Вы находитесь на странице: 1из 88

Axial Loading

JU. Dr. Ibrahim Abu-Alshaikh


Deformations Under Axial Loading
• From Hooke’s Law:

P
  E  
E AE
• From the definition of strain:


L
• Equating and solving for the deformation,
PL
 
AE
• With variations in loading, cross-section
or material properties,
L
Pi Li P( x)
    dx
i Ai Ei 0
EA( x)
Example
E  29 10  6 psi
D  1.07 in. d  0.618 in.

• Apply a free-body analysis on each component to determine the


internal force

SOLUTION:
• Divide the rod into components at the load application points.
• Evaluate the total of the component deflections.

Determine the deformation of the steel rod shown under the


given loads.
SOLUTION: Divide • Apply free-body analysis to each
the rod into three component to determine internal
components: forces,
P1  60 103 lb, P2  15 103 lb
P3  30 103 lb
• Evaluate total deflection,

Pi Li 1  P1L1 P2 L2 P3 L3 
      
A
i i iE E  1A A 2 A 3 


1      
 60  103 12  15  103 12 30  103 16 
 
6 
29  10  0. 9 0.9 0.3 

 75.9  10 3 in.

L1  L2  12 in. L3  16 in.   75.9  10 3 in.


A1  A2  0.9 in 2 A3  0.3 in 2
Sample Problem SOLUTION:
• Apply a free-body
analysis to the bar
BDE to find the forces
exerted by links AB
and DC.
• Evaluate the
The rigid bar BDE is supported by two links deformation of links AB
AB and CD. and DC or the
Link AB is made of aluminum (E = 70 GPa) displacements of B and
and has a cross-sectional area of 500 mm2. D.
Link CD is made of steel (E = 200 GPa) and
has a cross-sectional area of (600 mm2). • Work out the geometry
to find the deflection at
For the 30-kN force shown, determine the E given the deflections
deflection a) of B, b) of D, and c) of E. at B and D.
Sample Problem Displacement of B:
PL
B 
SOLUTION: AE

Free body: Bar BDE  60  103 N 0.3 m 



500 10-6 m2 70 109 Pa 
 514  10  6 m

 B  0.514 mm 

Displacement of D:
MB  0
PL
D 
AE
0  30 kN  0.6 m   FCD  0.2 m 90  103 N 0.4 m 

FCD  90 kN tension



600 10-6 m2 200 109 Pa 
 300  10  6 m
 MD  0
0  30 kN  0.4 m   FAB  0.2 m
FAB  60 kN compression  D  0.300 mm 
Sample Problem Displacement of D:

BB BH

DD HD
0.514 mm 200 mm   x

0.300 mm x
x  73.7 mm

EE  HE

DD HD
E

400  73.7 mm
0.300 mm 73.7 mm
 E  1.928 mm

 E  1.928 mm 
Static Indeterminacy
• Structures for which internal forces and reactions
cannot be determined from statics alone are said
to be statically indeterminate.
• A structure will be statically indeterminate
whenever it is held by more supports than are
required to maintain its equilibrium.
• Redundant reactions are replaced with
unknown loads which along with the other
loads must produce compatible deformations.

• Deformations due to actual loads and


redundant reactions are determined
separately and then added or superposed.

  L R  0
Example
Determine the reactions at A and B for the steel
bar and loading shown, assuming a close fit at
both supports before the loads are applied.
SOLUTION:
• Consider the reaction at B as redundant,
release the bar from that support, and solve
for the displacement at B due to the applied
loads.
• Solve for the displacement at B due to the
redundant reaction at B.
• Require that the displacements due to the
loads and due to the redundant reaction be
compatible, i.e., require that their sum be
zero.
• Solve for the reaction at A due to applied loads and the reaction found
at B.
Example SOLUTION:
• Solve for the displacement at B due to the
applied loads with the redundant constraint
released,
P1  0 P2  P3  600  103 N P4  900  103 N

A1  A2  400  10  6 m 2 A3  A4  250  10  6 m 2
L1  L2  L3  L4  0.150 m

Pi Li 1.125  109
L   
i Ai Ei E

• Solve for the displacement at B due to the


redundant constraint,
P1  P2   RB

A1  400 10  6 m 2 A2  250  10  6 m 2


L1  L2  0.300 m

δR  
Pi Li


1.95  103 RB 
A
i i iE E
Example
• Require that the displacements due to the loads
and due to the redundant reaction be compatible,

  L R  0

  

