Student Version Fall2020

Download as pdf or txt
Download as pdf or txt
You are on page 1of 432

AERO 234: Fundamentals of Aircraft

Structures

Chapter 1: Statics Analysis


Of Structures

Dr. Issam Tawk


In this Chapter

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

Chapter 1: Statics Analysis Of Structures 2


1.1 Introduction

 The full understanding of both the terminology in statics and


the fundamental principles of mechanics is an essential
prerequisite to the analysis and design of structures.

Chapter 1: Statics Analysis Of Structures 3


1.2 Structural System
Bar elements, such as shown in Fig. 1.1, are one-dimensional structural mem-
bers which are capable of carrying and transmitting bending, shearing, torsional,
and axial loads or a combination or all four.

Figure 1.1: Bar elements.


a.) General bar; b) axial rod;
c) torsional rod

Chapter 1: Statics Analysis Of Structures 4


1.2 Structural System
Bars which are capable of carrying only axial loads are referred to as axial rods
or two-force members.
Structural systems constructed entirely out of axial rods are called trusses and
frequently are used in many atmospheric, sea, and land-based structures, since
simple tension or compression members are usually the lightest for
transmitting forces.

Chapter 1: Statics Analysis Of Structures 5


1.2 Structural System
Plate elements, such as shown in Fig. 1.2, are two-dimensional extensions of
bar elements. Plates made to carry only in-plane axial loads are called
membranes.

Figure 1.2: Plate elements. a)General plate


element; b) membranes dement; c) shear panel.

Chapter 1: Statics Analysis Of Structures 6


1.2 Structural System
Those which are capable of carrying only in-plane shearing loads are
referred to as shear panels; frequently these are found in missile fins, aircraft
wings, and tail surfaces.
Shells are curved plate elements which occupy a space. Fuselages, building
domes, pressure vessels, etc., are typical examples of shells.

Chapter 1: Statics Analysis Of Structures 7


1.3 Load Classification
Loads which act on a structural system may be generally classified in accordance
with their causes. Those which are produced by surface contact are called surface
loads. Dynamic and/or static pressures are examples of surface loads. If the area of
contact is very small, then the load is said to be concentrated; otherwise, it is
called a distributed load. (See Fig. 1.3.)

Figure 1.3: Concentrated and


distributed loads. a)Actual loads: b)
idealized loads; c)wing pressure load.

Chapter 1: Statics Analysis Of Structures 8


1.3 Load Classification
Loads which depend on body volume are called body loads. Inertial, magnetic,
and gravitational forces are typical examples. Generally, these loads are assumed
to be distributed over the entire volume of the body.

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.

Regardless of the classification of the externally imposed loads, a structural


member, in general, resists these loads internally in the form of bending, axial,
shear, and torsional actions or a combination of the four.
Chapter 1: Statics Analysis Of Structures 9
1.3 Load Classification
 Thus, a bending moment may be defined as a force whose vector
representation lies in and parallel to the plane of the cut, while a torque is a force
whose vector representation is normal to that cut. On the other hand, shear load
is a force which lies in and is parallel to the plane of the cut, while axial load is a
force which acts normal to the plane of the cut (See Fig. 1.4).

Figure 1.4: Bending moment, torque,


shear load, axial load

Chapter 1: Statics Analysis Of Structures 10


1.4 Supports And Reactions
The primary function of supports is to provide, at some points of a structural
system, physical restraints that limit the freedom of movement to only that
intended in the design. The types of supports that occur in ordinary practice are
shown in Fig. 1.5.

Figure1.5: Support Types.


(a) Hinge support;
(b) hinge-roller support;
(c) fixed support; (d) fixed-
roller support.
.

Chapter 1: Statics Analysis Of Structures 11


1.4 Supports And Reactions
Likewise, a hinge-roller support allows rotation and a translation in only the x
direction, and hence there exists one reactive force (reaction) in the y direction.

A fixed support normally is designed to provide restraints against rotation and


all translations; therefore, reactive forces and moments (reactions) are developed
along the directions where movements are not permitted.

Chapter 1: Statics Analysis Of Structures 12


1.5 Equations Of Static Equilibrium
Any solid body in space or any part cut out of the body is said to be in a state of
stable static equilibrium if it simultaneously satisfies:

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.

Chapter 1: Statics Analysis Of Structures 13


1.6 Statically Determinate And Indeterminate Structures

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.

Chapter 1: Statics Analysis Of Structures 14


1.6 Statically Determinate And Indeterminate Structures
For example, in Fig. 1.6a if member 1-3 and/or reaction R4x is removed, then the
structure becomes statically determinate and maintains its stability. Likewise, in Fig.
1.6b if members 2-4 and 2-6 and either reaction R2y or R3y are removed, then the
structure becomes determinate and stable. If, on the other hand, additional
members such as 3-5 and 1-5 are removed, then the structure becomes a
mechanism or unstable.

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

Chapter 1: Statics Analysis Of Structures 16


1.7 Applications
Solution of example 1.1 :

Figure 1.8

Chapter 1: Statics Analysis Of Structures 17


1.7 Applications
Solution of example 1.1 :

For the pulley

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

Chapter 1: Statics Analysis Of Structures 21


1.7 Applications
Solution of example 1.2 :
 To Find the unknown reactions, consider the entire structure as a free body.

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

R4 y  4000  1000  45000  0


R4 y  500lb
Chapter 1: Statics Analysis Of Structures 22
1.7 Applications
Solution of example 1.2 :
At the joint 4

F x  0( )
F2  2000  0
F2  2000lb

F y  0( )
500  F1  0
F1  500lb

Chapter 1: Statics Analysis Of Structures 23


1.7 Applications
Solution of example 1.2 :
At the joint 1

F y  0( )
500  F4 sin 45o  0
F4  707lb

F x  0( )
2000  707 cos 45o  F3  0
F3  2500lb

Chapter 1: Statics Analysis Of Structures 24


1.7 Applications
Solution of example 1.2 :

Figure 1.10

Finally, isolation joint 5

F y  0( )
707 sin 45o  F5  0
F5  500lb
Chapter 1: Statics Analysis Of Structures 25
1.7 Applications

(b) Method of sections

Figure 1.11
F y  0( )
500  F5  0
F5  500lb

Chapter 1: Statics Analysis Of Structures 26


1.7 Applications
Example 1.3 :Find the loads on the lift and drag-truss members of the externally
braced monoplane wing shown in Fig. 1.12. The load is assumed to be uniformly
distributed along the span of the wing. The diagonal drag-truss members are wires,
with the tension diagonal effective and the other diagonal carrying no load.

Figure 1.12

Chapter 1: Statics Analysis Of Structures 27


1.7 Applications
Solution of example 1.3 :
The vertical load or 20 lb/in is distributed to the spars in inverse proportion to
the distance between the center of pressure and the spars. The load on the front
spar, is therefore 16 lb/in, and that on the rear spar is 4 lb/in. If the front spar is
considered as a free body, as shown in Fig. 1.13a, the vertical forces at A and G
may be obtained.

M A  16  180  90  100Gz  0


Gz  2590lb
Gy 2590

100 60
G y  4320lb
F z  16  180  2590  Az  0
Az  290lb
Chapter 1: Statics Analysis Of Structures 28
1.7 Applications

Solution of example 1.3 :


Force Ay cannot be found at this point in
the analysis, since the drag-truss members
exert forces on the front spar which are not
shown in Fig. 1.19a. If the rear spar is
considered as a free body, as shown in Fig.
1.19b, the vertical forces at B and E may be
obtained:
M B  4  180  90  100 Ez  0
Ez  648lb
Ex Ey 648
 
30 100 60
Ex  324lb, E y  1080lb
F z  4  180  648  Bz  0
Bz  72lb Figure 1.13
Chapter 1: Statics Analysis Of Structures 29
1.7 Applications
Solution of example 1.3 :
The loads in the plane of the drag truss
can be obtained now. The forward load of 5
lb/in is applied as concentrated loads at the
panel points, as shown in Fig. 1.19c. The
components of the forces at G and E which
lie in the plane of the truss also must be
considered. The remaining reactions at A
and B and the forces in all drag-truss
members now can be obtained by the
methods of analysis for coplanar trusses,
shown in Fig. 1.19c.

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

Chapter 1: Statics Analysis Of Structures 31


Problems
Prob. 1.2:

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

Chapter 1: Statics Analysis Of Structures 32


Problems
Prob. 1.3:

Find the forces at points A and B of the landing gear shown in Fig P1.3.

10

Figure P1.3

Chapter 1: Statics Analysis Of Structures 33


Further Readings

David J. Peery & J.J. Azar: AIRCRAFT STRUCTURES,


2nd Edition

Chapter 1: Fundamentals of Aircraft Structures

Chapter 1: Statics Analysis Of Structures 34


AERO 234: Fundamentals of Aircraft
Structures

Chapter:2 FLIGHT -VEHICLE


IMPOSED LOADS

Dr. Issam Tawk


In this Chapter

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 2


2.1 Introduction
 Before the final selection of member sizes on flight vehicles can be made, all
load conditions imposed on the structure must be known.
The load conditions are those which are encountered both in flight and on the
ground.

Since it is impossible to investigate every loading condition which a flight vehicle


might encounter in its service lifetime, it is normal practice to select only those
conditions that will be critical for every structural member of the vehicle. These
conditions usually are determined from past investigation and experience and are
definitely specified by the licensing or procuring agencies.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 3


2.1 Introduction
 Although the calculations of loads imposed on flight-vehicle structures are the
prime responsibility of a special group in an engineering organization called the
loads group, a basic general overall knowledge of the loads on vehicles is
essential to stress analysts.

Therefore, in this chapter we present the fundamentals and terminology


pertaining to flight-vehicle loads.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 4


2.2 General Considerations
Every flight vehicle is designed to safely carry out specific missions. This results
in a wide variety of vehicles relative to size, configuration, and performance.

Commercial transport aircraft are specifically designed to transport passengers


from one airport to another. These types of aircraft are never subjected to violent
intentional maneuvers.
Military aircraft, however, used in fighter or dive-bomber operations, are
designed to resist violent maneuvers.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 5


2.2 General Considerations
The design conditions usually are determined from the maximum acceleration
which the human body can withstand, and the pilot will lose consciousness
before reaching the load factor (load factor is related to acceleration) which
would cause structural failure of the aircraft.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 6


2.2 General Considerations
Terms are defined below which are generally used in the specification of loads
on flight vehicles.
 The limit loads used by civil agencies or applied loads used by military
agencies are the maximum anticipated loads in the entire service life-span of
the vehicle.
 The ultimate loads, commonly referred to as design loads, are the limit loads
multiplied by a factor of safety (FS):

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 7


2.2 General Considerations
 Generally, a factor of safety which varies from 1.25 for missile structures to
1.5 for aircraft structures is used in practically every design because of the
uncertainties involving:
– The simplifying assumption used in the theoretical analyses
– The variations in material properties and in the standards of quality control
– The emergency actions which might have to be taken by the pilot, resulting in loads
on the vehicle larger than the specified limit loads.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 8


2.2 General Considerations
The limit loads and ultimate loads quite often are prescribed by specifying
certain load factors.
 The limit-load factor is a factor by which basic loads on a vehicle are
multiplied to obtain the limit loads.
 Likewise, the ultimate load factor is a factor by which basic vehicle loads
are multiplied to obtain the ultimate loads; in other words, it is the
product of the limit load factor and the factor of safety.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 9


2.3 BASIC FLIGHT LOADING CONDITIONS
One of four basic conditions will probably produce the highest load in any
part of the airplane for any flight condition. Usually these conditions are called
positive high angle of attack, positive low angle of attack, negative high angle of
attack, and negative low angle of attack.
All these conditions represent symmetrical flight maneuvers; i.e., there is no
motion normal to the plane of symmetry of the airplane.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 10


2.3 BASIC FLIGHT LOADING CONDITIONS
The positive high angle of attack (PHAA) condition is obtained in a pullout at the
highest possible angle of attack on the wing. The lift and drag forces are
perpendicular and parallel respectively, to the relative wind, which is shown as
horizontal in Fig. 2.1a.

Figure 2.1a

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 11


2.3 BASIC FLIGHT LOADING CONDITIONS
The resultant R of these forces always has an aft component with respect to the
relative wind, but will usually have a forward component C with respect to the
wing chord line, because of the high angle of attack α.

