessary to calculate gust load factors at the minimum weight at
which the aircraft will be flown.
These conditions usually are not critical for wing bending stresses, since the,
specified load factors are not large, but may be critical for wing torsion, shear in
the rear spar, or down tail loads, since the negative pitching moments may be
quite high.
The aft portion of the wing, which forms the flap supporting structure, will be
critical for the condition with flaps extended.
The first aerodynamic data required for the structural analysis are the lift, drag,
and pitching-moment force distributions for the complete aircraft with the
horizontal tail removed, through the range of angles of attack from the negative
stalling angle to the positive stalling angle.
It is therefore desirable to obtain wind tunnel data on a model of the complete
airplane less horizontal tail.
Figure 2.3
The angle ϴ is measured from the flight path to the x axis, as shown in Fig. 2.3,
and is equal to the difference between the angle of attack α and the angle of wing
incidence i.
The Mean Aerodynamic Chord (MAC) is a wing reference chord which usually is
calculated from the wing planform. If every, airfoil section along the wing span has
the same pitching-moment coefficient cm, the MAC is determined so that the total
wing pitching moment is cm𝒄qS.
For a rectangular wing planform the value of 𝒄 (the MAC) is equal to the wing
chord; for a trapezoidal planform of the semiwing, the value of 𝒄 is equal to the
chord at the centroid of the trapezoid.
Figure 2.3
Eq. (2.3)
The total aerodynamic force on the airplane in the z direction, CzaqS, is equal to
the sum of the force CzqS on the airplane less tail and the balancing tail load CtqS:
Eq. (2.4)
Figure 2.4
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 37
2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
Since stalling reduces the air loads on the wing, these portions of the curves are
not used. Instead, the curves are extrapolated, as shown by the dotted lines, in
order to approximate the conditions of a sudden pull-up, in which high lift
coefficients may exist for a short time.
For the PHAA condition, the angle of attack corresponding to the force coefficient
of 1.25 times the maximum value of Cza is used, and the curves are extrapolated to
this value, as shown in Fig. 2.4
The loads produced by landing impact, maneuvering, gusts, boost and staging
operations, and launches are always greater than the loads occurring when all the
forces on the vehicle are in equilibrium.
The inertia force on any element of mass is equal to the product of the mass and
the acceleration and acts in the direction opposite to the acceleration.
If the applied loads and inertia forces act on an element as a free body, these
forces are in equilibrium. For example, a body of mass m under the action of a force
vector F moves so as to satisfy the equation
F= ma Eq. (2.5)
Eq. (2.6)
x’’,y’’and z’’ are the components of acceleration along the x, y, and z axes.
In many engineering problems, it is necessary to consider the inertia forces acting
on a rigid body which has other types of motion. In many cases where the elements
of 2 rigid body are moving in curved paths, they are moving in such a way that each
element moves in only one plane and all elements move in parallel planes.
Figure 2.5
If the x axis is chosen through the center of gravity C, the forces are simplified. The
resultant inertia force in the y direction for the entire body is found as follows:
The angular velocity ω is constant for all elements of the body, and the integral is
zero because the x axis was chosen through the center of gravity.
Eq. (2.7)
The term 𝑥 is the distance from the axis or rotation O to the center of gravity C as
shown in fig.2.5
Figure 2.5
Figure 2.6
Eq. (2.8)
The resultant inertia torque about the axis of rotation is found by integrating the
terms representing the product of the tangential force on each element αrdM and
its moment arm r:
Eq. (2.9)
104 47
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 47
2.5 FLIGHT VEHICLE INERTIA LOADS
The term I0 represents the moment of inertia of the mass about the axis of
rotation. It can be shown that this moment of inertia can be transferred to a parallel
axis through the center of gravity by use of the following relationship:
Eq. (2.10)
Where Ic is the moment of inertia of the mass about an axis through the center of
gravity obtained as the sum of the products of mass elements dM and the square of
their distances rc , from the center of gravity:
Eq. (2.11)
Figure 2.7
Forces at the centroid represent the product of the mass of the body and the
components of acceleration of the center of gravity. In many cases, the axis of
rotation is not known, but the components of acceleration of the center of gravity
can be obtained. In other cases, the acceleration of one point of the body and the
angular velocity and angular acceleration are known.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 50
2.5 FLIGHT VEHICLE INERTIA LOADS
If the point O in Fig. 2.8 has an acceleration a0, an inertia force at the center of
gravity of Mao, opposite to the direction of a0, must be considered in addition to
those previously taken into account.
Figure 2.8
For purposes of analysis, it is convenient to combine these inertia forces with the
forces of gravity, by multiplying the weight of each part by a load factor n, and thus
to consider the combined weight and inertia forces.
When the vehicle is being accelerated upward, the weight and inertia forms add
directly. The weight of w of any part and the inertia force wa/g have a sum nw:
Eq. (2.12)
Figure 2.9
Eq. (2.13)
This value for the load factor can be shown to be the same as that given by Eq.
(2.12) by equating the lift nW to the sum of the weight and inertia forces:
or
which corresponds to Eq. (2,12).
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 54
2.6 Load Factors for Translational Acceleration
Flight vehicles frequently have horizontal acceleration as well as vertical
acceleration.
The airplane shown in Fig. 2.10 is being accelerated forward, since the engine
thrust T is greater than the airplane drag D. Every element of mass in the airplane is
thus under the action of a horizontal inertia force equal to the product of its mass
and the horizontal acceleration.
Figure 2.10
Eq. (2.14)
or
or Eq. (2.15)
104
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 57
2.6 Load Factors for Translational Acceleration
Eq. (2.16)
or
which corresponds with the value used in Eq. (2.13) for a level attitude of the
airplane.
104
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 58
2.6 Load Factors for Translational Acceleration
The thrust load factor for the condition shown in Fig. 2.11 is also similar to that
obtained for the airplane in level attitude.
Since the thrust and drag forces must be in equilibrium with the components
of weight and inertia forces along the x axes, the thrust load factor is obtained as
Follows:
or
This value is the same as that obtained in Eq. (2.14) for a level attitude of the
airplane.
In the case of the airplane landing as shown in Fig. 2.12, the landing load
factor is defined as the vertical ground reaction divided by the airplane weight.
Figure 2.12
Eq. (2.17)
If load factors are obtained first along lift and drag axes, they may be resolved
into components along other axes, in the same manner as forces are resolved into
components.
The force acting on any weight w is wn, and the component of this force along
any axis at an angle ϴ to the force is wncosϴ. The component of the load factor is
then ncos ϴ.
As a general definition, the load factor n along any axis i is such that the
product of the load factor and the weight of an element is equal to the sum of the
components of the weight and inertia forces along that axis.
The weight and inertia forces are always in equilibrium with the external
forces acting on the airplane, and the sum of the components of the weight and
inertia forces along any axis must be equal and opposite to the sum of the
components of the external forces along the axis .
Eq. (2.19)
The V-n diagram is therefore the same for all altitudes if indicated airspeeds
are used. Where compressibility effects are considered, they depend on actual
airspeed rather than indicated airspeed and consequently are more pronounced
at altitude. Compressibility effects are not considered here.
The z component of the resultant gravity and inertia force is the force nW
acting at the center of gravity of the airplane, as shown in Fig. 2.13.
or Eq. (2.20)
Figure 2.13
The maximum value of the normal force coefficient Cza may he obtained at
various airplane speeds. For level flight at a unit-load factor, the value of V
corresponding to Cza,max would be the stalling speed of the airplane.
Figure 2.13
Some types of airplanes may be designed so that the pilot would have to exert
large forces on the controls in order to exceed the limit-load factor.
Line CD in Fig. 2.13 represents the limit on the permissible diving speed for the
airplane. This value is usually specified as 1.2 to 1.5 times the maximum indicated
airspeed in level flight.
Line OB corresponds to line OA, except that the wing is at the negative stalling
angle of attack, and the air load is down on the wing.
Figure 2.13
The method of obtaining the gust load factors, represented by points E and F, is
explained in the following section.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 73
2.8 Gust Load Factor
When an airplane is in level flight in calm air, the angle of attack Δα is measured
from the wing chord line to the horizontal.
If the airplane suddenly strikes an ascending air current which has a vertical
velocity KU, the angle of attack is increased by the angle Δα, as shown in Fig. 2.14.
Figure 2.14
The angle Δα is small, and the angle in radians may be considered as equal to its
tangent:
Eq. (2.21)
The change in the airplane normal force coefficient Cza resulting from a change
in angle of attack Δα may be obtained from the curve of Cza versus α of Fig. 2.4.
This curve is approximately a straight line, and it has a slope β which may be
considered constant:
Eq. (2.22)
After striking the gust, the airplane normal force coefficient increases by an
amount determined from Eqs. (2.21) and (2.22):
Eq. (2.23)
Eq. (2.25)
Where β is the slope of Cza versus α per degree, V is the indicated airspeed in
miles per hour, and other terms correspond to those in Eq. (2.24).