1.125  109 1.95  103 RB 
0
E E
RB  577  103 N  577 kN

• Find the reaction at A due to the loads and the


reaction at B

 Fy  0  R A  300 kN  600 kN  577 kN


R A  323 kN
R A  323 kN
RB  577 kN
Example
A horizontal rigid bar ABC is pinned at end A and supported by two
wires (BD and CD) at points B and C . A vertical load P acts at end C
of the bar. The bar has a length of 2b and wires BD and CD have
lengths of L1 and L2, respectively. Also, wire BD has a diameter of d1
and modulus of elasticity E1; wire CD has a diameter of d2 and
modulus E2.
(a) Obtain formulas for the allowable load P if the allowable
stresses in wires BD and CD, respectively, are σ1 and σ2. (Disregard
the weight of the bar and cables.)
(b) Calculate the allowable load P for the following conditions:
Wire BD is made of aluminum with a modulus E1=72 GPa and a
diameter of d1=4.2 mm. Wire CD is made of magnesium with a
modulus E2 = 45 Gpa and a diameter of d2 = 3.2 mm. The allowable
stresses in the aluminum and magnesium wires are σ1 = 200 MPa and
σ2 = 172 MPa, respectively. Dimensions are a = 1.8 m and b = 1.2 m
JU. Dr. Ibrahim Abu-Alshaikh
Thermal Stresses
• A temperature change results in a change in
length or thermal strain. There is no stress
associated with the thermal strain unless the
elongation is restrained by the supports.

• Treat the additional support as redundant


and apply the principle of superposition.
PL
 T   T L P 
AE
  thermal expansion coef.
• The thermal deformation and the
deformation from the redundant support
must be compatible.   T   P  0
  T   P  0 P   AE T 
P
  E T 
PL
 T L  0 
AE A
Example SOLUTION:
• Solve for the displacement at B due to the
applied loads with the redundant constraint
released,
P1  0 P2  P3  600  103 N P4  900  103 N

A1  A2  400  10  6 m 2 A3  A4  250  10  6 m 2
L1  L2  L3  L4  0.150 m

Pi Li 1.125  109
L   
i Ai Ei E
• Solve for the displacement at B due to the
redundant constraint,
P1  P2   RB

A1  400 10  6 m 2 A2  250  10  6 m 2


L1  L2  0.300 m

δR  
Pi Li


1.95  103 RB 
i Ai Ei E
Example
• Require that the displacements due to the loads
and due to the redundant reaction be compatible,
  L R  0

  

1.125  109 1.95  103 RB0
E E
RB  577  103 N  577 kN
• Find the reaction at A due to the loads and the
reaction at B

 Fy  0  R A  300 kN  600 kN  577 kN


R A  323 kN

R A  323 kN
RB  577 kN
Example
A sleeve in the form of a circular tube of length L is placed around a
bolt and fitted between washers at each end. The nut is then turned
until it is just snug. The sleeve and bolt are made of different
materials and have different cross-sectional areas. (Assume that the
coefficient of thermal expansion αS of the sleeve is greater than the
coefficient αB of the bolt.)
(a) If the temperature of the entire assembly is raised by an amount
ΔT, what stresses σS and σB are developed in the sleeve and bolt,
respectively?
(b) What is the increase δ in the length L of the sleeve and bolt?
JU. Dr. Ibrahim Abu-Alshaikh
Bolts and Turnbuckles
Pres-tressing a structure requires that one or more parts of the
structure be stretched or compressed from their theoretical lengths. A
simple way to produce a change in length is to tighten a bolt or a
turnbuckle. In the case of a bolt each turn of the nut will cause the
nut to travel along the bolt a distance equal to the spacing p of the
threads (called the pitch of the threads). Thus, the distance traveled
by the nut is
  np
in which n is the number of
revolutions of the nut (not
necessarily an integer). Depending
upon how the structure is arranged,
turning the nut can stretch or
compress a member.
Turnbuckles
In the case of a double-acting turnbuckle, there are two end screws.
Because a right-hand thread is used at one end and a left-hand thread
at the other, the device either lengthens or shortens when the buckle
is rotated. Each full turn of the buckle causes it to travel a distance p
along each screw, where again p is the pitch of the threads.
Therefore, if the turnbuckle is tightened by one turn, the screws are
drawn closer together by a distance 2p and the effect is to shorten the

  2np
device by 2p. For n turns, we have

Double-acting turnbuckle (Each full turn of the turnbuckle


shortens or lengthens the cable by 2p, where p is the pitch of the
screw threads.)
Example
The mechanical assembly shown in Figure consists of a copper tube, a
rigid end plate, and two steel cables with turnbuckles. The slack is
removed from the cables by rotating the turnbuckles until the assembly
is snug but with no initial stresses. (Further tightening of the
turnbuckles will produce a pre-stressed condition in which the cables
are in tension and the tube is in compression.)
a) Determine the forces in the tube and cables when the turnbuckles
are tightened by n turns.
b) Determine the shortening of the tube.
STRAIN ENERGY
Strain energy is a fundamental concept in applied mechanics, and strain
energy principles are widely used for determining the response of
machines and structures to both static and dynamic loads. In this section
we introduce the subject of strain energy in its simplest form by
considering only axially loaded members subjected to static loads.
More complicated structural elements will be discussed in later
chapters— like: bars in torsion, and beams in bending etc.
Consider a prismatic bar of length L
subjected to a tensile force P which
is applied slowly (static load). The
bar gradually elongates as the load is
applied, eventually reaching its
maximum elongation δ at the same
time that the load reaches its full
value P. Thereafter, the load and
elongation remain unchanged.
To evaluate work (in J where 1 J= 1 Nm), we recall from elementary
mechanics that a constant force does work equal to the product of
the force and the distance through which it moves.
Let us denote P1 by any value
where 0 < P1 < P , and its
corresponding elongation of
the bar by δ1, 0< δ1 < δ . Then
an increment dP1 in the load
will produce an increment dδ1
in the elongation. The work
done by the load during this
incremental elongation is 
W   P1 d1
0

In geometric terms, the work done by the load is equal to the area
below the load-displacement curve.
The strain energy, is defined as the energy absorbed by the bar
during the loading process. From the principle of conservation of
energy, we know that this strain energy is equal to the work done by
the load provided no energy is added or subtracted in the form of heat.
Therefore, 
U  W   P1 d1
0

in which U is the symbol for strain


energy. Sometimes strain energy is
referred to as internal work to
distinguish it from the external
work done by the load.