 The maximum forward component C will be obtained when α has a maximum


value. In order to account for uncertainties in obtaining the stalling angle of attack
under unsteady flow conditions, most specifications arbitrarily require that a value
of α be used which is higher than the wing stalling angle under steady flow
conditions.

Figure 2.1a

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 12


2.3 BASIC FLIGHT LOADING CONDITIONS
An angle of attack corresponding to a coefficient of lift of 1.25 times the
maximum coefficient of lift for steady flow conditions is often used, and
aerodynamic data are extrapolated from data measured for steady flow
conditions.

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 13


2.3 BASIC FLIGHT LOADING CONDITIONS
In the PHAA condition, the bending moments from the normal forces N, shown
in Fig. 2.1a, produce compressive stresses on the upper side of the wing and the
moments from the chord wise forces C produce compressive stresses on the
leading edge of the wing, These compressive stresses will be additive in the upper
flange of the front spar and the stringers adjacent to it.
The PHAA condition, therefore, will be critical for compressive stresses in the
upper forward region of the wing cross section and for tensile stresses in the
lower aft region of the wing cross section.

Figure 2.1a

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 14


2.3 BASIC FLIGHT LOADING CONDITIONS
In the positive low angle of attack (PLAA) condition, the wing has the smallest
possible angle of attack at which the lift corresponding to the limit-load factor may
be developed.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 15


2.3 BASIC FLIGHT LOADING CONDITIONS
Some specifications require that the terminal velocity of the aircraft (the velocity
obtained in a vertical dive sustained until the drag equals the airplane weight) be
calculated and the limit on the diving speed be determined as a function of the
terminal velocity.

 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 16


2.3 BASIC FLIGHT LOADING CONDITIONS
In the PLAA condition, shown in Fig. 2.1b, the chord wise force C is the largest
force acting aft on the wing for any positive flight attitude.

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 17


2.3 BASIC FLIGHT LOADING CONDITIONS
 In this condition, the line of action of the resultant wing force R is farther aft
than for any other positive flight condition.

 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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 18


2.3 BASIC FLIGHT LOADING CONDITIONS
The negative high angle of attack (NHAA) condition, shown in Fig. 2.1c, occurs in
intentional flight maneuvers in which the air loads on the wing are down or when
the airplane strikes sudden downdrafts while in level flight.
The load factors for intentional negative flight attitudes are considerably
smaller than for positive flight attitudes, because conventional aircraft engines
cannot be operated under a negative load factor for very long and because the pilot
is in the uncomfortable position of being suspended from the safety belt or
harness.

Figure 2.1c

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 19


2.3 BASIC FLIGHT LOADING CONDITIONS
In the NHAA condition, usually the wing is assumed to be at the negative stalling
angle of attack for steady flow conditions. The assumption used in the PHAA
condition—the maximum lift coefficient momentarily exceeds that for steady flow is
seldom used because it is improbable that negative maneuvers will be entered
suddenly.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 20


2.3 BASIC FLIGHT LOADING CONDITIONS
The negative low angle of attack (NLAA) condition, shown in Fig. 2.1d, occurs at the
diving-speed limit or the airplane. This condition may occur in an intentional
maneuver producing a negative load factor or in a negative gust condition.

Figure 2.1d

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 21


2.3 BASIC FLIGHT LOADING CONDITIONS
The aft load C is a maximum for any negative flight attitude, the compressive
bending stresses have a maximum value in the lower aft region of the wing cross
section, and the tensile bending stresses have a maximum value in the upper
forward region of the wing cross section.
The resultant force R is farther forward than in any other flight attitude, and the
download on the horizontal tail will probably be larger than in any other negative
flight attitude.

Figure 2.1d

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 22


2.3 BASIC FLIGHT LOADING CONDITIONS
In summary, one of the four basic symmetrical flight conditions is critical for the
design of almost every part of the airplane structure.

In the stress analysis of a conventional wing, it is necessary to investigate each


cross section for each of the four conditions. Then each stringer or spar flange is
designed for the maximum tension and the maximum compression obtained in any
of the conditions.

Chapter:2 FLIGHT -VEHICLE IMPOSED


23
LOADS
2.3 BASIC FLIGHT LOADING CONDITIONS
 The probable critical conditions for each region or the cross section are shown in
Fig. 2.2.

Figure 2.2

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 24


2.3 BASIC FLIGHT LOADING CONDITIONS

Some specifications require the investigation of additional conditions of medium-


high angle of attack and medium-low angle of attack which may be critical for
stringers midway between the spars, but usually these conditions are not
considered of sufficient importance to justify the additional work required for the
analysis.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 25


2.3 BASIC FLIGHT LOADING CONDITIONS
For aircrafts such as transport or cargo aircrafts, in which the load may be placed
in various positions in the gross-weight condition, it is necessary to determine the
balancing tail loads for the most forward and most rearward center of gravity
positions at which the airplane may be flown at the gross weight.

 Each of the four flight conditions must be investigated for each extreme
position of the center of gravity.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 26


2.3 BASIC FLIGHT LOADING CONDITIONS
For smaller aircraft, in which the useful load cannot be shifted appreciably, there
may be only one position of the center of gravity at the gross-weight condition.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 27


2.3 BASIC FLIGHT LOADING CONDITIONS
The gust load factors on an aircraft are greater when it is flying at the minimum
flying weight than they are at the gross-weight condition. While this is seldom
critical for the wings, since they have less weight to carry, it is critical for a structure
such as the engine mount which carries the same weight at a higher load factor.

It is therefore necessary to calculate gust load factors at the minimum weight at
which the aircraft will be flown.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 28


2.3 BASIC FLIGHT LOADING CONDITIONS
For aircraft equipped with wing flaps, other high lift devices, or dive brakes,
additional flight loading conditions must be investigated for the flaps extended.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 29


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
Extensive aerodynamic information is required to investigate the performance,
control, and stability of a proposed aircraft. Only the information which is required
for the structural analysis is considered here.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 30


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
While these data can be calculated accurately for a wing with a conventional
airfoil section, similar data for the combination of the wing and fuselage or the
wing, fuselage, and nacelles are more difficult to calculate accurately from
published information because of the uncertain effects of the aerodynamic
interference of various components.

It is therefore desirable to obtain wind tunnel data on a model of the complete
airplane less horizontal tail.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 31


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
Wind tunnel tests of a model of the complete airplane with the horizontal tail
removed provide values of the lift, drag, and pitching moment for all angles of
attack. Then components of the lift and drag forces with respect to airplane
reference axes are determined. The aircraft reference axes may be chosen as
shown in Fig. 2.3

Figure 2.3

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 32


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
The force components are CzqS and CxqS along these axes, where q=ρV2/2 is the
dynamic pressure and S is the surface wing area. The non dimensional force
coefficients Cz and Cx are obtained by projecting the lift and drag coefficients,
respectively, for the airplane less horizontal tail along the reference axes by the
following equations:

Cz= CL cos θ + CD sinθ Eq. (2.1)

Cx= CD cos θ - CL sinθ Eq. (2.2)

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 33


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
The pitching moment about the airplane's center of gravity is obtained from
wind tunnel data and is 𝑪𝒎𝒂−𝒕 𝒄𝒒𝑺
• where 𝑪𝒎𝒂−𝒕 is the dimensionless pitching-moment coefficient of the airplane less tail and 𝒄 is
the mean aerodynamic chord of the wing.

 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 34


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
The balancing air load on the horizontal tail, CtqS, is obtained from the
assumption that there is no angular acceleration of the airplane. The moments
of the forces shown in Fig. 2.3 about the center of gravity are therefore in
equilibrium:

Figure 2.3

Eq. (2.3)

 where Ct is a dimensionless tail force coefficient expressed in terms of the


wing area and Lt, is the distance from the airplane's center of gravity to the
resultant air load on the horizontal tail, as shown in Fig. 2.3.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 35


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
Since the pressure distribution on the horizontal tail varies according to the
attitude of the airplane. Lt theoretically varies for different loading conditions. This
variation is not great, and it is customary to assume Lt Constant, by using a
conservative forward position of the center of pressure on the horizontal tail.

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)

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 36


2.4 FLIGHT VEHICLE AERODYNAMLC LOADS
For power-on flight conditions, the moment of the propeller or jet thrust about
the center of gravity of the airplane should also be considered. This adds another
term to the Eq. (2.3).
Now the aerodynamic coefficients can be plotted against the angle of attack α, as
shown in Fig.2.4, if more than one position of the center of gravity is considered in
the analysis, it is necessary to calculate the curves for 𝑪𝒎𝒂−𝒕 , 𝑪𝒕 ,and 𝑪𝒛𝒂 , for each
center-of-gravity position.

The right-hand portions of the solid curves


shown in Fig. 2.4 represent the aerodynamic
characteristics after stalling of the wing.

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 38


2.5 FLIGHT VEHICLE INERTIA LOADS
The maximum load on any part of a flight-vehicle structure occurs when the
vehicle is being accelerated.

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.

Before any structural component can be designed, it is necessary to determine


the inertia loads acting on the vehicle.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 39


2.5 FLIGHT VEHICLE INERTIA LOADS
 In many of the loading conditions, a flight vehicle may be considered as being in
pure translation or pure rotation.

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)

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 40


2.5 FLIGHT VEHICLE INERTIA LOADS
If a Cartesian system of x, y, and z axes is chosen in this frame, then Eq. (2.5) gives,
upon resolving into components:

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 41


2.5 FLIGHT VEHICLE INERTIA LOADS
 This type of motion is called plane motion and it occurs, for example, when a
vehicle is pitching and yet has no rolling or yawing motion. All elements of the
vehicle move in planes parallel to the plane of symmetry.

Any type of plane motion can be considered as a rotation about some


instantaneous axis perpendicular to the planes of motion, and the following
equations for inertia forces are derived on the assumption that the rigid body is
rotating about an instantaneous axis perpendicular to a plane of symmetry of the
body.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 42


2.5 FLIGHT VEHICLE INERTIA LOADS
The rigid mass shown in Fig. 2.5 is rotating about point O with a constant angular
velocity ω.
The acceleration of any point a distance r from the center of rotation is ωr2 and is
directed toward the center of rotation.
The inertia force acting on an element of mass dM is the product of the mass and
the acceleration, or ω2rdM, and is directed away from the axis of rotation.

Figure 2.5

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 43


2.5 FLIGHT VEHICLE INERTIA LOADS
This inertia force has components ω2xdM, parallel to the x axis and ω2ydM
parallel to the y axis.

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 44


2.5 FLIGHT VEHICLE INERTIA LOADS
The inertia force in the x direction is found in the same manner:

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 45


2.5 FLIGHT VEHICLE INERTIA LOADS
If the body has an angular acceleration α, the element of mass dM has an
additional inertia force αrdM acting perpendicular to r and opposite to the
direction of acceleration. This force has components αx dM in the y direction and
αy dM in the x direction, as shown in Fig. 2.6.

Figure 2.6

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 46


2.5 FLIGHT VEHICLE INERTIA LOADS
The resultant inertia force on the entire body in the x direction is:

The resultant inertia force in the y direction is

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:

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 48


2.5 FLIGHT VEHICLE INERTIA LOADS
By substituting the value of I0 from Eq. (2.10) in Eq. (2.9), the following expression
for the inertia torque is obtained:

Eq. (2.11)

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 49


2.5 FLIGHT VEHICLE INERTIA LOADS
The inertia forces obtained in Eqs. (2.7), (2.8), and (2.11) may be represented as
forces acting at the center of gravity and the couple Icα, as shown in Fig. 2.7, The
force and the couple Icα must both produce moments about point O which are
opposite to the direction of α. The force ω2 must act away from point O.

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 51


2.6 Load Factors for Translational Acceleration
For flight or landing conditions in which the vehicle has only translational
acceleration, every part of the vehicle is acted on by parallel inertia forces which are
proportional to the weight of the part.

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)

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 52


2.6 Load Factors for Translational Acceleration
The combined inertia and gravity forces are considered in the analysis in the
same manner as weights which are multiplied by the load factor n.

In the case of an airplane in flight with no horizontal acceleration, as shown in


Fig. 2.9, the engine thrust is equal to the airplane drag, and the horizontal
components of the inertia and gravity forces are zero.