When the airplane is in level flight, the load factor is unity before the plane
strikes the gust. The change in load factor Δn from Eq. (2.25) must be combined
with the unit-load Factor in order to obtain the total gust load factor:
Eq. (2.26)
Equation (2.26) may be plotted on the V-n diagram, as shown by the inclined
straight lines through points F and H of Fig. 2.13. These lines represent load
factors obtained when the airplane is in a horizontal attitude and strikes positive
or negative gusts.
Figure 2.13
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 79
2.8 Gust Load Factor
The negative gust load factor represented by point E is greater than the
negative maneuvering load factor represented by point and will determine the
NLAA condition.
It might seem that the gust load factors should be added to the maneuvering
load factors, in order to provide for the possibility of the airplane's striking a
severe gust during a violent maneuver. While this condition is possible, it is
improbable because the maneuvering load factors are under the pilot's control,
and the pilot will restrict maneuvers in gusty weather.
Both the maneuvering and gust load factors correspond to the most severe
conditions expected during the life of the airplane, and there is little probability
of a combined gust and maneuver producing a condition which would exceed
the limit-load factor for the design condition.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 80
2.9 Examples
Example 2.1 When landing on a carrier, a 10,000-lb airplane is given
a deceleration of 3g (96.6 ft/s2) by means of a cable engaged by an
arresting hook as shown in Fig. 2.15.
Figure 2.15
a) Find the tension in the cable, the wheel reaction R, and the distance e from
the center of gravity to the line of action of the cable.
b)Find the tension in the fuselage at vertical sections AA and BB if the portion of
the airplane forward section AA weighs 3000 lb and the portion aft of section
BB weighs 1000 lb.
c) Find the landing run if the landing speed is 80 ft/s.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 81
2.9 Examples
Figure 2.17
The forces T1, T2 and V2 may be checked by considering the equilibrium of the
center portion of the airplane, as shown in Fig. 2.18.
Figure 2.18
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 85
2.9 Examples
Figure 2.19c
a) If one wheel and tire weighs 500 lb, find the compression C and bending moment m in
the oleo strut if the strut is vertical and is 6 in from the centerline of the wheel, as shown
in Fig. 2.19b.
b) Find the shear and bending moment at section AA of the wing if the wing outboard of
this section weighs 1500 lb and has its center of gravity 120 in outboard of section AA.
c) Find the required shock strut deflection if the airplane strikes the ground with a vertical
velocity of 12 ft/s and has a constant vertical deceleration until the vertical velocity is
zero. This neglects the energy absorbed by the tire deflection, which may be large in
some cases.
d) Find the time required for the vertical velocity to become zero.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 87
2.9 Examples
d) The time required to absorb the landing shock is found based on elementary
dynamics.
Since the landing shock occurs for such a short time, it may be less injurious to
the structure and less disagreeable to the passengers than would a sustained
load.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 91
2.9 Examples
Example 2.3 A 60,000-lb airplane with a tricycle landing gear makes a hard
two-wheel landing in soft ground so that the vertical ground reaction is
270,000 lb and the horizontal ground reaction is 90,000 lb. The moment of
inertia about the center of gravity is 5,000,000 lb .s2 .in, and the dimensions
are shown in Fig. 2.20.
Figure 2.20
c) If the nose wheel is 40 in from the ground when the main wheels touch the
ground, find the angular velocity of the airplane and the vertical velocity of
the nose wheel when the nose wheel reaches the ground, assuming no
appreciable change in the moment arms. The airplane's center of gravity has
a vertical velocity of 12 ft/s at the moment of impact, and the ground
reactions are assumed constant until the vertical velocity reaches zero, at
which time the vertical ground reaction becomes 60,000 lb and the horizontal
ground reaction becomes 20,000 lb.
Figure 2.21
In calculating the term α𝑥 M, 𝑥 is in inches and g is used as 386 in/s2 . If 𝑥 is in feet, g will be
32,2 ft/s2 . The total force on the turret is 600 lb forward and 3850 lb down. This total force
is seen to be almost 10 times the weight of the turret.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 95
2.9 Examples
Solution of example 2.3 :
c) The center of gravity of the airplane is decelerated vertically at 3.5g, or 112.7
ft/s2. The time of deceleration from an initial velocity of 12 ft/s to a zero vertical
velocity is found from the following.
During this time, the center of gravity moves through a distance found from
The angular velocity of the airplane at the end 0.106 s after the landing is found
from
The vertical motion of the nose wheel resulting from this rotation, shown in Fig.
2.22, is
Figure 2.22
The distance of the nose wheel from the ground, after the vertical velocity of
the center of gravity of the airplane has become zero, is
60,000
Since the ground reaction decreases by the ratio of vertical acceleration
270,000
of the airplane becomes zero, the angular acceleration decreases in the same
proportion, as found by equating moments about the center of gravity.
Since at this time the motion is rotation, with no vertical motion of the center
of gravity, the vertical velocity of the nose wheel is found as follows:
Note: 1 ft=12in
This velocity is smaller than the initial sinking velocity of the airplane.
Consequently, the nose wheel would strike the ground with a higher velocity in a
three-wheel level landing.
It is of interest to find the centrifugal force on the turret ω2xM at the time the
nose wheel strikes the ground. This force was zero when the main wheels hit
because the angular velocity ω was zero. For the final value of ω, the following
value is obtained:
An airplane weighing 5000 lb strikes an upward gust of air which produces a Wing
lift of 25,000 lb see Fig. P2.1. What tail load P is required to prevent a pitching
acceleration if the dimensions are as shown? What will be the vertical
acceleration of the airplane? if this lift force acts until the airplane obtains a
vertical velocity of 20 ft/s, how much time is required?
Figure P2.1
Figure P2.2
Figure P2.3
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 103
Problems
Prob. 2.4:
An airplane is flying at 550 mi/h in level flight when it is suddenly pulled upward
into a curved path of 2000-ft radius. (See Fig. P2.4.) Find the load factor of the
airplane.
57
Figure P2.4
Prob. 2.5:
If the airplane in Prob. 2.4 is given a pitching acceleration of 2 rad/s2, find its load
factor, assuming that the change in lift due to pitching may be neglected.(Assume
moment of inertia Iy=300,000 lb. s2.in.
47
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 104
Problems
Prob. 2.6:
A large transport aircraft is making a level landing., as shown in Fig. P2.6. The
gross weight of the aircraft is 150,000 lb, and its pitching mass moment of inertia
is 50x106 lb.in.s2 about the center of gravity. The landing rear-wheel reaction is
350,000 lb at an angle of 15° with the vertical. Determine whether passenger A or
B will receive the most load. Assume that each passenger weights 170 lb and
neglect the airplane lift.
Figure P2.6
Chapter 3: Elasticity Of
Structures
Elasticity Of Structures 2
3.1 Introduction
This chapter defines stresses and strains and their fundamental
relationships. The stress behavior of structures undergoing elastic
deformation that is due to the action of external applied loads is
also discussed. The term elasticity or elastic behavior is used here
to imply a recovery property of an original size and shape.
Elasticity Of Structures 3
3.2 Stresses
Consider the solid body shown in Fig. 3.1a which is acted on by a
set of external forces Qi , as indicated. If we assume that rigid-body
motion is prevented, the solid will deform in accordance with the
external applied forces; as a result, internal loads between all
parts of the body will be produced.
Figure 3.1a
Elasticity Of Structures 4
3.2 Stresses
If the solid is separated into two parts by passing a hypothetical
plane, as shown in Fig. 3.1b, then there exist internal forces whose
resultants are indicated by QI and QII acting on parts I and II
respectively.
Figure 3.1b
Elasticity Of Structures 5
3.2 Stresses
The forces which hold together the two parts of the body are
normally distributed over the entire surface of the cut plane. If we
consider only an infinitesimal area 𝜹𝑨 acted on by a resultant force
𝜹𝑸, then an average force per unit may be expressed as
Q
AV Eq. (3.1)
A
In the limit as δA approaches zero, Eq. (3.1) becomes
dQ
Eq. (3.2)
dA
Where σ now is the limiting value of the average force per unit
area and, by definition, the stress at that point. A stress is
completely defined if its magnitude and direction and the plane on
which it acts are all known.
Elasticity Of Structures 6
3.2 Stresses
For instance, it is not appropriate to ask about the stress at point
0 of the solid shown in figure 3.2 unless the plane on which the
stress is acting is specified. An infinite number of planes may be
passed through point 0, thus resulting in an infinite number of
different stresses.
Figure 3.2
Elasticity Of Structures 7
3.2 Stresses
In the most general three-dimensional state of stress, nine stress
components may exist:
xx xy xz
yx
yy yz Eq. (3.3)
zx zy zz
Elasticity Of Structures 8
3.2 Stresses
Figure 3.3 illustrates the stress notation.