When the load is removed, the load-displacement diagram follows


line BD if point B is beyond the elastic limit, and a permanent
elongation OD remains.
Linearly Elastic Behavior
Let us now assume that the material of
the bar follows Hooke’s law, so that the
load-displacement curve is a straight
line. Then the strain energy U stored in
the bar (equal to the work W done by
the load) is P
U W 
2
which is the area of the shaded triangle OAB in the figure, the
previous relation may be expressed as

P 2 L EA 2 P 2 k 2
U W    
2 EA 2L 2k 2
When replacing the stiffness EA/L of the prismatic bar by the stiffness
k of the spring.
Non-uniform Bars
The total strain energy U of a bar
consisting of several segments is
equal to the sum of the strain
energies of the individual segments.
That is
n
Pi 2 Li
n
U  U i  
i 1 i 1 2 Ei Ai

in which Ui is the strain energy of


segment i of the bar and n is the number
of segments and we assume that the
material of the bar is linearly elastic and
that the internal axial force is constant
within each segment.
We can further obtain the strain energy of a non-prismatic bar with
continuously varying axial force by using the differential element
(shown shaded in the figure) and then integrating along the length of
the bar

 P( x)
L 2

U  dx
0
2 EA( x)

In this equation, N(x) and A(x) are


the axial force and cross-sectional
area at distance x from the end of
the bar.
Strain-Energy Density
In many situations it is convenient to use a quantity called strain-
energy density (J/m3), denoted by u defined as the strain energy per
unit volume of material, that is
P2 L  1  P2 E 2 E 2  2
u    2  
2 EA  AL  2 EA 2
2L 2 2E
The strain-energy density of the material when it is stressed to the
proportional limit is called the modulus of resilience ur. It is found by
substituting the proportional limit σpl into the last equation
 pl
2

ur 
2E
Another quantity, called toughness, refers to the ability of a material to
absorb energy without fracturing. The corresponding modulus, called
the modulus of toughness ut, is the strain-energy density when the
material is stressed to the point of failure. It is equal to the area below
the entire stress-strain curve. The higher the modulus of toughness, the
greater the ability of the material to absorb energy without failing.
Example
Three round bars having the
same length L but different
shapes are shown. The first bar
has diameter d over its entire
length, the second has
diameter d over one-fifth of its
length, and the third has
diameter d over one-fifteenth
of its length. Elsewhere, the
second and third bars have
diameter 2d.
All three bars are subjected to the same axial load P. Compare the
amounts of strain energy stored in the bars, assuming linearly elastic
behavior. (Disregard the effects of stress concentrations and the
weights of the bars.)
a) Strain energy U1 of the first bar P2 L 2P2 L
U1  
2 EA E d 2
b) Strain energy U2 of the second bar. The strain energy is found by
summing the strain energies in the three segments of the bar
3
Pi 2 Li 2P2 L 2P 2 L 4P 2 L 2U1
U2      
i 1 2 Ei Ai 5E d 2
5E d 2
5E d 2
5
which is only 40% of the strain energy of the first bar. Thus,
increasing the cross-sectional area over part of the length has greatly
reduced the amount of strain energy that can be stored in the bar.
(c) The strain energy U3 of the third bar is
3
Pi 2 Li 2P2 L 7P2 L 9P2 L 3U1
U2      
i 1 2 Ei Ai 15E d 15E d
2 2
15E d 2
10
The strain energy has now decreased to 30% of the strain energy of the
first bar.
Example

Determine the strain energy


of a prismatic bar
suspended from its upper
end. Consider the following
loads:
(a) the weight of the bar
itself, and
(b) the weight of the bar
plus a load P at the
lower end. (Assume
linearly elastic
behavior.)
(a) Strain energy due to the weight of the bar itself . The bar is
subjected to a varying axial force, the internal force being zero at the
lower end and maximum at the upper end. Thus, the internal axial
force N(x) acting on this element is equal to the weight of the bar
below the element:
N ( x)   A( L  x)