Figure 2.9

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 53


2.6 Load Factors for Translational Acceleration
The weight and the inertia force on the airplane act down and will be equal to
the lift. The airplane lift L is the resultant of the wing and tail lift Forces. The load
factor is defined as follows :

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 55


2.6 Load Factors for Translational Acceleration

 It is also convenient to consider the horizontal inertia loads as equal to the


product of a load factor nx and the weights. This horizontal load factor, often
called the thrust load factor, is obtained from the equilibrium of the horizontal
forces shown in Fig. 2.10:

Eq. (2.14)
or

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 56


2.6 Load Factors for Translational Acceleration
 A more general case of translational acceleration is shown in Fig. 2.11, in which
the airplane thrust line is not horizontal. It is usually convenient to obtain
components of forces along x and z axes which are parallel and perpendicular to
the airplane thrust line.

 The combined weight and inertia load on any


element has a component along the z axis of the
following magnitude:
Figure 2.11

or Eq. (2.15)

104
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 57
2.6 Load Factors for Translational Acceleration

 From summation of all forces along the z axis

Eq. (2.16)

 By combining Eqs. (2.15) and (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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 59


2.6 Load Factors for Translational Acceleration

 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.

 The load factor in the horizontal direction


is similarly defined as the horizontal ground
reaction divided by the airplane weight:

Figure 2.12
Eq. (2.17)

and Eq. (2.18)

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 60


2.6 Load Factors for Translational Acceleration

 In the airplane analysis, it is necessary to obtain the components of the load


Factor along axes parallel and perpendicular to the propeller thrust line. However,
aerodynamic forces are usually obtained first as lift and drag forces perpendicular
and parallel to the direction of flight.

 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 61


2.6 Load Factors for Translational Acceleration

 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 .

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 62


2.6 Load Factors for Translational Acceleration

 The load factor is therefore defined as:

Eq. (2.19)

• Where includes all forces except weight and inertia forces.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 63


2.7 Velocity Load Factor Diagram

 The various loading conditions for an airplane usually are represented on a


graph of the limit-load factor n plotted against the indicated airspeed V.
 This diagram is often called a V-n diagram, since the load factor n is related to
the acceleration of gravity g. In all such diagrams, the indicated airspeed is used,
since all air loads are proportional to q or ρV2/2.
 The value of q is the same for the air density ρ and the actual airspeed at
altitude as it is for the standard sea-level density ρ0 and the indicated airspeed,
since the indicated airspeed is defined by this relationship.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 64


2.7 Velocity Load Factor Diagram

 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 aerodynamic forces on an airplane are in equilibrium with the forces of


gravity and inertia.
 If the airplane has no angular acceleration, both the inertia and gravity forces
will be distributed in the same manner as the weight of various items of the
aircraft and will have resultants acting through the center of gravity of the aircraft.
 It is convenient to combine the inertia and gravity forces as the product of a
load factor n and the weight W.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 65


2.7 Velocity Load Factor Diagram

 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.

 The load factor n is obtained from


a summation of forces along the z
axis:

or Eq. (2.20)

Figure 2.13

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 66


2.7 Velocity Load Factor Diagram

 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.

 In accelerated flight, the maximum coefficient might be obtained at higher


speeds. For Cza,max to be obtained at twice the stalling speed, a load factor n=4
would be developed as shown by Eq. (2.20).

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 67


2.7 Velocity Load Factor Diagram
 For a force coefficient of 1.25Cza,max representing the highest angle of attack for
which the wing is analyzed, the value of the load factor n is obtained from Eq.
(2.20) and may be plotted against the airplane velocity V, as shown by line OA in
Fig. 2.13.

 This line OA represents a limiting condition, since it is possible to maneuver


the airplane at speeds and load factors corresponding to points below or to the
right of line OA, but it is impossible to maneuver at speeds and load factors
corresponding to points above or to the left of line OA because this would
represent angles of attack much higher than the stalling angle.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 68


2.7 Velocity Load Factor Diagram
 The line AC in Fig. 2.13 represents the limit on the maximum maneuvering load
factor for which the airplane is designed. This load factor is determined from the
specifications for which the airplane is designed, and the pilot must restrict
maneuvers so as not to exceed this load factor.

Figure 2.13

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 69


2.7 Velocity Load Factor Diagram

 At speeds below that corresponding to point A, it is impossible for the pilot to


exceed the limit load factor in any symmetrical maneuver, because the wing will
stall at a lower load factor.

 For airspeeds between those corresponding to points A and C, it is not


practical to design the airplane structure so that it could not be overstressed by
violent maneuvers.

 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 70


2.7 Velocity Load Factor Diagram

 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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 71


2.7 Velocity Load Factor Diagram

 The equation for line OB is obtained by substituting the maximum negative


value of Cza, into Eq. (2.20). Similarly, line BD corresponds to line AC, except that
the limit-load factor specified for negative maneuvers is considerably less than for
positive maneuvers.

 The aircraft may therefore be maneuvered in such a manner that velocities


and load factors corresponding to the coordinates of points within the area
OACDB may be obtained.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 72


2.7 Velocity Load Factor Diagram

 The most severe structural load ng conditions will be represented by the


corners of the diagram, points A, B, C and D. Points A and B represent PHAA and
NHAA conditions. Point C represents the PLAA condition in most cases, although
the positive gust load condition, represented by point F, may occasionally be
more severe.

 The NLAA condition is represented by point D or by the negative gust


condition, point E, depending on which condition produces the greatest negative
load factor.

 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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 74


2.8 Gust Load Factor

 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)

Figure 2.4 Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 75


2.8 Gust Load Factor

 After striking the gust, the airplane normal force coefficient increases by an
amount determined from Eqs. (2.21) and (2.22):

Eq. (2.23)

 The increase in the airplane load factor Δn may be obtained by substituting


the value of Δcza from Eq. (2.23) into Eq. (2.20):

• ρ= standard sea-level air density,


0.002378 slug/ ft3
• S =wing area, ft2 Eq. (2.24)
• β= slope of the curve of Cza versus α , rad
• KU = effective gust velocity, ft/s
• V = indicated airspeed, ft/s or
• W=gross weight of airplane, lb

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 76


2.8 Gust Load Factor

 For purposes of calculation, it is more convenient to determine the slope β


per degree and the airspeed V in miles per hour.
 Introducing the necessary constants in Eq. (2.24) yields

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).

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 77


2.8 Gust Load Factor

 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 78


2.8 Gust Load Factor
 Equation (2.25) is similarly plotted, as shown by the inclined lines through
points G and E of Fig. 2.13. These lines represent load factors obtained when the
airplane is in a vertical attitude and strikes positive or negative gusts in
directions normal to the thrust line.
 The gust load factor represented by point F of Fig. 2.13 may be more severe
than the maneuvering load factor represented by point C. In the case shown,
however, the maneuvering load factor is obviously greater and will represent the
PLAA condition.

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

Solution of example 2.1 :


a) First consider the entire airplane as a free body

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 82


2.9 Examples

Solution of example 2.1 :


b) Consider the aft section of the fuselage as a free body, as shown in Fig.
2.16.
Figure 2.16

It is acted on by an inertia force of

The tension on section BB is found as follows:

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 83


2.9 Examples

Solution of example 2.1 :


Since there is no vertical acceleration, there is no vertical inertia force. Section
BB has a shear force V1 of 6300 Ib, which is equal to the sum of the weight and
the vertical component of the cable force.

Consider the portion of the airplane forward of section AA as a free body, as


shown in Fig. 2.17. The inertia force is the following.

Figure 2.17

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 84


2.9 Examples

Solution of example 2.1 :


The section AA must also resist a shearing force V2 of 3000 lb and a bending
moment obtained by taking moments of the forces shown in Fig. 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

Solution of example 2.1 :


c) From elementary dynamics, the landing run s is obtained as follows:

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 86


2.9 Examples
Example 2.2 A 30,000 lb airplane is shown in Fig. 2.19a at the time of
landing impact, when the ground reaction on each main wheel is 45,000 Ib.
Figure 2.19a Figure 2.19b

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

Solution of example 2.2 :


a) Considering the entire airplane as a free body and taking a summation of
vertical forces yield

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 88


2.9 Examples

Solution of example 2.2 :


Consider the landing gear as a free body, as shown in Fig. 2.19b.

The inertia force is:

The compression load C in the oleo strut is found


Figure 2.19b
from a summation of vertical forces

The bending moment m is found as follows:

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 89


2.9 Examples

Solution of example 2.2 :


b)The inertia force acting on the portion of the wing shown in Fig. 2.19c is

The wing shear at section AA is found from a


summation or vertical forces
Figure 2.19c

The wing bending moment is found by taking moments about section AA

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS


90
2.9 Examples

Solution of example 2.2 :


c) The shock strut deflection is found by assuming a constant vertical
acceleration or -2g, or -64.4 ft/s2, from an initial vertical velocity of 12 ft/s to a
final zero vertical velocity.

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

a) Find the inertia forces on the airplane.


b)Find the inertia forces on a 400-lb gun turret in the tail which is 500 in aft of the
center of gravity. Neglect the moment of inertia of the turret about its own
center of gravity.
Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 92
2.9 Examples

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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 93


2.9 Examples
Solution of example 2.3 :
a) The inertia forces on the entire airplane may be considered as horizontal and
vertical forces Max and May respectively, at the center of gravity and a couple
𝐼𝑐 ∝ as shown in Fig. 2.20, These correspond to the inertia forces shown on the
mass of Fig. 2.7, since the forces at the center of gravity represent the product of
the mass and the acceleration components of the center of gravity.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 94


2.9 Examples
Solution of example 2.3 :
b) The acceleration of the center of gravity of the airplane is now known, and the
acceleration and inertia forces for the turret can be obtained by the method
shown in Fig. 2.8, where the center of gravity of the airplane corresponds to point
O of Fig. 2.8 and the center of gravity of the turret corresponds to point C. These
forces are shown in Fig. 2.21 and have the following values:

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 96


2.9 Examples
Solution of example 2.3 :
The angle of rotation during this time is found from

The vertical motion of the nose wheel resulting from this rotation, shown in Fig.
2.22, is

Figure 2.22

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 97


2.9 Examples
Solution of example 2.3 :

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

The remaining angle of rotation ϴ2 shown in Fig. 2.23, 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.

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 98


2.9 Examples
Solution of example 2.3 :
The angular velocity of the airplane at the time the nose wheel strikes the ground
is found from the following equation.

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 99


2.9 Examples
Solution of example 2.3 :

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:

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 100


Problems
Prob. 2.1:

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

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 101


Problems
Prob. 2.2:
An airplane weighing 8000 lb has an upward acceleration of 3g when landing. If
the dimensions are as shown in Fig. P2.2, what are the wheel reactions R1 and R2
? What time is required to decelerate the airplane from a vertical velocity of 12
ft/s? What is the vertical compression of the landing gear during this
deceleration? What is the shear and bending moment on a vertical section AA if
the weight forward of this section is 2000 lb and has a center of gravity 40 in from
this cross section?

Figure P2.2

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 102


Problems
Prob. 2.3:
-The airplane shown in Fig. P2.3 is making an arrested landing on a carrier deck.
Find the load factors nz and nx, perpendicular and parallel to the deck at the
center of gravity.
-Find the relative vertical velocity with which the nose wheel strikes the deck if
the vertical velocity of the center of gravity is 12ft/s and the angular velocity is
0.5rad/s counterclockwise for the position shown. The radius of gyration for the
mass of the airplane about the center of gravity is 60in. Assume no change in the
dimensions or loads shown.

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:2 FLIGHT -VEHICLE IMPOSED LOADS 105


Further Readings

David J. Peery & J.J. Azar: AIRCRAFT STRUCTURES,


2nd Edition

Chapter 2: FLIGHT -VEHICLE IMPOSED LOADS

Chapter:2 FLIGHT -VEHICLE IMPOSED LOADS 106


AERO 234: Fundamentals of Aircraft
Structures

Chapter 3: Elasticity Of
Structures

Dr. Issam Tawk


In this Chapter
3.1 Introduction
3.2 Stresses
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field
3.4 Strains And Strain-Displacement Relationships
3.5 Compatibility Eqs. For Plane-Stress &Plane-Strain Problems
3.6 Boundary Conditions
3.7 Stress-Strain Relationships
3.8 Transformations Of Stresses And Strains
3.9 Principal Stresses
3.10 Mohr’s circle of stress
3.11 Mohr’s circle of strain
3.12 Experimental measurement of surface strains

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 

where σ ii (i = x, y, z) are the normal stresses and σ ij=σ ji (i≠j=x, y,


z) are the shearing stresses. The first subscript on σ ij (i,j=x, y, z)
denotes the plane at a constant i on which the stress is acting, and
the second subscript denotes the positive direction of the stress.