Figure 3.3
Elasticity Of Structures 9
3.2 Stresses
In the case of the two-dimensional state of stress, or what is
commonly referred to as the plane stress problem (σ zx = σ zy = σ zz =
0), Eq. (3.3) becomes
xx xy
(σ xy = σ yx ) Eq. (3.4)
yx yy
Elasticity Of Structures 10
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field
Figure 3.4
Elasticity Of Structures 11
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field
F y 0 xy , x yy , y zy , z Y 0 Eq. (3.6)
Or
Figure 3.5
Elasticity Of Structures 13
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field
1 rr
rr ,r r , rz , z R 0 Eq. (3.8)
r r
1 2 r
, r ,r z , z 0 Eq. (3.9)
r r
1 rz
zz , z z , rz ,r Z 0 Eq. (3.10)
r r
Elasticity Of Structures 14
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field
xx, x xy , x X 0
Or Eq. (3.11)
xy , x yy , y Y 0
Or in cylindrical coordinates:
1 rr
rr ,r r , R0
r r Eq. (3.12)
1 2
, r ,r r 0
r r
Elasticity Of Structures 15
3.4 Strains And Strain-Displacement Relationships
Figure 3.6
Elasticity Of Structures 16
3.4 Strains And Strain-Displacement Relationships
If, after deformations take place, the displacements of point 0 are
denoted by qx , qy and qz, in the x, y, and z directions, respectively,
then the displacements of points A, B and C which are dx, dy, and dz
away from point 0 will be qx + qx,x dx, qy + qy,y dy and qz + qz,z dz,
respectively.
Figure 3.7
Elasticity Of Structures 17
3.4 Strains And Strain-Displacement Relationships
Figure 3.7 b
Elasticity Of Structures 18
3.4 Strains And Strain-Displacement Relationships
Elasticity Of Structures 19
3.4 Strains And Strain-Displacement Relationships
On the basis of Eq. (3.13), the normal strain in the x direction, for
instance, is obtained as follows:
1/2
Elasticity Of Structures 20
3.4 Strains And Strain-Displacement Relationships
Which reduces to the Following by using the binomial expansion
technique:
1
xx qx , x
( q x, x ) 2
( q y,x ) 2
( q z,x ) 2
...
2
For small displacements, the terms involving the squares of
derivatives may be neglected in comparison with the derivative in
the first term.
Elasticity Of Structures 21
3.4 Strains And Strain-Displacement Relationships
Elasticity Of Structures 22
3.4 Strains And Strain-Displacement Relationships
The shearing strains may be derived by finding the relative
change in the angle between a given pair of the three line
segments shown in Fig. 3.7.
xy qx , y q y , x
xz qx , z qz , x Eq. (3.15)
yz q y , z qz , y
Figure 3.7
Or
Elasticity Of Structures 23
3.4 Strains And Strain-Displacement Relationships
Thus, in a three-dimensional state of strain, the strain field
components may be compacted in a matrix form:
xx xy xz
yx yy yz Eq. (3.16)
zy zz
zx
• where εij = εji for i ≠ j is a shearing strain and for i = j is the normal strain.
• It must be emphasized that Eqs (3.14) and (3.15) are derived on the assumption that
the displacements involved are small. Normally these linearized equations are
adequate for most types of structural problem but in cases where deflections are
large, for example types of suspension cable, etc., the full, non-linear, large deflection
equations, given in many books on elasticity, must be employed.
Elasticity Of Structures 24
3.4 Strains And Strain-Displacement Relationships
xx xy
Eq. (3.17)
yx yy
Elasticity Of Structures 25
3.4 Strains And Strain-Displacement Relationships
rr qr , r z
q , z
q , z
r Eq. (3.18)
q , qr
zr qz ,r qr , z
r r qr , q
z q z , z r q ,r
r r
Elasticity Of Structures 26
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems
Elasticity Of Structures 28
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems
Elasticity Of Structures 29
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems
Elasticity Of Structures 30
3.6 Boundary Conditions
qx q x ( ) qy q y ( ) qz q z ( ) Eq. (3.22)
Elasticity Of Structures 31
3.6 Boundary Conditions
On the other hand, if σxx, σyy and σxy are prescribed at the
boundary, then the conditions are called force boundary conditions
and are normally expressed as
N xx xx xy zx x
yy xy
N yy yz y Eq. (3.23)
N zz
zx
yz zz z
where the Nii bar (i =x, y,z) are the surface boundary forces, the σij bar (i,j= x,y,z)
are the prescribed stresses at the boundary, and ɳx , ɳy , ɳz are the direction
cosines.
It is appropriate to note at this point that in order to obtain the
exact stress field in any given solid under the action of external
loads, the equations of equilibrium, compatibility equations, and the
boundary conditions must all be satisfied.
Elasticity Of Structures 32
3.7 Stress-Strain Relationships
Elasticity Of Structures 33
3.7 Stress-Strain Relationships
Figure 3.8
Figure 3.9b
Figure 3.9c
Elasticity Of Structures 35
3.7 Stress-Strain Relationships
In the most general case for a linear elastic anisotropic solid.
Hooke's law, which relates stresses to strains, may be written in a
matrix form as follows, in which ε = strain, σ = stress, and
aij(i,j =x,y,z) are the material elastic constants (aij = aji )
Elasticity Of Structures 37
3.7 Stress-Strain Relationships
Where Eij , Gi and vij (I,j= x,y,z) are the modulus of elasticity, shear modulus, and
Poisson's ratio, respectively.
Elasticity Of Structures 38
3.7 Stress-Strain Relationships
A body which obeys Eq. (3.25) i.e, at each point there exist three
mutually perpendicular planes of elastic symmetry is commonly
referred to as an orthotropic body.
Elasticity Of Structures 39
3.7 Stress-Strain Relationships
If all directions in a solid are elastically equivalent and any plane
which passes through any point of the body is a plane of elastic
symmetry, then it is called an isotropic body.
In this case. the elastic constants in Eq, (3.2) simplify to
1
a11 a22 a33
E
1
a44 a55 a66 Eq. (3.27)
G
v
a12 a21 a23 a32 a13 a31
E
E
G
2(1 y )
Elasticity Of Structures 40
3.7 Stress-Strain Relationships
Elasticity Of Structures 41
3.7 Stress-Strain Relationships
For plane-stress problems where σzz , σyz and σxz are zero, if we
assume isotropic material, Eq. (3.25) becomes:
xx 1 v 0
1 xx
yy E v 1 0 yy Eq. (3.28)
xy 1
0 0 xy
2(1 v)
By substituting Eq. (3.28) into Eq. (3.20) and utilizing Eq. (3.11), the
compatibility equation in terms of stresses alone becomes
Eq. (3.30a)
Eq. (3.31)
Elasticity Of Structures 44
3.7 Stress-Strain Relationships
Above Equations may be transposed to obtain expressions for each
stress in terms of the strains. The procedure adopted may be any of
the standard mathematical approaches and gives
Eq. (3.32)
For the case of plane stress in which σz =0, above Eqs reduce to
Eq. (3.33)
Elasticity Of Structures 45
3.7 Stress-Strain Relationships
Suppose now that, at some arbitrary point in a material, there are
principal strains εI and εII corresponding to principal stresses σI and
σII. If these stresses (and strains) are in the direction of the
coordinate axes x and y, respectively, then τxy =γxy =0 and the shear
strain on an arbitrary plane at the point inclined at an angle θ to the
principal planes is
Elasticity Of Structures 46
3.7 Stress-Strain Relationships
Noting that for this particular case τxy =0, σx =σI and σy =σII
and the shear strains γxy, γxz and γyz are expressed in terms of their
associated shear stresses as follows
Eq. (3.34b)
Elasticity Of Structures 47
3.7 Stress-Strain Relationships
Eq. (3.35)
It may be seen that the conditions of plane stress and plane strain
do not necessarily describe identical situations.
Elasticity Of Structures 48
3.7 Stress-Strain Relationships
Changes in the linear dimensions of a strained body may lead to a
change in volume. Suppose that a small element of a body has
dimensions δx, δy and δz. When subjected to a three-dimensional
stress system the element will sustain a volumetric strain e (change
in volume/unit volume) equal to
Eq. (3.36)
Elasticity Of Structures 49
3.7 Stress-Strain Relationships
Elasticity Of Structures 50
3.7 Stress-Strain Relationships
Elasticity Of Structures 51
3.8 Transformations Of Stresses And Strains
Figure 3.10
Elasticity Of Structures 52
3.8 Transformations Of Stresses And Strains
Figure 3.11a
Elasticity Of Structures 53
3.8 Transformations Of Stresses And Strains
Figure 3.11b
Elasticity Of Structures 54
3.8 Transformations Of Stresses And Strains
F 0Z
A xx A cos (sin ) yy A sin (cos ) xy A cos (cos )
yx A sin (sin ) 0
xx sin 2 yy sin 2
(1 2 cos 2 ) xy Eq. (3.38b)
2 2
Elasticity Of Structures 55
3.8 Transformations Of Stresses And Strains
Elasticity Of Structures 56
3.8 Transformations Of Stresses And Strains
,
T x, y
Eq. (3.39b)
Solution:
The expressions for the longitudinal and circumferential stresses produced by the internal pressure may
be found in any text on stress analysis and are
Note that there are no shear stresses acting on the x and y planes; in this case, σx and σy then form a
biaxial stress system. Elasticity Of Structures 58
3.8 Transformations Of Stresses And Strains
The negative sign for τ indicates that the shear stress is in the direction BA and not AB.