 A( L  x)
2 L
A    
L 2 L 2 3 2 3
A ( L x ) A L
U  dx   ( L  x) dx 
2
 
0
2 EA 2E 0
2E 3  0 6E
(b) Strain energy due to the weight of the bar plus the load P, however
in this case the axial force N(x) acting on the element is
N ( x)   A( L  x)  P
By integration we can write
A 2 L3  PL2 P 2 L
U  
6E 2E 2 EA
Example
The cylinder for a compressed air machine is clamped by bolts that pass
through the flanges of the cylinder. A detail of one of the bolts is shown
in the figure. The diameter d of the shank is 13 mm and the root
diameter dr of the threaded portion is 10 mm. The grip g of the bolts is
40 mm and the threads extend a distance t = 6.5 mm into the grip.
Under the action of repeated cycles of high and low pressure in the
chamber, the bolts may eventually break. To reduce the likelihood of
the bolts failing, the designers suggest two possible modifications:
(1) Machine down the shanks of the bolts so that the shank diameter is
the same as the thread diameter dr , as shown.
(2) Replace each pair of bolts by a single long bolt. The long bolts are
similar to the original bolts except that the grip is increased to the
distance L=340 mm. Compare the energy-absorbing capacity of the
three bolt configurations: (a) original bolts, (b) bolts with reduced shank
diameter, and (c) long bolts. (Assume linearly elastic behavior and
disregard the effects of stress concentrations)
shank diameter (d) >
thread diameter (dr)

shank diameter (d) =


thread diameter (dr)
long bolt
(a) Original bolts. The original bolts can be idealized as bars
consisting of two segments. The left-hand segment has length (g-
t) and diameter d, and the right-hand segment has length t and
diameter dr. The strain energy of one bolt under a tensile load P
can be obtained by adding the strain energies of the two
segments
P (g  t) P t
2 2
U1  
2 EAs 2 EAr
In which As is the cross-sectional area of the shank and Ar is the
cross-sectional area at the root of the threads, thus,
d2  dr 2
As  , Ar 
4 4
Substituting these expressions into U, we get the following formula
for the strain energy of one of the original bolts:
2 P 2 ( g  t ) 2 P 2t
U1  
d E 2
 dr 2 E
(b) Bolts with reduced shank diameter. These bolts can be idealized
as prismatic bars having length g and diameter dr . Therefore, the
strain energy of one bolt is
2P2 g
U2 
 dr 2 E
The ratio of the strain energies for cases (1) and (2) is
U2 g
  1.52
U1 2  (g  t) t 
dr  2
 2
 d dr 

Thus, using bolts with reduced shank diameters results in a 52%


increase in the amount of strain energy that can be absorbed by the
bolts. If implemented, this scheme should reduce the number of
failures caused by the impact loads.
( c) Long bolts. The calculations for the long bolts are the same as for
the original bolts except the grip g is changed to the grip L. Therefore,
the strain energy of one long bolt is P 2 (L t ) P 2t
U3  
2 EAs 2 EAr
Since one long bolt replaces two of the original bolts, we must
compare the strain energies by taking the ratio of U3 to 2U1, as
 (L  t ) t 
 2
 2
U3
  d dr 
 3.87
2U1  (g  t) t 
2 2
 2
 d dr 
Thus, using long bolts increases the energy-absorbing capacity by
287% and achieves the greatest safety from the standpoint of strain
energy. Note: When designing bolts, designers must also consider the
maximum tensile stresses, maximum bearing stresses, stress
concentrations, and many other matters.
IMPACT LOADING
Loads can be classified as static or dynamic depending upon whether
they remain constant or vary with time. A static load is applied
slowly, so that it causes no vibrational or dynamic effects in the
structure. A dynamic load may take many forms—some loads are
applied and removed suddenly (impact loads), others persist for long
periods of time and continuously vary in intensity (fluctuating loads).
Impact loads are produced when two objects collide or when a falling
object strikes a structure. Fluctuating loads are produced by rotating
machinery, traffic, wind gusts, water waves, earthquakes, and
manufacturing processes.
The potential energy of the falling collar with respect to the elevation
of the flange is Mgh, where g=9.81 is the acceleration of gravity. This
potential energy is converted into kinetic energy as the collar falls. At
the instant the collar strikes the flange, its potential energy with respect
to the elevation of the flange is zero and its kinetic energy is Mv2/2,
where is its velocity.
During the ensuing impact, the kinetic energy of the collar is
transformed into other forms of energy. Part of the kinetic energy is
transformed into the strain energy of the stretched bar. Some of the
energy is dissipated in the production of heat and in causing localized
plastic deformations of the collar and flange.
A small part
remains as the
kinetic
energy of the
collar, which
either moves
further
downward
(while in contact
with the flange)
or else bounces
upward.
To make a simplified analysis of this very complex situation, we will
idealize the behavior by making the following assumptions:

1. The collar does not rebound after the collar sticks to the flange.
2. We disregard all energy losses and assume that the kinetic energy
of the falling mass is transformed entirely into strain energy of the
bar.
3. We disregard any change in the potential energy of the bar itself
(due to the vertical movement of elements of the bar)
4. We assume that the stresses in the bar remain within the linearly
elastic range.
5. We assume that the stress distribution throughout the bar is the
same as when the bar is loaded statically by a force at the lower
end, that is, we assume the stresses are uniform throughout the
volume of the bar. (In reality longitudinal stress waves will travel
through the bar, thereby causing distribution)
JU. Dr. Ibrahim Abu-Alshaikh
Maximum Elongation of the Bar
The maximum elongation δmax can be obtained from the principle of
conservation of energy by equating the potential energy lost by the
falling mass to the maximum strain energy acquired by the bar. The
potential energy lost is W(h + δmax), where W = Mg is the weight of
the collar and (h + δmax ) is the distance through which it moves. The
strain energy of the bar is EA 2
max