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

In general, a solid which is acted on by a set of external applied


loads experiences a state of stress that is not uniform throughout
the body. This condition gives rise to a set of equations which are
referred to as the equations of equilibrium.

Consider the three-


dimensional solid shown in
Fig. 3.4:

Figure 3.4
Elasticity Of Structures 11
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field

Using the equilibrium equations of statics yields the following,


where X,Y and Z are body forces.
 Fx  0  xx, x   xy , y   zx , z  X  0 Eq. (3.5)

F y 0  xy , x   yy , y   zy , z  Y  0 Eq. (3.6)

F z 0  zx, x   zy , y   zz , z  Z  0 Eq. (3.7)

Or

The equations of equilibrium must be satisfied at all interior


points in a deformable body under a three-dimensional force
system. Elasticity Of Structures 12
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field

The comma denotes partial differentiation with respect to the


following subscript:
 zx  yy
 zx, x   yy , y  etc.
x y
In a cylindrical coordinate set of axes, the equilibrium equations
may be easily derived from Figure 3.5

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

where again R, ϴ, and Z are body forces.

Elasticity Of Structures 14
3.3 Stress Equilibrium Equations In A Non-Uniform Stress Field

For plane stress problems, the equilibrium equations simplify to


the following:

 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 ,  R0
r r Eq. (3.12)

1 2
  ,   r ,r  r    0
r r

Elasticity Of Structures 15
3.4 Strains And Strain-Displacement Relationships

Strains are non-dimensional quantities associated with the


deformations (displacements) of an element in a solid body under
the action of external applied loads.
To arrive at a mathematical definition of the strain components,
take the solid body shown in Fig. 3.6 and consider only the
infinitesimal elements OA, 0B, and OC.

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 shows all the


displacements resulting from the
application of external applied
loads.

Figure 3.7
Elasticity Of Structures 17
3.4 Strains And Strain-Displacement Relationships

Figure 3.7b shows all the displacements resulting from the


application of external applied loads from another reference.

Figure 3.7 b
Elasticity Of Structures 18
3.4 Strains And Strain-Displacement Relationships

When these relative displacements occur in a solid body, the


body is said to be in a state of strain.
The strains associated with the relative change in length are
referred to as normal strains, and those related to relative change
in angles are called shearing strains.

The normal strain is defined as: L


 lim  Eq. (3.13)
L 0 L

Where ΔL is the change in length of an element whose original


length was L before deformation took place.

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:

(OA) O ' A'  OA


xx  
OA OA
OA  dx
1/2

O A  (dx  qx , x dx)  (q y , x dx)  (qz , x dx) 


' ' 2 2 2

The normal strain in the x direction is then:

1/2

xx  (1  qx , x )  (q y , x )  (qz , x ) 


2 2 2
1

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.

Therefore, the x-direction linear normal strain becomes

xx  qx, x Or Eq. (3.14a)

Elasticity Of Structures 21
3.4 Strains And Strain-Displacement Relationships

In similar manner, the linearized normal strains in the y and z


directions may be derived and are given by

yy  qy , y Or Eq. (3.14b)

zz  qz , z Or Eq. (3.14c)

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

In the case of two-dimensional state of strain (plane-strain


problem), εzz = εzx = εzy = 0 , Eq. (3.16) reduces to

 xx xy 
     Eq. (3.17)
 yx yy 

Elasticity Of Structures 25
3.4 Strains And Strain-Displacement Relationships

In cylindrical coordinates, Eqs. (3.14) and (3,15) may be written


as

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

Where qr , qѳ , and qz are the displacements in the r , ѳ , and z directions,


respectively.

Elasticity Of Structures 26
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems

The strain-displacement relationships for plane-strain problems


are given by the following.
xx  qx , x
yy  q y , y Eq. (3.19)
xy  qx , y  q y , x

By examining Eq. (3.19), it is apparent that there exist three


components of strain which are expressed in terms of only two
components of displacements.

Thus, it may be concluded that not all these strain components


are independent of one another.
Elasticity Of Structures 27
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems

This may be easily verified by differentiating twice the first and


the second equations of (3.19) with respect to y and x,
respectively, and the last equation with respect to x only.
xx , yy  qx , xyy
yy , xx  q y , yxx
xy , xy  qx , xyy  q y , xxy

Substituting the first two equations into the third yields

xy , xy xx, yy  yy , xx Or Eq. (3.20)

Which is the compatibility equation of deformation.

Elasticity Of Structures 28
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems

Thus, the choice of the three strain components cannot be


arbitrary, but must be such that the compatibility equation is
satisfied.

 This compatibility equation ensures the existence of single-


valued displacement functions qx and qy for a solid.

 In Fact, when Eq. (3,20) is expressed in terms of stresses, it


ensures the existence of a unique solution for a stress problem.

Elasticity Of Structures 29
3.5 Compatibility Equations For Plane-Stress And Plane-Strain Problems

For three-dimensional state of stress, the compatibility


equations may be derived in similar manner to that of Eq. (3.20)
and are given by the following.

xy , xy xx , yy  yy , xx


xz , xz xx , zz  zz , xx
yz , yz yy , zz  zz , yy
2 xx , yz xy , xz  xz , xy  yz , xx
Eq. (3.21)
2 yy , xz xy , yz  xz , yy  yz , xy
2 zz , xy xy , zz  xz , yz  yz , xz

Elasticity Of Structures 30
3.6 Boundary Conditions

Boundary conditions are those conditions for which the


displacements and/or the surface forces are prescribed at the
boundary of a given solid.

For instance, if the displacements qx ,qy ,and qz are prescribed at


the boundary, then conditions are referred to as displacement
boundary conditions and may be written as

qx  q x ( ) qy  q y ( ) qz  q z ( ) Eq. (3.22)

where 𝑞 𝑥, 𝑞 𝑦, and 𝑞 𝑧 are known functions of displacements at


the boundary.

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

Structural behavior may be classified into three basic categories


based on the functional relationship between stresses and strains:
1. Inelastic nonlinear
2. Elastic nonlinear
3. Linear elastic

Elasticity Of Structures 33
3.7 Stress-Strain Relationships

Consider a steadily loaded bar as shown in Fig. 3.8.

Figure 3.8

If upon loading the functional


relationship between the stress and its
corresponding strain takes on a curved
path, and upon unloading it takes a
different curved path, as in Fig. 3.9a then
the structural behavior is referred to as
inelastic nonlinear behavior.
Figure 3.9a
Elasticity Of Structures 34
3.7 Stress-Strain Relationships

If such relationship follows the same


curved path upon both loading and
unloading the bar as in Fig. 3.9b, then
the behavior is said to be elastic
nonlinear behavior.

Figure 3.9b

Last, if such relationship takes on the


same straight path as in Fig. 3.9c, then
the behavior is termed linear elastic
behavior.

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 )

xx   a11 a12 a13 a14 a15 a16   xx 


     
 yy   a21 a22 a23 a24 a25 a26   yy 
zz   a31 a32 a33 a34 a35 a36   zz  Eq. (3.24)
   
yz   a41 a42 a43 a44 a45 a46   yz 
xz   a51 a52 a53 a54 a55 a56   xz 
    
xy   a61 a62 a63 a64 a65 a66   xy 
Elasticity Of Structures 36
3.7 Stress-Strain Relationships

If there exist three orthogonal planes of elastic symmetry


through every point of the solid body, then Eq. (3.24) becomes:

xx   a11 a12 a13 0 0 0   xx 


     
a a
 yy   21 22 23a 0 0 0   yy 
zz   a31 a32 a33 0 0 0   zz 
   
 yz   0 0 0 a44 0 0   yz 

xz   0 0 0 0 a55 0   xz 
    
xy   0 0 0 0 0 a66   xy 
Eq. (3.25)

Elasticity Of Structures 37
3.7 Stress-Strain Relationships

The elastic constants aij in Eq. (3.25) may be defined in terms of


the engineering-material constants as follows:
1 1 1
a11  a22  a33 
Exx E yy Ezz
v yx vxy
a12  a21   
Exx E yy
vzy v yz
a23  a32   
E yy Ezz Eq. (3.26)
vxz v
a13  a31     zx
Ezz Exx
1 1 1
a44  a55  a66 
Gyz Gxz Gxy

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.

Plywood, for instance, and most reinforced plastics may be


designated as orthotropic materials.

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 )

where E=Young's modulus, G=shear modulus, and v=Poisson's ratio.


Most metals, such as aluminum, steel, and titanium, are isotropic materials.

Elasticity Of Structures 40
3.7 Stress-Strain Relationships

It is important to note here that in an anisotropic body, normal


strains will induce not only normal stresses but also shearing
stresses; likewise, shearing strains will produce normal stresses, as
may easily be seen from Eq. (3.24).

However, normal strains in an isotropic or orthotropic body will


cause only normal stresses, while shearing strains will cause
shearing stresses.

This may be verified by examining Eq. (3.25).

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

 xx, xx  2 xy , xy   yy , yy  0 Eq. (3.29)

The body forces are assumed to be zero.


Elasticity Of Structures 42
3.7 Stress-Strain Relationships

Statically indeterminate systems require the use of some, if not all,


of the other equations involving strain–displacement and stress–strain
relationships.
However, whether the system be statically determinate or not,
stress–strain relationships are necessary to determine deflections.
Experiments show that the application of a uniform direct stress, σx,
does not produce any shear distortion of the material and that the
direct strain εx is given by the Equation

Eq. (3.30a)

where E is a constant known as the modulus of elasticity or Young’s modulus.


Elasticity Of Structures 43
3.7 Stress-Strain Relationships

The above Equation is an expression of Hooke’s law. Further, εx is


accompanied by lateral Strains
Eq. (3.30b)

in which ν is a constant termed Poisson’s ratio.

For a body subjected to direct stresses σx, σy and σz the direct


strains are

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

from which we may rewrite in terms of τ as

The term E/2(1+ν) is a constant known as the modulus of rigidity


G. Hence
Eq. (3.34a)

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

Strain Equations provide the additional six equations required to


determine the 15 unknowns in a general three-dimensional problem
in elasticity. They are, however, limited in use to a linearly elastic
isotropic body. For the case of plane stress they simplify to

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

Neglecting products of small quantities in the expansion of the


right-hand side of the above equation yields

Eq. (3.36)

Elasticity Of Structures 49
3.7 Stress-Strain Relationships

Substituting for εx, εy ,and εz we find, for a linearly elastic,


isotropic body

In the case of a uniform hydrostatic pressure, σx =σy =σz=−p and


Eq. (3.37)

The constant E/3(1−2ν) is known as the bulk modulus or modulus of volume


expansion and is often given the symbol K.

Elasticity Of Structures 50
3.7 Stress-Strain Relationships

An examination of the above equation shows that ν≤0.5 since a


body cannot increase in volume under pressure. Also the lateral
dimensions of a body subjected to uniaxial tension cannot increase
so that ν>0.
Therefore, for an isotropic material 0≤ν≤0.5 and for most isotropic
materials ν is in the range 0.25–0.33 below the elastic limit. Above
the limit of proportionality ν increases and approaches 0.5.

Elasticity Of Structures 51
3.8 Transformations Of Stresses And Strains

A stress field (σxx,σyy,σxy), such as shown in Fig. 3.10, which is


known in one set of systems axes may be transformed to any
other arbitrary set of axes such as ɳβ.

Figure 3.10

Elasticity Of Structures 52
3.8 Transformations Of Stresses And Strains

For instance, assume that at a given point in a solid, the


stresses (σxx,σyy, and σxy) are known in reference to the x and
y axes, as shown in Fig. 3.11a.
 The object is to find the set of stresses in reference to a
new set of axes ɳβ , which are rotated through an angle θ as
shown.

Figure 3.11a

Elasticity Of Structures 53
3.8 Transformations Of Stresses And Strains

The stress σɳɳ may be found by considering the free-body


diagram which is cut by a plane along the β axis at an angle θ
from the vertical, as shown in Fig. 3.11b.
If side OB is assumed to have area A, then sides CS and CB
will each have an area of Acosθ and Asinθ, respectively.