The maximum value of τ will therefore occur when sin 2θ is a maximum, i.e. When sin 2θ =1 and θ =-45◦.
Elasticity Of Structures 59
3.8 Transformations Of Stresses And Strains
Example 2
A cantilever beam of solid, circular cross-section supports a compressive load of 50 kN applied to its
free end at a point 1.5mm below a horizontal diameter in the vertical plane of symmetry together with a
torque of 1200Nm (Fig. 3.12). Calculate the direct and shear stresses on a plane inclined at 60◦ (ACW)
to the axis of the cantilever at a point on the lower edge of the vertical plane of symmetry.
Figure 3.12
Elasticity Of Structures 60
3.8 Transformations Of Stresses And Strains
The direct loading system is equivalent to an axial load of 50 kN together with a bending
moment of 50×103 ×1.5=75 000Nmm in a vertical plane. Therefore, at any point on the
lower edge of the vertical plane of symmetry there are compressive stresses due to the
axial load and bending moment which act on planes perpendicular to the axis of the beam
and are given
The shear stress, τxy, at the same point due to the torque is obtained from Eq. (iv) in
Example 3.1, i.e.
Elasticity Of Structures 61
3.8 Transformations Of Stresses And Strains
The stress system acting on a two-dimensional rectangular element at the point is shown in
Fig. . Note that since the element is positioned at the bottom of the beam the shear stress
due to the torque is in the direction shown and is negative . Again σn and τ may be found
from first principles or by direct substitution. Note that θ =30◦, σy =0 and τxy=−28.3 N/mm2
the negative sign arising from the fact that it is in the opposite direction to τxy in Fig.
Then
Elasticity Of Structures 62
3.9 Principal Stresses
σn varies with the angle θ and will attain a maximum or minimum value
when dσn/dθ =0.
Eq. (3.41)
Two solutions, θ and θ +π/2, are obtained from above Eq. so that there
are two mutually perpendicular planes on which the direct stress is either
a maximum or a minimum. Further, by comparison of Eqs it will be
observed that these planes correspond to those on which there is no shear
stress. The direct stresses on these planes are called principal stresses and
the planes themselves, principal planes.
Eq. (3.42)
where σI is the maximum or major principal stress and σII is the minimum or minor
principal stress. Therefore, when σII is negative, i.e. compressive, it is possible for
σII to be numerically greater than σI.
Elasticity Of Structures 63
3.9 Principal Stresses
The maximum shear stress at this point in the body may be determined in an identical
manner.
Giving:
Here, as in the case of principal stresses, we take the maximum value as being the greater
algebric value. We can deduce that
Eq. (3.44)
The above Equations give the maximum shear stress at the point in the body in the plane of
the given stresses.
Elasticity Of Structures 64
3.10 Mohr’s circle of stress
And
The positive directions of these stresses and the angle θ are defined in the figure below
Finally:
Elasticity Of Structures 65
3.10 Mohr’s circle of stress
clearly C is the point ((σx +σy)/2, 0) and the radius of the circle as required.
CQ’ is now set off at an angle 2θ (positive clockwise) to CQ1, Q’ is then the point (σn,−τ) as
demonstrated below.
The principal planes are then given by 2θ =β(σI) and 2θ =β+π(σII). Also the maximum and minimum
values of shear stress occur when Q coincides with D and E at the upper and lower extremities of the
circle. At these points QN is equal to the radius of the circle which is given by
Hence as before. The planes of maximum and minimum shear stress are
given by 2θ =β+π/2 and 2θ =β+3π/2, these being inclined at 45◦ to the principal planes.
Elasticity Of Structures 66
3.10 Principal Strains
If we compare the equations of the strains, there are two mutually perpendicular planes on
which the shear strain γ is zero and normal to which the direct strain is a maximum or
minimum. These strains are the principal strains at that point and are given by
Eq. (3.45)
If the shear strain is zero on these planes it follows that the shear stress must also be zero
and we deduce, that the directions of the principal strains and principal stresses coincide.
The related planes are then determined
Eq. (3.46)
Eq. (3.47)
Elasticity Of Structures 67
3.11 Mohr’s circle of strain
We now apply the arguments of the last section to the Mohr’s circle
of stress described in before. A circle of strain, analogous to that
shown for the stress, may be drawn when σx, σy, etc. are replaced by
εx, εy, etc.. The horizontal extremities of the circle represent the
principal strains, the radius of the circle, half the maximum shear
strain and so on.
Elasticity Of Structures 68
3.11 Mohr’s circle of strain
Example 3
At a particular point in a structural member a two-dimensional stress system exists where σx
=60 N/mm2, σy=−40 N/mm2 and τxy =50 N/mm2. If Young’s modulus E =200 000 N/mm2
and Poisson’s ratio ν=0.3 calculate the direct strain in the x and y directions and the shear
strain at the point. Also calculate the principal strains at the point and their inclination to the
plane on which σx acts; verify these answers using a graphical method.
Elasticity Of Structures 69
3.11 Mohr’s circle of strain
Elasticity Of Structures 70
3.11 Mohr’s circle of strain
Elasticity Of Structures 71
3.12 Experimental measurement of surface strains
Stresses at a point on the surface of a piece of material may be determined by measuring the strains
at the point, usually by electrical resistance strain gauges arranged in the form of a rosette, as shown
in Fig. 3.13. Suppose that εI and εII are the principal strains at the point, then if εa, εb and εc are the
measured strains in the directions θ, (θ +α), (θ +α+β) to εI we have, from the general direct strain
relationship
Eq. (3.48)
since εx becomes εI, εy becomes εII and γxy is zero since the x
and y directions have become principal directions. Rewriting Eq.
(3.48) we have
Eq. (3.49)
Eq. (3.50)
Figure 3.13
Eq. (3.51)
Elasticity Of Structures 72
3.12 Experimental measurement of surface strains
Therefore if εa, εb and εc are measured in given directions, i.e. given angles α and β, then εI, εII and
θ are the only unknowns in Eqs (3.49)–(3.52).
Eq. (3.53)
Eq. (3.54)
Eq. (3.55)
Figure 3.14
Elasticity Of Structures 73
3.12 Experimental measurement of surface strains
A typical rosette would have α=β=45◦ in which case the principal strains are most conveniently
found using the geometry of Mohr’s circle of strain. Suppose that the arm a of the rosette is inclined
at some unknown angle θ to the maximum principal strain as in Fig. 3.13. Then Mohr’s circle of
strain is as shown in Fig. 3.14; the shear strains γa, γb and γc do not feature in the analysis and are
therefore ignored. From Fig. 3.14
Elasticity Of Structures 74
3.12 Experimental measurement of surface strains
Eq. (3.56)
Eq. (3.57)
Eq. (3.58)
Elasticity Of Structures 75
3.12 Experimental measurement of surface strains
Elasticity Of Structures 76
3.12 Experimental measurement of surface strains
Elasticity Of Structures 77
3.12 Experimental measurement of surface strains
Elasticity Of Structures 78
Problems
Prob. 3.1:
A structural member supports loads which produce, at a particular point, a
direct tensile stress of 80 N/mm2 and a shear stress of 45 N/mm2 on the
same plane.
-Calculate the values and directions of the principal stresses at the point
and also the maximum shear stress, stating on which planes this will act.
Prob. 3.2:
Prob. 3.4:
The direct strains in the directions a, b, c are −0.002, −0.002 and +0.002,
respectively. If I and II denote principal directions, find εI, εII and θ.
Elasticity Of Structures 80
Further Readings
Elasticity Of Structures 81
AERO 234: Fundamentals of Aircraft
Structures
4.1 Introduction
4.2 Mechanical Properties Of Materials
4.3 Equations for Stress-Strain Curve Idealization
4.4 Fatigue
4.5 Strength-Weight Comparisons of Materials
4.6 Sandwich Construction
4.7 Typical Design Data For Materials
4.8 Problems
Figure 4.1
Figure 4.2a
Figure 4.2b
Figure 4.2a
Figure 4.3
S
s Eq. (4.4) Where єs=shear strain
L ΔS= change in arc length
L =effective gage length
s
G Eq. (4.7)
s
Figure 4.4
Figure 4.5
Figure 4.6
Figure 4.8
Figure 4.7
Table 4.1
1
2 /2
probability (log N ) P(log N ) e d Eq. (4.13)
2
Figure 4.9
Mt
Bending: b Eq. (4.19)
2I
Solving for the free variable t from Eqs. (4.17) through (4.19)
and substituting into Eq. (4.20) yields the material weight
required to meet each specified design criterion. Thus
PL
W (tension) Eq. (4.21)
t
L2b 12 c
1/2
Table 4.3
Figure 4.10
M t f (t f tc )
W 4 t Eq. (4.31)
The weight Ws, of a solid beam element from Eq. (4,23) is:
6M
Ws Eq. (4.33)
Hence, the ratio of the weight of a sandwich beam element
to that of a solid element of the corresponding sandwich face
material is W 4 M /
1.63 Eq. (4.34)
Ws 6M /
Behavior & Evaluation of vehicle materials 55
4.6 Sandwich Construction
It is important to note that Eq. (4.34) is valid only for a
sandwich in which the total weight of the facings is equal to
the weight of the core.