2L
where EA is the axial rigidity and L is the length of the bar. Thus, we
obtain the following equation:
EA max
2
W (h   max ) 
2L
This equation is quadratic in δmax and can be solved for the positive
root; the result is 1/2
WL  WL   WL  
2

 max      2h  
EA  EA   EA  
The preceding equation can be written in simpler form by
introducing the notation WL MgL
 st  
EA EA
in which δst is the elongation of the bar due to the weight of the collar
under static loading conditions. Thus, δmax becomes
1/2   2h 1/2 
 max   st   st   2h st    st 1  1   
2
     st  
 
When the height h is large compared to the static elongation (h>> δst ),
we can obtain
2
Mv L
 max  2h st  , where v  2 gh
EA
in which M = W/g
Maximum Stress in the Bar
The maximum stress can be calculated easily from the maximum
elongation, hence we assume that the stress distribution is uniform
throughout the length of the bar, we get .
PL  L 1/2  max L
    max   st   st   2h st
  
2

EA E   E
E 1/2
   2 hE 
1/2

 max    st   st   2h st      st   st    st  
2 2

L      L  
 
  2 Eh 1/2 
 max   st 1  1   
  L st  
 
E st
where  st 
L
Again considering the case where
2hE st 2
Mv E
the height h is large compared to the  max  
elongation of the bar, we obtain L AL
Impact Factor
The ratio of the dynamic response of a structure to the static response
(for the same load) is known as an impact factor. For instance, the
impact factor for the elongation of the bar is the ratio of the maximum
elongation to the static elongation:
 max
Impact Factor 
 st
Suddenly Applied Load
A special case of impact occurs when a load is applied suddenly with
no initial velocity. To explain this, assume that the sliding collar is
lowered gently until it just touches the flange, that is h=0 which leads
that 1/2
 max   st   st   2h st   2 st
2
 
Limitations

The preceding analyses were based upon the assumptions:


1. No energy losses occur during impact.

2. The stresses in the bar remain within the proportional limit.

3. Materials that exhibit considerable ductility beyond the


proportional limit generally offer much greater resistance to impact
loads than do brittle materials. Also, bars with grooves, holes, and
other forms of stress concentrations are very weak against
impact—a slight shock may produce fracture, even when the
material itself is ductile under static loading.

JU. Dr. Ibrahim Abu-Alshaikh


Example
A round, prismatic steel bar (E = 210 GPa) of length L = 2 m and
diameter d = 15 mm hangs vertically from a support at its upper end. A
sliding collar of mass M = 20 kg drops from a height h = 150 mm onto
the flange at the lower end of the bar without rebounding.

a) Calculate the maximum


elongation of the bar due to
the impact and determine the
corresponding impact factor.

b) Calculate the maximum


tensile stress in the bar and
determine the corresponding
impact factor.
Solution
(a) Maximum elongation. The elongation of the bar produced by the
falling collar can be determined as follows:

WL MgL 20(9.81)2
 st     0.0106 mm
EA EA 210 109  (0.015) 2
4
h 150
  14150
 st 0.0106

  2h 1/2 
 max   st 1  1     1.78 mm
   st  
 
The impact factor is equal to the ratio of the maximum elongation to
the static elongation  1.78
Impact Factor  max
  169
 st 0.0106
(b) Maximum tensile stress. The maximum stress produced by the
falling collar is obtained as follows:
 max E 1.78  210 109
 max    188MPa
L 2
This stress may be compared with the static stress, which is

W
 st   1.1 MPa
A
The ratio of σmax to σst is 188/1.11 = 169, which is the same impact
factor as for the elongations.

JU. Dr. Ibrahim Abu-Alshaikh


Example
A horizontal bar AB of length L is struck at its free end by a heavy
block of mass M moving horizontally with velocity v .

1. Determine the maximum shortening δmax of the bar due to the


impact and determine the corresponding impact factor.

2. Determine the maximum compressive stress σmax and the


corresponding impact factor. (Let EA represent the axial rigidity of
the bar.)
Solution
The kinetic energy of the block at the instant of impact is Mv2/2. The
strain energy of the bar when the block comes to rest at the instant of
maximum shortening is (EA δ2max/2L), therefore, we can write the
following equation of conservation of energy:

2
Mv L
 max 
EA
The maximum stress in the bar is found from the maximum shortening
by means of
2
Mv E
 max 
AL End of IMPACT LOADING
JU. Dr. Ibrahim Abu-Alshaikh
STRESS CONCENTRATIONS
• Loads transmitted through rigid
plates result in uniform
distribution of stress and strain.
• Concentrated loads result in
large stresses in the vicinity of
the load application point.
• Stress and strain distributions
become uniform at a relatively
short distance from the load
application points.
• Saint-Venant’s Principle:
Stress distribution may be
assumed independent of the
mode of load application
except in the immediate
vicinity of load application
points.
Of course, this “principle” is not a
rigorous law of mechanics but is a
common-sense observation based
upon theoretical and practical
experience.