Figure 3.11b

Elasticity Of Structures 54
3.8 Transformations Of Stresses And Strains

Since equilibrium conditions prevail, the summation of forces


along each of the ɳ and β directions must be zero, or

 F  0 ]
 A   xx A cos cos   yy sin  sin    xy cos sin 
 yx A sin  cos   0
By simplifying and noting that σxy = σyx the following is
obtained: 𝜎𝜂𝜂 = 𝜎𝑥𝑥 𝑐𝑜𝑠 2 𝜃 + 𝜎𝑦𝑦 𝑠𝑖𝑛2 𝜃 − 𝜎𝑥𝑦 𝑠𝑖𝑛2𝜃 Eq. (3.38a)

 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

In a similar manner σββ may he obtained and is given by

    xx sin 2    yy cos 2    xy sin 2 Eq. (3.38c)

Equations (3.38a),(3.38b), and (3.38c) may be assembled in


a matrix as
 
    cos 2  sin 
2
 sin 2   xx 
     Eq. (3.39a)

     sin 2  cos 2
sin 2   yy 
    sin 2 sin 2   xy 
   1  2 cos   
2 
 2 2 

• Where θ is the angle of rotation and is positive in the clockwise direction.

Elasticity Of Structures 56
3.8 Transformations Of Stresses And Strains

In compact matrix form Eq.(3.28) becomes

   ,
 T   x, y
Eq. (3.39b)

Where [T] is referred to as the transformation matrix.


In similar manner, the strains may be transformed as
follows:    
    xx 
   
    T   yy 
Eq. (3.40)
1  1 
      xy 
2  2 
Where γij (tensor shearing strain) = 2 εij (engineering shearing
strain) and the matrix [T] is the same as that in Eq. (3.39).
Elasticity Of Structures 57
3.8 Transformations Of Stresses And Strains
Example 1
A cylindrical pressure vessel has an internal diameter of 2m and is fabricated from plates
20mm thick. If the pressure inside the vessel is 1.5 N/mm2 and, in addition, the vessel is
subjected to an axial tensile load of 2500 kN, calculate the direct and shear stresses on a
plane inclined at an angle of 60◦ (ACW) to the axis of the vessel. Calculate also the
maximum shear stress.

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

Note that θ =-30◦ and τxy =0.

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:

Finally: Eq. (3.43)

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

The state of stress at a point in a deformable body may be determined graphically by


Mohr’s circle of stress. The direct and shear stresses on an inclined plane were shown to be
given by

And

The positive directions of these stresses and the angle θ are defined in the figure below
Finally:

which represents the equation of a circle of radius


and having its centre at the point ((σx +σy)/2, 0).

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)

In addition the maximum shear strain at the point is

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).

The principal stresses are now obtained by substitution of εI and εII.


Thus
Eq. (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:

At a point in an elastic material there are two mutually perpendicular


planes, one of which carries a direct tensile stress at 50 N/mm2 and a
shear stress of 40 N/mm2, while the other plane is subjected to a direct
compressive stress of 35 N/mm2 and a complementary shear stress of
40 N/mm2.
-Determine the principal stresses at the point, the position of the planes
on which they act and the position of the planes on which there is no
normal stress.
Elasticity Of Structures 79
Problems
Prob. 3.3:
Listed below are varying combinations of stresses acting at a point and
referred to axes x and y in an elastic material. Using Mohr’s circle of stress
determine the principal stresses at the point and their directions for each
combination.

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

David J. Peery & J.J. Azar: AIRCRAFT STRUCTURES, 2nd


Edition

Chapter 3: Elasticity Of Structures

 Aircraft Structures for engineering students, Fourth


Edition United States 2007, T.H.G. Megson.

 Chapter 1: Basic elasticity

Elasticity Of Structures 81
AERO 234: Fundamentals of Aircraft
Structures

Chapter 4: Behavior And


Evaluation of Vehicle Material

Dr. Issam Tawk


In this Chapter

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

Behavior & Evaluation of vehicle materials 2


4.1 Introduction
 It is of utmost importance for the structural analyst to have
a full understanding of the behavior of vehicle materials and
be able to intelligently evaluate and select the material best
suited to the constraints and operational requirements of the
design.

The materials used in various parts of vehicle structures


generally are selected by different criteria. The criteria are
predicated on the constraints and operational requirements
of the vehicle and its various structural members.

Behavior & Evaluation of vehicle materials 3


4.1 Introduction
 Some of these more important requirements involve
1. Environment
2. Fatigue
3. Temperature
4. Corrosion
5. Creep
6. Strength and stiffness
7. Weight limitation
8. Cost
9. Human factor

Behavior & Evaluation of vehicle materials 4


4.2 Mechanical Properties Of Materials
Materials, in general, may be classified according to their
constituent composition as single-phase or multiphase.
 All metals, such as aluminum, steel, and titanium, are
referred to as single-phase materials.
 All composites, which are made out of filaments (fibers)
embedded in a matrix (binder), are called multiphase
materials. Plywood and reinforced fiber glass are
examples of multiphase materials.

Almost all important structural properties of single-or


multiphase materials are obtained by three basic tests: tensile
test, compression test, and shear test. The American Society
for Testing Materials (ASTM) sets all the specifications and test
procedures for materials testing.
Behavior & Evaluation of vehicle materials 5
4.2 Mechanical Properties Of Materials
Tensile Test
Figure 4.1 shows the basic configuration of a tensile test
specimen. The load P is applied gradually through the use of a
tensile testing machine.

Figure 4.1

Behavior & Evaluation of vehicle materials 6


4.2 Mechanical Properties Of Materials
The normal strain єn usually is measured either by utilizing
electrical strain gage techniques or by measuring the total
elongation δ in an effective gage length L for various values of
the tension load P.
For small loads, the elongation is assumed to be uniform
over the entire gage length and therefore the normal strain
may be mathematically expressed in the form

n  Eq. (4.1)
L

Where δ and L are both measured in the same units of length.

Behavior & Evaluation of vehicle materials 7


4.2 Mechanical Properties Of Materials
The corresponding normal stress σn is also assumed to be
uniformly distributed over the cross-sectional area A of the lest
specimen and is obtained as follows:
P
n  Eq. (4.2)
A
For common engineering units, the load P is in pounds, the area
A is in square inches, and the stress σ is in pounds per square
inch.

Behavior & Evaluation of vehicle materials 8


4.2 Mechanical Properties Of Materials
 The stress-strain diagram for
a material is obtained by
plotting values of the stress σ
against corresponding values of
Figure 4.2a
the strain є, as shown in Fig.
4.2a.

 For small values of the stress,


the stress-strain curve is a
straight line, as shown by line
OA of Fig. 4.2b Figure 4.2b

Behavior & Evaluation of vehicle materials 9


4.2 Mechanical Properties Of Materials
The constant ratio of stress to strain for this portion of the
curve is called the modulus of elasticity E, as defined in the
following equation:
n
E Eq. (4.3)
n
Where E has units of pounds per square inch.

Behavior & Evaluation of vehicle materials 10


4.2 Mechanical Properties Of Materials
A material such as plain low-carbon steel, which is commonly
used for bridge and building structures, has a stress-strain
diagram such as that shown in Fig. 4.2a.
At point B, the elongation increases with no increase in load.
This stress at this point is called the yield point, or yield stress,
σty and is very easy to detect when such materials are tested.

Figure 4.2a

Behavior & Evaluation of vehicle materials 11


4.2 Mechanical Properties Of Materials
The stress at point A, where the stress-strain curve first
deviates from a straight line, is called the proportional limit σtp
and is much more difficult to measure while a test is being
conducted.
Specifications for structural steel usually are based on the
yield stress rather than the proportional limit, because of the
ease in obtaining this value.

Flight-vehicle structures are made of materials such as


aluminum alloys, high-carbon steels, and composites which do
not have a definite yield point, but which do exhibit stress-
strain behavior similar to that shown in Fig. 4.2b.

Behavior & Evaluation of vehicle materials 12


4.2 Mechanical Properties Of Materials
It is convenient to specify arbitrarily the yield stress for
such materials as the stress at which a permanent strain of
0.002 in/in is obtained.

 Point B of Fig. 4.2b represents this yield stress and is


obtained by drawing line BD parallel to OA through point D,
representing zero stress and 0.002-in/in strain, as shown.

Figure 4.2b

Behavior & Evaluation of vehicle materials 13


4.2 Mechanical Properties Of Materials
When the load is removed from a test specimen which has
passed the proportional limit, the specimen does not return to
its original length, but retains a permanent strain.

For the material represented by Fig. 4.2a, the load might be


removed gradually at point C. The stress-strain curve would then
follow line CD, parallel to OA, until at point D a permanent strain
equal to OD were obtained for no stress.

Figure 4.2a

Behavior & Evaluation of vehicle materials 14


4.2 Mechanical Properties Of Materials
Upon a subsequent application of load, the stress-strain
curve would follow lines DC and CG. Similarly, if the specimen
represented by Fig. 4.2b were unloaded at point B, the stress-
strain curve would follow line BD until a permanent strain of
0.002 in/in were obtained for no stress.

The actual failure occurs at point H, but the maximum


apparent stress, represented by point G, is the more
important stress to use in design calculations. This value
is defined as the ultimate strength σtu.

Behavior & Evaluation of vehicle materials 15


4.2 Mechanical Properties Of Materials
In the design of tension members for vehicle structures. It is
accurate to employ the initial area of the member and the
apparent ultimate tensile strength σtu.

In using the stress-strain curve to calculate the ultimate


bending strength of beams, as shown in a later chapter, the
results are slightly conservative because the beams do not neck
down in the same manner as tension members.

Behavior & Evaluation of vehicle materials 16


4.2 Mechanical Properties Of Materials
Compression Test
The compressive strength of materials is more difficult to
identify from stress-strain curves than the corresponding tensile
strength.

Compressive failures for most structural designs in


engineering applications are associated with instabilities which
are related to yield stress rather than ultimate stress. Yield-
stress values which are obtained based on the 0.2 percent offset
method (0.002-in/in permanent strain) have been proved to be
relatively successful for correlating instabilities in most metals;
however, the correlation is less satisfactory for non-metals,
specifically composites.
Behavior & Evaluation of vehicle materials 17
4.2 Mechanical Properties Of Materials
Compression Test
 Specimen geometry and means of supports seem to have
considerable effect on compression test results. The
compressive stress-strain diagrams for most materials are
similar to the tensile stress-strain curves.

Behavior & Evaluation of vehicle materials 18


4.2 Mechanical Properties Of Materials
Shear Test
In plane shear properties are the most difficult to obtain and
have the least standardized testing procedures of all major
mechanical properties.
 The design of a shear-test specimen having a test section
subjected to a uniform shearing stress is impossible, the
closest practical approach is probably a thin-walled circular
cylinder loaded in torsion, as shown in Fig. 4.3.

Figure 4.3

Behavior & Evaluation of vehicle materials 19


4.2 Mechanical Properties Of Materials
Shear Test
For small displacements, the shear strain is expressed as

S
s  Eq. (4.4) Where єs=shear strain
L ΔS= change in arc length
L =effective gage length

The shear stress may be calculated from the well-known


strength of material torsional equation as
Tr
s  Eq. (4.5)
J
where σs= shear stress
r= cylinder radius
J = cross-sectional polar moment of inertia
Behavior & Evaluation of vehicle materials 20
4.2 Mechanical Properties Of Materials
For very thin-walled cylindrical test specimens, Eq. (4.5) may
be written as
TRm
s  Eq. (4.6)
Jm
where Rm= mean radius =(R0+Ri)/2 and Jm=2πRm3t.

The shear modulus of elasticity may be obtained from the


shear stress–shear strain diagram

s
G Eq. (4.7)
s

Behavior & Evaluation of vehicle materials 21


4.3 Equations for Stress-Strain Curve Idealization
In the design of vehicle structural members, it is necessary
to consider the properties of the stress-strain curve at stresses
higher than the elastic limit.

In other types of structural and machine designs, it is


customary to consider only stresses below the elastic limit; but
weight considerations are so important in flight vehicle design
that it is necessary to calculate the ultimate strength of each
member and to provide the same factor of safety against
failure for each part of the entire structure.