In order to compare the weights of a sandwich with solid
elements studied in Table 4.3, consider a sandwich whose
facings are made of 2024-T3 aluminum alloy and a core
material whose density is 0.01 lb/in3. From Eq. (4.30)
c 0.01
0.05
2 f 2(0.1)
From Eq. (4.34)
W
1.63 0.05 0.37
Ws
Also note that the value of 0.37 is less than any of the other
values in column 6 of Table 4.3.
Table 4.4
4.3 Work Prob. 4.2 for a core density of 0.015 lb/in3 and the fallowing
specific cases of facing materials:
(a) 2024-T3 aluminum alloy (density=0.1 lb/in3)
(b) 6A I-4V titanium (density = 0.16 lb/in3)
(c) 321 stainless steel (density=0.286 lb/in3)
(d) Inconel (density = 0.3 lb/in3)
(e) Beryllium (density = 0.069 Ibirin3)
(f) Reinforced composite (unidirection)
(1)Glass fiber (density= 0.09 lb/in3)
(2)Boron fiber (density = 0.095 lb/in3)
(3)Graphite (density = 1.053 lb/in3)
Figure 4.8
After how many cycles should the part be replaced so that only the
(allowing percentage of the parts in service fails before replacement?
a)1
b)5
c)10
Behavior & Evaluation of vehicle materials 66
Further Readings
5.1 Introduction
5.2 Force-Stress Relationships
5.3 Normal Stress in Beams
5.4 Shear Stresses in Beams
5.5 Shear Flow in Thin Webs
5.6 Shear Center
5.7 Torsion Of Closed-Section Box Beams
5.8 Shear Flow in Closed-Section Box Beams
5.9 Spanwise Taper Effect
5.10 Beams With Variable Stringer Areas
This chapter discusses the theory and the application of these two
fundamental stress components.
Vy xy dA M y z xx dA Eq. (5.1)
A
A
Vz xz dA T ( y xz z xy )dA
A A
The latter, which is used here, assumes that plane sections remain
plane after extensional-bending deformation takes place. This
assumption implies that the deformations due to transverse shear
forces ( Vz and Vy ) are very small and therefore may be neglected. In
addition, this assumption allows the displacements (deflections) of
any point in the beam to be expressed in terms of the
displacements of points located on the beam axis.
є xx B1 B2 y B3 z Eq. (5.4)
Constants B1, B2, and B3 may now be determined through the use
of Eq. (5.1), or
P E ( B1 B2 y B3 z )dA
A
M y Ez ( B1 B2 y B3 z )dA
A
11
Chapter 5:Stress Analysis
5.3 Normal Stress in Beams
Carrying out the integrations yields
P
B1 A B2 y B3 z
E
Mz
B1 y I z B2 I yz B3 Eq. (5.7)
E
My
B1 z I yz B2 I y B3
E
12
Chapter 5:Stress Analysis
5.3 Normal Stress in Beams
y ydA
A
P
B1 A
E
M z
I z B2 I yz B3 Eq. (5.10)
E
My
I yz B2 I y B3
E
P
B1
AE
I y M z I yz M z
B2 Eq. (5.11)
E ( I y I z I yz )
2
I z M y I yz M z
B3
E ( I y I z I yz 2 )
M z y
xx Eq. (5.14)
Iz
Thus, the expression of the vertical shear stress σxy at any point in
the cross section is obtained by determining the shear stress σyx on a
horizontal plane through the point.
19
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The bending stresses on the left and right sections of the beam
element (Fig. 5.2) are shown in Fig. 5.3. At any point a distance y
from the neutral axis, the bending stress will be Mzy/Iz on the left
face and Mzy/Iz +Vyηy/Iz on the right face.
Figure 5.3
20
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
In order to obtain the shear stress at a distance y = y1 above the
neutral axis, the portion of the beam above that point is considered
as a free body, as shown in Fig. 5.3c.
y1
Iz
21
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Equation (5.15) may be written in standard form as
c
Vy
yx
It
z
ydA
y1
Eq. (5.16)
Where the integral represents the moment of the area of the cross
section above the point where the shear stress is being determined,
with the moment arms measured from the neutral axis. The cross-
sectional area considered is shown by the shaded portion in Fig.
5.3a.
22
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Example 5.1: Find the maximum normal stress in the beam in
Fig.5.4 and the shear stress distribution over the cross section.
Figure 5.4
23
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Solution: The maximum normal stress due to bending will occur at
the point of maximum bending moment, or at the fixed end of the
beam. Since the shear force is constant throughout the beam span,
the shear stress distribution will be the same at any cross section.
The moment of inertia for the cross is obtained as follows:
13 43
I z 2 3 3 2.5 1 43.3in4
2
12 12
I yz P M y 0
24
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The maximum normal stress is
Mz y 40 20(3)
xx 55.4kips / in 2
Iz 43.3
For a point 1 in below the top of the beam, the integral of Eq. (5.16)
is equal to the moment of the area of the upper rectangle about the
neutral axis:
y1
ydA 2.5(3) 7.5in3
25
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The average shear stress just above this point, where t = 3 in is
c
Vy 40, 000
I z t y1
yx ydA 7.5 2310lb / in 2
43.3 3
43.3 1
26
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
For a point 2 in below the top of the beam, the integral of Eq. (5.16)
is c
c
Vy 40, 000
yx ydA (2.5 3 1 2) 8780lb / in 2
I z t y1 43.3 1
27
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The distribution of shear stress over the cross section is shown in Fig.
5.5. The stress distribution over the lower half of the beam is
similar to the distribution over the upper half because of the
symmetry of the cross section about the neutral axis.
Figure 5.5
28
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Alternative Solutions for Shear Stresses
In some problems it is more convenient to find shear stresses by
obtaining the forces resulting from the change in bending stresses
between two cross sections than it is to use Eq. (5.16). Portions of
the beam between two cross sections a unit distance apart are
shown in Fig. 5.6.
Figure 5.6
29
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The bending moment increases by Vy, in this unit distance, and the
bending stresses on the left face of the beam are larger than those
on the right face by an amount Vyηy/Iz, where η=1. At the top of the
beam, this difference is
Vy y 40(3)
2.77kips / in 2
Iz 43.3
The differences in bending stresses at other points of the cross
section are obtained by substituting various values of y and are
shown in Fig. 5.6b. Cutting sections and utilizing the equations of
static equilibrium in each case (Fig. 5.6c, d, and e) yield the shearing
stresses at these various points:
30
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
σ yx = 6930/3 = 2310 lb/in2 at 1 in below top of beam
Note that these shear stress values are the same as shown in Fig.5.5.
31
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Example 5.2 : In the beam cross section shown in Fig. 5.7, the
webs are considered to be ineffective in resisting normal stresses but
capable of transmitting shear. Each stringer area of 0.5 in2 is
assumed to be lumped at a point. Find the shear stress distribution
in the webs.
Figure 5.7
32
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
SOLUTION: If we neglect the moments of inertia of the webs and of
the stringers about their own centroids, the cross-sectional moment
of inertia about the neutral axis is
34
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The shear stress in the web between the two middle stringers is
found by considering spanwise forces on the two upper stringers:
35
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
In problems involving shear stresses in thin webs, the shear force
per inch length of web often is obtained rather than the shear
stress.
The shear per inch, or shear flow, is equal to the product of the
shear stress and the web thickness.
The shear flow for each web, shown in Fig. 5.7c, is equal to the
sum of the longitudinal loads above the web.
36
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The shear stresses may also be obtained by using Eq. (5.16). For a
point between the two upper stringers
c
Vy 8000
I z t y1
yx ydA (0.5 6) 15, 000lb / in 2
40 0.040
c
Vy 8000
I z t y1
yx ydA (0.5 6 0.5 2) 20, 000lb / in 2
40 0.040
37
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Example 5.3: Find expressions for the normal stress for all beams
whose unsymmetrical cross sections are given in Fig. 5.8a and b.
135 y 494 z
Similarly, for the beam in Fig, 5.b, the normal stress expression is
xx 6457y 9.06z
39
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
Shear flow is defined as the product of the shear stress and the
thickness of the web. For all practical purposes, it is sufficiently
accurate to assume that shear stresses in thin webs are always
parallel to the surfaces for the entire thickness of the web.