Thus, we can make a general


statement that the equation σ
P/A gives
the axial stresses on a cross
section only when the cross
section is at least a distance b
away from any concentrated
load or discontinuity in shape,
where b is the largest lateral
dimension of the bar (such as the
width or diameter).
JU. Dr. Ibrahim Abu-Alshaikh
Stress Concentration: Hole

 max
For a bar in tension, the nominal stress is the average K
stress based upon the net cross-sectional area.  nom
Stress Concentration: Fillet
Stress-concentration factor K for flat bars with shoulder fillets. The
dashed line is for a full quarter-circular fillet
Stress-concentration factor K for round bars with shoulder fillets.
The dashed line is for a full quarter-circular fillet.
Example SOLUTION:
• Determine the geometric ratios
and find the stress concentration
factor from the Figure in the
previous slide .
Determine the largest axial
load P that can be safely • Find the allowable average
supported by a flat steel bar normal stress using the material
consisting of two portions, allowable normal stress and the
both 10 mm thick, and stress concentration factor.
respectively 40 and 60 mm
wide, connected by fillets of
• Apply the definition of normal
radius R = 8 mm. Assume an
stress to find the allowable load.
allowable normal stress of
165 MPa.
• Determine the geometric ratios
and find the stress concentration
factor from Fig. 2.64b.
b 60 mm R 8 mm
  1.50   0.20
c 40 mm c 40 mm
K  1.82

• Find the allowable average


normal stress using the material
allowable normal stress and the
stress concentration factor.
 max165 MPa
 nom    90.7 MPa
K 1.82
• Apply the definition of normal stress to find the allowable load.
P  A ave  40 mm 10 mm 90.7 MPa 

 36.3 103 N P  36.3 kN


Example
A stepped brass bar with a hole (Fig. 2-67a) has widths of b 9.0 cm
and c = 6.0 cm and a thickness of t = 1.0 cm. The fillets have radii
equal to 0.5 cm and the hole has a diameter of d = 1.8 cm. The
ultimate strength of the brass is 200 MPa.
(a) If a factor of safety of 2.8 is required, what is the maximum
allowable tensile load Pmax? (b) Find the hole diameter dmax at which
the two segments of the bar have tensile load carrying capacity equal
to that for the fillet region of the stepped bar.

JU. Dr. Ibrahim Abu-Alshaikh


K  2.51

K  2.35
(a) Determine the maximum allowable tensile load. Pmax
For the segment of the bar of width b and thickness t and having a
hole of diameter d, the net cross-sectional area is (b - d)(t), and the
nominal axial stress may be computed as
Pmax  allow  u / FS 
 nom  and  nom  
(b  d )t K hole K hole

Pmax 
 200 10 6
/ 2.8 
(9  1.8)(1)  20.5 kN
2.51
Next, we must investigate the tensile load carrying capacity of the
stepped bar in the segment having fillets of radius R =0.5 cm.
Following the same procedure we can write
Pmax  allow  u / FS 
 nom   
ct K fillet K fillet

Pmax 
 200  10 6
/ 2.8 
(6)(1)  18.24 kN
2.35
we see that the lesser value of the maximum allowable tensile load
Pmax is 18.24
(b) Determine the maximum hole diameter.
Comparing results of part a, we see that the segment of the stepped bar
with a hole has greater tensile load capacity Pmax than the segment with
the fillets. Thus, if we enlarge the hole, we will reduce the net cross
sectional area, Anet = (b - d)(t), (note that width b and thickness t
remain unchanged), but at the same time, we will reduce the stress
concentration factor Khole because ratio d/b increases. Rearranging
previous equations, we get
Pmax 
 u / FS (b  d )t  18.24 10
3

K hole
d

 u / FS  b(1  d )t  18.24 103  b )  0.284
(1
K hole b K hole

By trial-and-error d  2.97  0.33 so 1  0.33  0.284  d max  2.97


b 9 2.36
Hence, the maximum hole diameter dmax is approximately 3.0 cm if the
two segments of the stepped bar are to have equal tensile load
capacities.
NONLINEAR BEHAVIOR
Up to this point, our discussions have dealt primarily with members
and structures composed of materials that follow Hooke’s law. Now
we will consider the behavior of axially loaded members when the
stresses exceed the proportional limit. In such cases the stresses,
strains, and displacements depend upon the shape of the stress-strain
curve in the region beyond the proportional limit.

