You are on page 1of 36

MODULE 2 – MECHANICAL STABILITY MODELS

Before undertaking the study of stability of elastic structures, the different


methods available for understanding and obtaining critical conditions will be
demonstrated through the use of simple mechanical models. The discussion will be
limited to conservative systems. It is also intended to demonstrate the effect of geometric
imperfections and load eccentricity on the response of the system. For a number of
models, both the small-deflection (linear) and large-deflection (nonlinear) theories will be
used for the sake of comparison. Finally, a comprehensive discussion of the different
types of behaviors will be given to enhance understanding of buckling, critical
conditions, and advantages or disadvantages of the approaches used.

2.1 ONE-DEGREE-OF- FREEDOM MODELS

Example 1: Consider a rigid bar of length L, hinged at one end, free at the other, and
supported through a frictionless ring connected to a spring that can move only
horizontally. The free-end is loaded with a force P in the direction of the bar. It is
assumed that the direction of the force remains unchanged. What is the critical load of the
bar?

P L
P

L(1-cos)=Z
Rigid Bar
(a
K
K Frictionless Ring
Spring
L K(a

a a 
Small  deflected
position
O
O
Fig. 2.1 Fig. 2.2
(A) Small -  Analysis

In the casual small -  analysis, we make the usual assumption that is so small
that   sin   tan  . With this restriction, we can only investigate the stability of
the equilibrium corresponding to = 0

1. The Classical or Equilibrium Method

M O 0 PL   K (a )a  0

PL  Ka 2   0 (2.1)

Thus, a nontrivial solution exists if

PL  Ka 2

and the bifurcation is located by

Ka 2 Ka 2
P or Pcr  (2.2)
L L

From Eq. 2.1, divide each term by Ka 2

 PL 
 2  1  0
 Ka 

PL
let p (2.3)
Ka 2

The bifurcation point is located at P = 1


PL
p
x Ka 2
x
1.0
x x x x x

x Bifurcation points
x

x

Fig. 2.3

2. The Energy Method


Since the system is conservative, the externally applied load P can be derived
from a potential. Thus (Fig 2.2)

U p  PZ o  Z (2.4)

and letting Z o  0 , then

U p  PZ

On the other hand, the energy stored in the system is given by

Ui 
1
Ka a
2

K a 
1
Ui 
2
(2.5)
2

Thus the total potential UT is given by

UT  Ui  U p

K a  PZ
1
UT 
2
(2.6)
2

But Z  L1  cos  , then


K a  PL 1  cos 
1
UT 
2
(2.7)
2

For static equilibrium, the total potential must have a stationary value. Thus

dU T
0
d

1
Ka 2 .2a  PL sin   0
2

and since sin   0 , then

 PL  Ka 2   0 (2.8)

Ka 2
P
L

This the same equilibrium equation derived previously. Furthermore, if the second
variation is positive definite, the static equilibrium is stable. If the second variation is
negative definite, the static equilibrium is unstable; if it is zero, no conclusion can be
drawn.
It is seen in this case that
d2UT
 Ka 2  PL (2.9)
d 2

Ka 2
Therefore, for P  the static equilibrium positions (  = 0) are stable, while
L
Ka 2
for P  they are unstable. Thus, as before
L

Ka 2
Pcr 
L
(B) Large -  Analysis

Lsin
P

L(1-cos

atan

L Katan Frictionless ring


 
R
a
a/cos L
O

Fig 2.4

In this particular approach, the only limitation on  is dictated from geometrical


a a
considerations. Note from Fig 2.4 that  cos 1    cos 1 .
L L
For  values outside this range, the ring will fly off the rigid bar.

1. The Classical or Equilibrium or Bifurcation Method


Since the ring is frictionless, the force R, normal to the rigid bar, is related to
the spring force through the following expression (see Fig 2.4)

Ka tan   R cos 

M O 0
a
PL sin   R
cos 

Ka tan   a  Ka 2 sin 
PL sin     (2.10)
cos   cos   cos3 

Thus, the equilibrium positions are characterized by

 Ka 2 
  PL  sin   0 (2.11)
 cos 
3

Which implies that either

 0 (2.12a)

PL
or 2
 sec 3  (2.12b)
Ka

It is clearly seen that a nontrivial solution   0 can exist for


PL
 1 and a
Ka 2
PL Ka 2
bifurcation point exists at  1 . Hence Pcr 
Ka 2 L

PL
x - stable equilibrium o Ka 2
o – unstable equilibrium Bifurcation point
x o x
x x
x x
o
1.0
x

 a  
  cos 1   O a
1
2 L cos   2
 L

Fig 2.4 (a )

2. Energy Method
The total potential of the system is given by

U T  PL 1  cos  
1
Ka tan (a tan )
2

U T  PL 1  cos  
1
Ka 2 tan 2  (see note below) (2.13)
2

The static equilibrium positions are characterized by the equation

dU T
0
d
Ka 2 sin 
 PL sin   0 (2.14)
cos3 

Furthermore, the second variation is given by

d 2 U T  Ka 2  2 sin 
2

 
 PL  cos   3Ka (see note below) (2.15)
d 2  cos 
3
 cos 4 

It is easily concluded that the static equilibrium positions characterized by


  0 are stable. Similarly, the positions   0 or PL  Ka 2 are unstable. It can
also be concluded, by considering higher variations, that the position denoted by
 o  0 and PL  Ka 2 is stable. The critical load is

Ka 2
Pcr 
L

NOTE:

d d  sin 2   cos 2 .2 sin  cos   sin 2 .2 cos ( sin )
(tan 2 )   
d d  cos 2   cos 4 

d
(tan 2 ) 
 
2 sin  cos  cos 2   sin 2  2 sin 

d cos 4  cos 3 

d2UT 2  cos  cos   sin 3 cos  sin  


3 2
  PL cos   Ka 
 
d 2  cos 6  

d2UT Ka 2 cos  2 sin 


2
 PL cos    3Ka
d 2 cos 3  cos 4 
Example 2:

1
a a

B’ B
P P

2

Fig. 2.5
(1) By Classical or Equilibrium or Bifurcation Method

a
P 

 2  a tan   a
P 

Fig. 2.5a

` P 2  M (2.16)

Pa  c(2 )

2c
Pcr 
a

(2) By Energy Method

The strain energy stored in the spring is the product of the angle of rotation of one
rod, measured with respect to the other, 2 and the mean value of the couple exerted by
1
the spring, (c 2 ) . The strain energy is
2

1
U c(2 )( 2 )  2c 2 (2.17)
2
The constant force P must be regarded as having been applied when the structure was
straight, and then maintained in position as the structure moved slowly to its deformed
state. While this happens, P losses potential energy, moving as it does along its own line
of action from B to B’. If the potential energy of P is taken as zero at B, then its potential
energy of position at B’ is

V p   P1 (2.19)

From the geometry of the diagram it is easy to see that

1  2a(1  cos ) (2.20)

 2  a sin  (2.21)

The total potential energy in terms of the displacement  is:

V  U  V p  2c 2  2Pa(1  cos ) (2.22)

dV
Hence,  4c  2 Pa sin  ` (2.23)
d

d 2V
and  4c  2 Pa cos (2.24)
d 2

dV
From the condition that  0;
d

2c 
P . (2.25)
a sin 
P D E
Pcr

stable

Unstable

E’
A C
stable

O 100○ 180○

Fig. 2.6

 = 0○  = 90○  → 180○

The behavior of the structure may be summarized by referring to Fig. 2.6 in which the
2c P
quantity has been denoted by Pcr and a non-dimensional load parameter, has been
a Pcr
plotted against the angular displacement .
2c P
When the end load P is less than the critical value ( i.e when  1 ) the only
a Pcr
position of equilibrium is the straight one (  = 0 , line OA ). The equation

d2V
 4c  2Pa cos 
d 2

d2V
may be used to verify the fact that is positive so V is minimum and the equilibrium
d 2
is stable. If the straight structure is deformed slightly and then released it returns to its
original straight configuration.
P
When the end load P is greater than the critical value (  1 ), the structure is still in
Pcr
equilibrium when it is straight (  = 0 , line AD ) but it is unstable. This is because
d2V
is negative and V is therefore maximum. If the straight is disturbed slightly, it does
d 2
not return to the original straight position. However, there is another position of
equilibrium defined by

2c 
P .
a sin 

d2V
one which describes the curve AE. It can be shown that is positive along AE and the
d 2
equilibrium represented by it is therefore stable. The curve AE’ is merely a reflection of
AE, indicating that there is a similar position of stable equilibrium with negative values
of .
If the load P is increased from zero, the load displacement relationship follows the
2c
line OA. When P reaches the critical value Pcr  (which is the elastic critical load or
a
elastic buckling load) the structure has a choice of equilibrium paths – AD, AE, oe AE’.
This is often referred to as bifurcation of equilibrium path. The structure will normally
follow the stable path AE or AE’, and it is therefore said to exhibit stable-symmetric
bifurcation.
The angular rotation  may be treated as an independent parameter in the following
equations:

1  2a(1  cos )

 2  a sin 

2c 
And P .
a sin 

In this way, relationships between P and 1 and between P and 2 ,may be obtained as
illustrated in Fig. 2.7
In Fig. 2.6 it is clear that P must increase as the deflection increases.
P
Pcr 1
2
a
a

1
Tangent to curve
a

1  2
,
O a a
P
Fig. 2.7 Relationships between and 1 and 2
Pcr

Example3:
Consider two rigid links pinned together and supported by hinges on rollers at the free
ends ( Fig. 2.8 ). The system is supported at the middle hinge by a vertical linear spring
and is acted upon by two collinear horizontal forces of equal intensity. The two links are
initially horizontal. Can this system, buckle? What is the critical load?

1. The Classical or Bifurcation Method


Using the casual small-deflection theory, we put the system into a deflected
position (Fig 2.8) and write the equilibrium equations.
L L
P P

(a) Geometry

L2   2 L2   2

P P

L

K K
K
2 2

(b) Forces and displacements

2P
x - stable equilibrium o KL
o – unstable equilibrium Bifurcation point
o
o
1.0
x


O

(c) Load-deflection curve

K
Since the system is symmetric, the vertical reactions at the hinges are .
2
Furthermore, the moment about the middle hinge must vanish. This requirement
leads to the equilibrium equation
1
P  KL  0
2

 1 
 P  KL   0 (2.26)
 2 

Thus, the equilibrium positions are defined by either   0 (trivial solution) or

KL 2P
P . In plotting versus , we notice that the bifurcation point exists at
2 KL
2P
 1 and
KL

KL
Pcr  (2.27)
2

2. By Energy Method
The total potential is the sum of the energy stored in the spring and the
potential of the external forces. Thus

UT 
K.
2

 P L  L2   2 2  (2.28)

dU T
For static equilibrium 0
d

dU T 1 2
 K  2P. 0 (2.29)
d
L 
1
2 2
 2 2

which under the assumption  2  L2 , is identical to

 KL 
P    0 (2.30)
 2 

For the static equilibrium positions to be stable, the second variation must be
positive definite,

d2UT 2P 2 P 2  2P 
K   K   (2.31)
d 2
L  L   L 
1 3
2
 2 2 2
 2 2
KL
Thus the equilibrium positions denoted by  = 0 and P  are stable and the
2
critical load

KL
Pcr  (2.32)
2

2a. By Energy Method (another approach)

1 1
2 L L 2
P P
 
2

Fig. 2.9
The strain energy in the spring is

1
U K 2
2
(2.33)
2

where  2  L sin 

1
 L1  cos 
2

The potential energy is


1
V  U  Vp  KL2 sin 2   2PL (1  cos ) (2.34)
2

And for equilibrium

dV
 KL2 sin  cos   2PL sin   0 (2.35)
d

KL 2

cos   2PL sin   0

This is satisfied when  = 0 or , or when

P
 cos 
Pcr

KL
where Pcr  (2.36)
2

The complete solution is depicted graphically in Fig. 2.10; the type of equilibrium
d2V
represented by each curve may be verified by finding the sign of as before.
d 2
As in the previous example, a bifurcation of the equilibrium path occurs at a
critical value of P (point A). In this case, however, each of the three possible paths
(AE, AB, AB’) represents unstable equilibrium and the structure is said to exhibit
unstable-symmetric bifurcation. If the end-load is held constant at P=Pcr, the structure
collapses (after some oscillation) to the point G on the stable equilibrium path CFGH.
unstable
P
Pcr E H

1.0 A G

stable

stable
unstable

F
B
0○ 90○ 180○
stable

stable
-1.0
C

unstable
Fig. 2.10 Force-displacement relations
eeeeee

 = 0○

P P
eeeeee

 = 90○
eeeeee
 →180○

2.2 TWO-DEGREE-OF-FREEDOM MODELS


Consider the system shown in Fig. 2.11, composed of three rigid bars of equal length
hinged together as shown. The linear springs are of equal stiffness. This is a two-degree-
of-freedom system and it is acted upon by a horizontal force, P, applied quasistatically.
We must determine whether or not the system will buckle and the critical value of the
applied load. The load is assumed to remain horizontal.

P L L L P

K K

Fig. 2.11

R1 R2
L L L
A B
P P
   

C
D
KL KL

1  L sin   L 1  L sin   L
(1) By Classical or Bifurcation Method

M B 0

R 1 (3L)  KL(2L)  KL(L)  0 (2.37)

M A 0

R 2 (3L)  KL(L)  KL(2L)  0 (2.38)

M C 0

R 1L  PL  0 (2.39)

M D 0
R 2 L  PL  0 (2.40)

Eliminating R1 and R2 yields the following system of linear homogenous algebraic


equations

 2  KL
 P  KL   0
 3  3
(2.41)
KL  2 
  P  KL   0
3  3 

The critical condition is derived if we require the existence of a non-trivial solution.


This leads to the characteristic equation

 2KL   KL 
 P  3  
  3 0 (2.42)
 KL  2KL 
 P  
 3  3 

Thus,

 2  2   KL  KL 
  P  KL  P  KL      0
 3  3   3  3 

2 2
 2   KL 
  P  KL     0
 3   3 
 2
  P 2  PKL   KL   
4 4 KL   0 2

 3 9  9

 2 4 2 KL   0 2

 P  PKL   KL   
4
 3 9  9

PKL   ( KL) 2  0
4 3
P2 
3 9

PKL   ( KL ) 2  0
4 1
P2  (2.43)
3 3

 KL 
P  P  KL   0
 3 

KL
P
3
(2.44)
or P  KL

(2) By Energy Method


The total potential for the system which consists of the energy stored in the
springs and the potential of the external forces, is given by

U T  U1  U P

KL2  2  KL2  2  PL1  cos   (1  cos )  1  cos(  ) (2.45)


1 1
UT 
2 2

By assuming that the angles  and  can be made as small as desired, we can write U T
as

UT 
1
2
1
 
KL2  2  KL2  2  PL  2   2   (see note below)
2
(2.46)

For static equilibrium, the total potential must be stationary; therefore

U T U T
 0 (2.47)
 
Which leads to the following equilibrium equations:

KL 2

 2PL   PL  0
(2.48)

PL  KL  2PL   0
2

The nontrivial solution is the same as the one obtained by the classical approach.

KL
P
3

And P  KL

Study of the stability of the equilibrium positions characterized by     0 for the


entire range of values of P requires knowledge of the second variations

2UT
 KL2  2PL (2.49a)
 2

 2UT
 KL2  2PL (2.49b)
 2

2UT
 PL (2.49c)


The equilibrium positions are stable if and only if both of the following inequalities
are satisfied.

2UT
0 (2.50)
 2
2
2UT 2UT 2UT 
.   (2.51)
 2  2   

In terms of the applied load and the structural geometry, these inequalities are

KL  2P (2.52)

 KL 
(KL  P)  P  0 (2.53)
 3 
KL
From these expressions, we see that equilibrium positions for which P  are
3
KL
stable, while all equilibrium positions for which P  are unstable. Therefore
3

KL
Pcr 
3

2 2
Note: cos   1  ; cos   1 
2 2

cos    cos  cos   sin  sin 

 2   2 
cos    1  1    
 2  2 

2 2 22
cos    1     
2 2 4

Therefore:

1  cos   (1  cos )  1  cos(   )   2 2 2  2 2


2
 11    
2 2 2 2 4

 22   22 
  2   2     note   is HOT
 4   4 


  2   2   
2.3 THREE-DEGREE-OF-FREEDOM MODELS
The system of three light rigid rods in Fig. 2.12 is a special case to be considered
here. The rods are freely hinged at A, B, and C. The length of each rod is “a”. The
restraining springs at B, C, and D are each of stiffness K and they are unstressed when
the rods are vertical. A vertical load P is applied at D and a horizontal force Q is applied
at the midpoint of each rod. It is required to find the displacements x 1, x2 and x3,
assuming that the springs remain nearly horizontal as the structure deforms.

K x3
D
eeeeeeeeeeeeeeeeee x1

Q a
K x2 B2
C
eeeeeeeeeeeeee Q B
B1
a
Q
K x1
B
eeeeeeeeeee
a
a
Q A
A

Rod No. 1
Fig. 2.12

From the figure:


1
  x1  2  2
AB1  a 2  x 1  a 1    
2
(2.54)
  a  

The term inside the bracket may be expanded by means of binomial theorem. Only the
x
first two terms of the expansion are required because 1 is considered to be small; this is
a
 2
equivalent to the approximation cos   1 
2

 1  x 2 
Hence AB1  a 1   1   (2.55)
 2  a  
The distance B2B1, by which the hinge descends is

2 2
1 x1 1 x1
B 2 B1  a  AB1  a. 2
 (2.56)
2 a 2 a

A similar argument may be used to show that hinge 2 descends (relative to 1) through a
x  x 1 2 . Similarly, hinge 3 moves relative to hinge 2 by x 3  x 2 2
distance 2
2a 2a

The strain energy absorbed in the springs is

U
2

K 2
x1  x 2  x 3
2 2
 (2.57)

The change in the potential energy of P is

VP  Px (descent of D)

 x 2 x  x 1 2 x 3  x 2 2 
VP  P  1  2   (2.58)
 2a 2a 2a 

The work done by the horizontal forces Q can be evaluated as follows:


1 1
x3 Q Q
2 2
Q
1
Q
2
x2 Q
1
Q
2
Q
1
Q
x1 Q 2
1
Q
2
Q
1
Q
2

x3
Thus Vq  Q( x 1 )  Q( x 2 )  Q (2.59)
2
The total potential energy of the system is

V  U  VP  Vq
V
K 2
2

x1  x 2  x3 
2 2 P
2a
 
2 x1  2 x 2  x3  2 x1 x 2  2 x 2 x3
2 2 2

(2.60)
 x 
 Q x1  x 2  3 
 2

For equilibrium, V must be stationary with respect to variations in the displacements x 1,


x2, x3:

V P
 Kx 1  (4x 1  2x 2 )  Q  0
x 1 2a

V P
 Kx 2  (4x 2  2x 1  2x 3 )  Q  0 (2.61)
x 2 2a

V P
 Kx3  ( 2 x3  2 x 2 )  Q  0
x 2 2a

or
P P
Kx 1  (4 x 1 )  .2x 2  0  Q  0
2a 2a

(2 x 1 )  Kx 2  4 x 2   2 x 3   Q  0
P P P
(2.62)
2a 2a 2a

0
P
2x 2   Kx 3  P 2x 3   Q  0
2a 2a 2

This may be written as

1  2p  p 0  x1   1 
 p    Q 
 1  2p  p  x 2    1  (2.63)
K 
 0 p 1  p   x 3  1 / 2

P
Where p
Ka

The determinant of the above equation can be expanded as:

(1  2 p)(1  2 p)(1  p)  p. p(0)  0( p. p)  0(1  2 p)0  p. p(1  2 p)  (1  p) p. p  0

or p 3  6p 2  5p  1  0 (2.64)
The three roots are: p1 = 0.3080 ; p2= 0.6431; p3 = 5. 049

These values (multiplied by Ka) represent the three elastic critical loads at which nonzero
displacement can occur.
It is not possible to find the buckled shape of the structure by substituting p = 0.3080
in the characteristic equation and solving for x 1, x2, and x3. Indeed , x is indeterminate
when the determinant, K = 0. Nevertheless, some information can be obtained about the
buckled shape by dividing x1, x2, and x3 by x1, thus

1  2p  p 0   1  0
 p    
 1  2p  p  x 2 / x 1   0 (2.65)
 0 p 1  p   x 3 / x 1  0

x2 x
Any two of these three equations may be solved for the ratios and 3 provided p is
x1 x1
made equal to one of the three critical loads. For example, where p = p 1 =0.3080, the
result is

x2 x3
 1.25 ;  0.555
x1 x1

It might be thought that, with three equations to be solved for two unknowns, some
difficulty could be encountered. However, provided p is given one of the three critical
values, any two will give the same result as any other two.
If x1 is arbitrarily made equal to +1, then x 2 = -1.25 and x3 = +0.555; this defines the
configuration of the buckled structure at the first critical load.
x3 +0.555 -0.802 +2.250

x2 -1.250 +0.445 +1.800

x1 +1.000 +1.000 +1.000

p = 0.3080 p = 0.6431 p = 5.049

Fig. 2.13 Buckling modes for the system

Now let us return to consider the case in which the structure carries both vertical load P
and horizontal forces Q. The three equations in terms of x 1, x2, and x3 derived earlier
govern the behavior of the structure. The solution is obtained by inverting the stiffness
matrix. This can always be done except when p has one of the three critical values, in
which case K cannot be inverted because it is singular.
For any given value of P, the displacements “x” are proportional to the transverse
loads, Q.
For any given value of Q, it can be shown that the displacements “x” increase more
rapidly with p as p rises, until, as p approaches the first critical load p1, the displacements
approach infinity.
Alternative Approach: Solution by Equilibrium Method

P
x1 A
Kx1

a
Q
x2
Kx2 B

Q a
x3
Kx3 C

a
Q
D H

M B 0

Qa
(1) P( x 1  x 2 )  Kx 1a  0
2

M C 0

3a Qa
(2) P( x 1  x 3 )  Kx 1 2a  Kx 2 a  Q  0
2 2

M D 0

5a 3a Qa
(3) Px 1  Kx 1 3a  Kx 2 2a  Kx 3 a  Q Q  0
2 2 2

Qa
From (1) (P  Ka ) x 1  Px 2  0x 3  
2

From (2) (P  2Ka ) x 1  Kax 2  Px 3  2Qa


9
From (3) (P  3Ka ) x 1  K 2ax 2  Kax 3   Qa
2

Hence,

 P  Ka  P 0  x1   Qa 
(P  2Ka )  Ka     1 
  P  x 2    4Qa 
2
 (P  3Ka )  2Ka  Ka   x 3  
9Qa 

(P  Ka )( Ka ) 2  P 2 (P  3Ka )  0  0  (2PKa )( P  Ka )  (PKa )( P  2Ka ) 


P(Ka ) 2  (Ka ) 3  P 3  3P 2 (Ka )  2P 2 (Ka )  2P(Ka ) 2  P 2 Ka  2P(Ka ) 2  0

P 3  6P 2 (Ka )  5P(Ka ) 2  (Ka ) 3  0

Divide by (Ka)3

P3 6P 2 5P
3
 2
 1  0
(Ka ) (Ka ) (Ka )

P
Let  
Ka

3  62  5  1  0 This is the transcendental equation

The roots are:

1  0.3080 ; Pcr = 0.308Ka

 2  0.6431 ; P2 = 0.6431Ka

 3  5.049 ; P3 = 5.049Ka
2.4 SNAPTHROUGH MODEL
In the analysis of this model, we will demonstrate the type of buckling known as
snapthrough or oil-canning.
Consider two rigid bars of length L pinned together, with one end of the system
pinned to an immovable support, and the other pinned to a linear horizontal spring. The
rigid bars make an angle with the horizontal when the spring is unstretched and the
system is loaded laterally through a force P applied quasistatically from zero at the
connection of the two rigid bars. Is it possible for the system to snapthrough toward the
other side at some value of the applied load?

P
P
deflected position
B

L
Lsin

Lsin
 
A C K
eeeee

2Lcos
 = 2L(cos-cos

2Lcos

Fig. 2.14 Snapthrough Model


P

F = 2KL(cos- cos
L


F F
P P
2 2
P P
(1) The Equilibrium Method (Classical or Bifurcation Method)
Let the horizontal reaction of the spring be F. This force is equal to K times the
compression in the spring, or

F  2KL(cos   cos ) (2.66)

M B 0

P
L cos   FL sin  (2.67)
2

Substituting Eq. 2.66,

P sin 
 sin   cos  (2.68)
4KL cos 

P
 sin   cos  tan 
4KL

 
Note from the figure that  
2 2
The equilibrium state, Eq 2.68 are plotted in Fig. 2.15. Note that loading starts at point A
and it is increasing quasistatically. When point B is reached, we see that no appreciable
change in the load the system will tend to snapthrough toward the CD portion of the
curve. The load corresponding to position B is the critical one, and its magnitude may be
obtained from the fact that
dP
0 (2.69)
d

Note that the right side of Eq. 2.68 is a continuous function of  with continuous first
derivatives.
If we denote by B the angles corresponding to positions B and B’, then

 B   cos 1 (cos ) 3 (2.70)

 P 
and    sin  B  cos tan  B (2.71)
 4 KL  cr
P
4KL

B
C

 A 
 
2   2

B’

Fig. 2.15 Load-deflection Curve


(2) Energy Approach
The total potential, UT, for the system, which is equal to the potential of the external
force and the energy stored in the spring, is given by

U T  2KL2 (cos   cos ) 2  PLsin   sin   0 (2.72)

Static equilibrium positions are characterized by the vanishing of the first variation of the
total potential, or

dU T
 2KL2 .2(cos   cos )( sin )  PL cos   0
d

This leads to the equilibrium equation

P
 sin   cos  tan  (2.73)
4KL

The character of the equilibrium positions is governed by the second variation, or

d2UT cos 
 4K 2 (  cos 2 ) (2.74)
d 2
cos 

Thus, in the region

1 1

 cos 1 (cos ) 3    cos 1 (cos ) 3 (2.75)

The second derivative is negative and the equilibrium positions are unstable. Outside this
region, the second derivative is positive and the equilibrium positions are stable. Thus, P cr
is given by Eq (2.71). Note that points between B and B’ represent “hills”, on the total
potential curve, while points outside this region represent “valleys”.
The critical condition is reached when the load is such that the near equilibrium point
coincides with the unstable point.
2.5 MODELS OF IMPERFECT GEOMETRIES
Consider, for instance the model shown in Fig 2.8 with a small imperfection 0 when
the spring is unstretched. The problem is to find the behavior of the imperfect system
under the quasistatic application of the horizontal forces.

L L
P P
0

Fig. 2.16

K K
2 2

L2     0  L2     0 
2 2

P P

L  L

K

Fig. 2.17 Geometry

 M(middle hinge)  0

P (   0 ) 
K
2
 L  (   )   K2L
2
0
2
(2.76)

This equation can be written as

 KL  KL
P  (   0 )   0 (2.77)
 2  2
KL
If we divide both sides by  0 , the equilibrium equation becomes
2

 2P  
  1(1  )  1 (2.78)
 KL  0

 2P  
This represents a hyperbola in the coordinate system  1 and (1  ) .
 KL  0

2P 2P
 1`
KL KL

  
1  
 0 

0

Fig. 2.18 Load-deflection curve

You might also like