In Fig.5.9. a curved web representing the leading edge of a wing is
shown, and the shear stresses are parallel to the surfaces of the web
at all points.
Figure 5.9
40
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
Air loads normal to the surface must, of course, be resisted by
shear stresses perpendicular to the web, but these stresses usually
are negligible and are not considered here. it might appear that a
thin, curved web is not an efficient structure for resisting shearing
stresses, but this is not the case. The diagonal tensile and
compressive stresses σt and σc , are shown in Fig. 5.9 on principal
planes at 45° from the planes of maximum shear σs.
41
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
If the diagonal compression alone were acting on the curved web,
it would bend the web to an increased curvature. The diagonal
tensile stress, however, tends to decrease the curvature, the two
effects counteract each other. Consequently, the curved web will
resist high shear stress without deforming from its original
curvature.
42
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
The shear flow q, which is the product of the shear stress σs and
the web thickness t, usually is more convenient to use than the shear
stress.
The shear flow may be obtained before the web thickness is
determined, but the shear stress depends on the web thickness.
Often it is necessary to obtain the resultant force on a curved
web in which the shear flow q is constant for the length of the web.
43
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
The element of the web shown in Fig. 5.10 has length ds, and the
horizontal and vertical components of this length are dz and dy,
respectively.
The force on this element of length is qds, and the components of
the force are qdz horizontally and qdy vertically.
Figure 5.10
44
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
The total horizontal force is:
z
Fz qdz qz Eq. (5.17)
0
where z is the horizontal distance between the ends of the web. The
total vertical force on the web is:
45
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
The resultant force is qL where L is the length of the straight line
joining the ends of the web, and the resultant force is parallel to this
line.
Equations (5.17) and (5.18) are independent of the shape of the
web, but depend on the components of the distance between the
ends of the web. The induced torsional moment of the resultant
force depends on the shape of the web. The torque induced at any
point such as O, shown in Fig. 5.11a, is equal to rq ds.
Figure 5.11-b
47
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
The distance e, shown in Fig. 5.11 b, of the resultant force from
point O may be obtained by dividing the torque by the force:
2 Aq 2 A
e Eq. (5.20)
qL L
It is important to note that the shear flow q is assumed to be
constant in the derivation of Eqs.(5.19) and (5.20).
48
Chapter 5:Stress Analysis
5.6 Shear Center
Open-section thin web beams, such as in Fig. 5.12, are unstable in
carrying torsional loads. Thus if a beam cross section is symmetrical
about a vertical axis, then the vertical loads must be applied in the
plane of symmetry in order to produce no torsion on the cross
section.
Figure 5.12
49
Chapter 5:Stress Analysis
5.6 Shear Center
However, if the beam cross section is not symmetrical, then the
loads must be applied at a point such that they produce no torsion.
This point is called the shear center and may be obtained by finding
the position of the resultant of the shear stresses on any cross
section.
The simplest type of beam for which the shear center may be
calculated is made of two concentrated flange areas joined by a
curved shear web, as shown in Fig. 5.12, The two flanges must lie in
the same vertical plane if the beam carries a vertical load. lf the
web resists no bending, the shear flow in the web will have a
constant value q.
50
Chapter 5:Stress Analysis
5.6 Shear Center
The resultant of the shear flow will be qL=V, and the position of
this resultant from Eq.(5.20) will be a distance e=2A/L to the left of
the flanges, as shown in Fig. 5.12a. Therefore, all loads must be
applied in a vertical plane which is a distance e from the plane of
the flanges.
A beam with only two flanges that are in a vertical plane is not
stable for horizontal loads. The vertical location of the shear center
would have no significance for this beam. For beams which resist
horizontal loads as well as vertical loads, it is necessary to determine
the vertical location of the shear center.
51
Chapter 5:Stress Analysis
5.6 Shear Center
If the cross section is symmetrical about a horizontal axis, the
shear center must lie on the axis of symmetry.
If the cross section is not symmetrical about a horizontal axis, the
vertical position of the shear center may be calculated by taking
moments of the shear forces produced by horizontal loads.
The method of calculating the shear center of a beam can be
illustrated best by numerical examples.
52
Chapter 5:Stress Analysis
5.6 Shear Center
Example 5.4: Find the shear flows in the webs of the beam shown
in Fig. 5,13a. Each of the four flange members has an area of 0.5 in2.
The webs are assumed to carry no bending stress. Find the shear
center for the area.
Figure 5.13-a
53
Chapter 5:Stress Analysis
5.6 Shear Center
Solution: Two cross sections 1 in apart are shown in Fig. 5.13b. The
increase in bending moment in the 1-in length is equal to the shear
V.
Figure 5.13-b
54
Chapter 5:Stress Analysis
5.6 Shear Center
The increase of bending stress on the flanges in the 1-in length is:
Vy y 10,000 5
1000lb / in2
Iz 50
The load on each 0.5-in2 area resulting from this stress is 500 lb
and is shown in Fig. 5.13b. The actual magnitude of the bending
stress is not needed in the shear-flow analysis, since the shear flow
depends on only the change in bending moment or the shear.
If each web is cut in the span wise direction, as shown, the shear
forces on the cut webs must balance the loads on the flanges.
55
Chapter 5:Stress Analysis
5.6 Shear Center
The force in web ab must balance the 500-lb force on flange a, and
since this spanwise force acts on a 1-in length, the shear flow in the
web will be 500 lb/in in the direction shown.
The shear flow in web bc must balance the 500-lb force on flange
b as well as the 500-lb spanwise force in web ab, and consequently
the shear flow has a value of 1000 lb/in.
The shear flow in web cd must balance the 1000-lb spanwise force
in web bc as well as the 500-lb force on flange c. which is in the
opposite direction. The shear flow in web cd is therefore 500 lb/in
and is checked by the equilibrium of flange d.
56
Chapter 5:Stress Analysis
5.6 Shear Center
The directions of the shear flow on the vertical beam cross section
are obtained from the directions of the spanwise forces.
Since each web has a constant thickness, the shear flow, like shear
stresses, must be equal on perpendicular planes.
57
Chapter 5:Stress Analysis
5.6 Shear Center
The shear center is found by taking moments about point c:
T c 0
10000e 500(4)(10) 0
e 2in
58
Chapter 5:Stress Analysis
5.6 Shear Center
Example 5.5: Find the shear flows in the webs of the beam shown
in Fig. 5.14a. Each of the four flanges has an area of 1.0 in2.
Find the shear center for the area.
Figure 5.14-a
59
Chapter 5:Stress Analysis
5.6 Shear Center
SOLUTION: The moment of inertia of the area about the horizontal
centroidal axis is
Iz 4(1 x 42 ) 64in 4
The change in axial load in each flange between the two cross
sections 1 in apart is
Vy 16,000
yA 4 1 1000lb
Iz 64
60
Chapter 5:Stress Analysis
5.6 Shear Center
The axial loads and shear flows are shown in Fig. 5.14b. The shear
flows in the webs are obtained by a summation of the spanwise
forces on the elements, as in Example 5.4.
Figure 5.14-b
61
Chapter 5:Stress Analysis
5.6 Shear Center
The distance e to shear center is found by taking moments about a
point below c, on the juncture of the webs. The shear flow in the
nose skin produces a moment equal to the product of the shear flow
and twice the area enclosed by the semicircle.
The shear flow in the upper horizontal web has a resultant force of
6000 lb and a moment arm of 10 in. The short vertical webs at a and
d each resist forces of 1000 lb with a moment arm of 6 in.
62
Chapter 5:Stress Analysis
5.6 Shear Center
The resultant forces on the other webs pass through the centers of
moment:
c
T 0
63
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
The thin-web, open-section box beams previously considered are
capable of resisting loads which are applied at the shear center but
become unstable under torsional loads. In many structures,
especially in aerospace vehicles, the resultant load takes on different
positions for different loading conditions and consequently may
produce torsion.
Figure 5.15-a
65
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
The wing section shown in Fig. 5.15b has two spars which form a
closed-section box beam. In some wings, two or more closed boxes
may act together in resisting torsion, but such sections are statically
indeterminate and are considered in a later chapter.
Figure 5.15-b
66
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
The box section shown in Fig. 5.16 is loaded only by a torsional
moment T. Since the axial loads in the stringers are produced by
wing bending, they are zero for the condition of pure torsion.
Figure 5.16-a
67
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
If the upper stringers are considered as a free body, as shown in
Fig. 5.16b, the spanwise forces must be in equilibrium; that is,
qa=q1a or q=q1.
Figure 5.16-b
68
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
If similar sections containing other flanges are considered, it
becomes obvious that the shear flow at any point must be equal to
q.
69
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
If we take point O in Eq.5.16c as a reference, the following may be
immediately written from Eq. (5.19):
Figure 5.16-c
70
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
The area A is the same regardless of the position of point O, since
the moment of a couple is the same about any point.