Behavior & Evaluation of vehicle materials 22


4.3 Equations for Stress-Strain Curve Idealization

The ultimate bending or compressive strengths of many


members are difficult to calculate, and it is necessary to obtain
information from destruction tests of complete members.

In order to apply the results on tests of members of one


material to similar members of another material, it is desirable
to obtain an idealized analytical expression for the stress-strain
diagrams of various materials.

Behavior & Evaluation of vehicle materials 23


4.4 Fatigue
Fatigue is a dynamic phenomenon which may be defined as
the initiation and propagation of microcracks into macrocracks
as a result of repeated applications of stresses.

It is a process of localized progressive structural fracture in


material under the action of dynamic stresses.

A structure which may not even fail under a single


application of load may very easily fail under the same load if it
is applied repeatedly.

This failure under repeated application of loads is termed


fatigue failure.
Behavior & Evaluation of vehicle materials 24
4.4 Fatigue
In spite of the many studies and vast amount of experimental
data accumulated over the years, fatigue is still the most
common cause of failure in machinery and various structures as
well as the least understood of all other structural behavior.

This lack of understanding is attributed to the fact that the


initiation and the propagation of microscopic cracking are
inherently statistical in nature.

The main objective of all fatigue analyses and testing is the


prediction of fatigue life of a given structure or machine part
subjected to repeated loading.

Behavior & Evaluation of vehicle materials 25


4.4 Fatigue
Such loads may have constant amplitude, as indicated in Fig.
4.4; however, in flight vehicle structures, the load history is
usually random in nature, as shown in Fig. 4.5.

Figure 4.4

Figure 4.5

Behavior & Evaluation of vehicle materials 26


4.4 Fatigue
Fatigue-life prediction: Most of the fatigue-life prediction
methods used in the design of structures have been based on
fatigue allowable data generated by sine wave excitation.
Only recently closed-loop, servo-controlled hydraulic
machines have become available for true random loading
testing. Fatigue test data usually are presented graphically (Fig.
4.6) and the curves are referred to as the allowable S-N curves.

Figure 4.6

Behavior & Evaluation of vehicle materials 27


4.4 Fatigue
Among the several theories proposed for fatigue-life
prediction, the Palmgren-Miner theory, because of its
simplicity, seems to be the most widely used.
The method hypothesizes that the useful life expended may
be expressed as the ratio of the number of applied cycles ηi to
the number of cycles Ni to failure at a given constant stress
level σi.
When the sum of all the fractions reaches 1, failure should
occur. Mathematically this failure criterion is written as:
 1   2   r 
      .....    1
 N1 1 const  N 2  2 const  N r  r const
r
 i 
   1 Eq. (4.8)
Or i 1  N i   const
i

Behavior & Evaluation of vehicle materials 28


4.4 Fatigue
It is important to note that in Palmgren-Miner theory no
provisions are made to take into account the various effects on
fatigue life, such as notch sensitivity effect, loading sequence
effect (high-low or low-high), and the consequences of
different levels of mean stress.

In fact, the theory has been shown to yield unconservative


results in some test case studies conducted by Gassner,
Kowalewski, and Corten and Dolan.

Behavior & Evaluation of vehicle materials 29


4.4 Fatigue
To illustrate the use of Eq. (4.8), consider a bracket which
supports an electronic box in the aircraft cockpit. In a typical
mission, the bracket encounters a stress history spectrum
idealized as shown in Fig. 4.7.
The fatigue allowable of the bracket material is given in Fig.
4.8. The problem is to find the number of missions the aircraft
may accomplish before the bracket fails.

Figure 4.8
Figure 4.7

Behavior & Evaluation of vehicle materials 30


4.4 Fatigue
From Figs. 4.7 and 4.8. Table 4.1 may be easily constructed.
Therefore, in one mission 0.433 percent of the useful life of the
bracket is expended. This means that the aircraft might
accomplish about 200 missions without bracket failure, based
on Eq. (4.6).

Table 4.1

Behavior & Evaluation of vehicle materials 31


4.4 Fatigue
Fatigue test data : S-N curves Fatigue tests are conducted for
a wide variety of reasons. One of which is to establish materials'
fatigue allowable, or what is commonly referred to as the S-N
curves.
Contrary to results of static tests, it has been observed that
the scatter in fatigue test results can be quite large.
This inherent scatter characteristic leaves no choice but to
treat the results statistically. One of the more widely used
statistical distribution function is the log-normal distribution,
whose mean value is taken as
1 n
M   log Ni Eq. (4.9)
n i 1
where n= total number of specimens tested at the same stress
level and Ni = numberBehavior
of cycles to failure for specimen.
& Evaluation of vehicle materials 32
4.4 Fatigue
The unbiased standard deviation of M is defined by
t/2
 n 2 
 (log N i  M ) 
  i 1

Eq. (4.10)
 n  1 
 

In order to calculate the number of cycles to failure based on


some confidence level, the standard variable ξ is taken as
Eq. (4.11)
log N  M
 

Or
log N  M   Eq. (4.12)

Behavior & Evaluation of vehicle materials 33


4.4 Fatigue
The probability of surviving log N cycles is


1

 2 /2
probability (log N )  P(log N )  e d Eq. (4.13)
2 

Equation (4.13) may be used to tabulate the probabilities of


survival for various values of the standard variable ξ, as shown
in Table 4.2.

Behavior & Evaluation of vehicle materials 34


4.4 Fatigue
To illustrate the procedure, consider the following actual
fatigue test results for seven tested specimens:

 From Eq. (4.9) the mean value is:


1 n 1 7
M   log Ni   log Ni
n i 1 7 i 1
1
 (log 63,318  log39,695  ....  log96,500)
7
 4.76463
From Eq. (4.10) the unbiased standard deviation is
1/ 2
 7 (log Ni  M ) 2 
   
 i 1 6 
t /2
 (log 96,500  4.76463)  2
 (log 61,318  4.76463) 2  .....  
 6 
 0.0503
Behavior & Evaluation of vehicle materials 35
4.4 Fatigue
From Eqs. (4.12), if we assume a probability survival of 95
percent

log N  M    4.76463  (1.645)(0.0503)  4.68189

Therefore the number or cycles to failure for a 95 percent


probability or survival is

N = antilog 4.68189= 48,100 cycles

Behavior & Evaluation of vehicle materials 36


4.5 Strength-Weight Comparisons of Materials
The criterion commonly used in the selection of structural
materials for aerospace vehicle application is that which yields
minimum weight.
This involves selecting the proper combination of material
and structural proportions with the weight as the objective
function to be minimized to yield an optimum design.

Although weight comparisons of materials may be based on


several factors, such as resistance to corrosion, fatigue
behavior, creep characteristics, strength, and so on, the only
treatment given here is with respect to strength.

Behavior & Evaluation of vehicle materials 37


4.5 Strength-Weight Comparisons of Materials
As an illustration, let us consider the three loaded members
shown in Fig 4.9 , for simplicity, it is assumed that there exists
only one free variable (the thickness t, in this case) to be
chosen in the design.
The criteria which govern the design of members in Fig. 4.9
a, b, and c are ultimate uniaxial tension, ultimate uniaxial
compression or buckling, and ultimate bending.

Figure 4.9

Behavior & Evaluation of vehicle materials 38


4.5 Strength-Weight Comparisons of Materials
The expressions relating the applied external loads to the
induced actual stresses are
Tension: P
t  ( A  bt ) Eq. (4.17)
A
Compression (buckling):
 2 EI  bt 3 
c  I   Eq. (4.18)
AL 2
 12 

Mt
Bending: b  Eq. (4.19)
2I

Where σt , σc , σb = ultimate tensile, compression, and flexural


stresses.
A= cross-sectional area
I=moment of inertia of member
Behavior & Evaluation of vehicle materials 39
4.5 Strength-Weight Comparisons of Materials
The weight of the member may be expressed in terms of the
material density (ρ pounds per cubic inch) and the fixed and
free geometric dimensions as
W  Lbt  Eq. (4.20)

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

W (compression) Eq. (4.22)


  E 
1/2
 6Mb  Eq. (4.23)
W  L   (bending)

 b 
Behavior & Evaluation of vehicle materials 40
4.5 Strength-Weight Comparisons of Materials
With Eqs (4.21) to (4.23) available, weight comparisons of
different materials may be conducted. Thus the weights of
two different materials required to carry the axial load P may
be readily obtained from Eq.(4.21) as
PL 1 PL2
W1  W2 
 t1 t2
Where the subscripts 1 and 2 refer to materials 1 and 2,
respectively, and σt1 and σt1 are the ultimate tensile stresses of
materials 1 and 2 respectively.
W1 1  t 2
 Eq. (4.24)
W2  2  t1

Behavior & Evaluation of vehicle materials 41


4.5 Strength-Weight Comparisons of Materials
Similarly, the ratio of weights of two members of two
different materials resisting the same bending condition may
be easily obtained by utilizing Eq (4.23):
1/2
W1 1   b 2 
   Eq. (4.25)
W2  2   b1 

Likewise, the ratio of weights for two members of two


different materials resisting the same compressive (buckling)
load may be written immediately by using Eq. (4.22):
1/3
W1 1  E2 
   Eq. (4.26)
W2  2  E1 
Behavior & Evaluation of vehicle materials 42
4.5 Strength-Weight Comparisons of Materials
Typical aerospace vehicle sheet materials are compared in
Table 4.3 by means of Eqs (4.24) through (4.26) the weights of
the various materials are compared with the aluminum alloy
2024-T3. The weight ratios for tension members, shown in
column 5 do not vary greatly for the different materials. For
members in bending, however, the lower-density materials
have a distinct advantage, as shown in column 6.

Table 4.3

Behavior & Evaluation of vehicle materials 43


4.5 Strength-Weight Comparisons of Materials
Similarly, the lower-density materials have an even greater
advantage in compression buckling, as indicated in column 7.

Values of σ vary with sheet thickness, and those shown are


used only for comparison.

The computations of Table 4.3 indicate that the last three


materials, having lower densities, are superior to the aluminum
alloys.

However, it is important to note that magnesium alloys are


more subject to corrosion than aluminum alloys, while wood
and plastic materials are less ductile.
Behavior & Evaluation of vehicle materials 44
4.5 Strength-Weight Comparisons of Materials
 Brittle materials are undesirable for structures with
numerous bolted connections and cutouts which produce
local high-stress concentrations. Ductile materials, which
have a large unit elongation at the ultimate tensile
strength, will yield slightly at points of high local stress and
will thus relieve stress. Whereas brittle materials may fail
under the same conditions.

 Fiber-reinforced plastics have been used successfully for


aerospace vehicle structures as long ago as the late 1940s and
early 1950s. In those days, the main reinforcement was glass
fiber in fabric form with polyester resin as the bonding agent.

Behavior & Evaluation of vehicle materials 45


4.5 Strength-Weight Comparisons of Materials
Since then and primarily in the last few years, development
of new high-modulus fibers in combination with high-modulus,
high-temperature-resistant resins has added a new dimension
to materials for applications in aerospace and marine and land-
based structures.
These new fibers and resins aye being combined in a
unidirectional, pre-impregnated form which gives the analyst
complete freedom to tailor the composite (a composite is
made of a certain number of unidirectional plies) to meet the
imposed load requirements in both magnitude and direction.
Studies have indicated that through the use of composite
materials, the total weight of an aerospace vehicle could be
reduced by more than 35 percent.
Behavior & Evaluation of vehicle materials 46
4.6 Sandwich Construction

The problem of increasing weight which accompanies


increasing material thickness is being met frequently by the use
of sandwich construction in application to aerospace vehicles.
This type of construction consists of thin, outer- and inner-
facing layers of high-density material separated by a low-
density, thick core material.

Behavior & Evaluation of vehicle materials 47


4.6 Sandwich Construction

In aerospace applications, depending on the specific mission


requirements of the vehicle, the material of the sandwich
facings may be reinforced composites, titanium, aluminum,
steel, etc.

Several types of core shapes and core materials may be


utilized in the construction of the sandwich.

The most popular core has been the "honeycomb" core,


which consists of very thin foils in the form of hexagonal cells
perpendicular to the facings.

Behavior & Evaluation of vehicle materials 48


4.6 Sandwich Construction
Although the concept of sandwich construction is not new,
only in the last decade has it gained great prominence in the
construction or practically all aerospace vehicles, including
missiles, boosters, and spacecraft.