For purposes of analysis and design, we often represent the actual


stress-strain curve of a material by an idealized stress-strain curve
that can be expressed as a mathematical function. Some examples are
shown in the next Fig.
Types of idealized material behavior that can be expressed as a
mathematical function: (a) elastic-nonlinear stress-strain curve, (b)
general nonlinear stress-strain curve, (c) elastoplastic stress-strain
curve, and (d) bilinear stress-strain curve
Changes in Lengths of
Bars
The elongation or shortening
of a bar can be determined if
the stress-strain curve of the
material is known.
To illustrate the general procedure, we will consider the tapered bar AB
shown. Both the cross-sectional area and the axial force vary along the
length of the bar, and the material has a general nonlinear stress-strain
curve as shown.
• Because the bar is statically determinate, we can determine the
internal-axial forces at all cross sections from static equilibrium
alone.
• We can find the stresses by dividing the forces by the cross-sectional
areas, and we can find the strains from the stress-strain curve.
• we can determine the change in length from the strains, as described
in the following paragraph.
The change in length of
an element dx of the bar
is ε dx, where ε is the
strain at distance x from
the end. By integrating
this expression from one
end of the bar to the
other, we obtain the
change in length of the
entire bar :

L
    dx
0 where L is the length of the bar.
• If the strains are expressed analytically, that is, by algebraic
formulas, it may be possible to integrate the last equation by formal
mathematical means and thus obtain an expression for the change in
length.
• If the stresses and strains are expressed numerically, that is, by a
series of numerical values, we can proceed as follows. We can divide
the bar into small segments of length Δx, determine the average
stress and strain for each segment, and then calculate the elongation
of the entire bar by summing the elongations for the individual
segments. This process is equivalent to evaluating the integral of the
last equation by numerical methods instead of by formal integration.
• If the strains are uniform throughout the length of the bar, as in the
case of a prismatic bar with constant axial force, the integration of
the last Eq. is trivial and the change in length is

 L
JU. Dr. Ibrahim Abu-Alshaikh
Ramberg-Osgood Stress-Strain Law
Stress-strain curves for several metals, including aluminum and
magnesium, can be accurately represented by the Ramberg-Osgood
equation: m
   
   
0  0  0 
In this equation, σ and ε are the stress and strain, respectively, and ε0,
σ0, α, and m are constants of the material (obtained from tension
tests). An alternative form of this equation is
m
 0   

    
E E  0 
in which E=σ0/ε0 is the modulus of elasticity in the initial part of the
stress-strain curve.
JU. Dr. Ibrahim Abu-Alshaikh
Example
A graph for an aluminum
alloy for which the
constants are as follows:
E = 70 GPa, σ0 = 260
MPa, α = 3/7, and m =
10. The equation of this
particular stress-strain
curve is

 1   
10

   
70 103
628.2  260 

where σ has units of mega-pascals (MPa).


A prismatic bar AB of length L = 2.2 m and cross-
sectional area A = 480 mm2 supports two concentrated
loads P1 = 108 kN and P2 = 27 kN, as shown. The
material of the bar is an aluminum alloy having a
nonlinear stress-strain curve described by the
following Ramberg-Osgood equation:

 1   
10

   
70 103
628.2  260 

in which σ has units of MPa. Determine the


displacement δB of the lower end of the bar under each
of the following conditions:
1. the load P1 acts alone,
2. the load P2 acts alone, and
3. the loads P1 and P2 act simultaneously.
(a) Displacement due to the load P1 acting alone. The load P1
produces a uniform tensile stress throughout the length of the bar equal
to P1/A, or 225 MPa. Substituting this value into the stress-strain
relation gives ε = 0.003589. Therefore, the elongation of the bar, equal
to the displacement at point B δB = εL = (0.003589)(2.2 m) = 7.90 mm
(b) Displacement due to the load P2 acting alone. The stress in the
upper half of the bar is P2/A or 56.25 MPa, and there is no stress in the
lower half. Proceeding as in part (a), we obtain the following
elongation: δB = εL/2 = (0.0008036)(1.1 m) = 0.884 mm
(c) Displacement due to both loads acting simultaneously. The stress
in the lower half of the bar is P1/A and in the upper half is (P1 + P2)/A.
The corresponding stresses are 225 MPa and 281.25 MPa, and the
corresponding strains are 0.003589 and 0.007510 (from the Ramberg-
Osgood equation). Therefore, the elongation of the bar is
δB = (0.003589)(1.1 m) + (0.007510)(1.1 m) = 12.2
In a nonlinear structure, the displacement produced by two (or more)
loads acting simultaneously is not equal to the sum of the
displacements produced by the loads acting separately.
Elastoplastic Analysis
Now we will consider a material of considerable importance in
engineering design—steel, the most widely used structural metal. Mild
steel (or structural steel) can be modeled as an elastoplastic material
with a stress-strain diagram as shown in the Fig. An elastoplastic
material initially behaves in a linearly elastic manner with a modulus
of elasticity E. After plastic yielding begins, the strains increase at a
more-or-less constant stress, called the yield stress σY.
The strain at
the onset of
yielding is
known as the
yield strain
εY .
Initially, the bar elongates in a linearly elastic manner and Hooke’s
law is valid. Therefore, δ = PL/EA . The load at which yielding begins
is called the yield load PY and the corresponding elongation of the bar
is called the yield displacement δY = (PYL/EA), or (σY L/E). (Similar
comments apply to a bar in compression, provided buckling does not
occur). Thus, the structure behaves in a linearly elastic manner until
one of its members reaches the yield stress. Then that member will
begin to elongate (or shorten) with no further change in the axial load
in that member.
If a structure consisting only of axially loaded members is statically
determinate, its overall behavior follows the same pattern.

Statically Indeterminate Structures:


The situation is more complex if an elastoplastic structure is statically
indeterminate. If one member yields, other members will continue to
resist any increase in the load. However, eventually enough members
will yield to cause the entire structure to yield.
The situation is more complex if an
elastoplastic structure is statically
indeterminate. If one member yields,
other members will continue to resist
any increase in the load.
To illustrate the behavior of a statically
indeterminate structure, we will use the
simple arrangement shown in the Fig.
This structure consists of three steel bars
supporting a load P applied through a
rigid plate. The two outer bars have
length L1, the inner bar has length L2, and
all three bars have the same cross-
sectional area A.
The stress-strain diagram for the steel is
idealized as elastoplastic material with
modulus of elasticity E = σY /εY.
From equilibrium of the rigid plate in the vertical direction we obtain
2F1 + F2 = P
where F1 and F2 are the axial forces in the outer and inner bars,
respectively. Because the plate moves downward as a rigid body when
the load is applied, the compatibility equation is δ = δ
1 2
where δ1 and δ2 are the elongations of the outer and inner bars
δ1 = (F1 L1 /EA), δ2 = (F2 L2 /EA)
Solving the above equations simultaneously
PL2 PL1
F1  ; F2  ;
L1  2 L2 L1  2 L2
PL2 PL1
1  ; 2  ;
A  L1  2 L2  A  L1  2 L2 
These equations for the forces and stresses are valid
provided the stresses in all three bars remain below the
yield stress σY.
As the load P gradually increases, the stresses
in the bars increase until the yield stress is
reached in either the inner bar or the outer bars.
Let us assume that the outer bars are longer
than the inner bar as sketched.

L1  L2
Then the inner bar is more highly stressed than the outer bars and will
reach the yield stress first. When that happens, F2 = σY A and We can
determine PY as :  A  L  2L   A  2L 
PY  Y 1 2
 
Y
1  2

PL1 P  L1 
The downward displacement of the rigid bar at the yield load, called
the yield displacement δY, is equal to the elongation of the inner bar
when its stress first reaches the yield stress σY: FL  L
Y  2 2
 Y 2
AE E
Example
The structure consists of a horizontal beam AB (assumed to be rigid)
supported by two identical bars (bars 1 and 2) made of an elastoplastic
material. The bars have length L and cross-sectional area A, and the
material has yield stress σY, yield strain εY, and modulus of elasticity E
= σY/εY. The beam has length 3b and supports a load P at end B.
(a) Determine the yield load PY and the corresponding yield
displacement δY at the end of the bar (point B). (b) Determine the
plastic load PP and the corresponding plastic displacement δP at point
B.
(c) Construct a load-
displacement diagram
relating the load P to
the displacement δB of
point B.
Equations of equilibrium. Considering the equilibrium of beam AB,
we take moments about point A and obtain
MA  0 
F1(b) + F2(2b) - P(3b) = 0
in which F1 and F2 are the axial forces in bars 1 and 2, respectively.
This equation simplifies to
F1 + 2F2 = 3P
Equation of compatibility. The compatibility equation is based upon
the geometry of the structure. Under the action of the load P the rigid
beam rotates about point A, the downward displacement at every
point along the beam is proportional to its distance from point A.
Thus, the compatibility equation is
δ2 = 2δ1
where δ2 is the elongation of bar 2 and δ1 is the elongation of bar 1.
JU. Dr. Ibrahim Abu-Alshaikh
(a) Yield load and yield displacement. When the load P is small and
the stresses in the material are in the linearly elastic region, the force-
displacement relations for the two bars are
F1 L F2 L
1  ; 2 
EA EA
Combining these equations with the compatibility condition gives
F2 L F1 L
 2  2 2  2  F2  2 F1
EA EA
3P 6P
F1  2 F2  3P  F1  and F2 
5 5
Bar 2, which has the larger force, will be the first to reach the yield
stress, thus at that instant the force in bar 2 will be F2 = σY A:
5 Y A
FY 
6
The corresponding elongation of bar 2 is δ2 = σY L/E, and therefore the
yield displacement at point B is
3 2 3 Y L
Y  
2 2E
Both PY and δY are indicated on the load-displacement diagram
Point B
(b) Plastic load and plastic displacement. When the plastic load PP is
reached, both bars will be stretched to the yield stress and both forces
F1 and F2 will be equal to σY A. It follows from equilibrium
F1 + 2F2 = 3P
that the plastic load is
Pp   Y A
At this load, the left-hand bar (bar 1) has just reached the yield stress;
therefore, its elongation is δ1 = σY L/E, and the plastic displacement of
point B is
3 Y L
 P  31 
E
The ratio of the plastic load PP to the yield load PY is 6/5, and the ratio
of the plastic displacement δP to the yield displacement δY is 2. These
values are also shown on the load-displacement diagram.
(c) Load-displacement diagram. The complete load-displacement
behavior of the structure is pictured in Figure. The behavior is
linearly elastic in the region from O to A, partially plastic from A to
B, and fully plastic from B to C.

JU. Dr. Ibrahim Abu-Alshaikh

Вам также может понравиться