71
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Consider a box beam containing only two stringers, as shown in Fig. 5.17. Since
this section is stable under the action of torsional loads, the vertical shear force V
may be applied at any point in the cross section.
Note that this beam is unstable under the action of a horizontal load, since the
two stringers in the same vertical plane cannot resist a bending moment about a
vertical axis.
Figure 5.17-a
72
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
If two cross sections 1 in apart are considered, as shown in Fig.
5.17b, the difference in axial load on the stringers, ΔP, between the
two cross sections may be found from the difference in the bending
stress σxx=-Mzy/Iz=Vy(1)y/Iz, or |ΔP|=σxxAf=VyA1h1/Iz= VyA2h2/Iz where
A 1 and A2 are stringer areas.
Figure 5.17-b
73
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
These loads must be balanced by the shear flow shown in Fig.
5.17b. if we consider equilibrium, the summation of forces on the
upper stringer in the spanwise direction must be zero:
Vy A1h1
q1 (1 in) q0 (1 in) 0
Iz
Or
Vy A1h1
q1 q0 Eq. (5.22)
Iz
74
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
The shear flow q0 may be found by summing torsional moments
for the back section about a perpendicular axis through the lower
stringer:
Vy C 2Aq 0 0
Or
Vy C
q0
2A
Figure 5.17-a
where A is the total area enclosed by the box.
75
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Substituting this value in Eq. (5.22) yields
Vy h1 A1 Vy C
q1 Eq. (5.23)
Iz 2A
76
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
For the box beam shown in Fig. 5.18, all the shear flows q1,q2,….,qn
may be expressed in terms of the shear flow q0 by considering the
spanwise equilibrium of the stringers between web 0 and the web
under consideration:
Figure 5.18
q1 q 0 P1
q 2 q 0 P1 P2
Or n Eq. (5.24)
q n q 0 pn
n
0
For the case of general bending, the difference in axial load on the
stringers ΔP between two sections 1 in apart may be found from Eq.
(5.12). Making the substitutions Mz=Vy(1 in),My=Vz(1 in),
and P=0 yields :
I yVy I yzVz I zVz I yzVy
p1 y z Af Eq. (5.25)
I I I 2
I I I 2
y z yz y z yz
Where y and z are the coordinates of the stringer area Af.
78
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Example 5.6: Find the shear flow in all webs of the box beam
shown in Fig. 5.19a.
Figure 5.19a
79
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Solution: The moment of inertia of the beam cross section
about the neutral axis is l =(4 x 0.5 + 2 x 1)(52) = 100 in4. The
difference in bending stress between the two cross sections 1 in
apart is V(1)y/l = 10,000 (1)(5)/100 = 500 lb/in2. This produces
compressive loads ΔP of 500 lb on the 1-in2 upper stringer areas
and 250 lb on the 0.5 in2 stringer areas, as shown in Fig, 5.14b.
Figure 5.19b
80
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
The shear flow in the loading-edge skin is considered as the
unknown q0, although the shear flow in any other web could
have been considered as the unknown. Now the shear flow in
all other webs may be obtained in terms of q0, by considering
the equilibrium of the spanwise forces on the stringers, as
shown in Fig. 519.b.
81
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Taking the axis through point O, for example, as shown in Fig.
5.19c, and summing torques to zero yield:
Figure 5.19c
0
T 0
82
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Example 5.7: Find the shear flow in the webs of the box beam
shown in Fig.5.20.
83
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Solution: The change in bending stress between two cross
sections is obtained from Eq. (5.25). The terms to be used in
this equation are obtained as follows:
I z 2 x 3 2 x 1 52 200 in 4
I y 2 x 3 2 x 1 102 800 in 4
_ _
_ Figure 5.21a
85
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Now the shear flow in each web can be obtained from the
increments of the flange loads, as was done in Example 5.6 for
the symmetrical box beam .
86
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
Now the unknown shear flow q0 is obtained from the
equilibrium of torsional moments. Taking point 0 as a reference
point and summing moments about the axis through O yield
(q0 - 400) 100 q0 100 q0 600 100 (q0 1000)(100) 0
400q 0 200,00 0
q 0 500lb / in
The final shear-flow results are indicated on Fig. 5.21b. The
minus sign implies that the wrong direction of shear flow has
been assumed.
Figure 5.21b
87
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
In the preceding analysis of shear stresses in beams, we
assumed that the cross section of the beam remained
constant. Since in aerospace vehicle structures a minimum
weight is always sought, usually the beams are tapered in
order to achieve maximum structural efficiency.
88
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
As an illustration, consider the beam shown in Fig, 5.22
which, for simplicity, is assumed to consist of two stringers
joined by a vertical web that resists no bending. The resultant
axial loads in the stringers must be in the direction of the
stringers and must have horizontal components Px=Mz/h.
Figure 5.22a
89
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The vertical components of this load which act on the
stringers, P tan α1 and P tan α2, as shown in Fig. 5.22b, resist
part of the external applied shear Vy.
Figure 5.22b
90
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
By designating the shear resisted by the stringers as Vf and
that resisted by the webs as Vw
Vv V f Vw Eq. (5.26a)
Figure 5.23
94
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
Solution: The shear flows are obtained by the use of Eqs.
(5.30a) and (5.31). The solution of these equations is shown in
Table 5,1. While slide-rule accuracy is sufficient for shear-flow
calculations, the values in Table 5.1 are computed to four
significant figures for comparison with a method to be
developed later. Vy h0
Vw Eq. (5.30a)
h
Vw Vw Ah Vw
q
I ydA 2
Ah 2 2
h
Eq. (5.31)
Table 5.1
95
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
Example 5.9: Find the shear flows at section AA of the box
beam shown in Fig. 5.24.
Figure 5.24
96
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
SOLUTION: The moment of inertia of the cross section at AA
about the neutral axis is I 2 2+1 1 (52 ) 200in 4
97
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The horizontal components of the forces acting on the 2-in2
stringers are 20,000 lb, and the forces on the 1-in2 stringers are
10,000 lb, as shown in Fig. 525a. The vertical components are
obtained by multiplying the forces and the tangents of the
angles between stringers and the horizontal.
Figure 5.25a
98
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The sum of the vertical components of forces on all stringers
Vf is 4000 lb, and the remaining shear Vw 4000 lb is resisted by
the shear flows in the webs. If one of the upper webs is cut, as
shown in Fig. 5.25b, the shear flows in the webs may be
obtained from
Vw
q
Iz ydA
4000
q= *5*2=200
200
Figure 5.25b
where the integral represents the moment of the areas
between the cut web and the web under consideration.
99
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The change in bending stress on a stringer between the two
cross sections 1 in apart is Vwy/I when the effect of taper is
considered, and the change in axial load on a stringer of area Af
is
Vw
p yAf
I
These axial loads are shown in Fig. 5.25b in the same way as
they were shown previously for beams with no taper. The
equilibrium of forces in the spanwise direction yields the shear
flow in terms of q0 in all the webs, as shown in Fig. 5.25b.
100
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
Now the shear flow q0 can be found by summing torsional
moments about the z axis through point 0 for the back section, as
shown in Fig. 5.26a
Figure 5.26a
101
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The final shear flow in each web is shown in Fig. 5.26b.
Figure 5.26a
102
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
In Sec.5.9, beams were considered which varied in depth but
had stringers whose cross sections were constant. In many
aerospace structural beams, the cross-sectional area of the
stringer members varies as well as the depth of the beam .
103
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The airplane wing section in Fig. 5.27 represents a structure in
which the variation in stringer areas must be considered. The
stringer areas in this wing are designed in such a manner that
the bending stresses are constant along the span. In order to
resist the larger bending moments near the root of the wing, the
bending strength is augmented by increasing the depth of the
wing and the area of spar caps A and B.
Figure 5.27
104
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The stringers which resist the part of the bending moment
not resisted by the spar caps have the same area for the entire
span. Since the axial stresses on these stringers are the same
at every point along the span, the increments of load increase
ΔP will be zero except on spar caps A and B.
It may be seen from Eq. (5.24) that the shear flow must be
constant around the entire leading edge of the wing and
changes only at the spar caps. Consequently, the methods of
analysis previously used are not applicable to this problem.
q1 q 0 P1
q 2 q 0 P1 P2 Eq. (5.24)
n
q n q 0 pn
0
105
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The bending stresses and total stringer loads may be
calculated for two cross sections of the beam. The actual
dimensions and stringer areas for each cross section are used,
so that any changes between the cross sections are taken into
consideration.
The stringer loads pa and pb are shown in Fig. 5.28 for two
sections a distance apart.
Figure 5.28
106
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The increase in load in any stringer is assumed to be uniform
in length a. The increase in stringer load per unit length along
the span is P P
P b a
a Eq. (5.32)
This typical force is shown in Fig. 5.28b. Now the shear flow
can be obtained from these values of ΔP, as in the previous
analysis.
Figure 5.28b
107
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
It is seen that the shear force is not used in finding the
values of ΔP; consequently, it is not necessary to calculate the
vertical components of the stringer loads. The effect of beam
taper and changes in stringer area are implemented
automatically when the moments of inertia and bending
stresses are calculated.
109
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
Example 5.10: Find the shear flows in the beam of Fig. 5.23 by
the method of using differences bending stresses.
Figure 5.23
110
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
SOLUTION: For this two-flange beam, the axial load in the
flanges has a horizontal component P=M/h. The values of P for
various sections are calculated in column 4 of Table 5.2.
Table 5.2
111
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
In computing the shear at any cross section, values of the axial
loads at cross sections 10 in on either side are found. The free-
body diagrams are shown in Fig. 5.29.
Figure 5.29
112
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The circled numbers represent stations, or the distance from
the cross section to the left end of the beam. The difference in
horizontal loads on the upper part of the beam between the
cross sections 20 in apart must be balanced by the resultant of
the horizontal shear flow, 20q.
113
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
Even though the shear does not vary linearly along the span,
the error in this assumption is only 0.7 percent, as found by
comparison with the exact value obtained in Table 5.1. This
error is even smaller at the other stations.
114
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
Example 5.11: Find the shear flows at cross section AA of the
box beam shown in Fig. 5.24 by considering the difference in
bending stresses at cross sections 10 in on either side of AA.
Figure 5.24
115
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
SOLUTION: The moment of inertia at station 40 (40 in from
the left end) is found from the dimensions shown in Fig. 5.30a.
The bending stresses at station 40, resulting from the bending
moment of 320,000 in.lb, are
My 320,000(4.5)
xx 8888 lb in2
Iz 162
The loads on the 1-in2 areas are 8888 lb, and the loads on the
2-in2 areas are 17,777 lb, as shown in Fig. 5.31a.
Figure 5.31a
116
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The moment of inertia at station 60 is found from the
dimensions shown in Fig. 5.31b:
I z 8(5.52 ) 242in4
Figure 5.31b
21,818 17,777
P 202lb
20
10,909 8888
P 101lb
20
118
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The values of ΔP are shown in 5.31b. The remaining solution is
identical to that of Example 5,9. The values of ΔP are 1 percent
higher than the exact values shown in Fig. 5.25b.
The reason for this small discrepancy is that the average
shear flow between stations 40 and 60 is 1 percent higher than
the shear flow at station 50.
The other assumptions used in the two solutions are identical.
The method of using differences in bending stresses
automatically considers the effects of the shear carried by the
stringers, and it is not necessary to calculate the angles of
inclination of the stringers, It is, however, necessary to find the
torsional moments about the proper axis if the stringer forces
are omitted in the moment equation.
119
Chapter 5:Stress Analysis
Problems
Problem 5.1:Find the maximum tensile and maximum
compressive stresses resulting from bending of the beam
shown in Fig.P5.1. Find the distribution of shear stresses over
the cross section at the section where the Shear is a
maximum, considering points in the cross section at vertical
intervals of 1 in.
Figure P5.1
120
Chapter 5:Stress Analysis
Problems
Problem 5.2 : Find the maximum shear and bending stresses in
the beam cross section shown in Fig. P5.2 if the shear V is
10,000 lb and the bending moment M is 400,000 in.lb. Both
angles have the same cross section. Assume the web to be
effective in resisting bending stresses.
Figure P5.2
121
Chapter 5:Stress Analysis
Problems
Problem 5.3: Find the shear stress and the shear-flow
distribution over the cross section of the beam shown in Fig.
P5.3. Assume the web to be ineffective in resisting bending
and the stringer areas to be concentrated at points.
Figure P5.3
122
Chapter 5:Stress Analysis
Problems
Problem 5.4: Each of the five upper stringers has an area
0.4in2, and each of the five lower stringers has an area of 0.8
in2. Find the shear flows in all the webs if the vertical shearing
force is 12,000 lb.
Figure P5.4
123
Chapter 5:Stress Analysis
Problems
Problem 5.5: Each or the six stringers of the cross section
shown in Fig. P5.5 has an area of 0.5 in2. find the shear flows in
all webs and the location or the shear center for a vertical
shearing force of 10,000 lb.
Figure P5.5
124
Chapter 5:Stress Analysis
Problems
Problem 5.6: Find the shear flows in all webs in P5.6 for a
horizontal shearing force of 3000 lb. Each stringer has an area
a 0.5 in2.
Figure P5.6
125
Chapter 5:Stress Analysis
Problems
Problem 5.7: Find a general expression for the shear-flow
distribution around the circular tube shown. in Fig.5.7 Assume
the wall thickness t to be small compared with the radius R.
Figure P5.7
126
Chapter 5:Stress Analysis
Problems
Problem 5.8: Use Eqs.(5.17) and (5.18) to find the shear flow
in the webs of the two-stringer beams shown in Fig. P5.8
under the action of a vertical shear Vy.
Figure P5.8
127
Chapter 5:Stress Analysis
Problems
Problem 5.9:Find the shear-flow distribution for the section
shown in Fig. P5.9.
Figure P5.9
128
Chapter 5:Stress Analysis
Problems
Problem 5.10:Find the shear flows in the webs of the box
beam shown in Fig. P5.10 if the area is symmetrical about a
horizontal centerline
Figure P5.10
129
Chapter 5:Stress Analysis
Problems
Problem 5.11:Find the shear flows in the webs of the beam
shown in Fig. P5.11. all stringers have areas of 1.0 in2.
Figure P5.11
130
Chapter 5:Stress Analysis
Problems
Problem 5.12:Assume that the two right-hand stringers in Fig
P5.12 have arms of 3.0in2 and the other stringers have area of
1.0 in2. Find the shear flows in the webs by two methods.
Figure P5.12
131
Chapter 5:Stress Analysis
Problems
Problem 5.13:Find the shear flows in all webs if the two right-
hand stringers shown in Fig15.13 have areas of 1.5 in2 and the
other stringers have areas of 0.5 in2
Figure P5.13
132
Chapter 5:Stress Analysis
Problems
Problem 5.14: Find the shear-flow distribution. In all webs
shown in Fig. P5.14. All parts of the cross section resist
bending stresses.
Figure P5.14
133
Chapter 5:Stress Analysis
Problems
Problem 5.15: Solve example 5.8 if the beam depth varies from 5 in
at the free end to 15 in at the support .
(See Fig.P5.15).
Figure P5.15
134
Chapter 5:Stress Analysis
Problems
Problem 5.16:Find the shear flow for the cross section at x=50 in.
Consider only this one cross Section, but calculate the torsional
moments by two methods.
a- Select the torsional axis arbitrarily, and calculate the in-plane
components of the flange loads.
b- Take moments about a torsional axis joining the centroids attic
various crows sections.
135
Chapter 5:Stress Analysis
Problems
Problem 5.18: A cantilever beam 30 in long carries a vertical load of
1000 lb at the free end. The cross section is rectangular and -is 6 by
1 in . Find the maximum bending stress and the location of the
neutral axis if
(a) the 6-in side is vertical.
(b) (b) the 6-in side is tilted 50 from the vertical.
(c) (c) the 6-in side is tilted 300 from the vertical.
136
Chapter 5:Stress Analysis
Problems
Problem 5.20: Find the bending stresses and stringer loads for
the box beam whose cross section is shown in Fig. P5.20 if
Mz=10,000 and My=-40,000 in.lb. Assume the areas of the
stringer members are as follows:
a) a=b=c=d= 2 in2
b) a=b=3 in2 c=d=1 in2
c) a=d=3 in2 c=b=1 in2
d) a=c=3 in2 b=d=1 in2
e) a=c=1 in2 b=d=3 in2
Figure P5.20
137
Chapter 5:Stress Analysis
Problems
Problem 5.21: A beam with the cross section shown in Fig.
P5.21 resists a bending moment Mz=100 in.lb. Calculate the
bending stresses at points A,B, and C.
Figure P5.21
138
Chapter 5:Stress Analysis
Problems
Problem 5.22: The box beam shown in Fig. P5.22 resists bending
moments of Mz= 1,000,000 and My=120,000 in.lb. Find the bending
stress in each stringer member. Assume that the webs are
ineffective in bending and the areas and coordinates of the stringers
are as Follows:
Figure P5.22
139
Chapter 5:Stress Analysis
Problems
Problem 5.23: Find the shear flows at the cross section shown in
Fig. P5.23 for x= 50 in. Consider only the one cross section, and
calculate the in-plane components of the stringer loads.
Figure P5.23
Consider only the one cross section, and calculate the in-plane
components of the flange loads.
141
Chapter 5:Stress Analysis
Problems
Problem 5.27: Find the shear flows at station 100 of the fuselages
shown in Fig. P5.27. Assume all stringer areas to be 1 in2 .
Figure P5.27
142
Chapter 5:Stress Analysis
Further Readings