This is primarily a result of the high structural efficiency that


can be developed with sandwich construction.

 Other advantages offered by sandwich construction are its


excellent vibration characteristics, superior insulating qualities,
and design flexibility.

Behavior & Evaluation of vehicle materials 49


4.6 Sandwich Construction
An element of a sandwich beam is shown in Fig. 4.10. For
simplicity, the facings are assumed to have equal thickness tf
and the core thickness is tc .

Figure 4.10

It is also assumed that the core carries no longitudinal normal


stress σ.
Behavior & Evaluation of vehicle materials 50
4.6 Sandwich Construction
Let us consider the case where it is desired to find the
optimum facing thickness which results in a minimum weight
of the sandwich beam carrying a bending moment M.

Without any loss in generality, b and L may be taken as unit


values. And thus the resisting bending moment can be
expressed as

M   t f (t f  tc )

Behavior & Evaluation of vehicle materials 51


4.6 Sandwich Construction
If we assume that tf is small compared to tc, which is normally
the case, then the above equation reduces to
M   t f tc
Or
M   tc 2 Eq. (4.27)

where tf is expressed in terms of tc by the equation tf=βtc.

The weight of a unit element of the beam is approximately


W   c tc  2  f  t c Eq. (4.28)

Where ρc and ρf equal the core and facing densities, respectively.


Eliminating the variable tc from Eqs. (4.27) and (4.28) yields
M
W  ( c  2  f  ) Eq. (4.29)

Behavior & Evaluation of vehicle materials 52
4.6 Sandwich Construction
The value of β for the minimum weight may be obtained by
differentiating Eq.(4.29) with respect to β equating the
derivative to zero. Performing the differentiation and solving for
β yield c
 Eq. (4.30)
2 f

Equation (4.30) indicates that for a sandwich material


resisting bending moment, the minimum weight is obtained
when the two layers of the face material have approximately
the same total weight as the core.
 Note that this condition does not yield minimum weight if the
beam element is under the action, of compressive (buckling)
load.

Behavior & Evaluation of vehicle materials 53


4.6 Sandwich Construction

It is now possible to compare the weight of a sandwich-


construction beam element with that of a solid element
corresponding to the sandwich face material, if we assume that
they both resist the same loads.

A sandwich element designed to resist bending moments will


have a total weight equal to twice the total weight of the
facings, if the face and core materials have equal weights.
Thus, from Eq. (4.28)

W  4  t Eq. (4.31)

Behavior & Evaluation of vehicle materials 54


4.6 Sandwich Construction
By solving, for t from Eq.(4.27) and substituting into Eq.(4.31),
the following weight for the sandwich element is obtained:
M
W  4  Eq. (4.32)


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

Behavior & Evaluation of vehicle materials 56


4.6 Sandwich Construction

Thus, the sandwich has only 37 percent of the weight of a


solid element resisting the same bending moment.

 Also note that the value of 0.37 is less than any of the other
values in column 6 of Table 4.3.

Behavior & Evaluation of vehicle materials 57


4.6 Sandwich Construction
In the preceding discussion, it was assumed that the
proportions for the sandwich element were limited only by
theoretical considerations.
In actual structures, practical considerations are much more
important.

The thickness of the face material, for example, usually is


greater than the theoretical value, because it might not be
feasible to manufacture and form very thin sheets.
Likewise, the core was assumed to support the facings
sufficiently to develop the same unit stress as in a solid
element, whereas the actual low-density materials might not
provide such support.
Behavior & Evaluation of vehicle materials 58
4.7 Typical Design Data For Materials
In the manufacture of materials, it is not possible to obtain
exactly the same structural properties for all specimens of a
material.

 In a large number of tested specimens of the same material, the


ultimate strength may vary as much as 10 percent.

 In the design of an aerospace vehicle structure, therefore, it is


necessary to use stresses which are the minimum values that may
be obtained in any specimen of the material.

Behavior & Evaluation of vehicle materials 59


4.7 Typical Design Data For Materials

These values are termed the minimum guaranteed values of the


manufacturer.

 The licensing and procuring agencies specify the minimum


values to be used in the design of aerospace vehicles.

These values are contained in Military Handbooks, such as MIL-


HDBK-5A, MIL-HDBK-5, MIL-HDBK-17, MIL-HDBK-2B.

Table 4.4 shows the typical mechanical data required in the


design of aerospace structures.

Behavior & Evaluation of vehicle materials 60


4.7 Typical Design Data For Materials

Table 4.4

Behavior & Evaluation of vehicle materials 61


4.7 Typical Design Data For Materials

Normally these data are presented in accordance with one of


the following bases:

A basis: At least 99 percent of all mechanical property values


are expected to fall above the specified property values with a
confidence of 95 percent.
B basis: At least 90 percent of all mechanical property values are
expected to fall above the specified property values with a
confidence of 95 percent.
S basis: Minimum mechanical property values as specified by
various agencies.

Behavior & Evaluation of vehicle materials 62


4.8 Problems

4.1 The buckling load for a sandwich column is approximately given


by :𝑃 = 𝜋 2 𝐸𝐼/𝐿2 . Find the thickness ratio of facing to the core which
results in an optimum design (minimum weight) for the column. See
Fig. P 4.1.

Behavior & Evaluation of vehicle materials 63


4.8 Problems
4.2 Find the weight ratio of a sandwich column to that of a solid column
whose material is the same as that of the sandwich facings.

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)

Behavior & Evaluation of vehicle materials 64


4.8 Problems
4.4 A missile-holding fixture on an aircraft is subject during each flight
to the stress-load history shown in Fig. P4.4. After how many flights
will the fixture fail if the material fatigue allowable is that shown in
Fig. 4.8?

Figure 4.8

Behavior & Evaluation of vehicle materials 65


4.8 Problems
4.5 Ten parts were fatigue-tested at the same stress level, and the
following failure cycles were reported:

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

David J. Peery & J.J. Azar: AIRCRAFT STRUCTURES, 2nd


Edition

Chapter 4: Behavior & Evaluation of vehicle materials

Behavior & Evaluation of vehicle materials 67


AERO 234: Fundamentals of Aircraft
Structures

Chapter 5: Stress Analysis

Dr. Issam Tawk


In this Chapter

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

Chapter 5:Stress Analysis 2


5.1 Introduction
In order to select sizes of structural members to meet the design
load requirements on a specific aerospace vehicle, it is necessary to
find the unit stresses acting on the cross section of each structural
element.

Chapter 5:Stress Analysis 3


5.1 Introduction
It was shown in Chap. 3 that there exist two distinct components
of stress, normal and shear stress. A normal stress is a unit stress
which acts normal to the cross section on the structural element,
while the shear stress is parallel and in the plane of the cross
section.
 A normal stress is induced by bending moments and axial forces.
A shear stress, however, is caused by torsional moments and
shear forces.

This chapter discusses the theory and the application of these two
fundamental stress components.

Chapter 5:Stress Analysis 4


5.2 Force-Stress Relationships
The stress field at any chosen point in a solid beam may be
entirely defined by the components of force resultants or stresses
acting along the directions of some “gaussian" coordinate system,
as shown in Fig. 5.1. The forces and stresses are taken to be
positive if they act in the positive direction of the corresponding
coordinate axis.

Figure 5.1:Stress and force resultant

Chapter 5:Stress Analysis 5


5.2 Force-Stress Relationships
From Fig. 5.1 the force resultants may be related to the stresses
as follows:
P    xx dA M z   y xx dA
A A

Vy    xy dA M y   z xx dA Eq. (5.1)
A
A

Vz    xz dA T   ( y xz  z xy )dA
A A

where P = axial force


Vy,Vz = shear force
Mz, My = bending moments
T = torque
σxx = normal stress
σxy, σxz = shearing stresses
Chapter 5:Stress Analysis 6
5.3 Normal Stress in Beams
The normal stresses in beam elements are induced by bending
and/or extensional actions. Two approaches may be used to
determine stresses; the first is based on the theory of elasticity, and
the second is based on strength of materials theory.

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.

Chapter 5:Stress Analysis 7


5.3 Normal Stress in Beams
Assume that the displacement in the x direction of any point in the
beam is represented by qx (x, y, z). If we take ux(s) to be the
extensional displacement of any point on the beam axis (y = z= 0) and
ψz and ψy to be the rotational displacements of the beam cross
section, then
q x  x, y, z   u x  x   y z  x   z y  x  Eq. (5.2)

The axial strain from Eq. (3.14a) is defined by


є xx  q x, x

Chapter 5:Stress Analysis 8


5.3 Normal Stress in Beams
Hence, From Eq.(4.2)
єxx  u x,x  y z,x  z y,x Eq. (5.3)

At any given cross section x =x0


dux ( x0 )
 const  B1
dx
d z ( x0 ) d y ( x0 )
 const  B2  const  B3
dx dx

where Eq. (5.3) becomes

є xx  B1  B2 y  B3 z Eq. (5.4)

Chapter 5:Stress Analysis 9


5.3 Normal Stress in Beams
In order to determine the stresses which correspond to the strains
in Eq. (5.4), the stress-strain relationship in Chap. 3 is utilized. By
assuming that the stresses σzz , and σyy are negligible compared to σxx
the following relationship for an isotropic material may he obtained
easily from Eq. (3.25):

 xx  E єxx Eq. (5.5)

where E=modulus of elasticity of the material.

Chapter 5:Stress Analysis 10


5.3 Normal Stress in Beams
Substituting Eq. (5.4) into Eq. (5.5) yields
 xx  E (B1  B2 y  B3 z) Eq. (5.6)

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 z    Ey( 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

where A = cross-sectional area

12
Chapter 5:Stress Analysis
5.3 Normal Stress in Beams

I z   y 2 dA  moment of inertia of cross section about x axis Eq. (5.8a)


A

I y   z 2 dA  moment of inertia of cross section about y axis Eq. (5.8b)


A

I yz   yz dA  product moment of inertia of cross section Eq. (5.8c)


A

y   ydA
A

z   zdA Eq. (5.9)


A

Chapter 5:Stress Analysis 13


5.3 Normal Stress in Beams
If the z and y axes are taken through the geometric centroid of the
cross section, then 𝑦 and 𝑧 become identically zero. Hence Eq. (5.7)
reduces to

P
 B1 A
E
M z
 I z B2  I yz B3 Eq. (5.10)
E

My
 I yz B2  I y B3
E

Chapter 5:Stress Analysis 14


5.3 Normal Stress in Beams
Solving Eqs. (5.10) for the unknown constants yields

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 )

Chapter 5:Stress Analysis 15


5.3 Normal Stress in Beams
Substituting Eqs.(5.11) into Eq.(5.6) yields the general expression
of the normal stress:
P I y M z  I yz M y I z M y  I yz M z
 xx   y z Eq. (5.12)
A I y I z  I yz 2
I y I z  I yz 2

When Eq. (5.12) is used, it is important to observe the sign


convention used in the derivation. See Fig. 5.1. In cases where y and z
axes are principal axes of the cross-sectional area, the product of the
moment of inertia Iyz about these axes is zero. For this condition, Eq
(5.12) reduces to
P Mz My
 xx   y z Eq. (5.13)
A Iz Iy

Chapter 5:Stress Analysis 16


5.3 Normal Stress in Beams
If there is no axial force acting on the beam and bending occurs
about the z axis only, then Eq. (5.13) reduces to the familiar strength-
of-material pure bending equation

M z y
 xx  Eq. (5.14)
Iz

Chapter 5:Stress Analysis 17


5.4 Shear Stresses In Beams
The shear stresses in beams are induced by pure shear Force
action and/or torsional action. In this section, only shear stresses
due to shear forces are considered, while the latter are dealt with in
a separate section.
Consider a small section of a beam, as shown in Fig. 5.2. For
simplicity, assume that the beam cross section is symmetrical and
the theory of strength of materials holds.

Figure 5.2:Beam element


Chapter 5:Stress Analysis
18
5.4 Shear Stresses In Beams
The shear force Vy parallel to the beam cross section produces
shear stresses σxy of varying intensity over the cross-sectional area.

Corresponding to the vertical shear stress σxy, there exists a shear


stress σyx in the xz plane which is equal to σyx at the points of
intersection of the two planes.

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.

 For equilibrium of the horizontal forces, the force produced by the


shear stress σyx on the horizontal area of width t and length 𝛿x must
be equal to the difference in the normal forces on the two cross
sections. Summing forces in the horizontal direction yields
c
Vy y
 yx t  
Eq. (5.15)

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.

It is important to note that Eq. (5.16) is applicable only to beams of


uniform, symmetrical cross sections. Tapered beams and beams of
unsymmetrical cross section are considered later.

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 y1
 yx  ydA  7.5  2310lb / in 2

43.3  3

The average shear stress just below this point, where t = 1 in is


c
Vy 40, 000
I z t y1
 yx  ydA  7.5  6930lb / in 2

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

 ydA  2.5  3  1.5 1  9.0


y1

This shear stress at this point is


c
Vy 40, 000
 yx   ydA  9.0  8320lb / in 2
I z t y1 43.3 1

At a point on the neutral axis of the beam, the shear mess is

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

σ yx = 8320/1 = 8320 lb/in2 at 2 in below top of beam

σ yx = 8780/1 = 8780 lb/in2 at neutral axis

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

Iz= 2(0.5)(62)+ 2(0.5(22) = 40 in4

If two cross sections 1 in apart are considered, the difference in


bending stresses Vyy/Iz on the two cross sections will be 8(6/40) = 1.2
kips/in2 on the outside stringers and 8(2/40)= 0.4 kips/in2 on the
inside stringers.
The differences in axial, loads on the stringers at the two cross
sections are found as the product of these stresses and the stringer
areas and are shown in Fig. 5.7c.
33
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
The shear stress in the web at a point between the upper two
stringers is found from the equilibrium of spanwise forces on
the upper stringer.
σyx(0.04)(1) = 600
Or
σyx =15,000 lb/in2

If the webs resist no bending stress, the shear stress will be


constant along each web, as shown in Fig. 5.7d. If the webs
resist bending stresses, the shear stress in each web will vary
along the length of the web and will be greater at the end
nearer the neutral axis.

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:

σyx(0.04)(1)= 600 + 200


Or
σyx=20,000lb/in2

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 y1
 yx  ydA  (0.5  6)  15, 000lb / in 2

40  0.040

For a point between the two middle stringers.

c
Vy 8000
I z t y1
 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.

Figure 5.8 a Figure 5.8 b


38
Chapter 5:Stress Analysis
5.4 Shear Stresses In Beams
Solution From Eq.(5.12) with P set to zero, the normal stress for
the beam in Fig. 5.8a is
I y M z  I yz M y I z M y  I yz M z
 xx   y z
I y I z  I yz 2
I y I z  I yz 2

173.3(100,000)  240(10,000) 693.3(10,000)  240(100,000)


 y  z
(693.3)(173.3)  240 2
(693.3)(173.3)  240 2

 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.

From Mohr's circle for a condition of pure shear, it may be shown


that the diagonal compressive stress σc , and the diagonal tensile
stress σt are both equal to the maximum shear stress 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:

Fy   qdy  qy Eq. (5.18)


0

Where y is the vertical distance between the ends of the web.

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.

The area dA of the triangle formed by


joining point O and the extremities of the
element of length ds is r ds/2.
Figure 5.11-a 46
Chapter 5:Stress Analysis
5.5 Shear Flow In Thin Webs
Then the torque induced by the shear flow along the entire web
may be obtained as follows:
T   qrds   2qdA  2q  dA
s A A Eq. (5.19)
Or
T  2Aq
where A is the area enclosed by the web and the lines joining the
ends of the web with point O, as shown hi Fig. 5.11b.

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.

The shear flow on a rectangular element must form two equal


and opposite couples. The directions of all shear flows are shown in
Fig.5.13b and the back section in Fig. 5.13a.

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

The shear center will be on a horizontal axis of symmetry, since


a horizontal force along this axis will produce no twisting of the
beam.

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 

— 16,000e  2 39.27  2000  6000 10  2 1000  6   0


Or
e  14.32 in

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.

On an aircraft wing, for example, the resultant aerodynamic load


is farther forward on the wing at high angles of attack than at low
angles of attack. The position of this load also changes when the
ailerons or wing flaps are deflected. Thus a closed-section box beam,
which is capable of resisting torsion, is used for aircraft wings and
similar structures.
64
Chapter 5:Stress Analysis
5.7 Torsion Of Closed-Section Box Beams
Typical types of wing construction are shown in Fig. 5.15. The wing
section of Fig. 5.15a has only one spar, and the skin forward of this
spar forms a closed section which is designed to resist the wing
torsion, whereas the portion aft of the spar is lighter and is designed
not to resist any loads on the wing but to act as an aerodynamic
surface.

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.

The constant shear flow q around the circumference has no


resultant horizontal or vertical force, since in the application of
Eqs.(5.11) and (5.18) the horizontal and vertical distances between
the endpoints of the closed web are zero.

The resultant of the shear flow is thus a torque equal to the


applied external torque T, taken about any axis perpendicular to
the cross section.

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):

T   2(A)q  2 Aq Eq. (5.21)

where A is the sum of the triangular areas ΔA and is equal to the


total area enclosed by the box section.

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.

 If point O is chosen outside the section some of the triangular


areas ΔA will be negative, corresponding to the direction of the
moment of the shear flow about point O, but the algebraic sum of all
areas ΔA will be equal to the enclosed area A.

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

The shear flows in box beams with several stringers may be


obtained by a method similar to that previously used. From a
summation of spanwise loads on various stringers the shear flows
may all be expressed in terms of one unknown shear flow. Then this
shear flow may be obtained by equating the moments of the shear
flows to the external torsional moment about a spanwise axis.

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

Where  pn represents the summation of loads ΔP between 0


0
and any web n. 77
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
After all the shear flows are expressed in terms of the unknown
qo, the value of qo may be obtained from the summation of the
torsional moments. Note that the shear flow in any other web could
have been considered as the unknown qo.

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.

The value of qo is obtained by considering the equilibrium of


torsional moments due to the shear flow and the external
applied loads in reference to any axis normal to the back cross
section of the beam.

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 

-10, 000  8    q 0 -500 100    q 0 -250 100   q 0  278.5   0


Or q  324lb / in
0

The shear flow in the rest of the webs may be computed


easily from Fig. 5.19c.

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.

Figure 5.20: Box beam

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

I yz =1 5  10   3  5  10   110  5   3  5 10   200 in 4


Vz  4kips
Vy  10kips
 I yVy  I yzVz I zVz  I yzVy 
p1    y z  Af Eq. (5.25)
 I I  I2 I I  I 2 
 y z yz y z yz 
84
Chapter 5:Stress Analysis
5.8 Shear Flow In Closed-Section Box Beams
The substitution in Eq (5.25) yields:
P   23.33z  73.33y  Af
With the above equation, the ΔP on each stringer may be
obtained easily by making the proper substitution for the flange
area and its corresponding coordinates. The results are shown
in Fig. 5.21a.

_ _

_ 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 .

The shear flow in the left-hand web is designated q0. The


shear flows in the rest of the webs are obtained by considering
the equilibrium of forces in the spanwise direction and are
given in Fig. 5.21a.

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.

While this variation in cross section may not cause


appreciable errors in the application of the flexure formula
for bending stresses, often it causes large errors in the shear
stresses determined from Eq.(5.16).
c
Vy
 yx 
It
z
 ydA
y1
Eq. (5.16)

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)

V f  P  tan 1  tan  2  Eq. (5.26b)

From the geometry of the beam, tanα1=h1/C, tanα2=h2/C, and


tanα1+ tanα2=(h1+ h2)/C=h/C.

Substituting this value into Eq. (5.26b) yields


h
Vf  p Eq. (5.27)
C
91
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
Equation (5.27) will apply for a beam with any system of
vertical loads. For the present loading, the value of P is Vyb/h.
Substituting this value for P into Eq. (5,27) yields
b
Vf  V Eq. (5.28)
C
From Eqs. (5.26a) and (5.28) and from the geometry,
a
Vw  V Eq. (5.29)
C
Equations (5.28) and (5.29) can be expressed in terms of the
depths h0 and h of the beam by making use of the proportion
a/c=h0/h: Vy h0
Vw  Eq. (5.30a)
h
and h  h0
Vf  V Eq. (5.30b)
h 92
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
The shear flow in the webs now can he found by using Eq.
(5.16) in conjunction with the shear Vw in Eq. (5.30a). For
instance, if we assume that the areas of both stringers in
Fig.5.22 are the same, the shear flow at a section where the
distance between the stringers is h may be calculated as
follows:
Vw Vw Ah Vw
I 
q ydA   Eq. (5.31)
Ah 2 2 2 h
When the beam has several stringers, the shear flow may be
obtained in a manner similar to that for the two-stringer beam
as long as the stringer areas remain constant along the span. If
the stringer areas vary along the span and not all vary in the
same proportion, Eq. (5.16) cannot be applied.
93
Chapter 5:Stress Analysis
5.9 Spanwise Taper Effect
Example 5.8: Find the shear flows in the web of the beam
shown in Fig. 5.23 at 20-in intervals along the span.

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

The bending stresses at section AA are


M z y 400, 000  5
 xx    10, 000 lb in 2
I 200

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 .

If the areas of all the stringer members are increased by a


constant ratio, the method of Sec. 5.9 can be used; if the areas
at one cross section are not proportional to the areas at
another cross section, the method would be considerably in
error.

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.

Since it is necessary to determine the wing bending stresses at


frequent stations along the wing span in order to design the
stringers, the terms Pa and Pb can be obtained without too many
additional calculations. Thus this method of analysis is often
simpler and more accurate than the method which considered
variations in depth but not variations in stringer area.
108
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The distance a between two cross sections may be any
convenient value. It is common practice to calculate wing
bending stresses at intervals of 15 to 30 in along the span. The
intervals are quite satisfactory for shear-flow calculations.

Note that for very small values of a, small percentage errors in


Pa and Pb ,result in large percentage errors in ΔP. However, if a is
too large, the average shear flow obtained between two
sections may not be quite the same as the shear flow midway
between the sections.

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.

The differences in axial loads are tabulated in Column 5, and


the shear flows q=(Pb-Pa)/20 are shown in column 6. The value
of the shear flow at station 20 thus is assumed to be equal to
the average horizontal shear between stations 10 and 30.

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

The bending stresses resulting from the bending moment of


480,000 in.lb
M z y 480,000(5.5)
 xx    10,909 lb in2
Iz 242
The loads on the stringers are 10.909 and 21,818 lb, as shown in
Fig. 5.31a. 117
Chapter 5:Stress Analysis
5.10 Beams With Variable Stringer Areas
The increments or flange loads ΔP in a 1-in length are found
from Eq. (5.32).

For the area of 2 in2

21,818  17,777
P   202lb
20

For the area of 1.0 in2

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.

Problem 5.17:Repeat Prob. 5.16 if there is an additional chordwise


load or 6000 Lb acting to the left at the center of the tip cross
section.

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.

Problem 5.19: A horizontal beam with a square cross section resists


vertical loads. Find the angle of the neutral axis with the horizontal
if one side of the beam makes an angle θ with the horizontal. At
what angle should the beam be placed for the bending stress to
have a minimum value?

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

Problem 5.24: Repeat Prob. 5.23, using the differences in


stringer loads shown in Fig. P5.24 at the cross section for x=40
and x=60 in.
140
Chapter 5:Stress Analysis
Problems
Problem 5.25: Calculate the shear flows in the webs of the cross
section shown in Fig. P.5.25 at x= 50 in. Assume the flange areas as
follows :
a) a=b=3 in2, c=d=1in2
b) a=c=1 in2, b=d=3in2 Figure P5.25
c) a=c=3 in2, b=d=1in2

Consider only the one cross section, and calculate the in-plane
components of the flange loads.

Problem 5.26: Repeat Prob. 5.25. using the differences in flange


loads at the cross section shown in Fig. P5.26 for x=40 in and 60in.
Use a torsional axis joining the centroids of the cross sections.

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

David J. Peery & J.J. Azar: AIRCRAFT STRUCTURES,


2nd Edition

Chapter 5: Stress Analysis

Chapter 5:Stress Analysis 143


p
ρ= (9)
RT

2. International Standard Atmosphere (ISA) Table [2]

The International Standard Atmosphere parameters (temperature, pressure, density) can


be provided as a function of the altitude under a tabulated form, as given in Table 3:

Table 3 International Standard Atmosphere [2]

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy