Anda di halaman 1dari 19

Annexure-I

TERM PAPER

MECHANICS

PHY101

Topic: Buckling (mechanical failure like corrosion)

DOA:30/08/10

DOR: 28/09/10

DOS: 10/11/10

Submitted to: Submitted by:

Mr. BHARPUR SINGH Mr. ASHUTOSH KUMAR

Deptt. Of Physics Roll. No. RG4003 B37

Reg.No11006427

Section G4003
ACKNOWLEDGEMENT

I take this opportunity to present my votes of thanks to all those guidepost


who really acted as lightening pillars to enlighten our way throughout this
project that has led to successful and satisfactory completion of this study.

I am really grateful to our MECHANICS teacher and HOD. For providing us


with an opportunity to undertake this project in this university and providing
us with all the facilities.

Lastly, I am thankful to all those, particularly the various friends , who have
been instrumental in creating proper, healthy and conductive environment and
including new and fresh innovative ideas for me during the project, their help,
it would have been extremely difficult for me to prepare the project in a time
bound framework.

Contents

Definition
Types
Formula
Other form
Design
Theory
Equation
Design consideration
Some pictures
BUCKLING

A structure may fail to support its load when a connection snaps, or it bends
until it is useless, or a member in tension either pulls apart or a crack
forms that divides it, or a member in compression crushes and crumbles,
or, finally, if a member in compression buckles, that is, moves laterally and
shortens under a load it can no longer support. Of all of these modes of
failure, buckling is probably the most common and most catastrophic.

Leonhard Euler long ago showed that there was a critical load for buckling of a
slender column. A column, of course, is simply a common case of a
compression member. With any smaller load, the column would remain
straight and support it. With any larger load, the least disturbance would
cause the column to bend sideways with an indefinitely large displacement;
that is, it would buckle. The simplest case is that of a column pinned at
both ends, that is, free to rotate, under a load P. Let x be the distance along
the column, and y its displacement to the side. The bending moment at any
point x is Py, which increases with the displacement, it should be noticed.
This means that any buckling merely encourages further buckling,
explaining why such failure is catastrophic.

The bending moment is related to the curvature by Py = M = -YI d2y/dx2, since


for small y the second derivative is the reciprocal of the radius of
curvature. Y is Young's modulus, and I is the moment of inertia of area of
the cross-section of the column. I is the sum of elements of area times the
square of their distances from the neutral axis, as in the bending of a beam.
The differential equation is easy to solve, with the result that y = A sin ax +
B cos ax, where a2 = P/YI. B = 0, since y(0) = 0, and the condition A sin aL
= 0 must also be satisfied, where L is the length of the column. This means
that aL = π, 2π, ... , or the smallest value of P is given by P' = π2YI/L2. The
constant A itself remains indeterminate.

The interpretation of this result is that for P < P' the column remains straight
and A = 0. For P > P', the column is unstable and buckles. P' is the critical
load for buckling. It is found from practice that this theory gives good
results for columns that are more than 30 times longer than wide. P' is
proportional to 1/L2, so the supporting capacity of a column decreases
rapidly with an increase in length. We also notice that an eccentric load
will cause a decrease in strength, since there will be a bending moment and
a curvature from the start. The column will be stable under sideways
impacts and will return to straightness when such load is removed. The
moment of inertia I will be different for different directions of buckling if
the cross-section is not axially symmetrical. The direction with the
minimum I will be the critical one.

There is a good incentive to look for ways to increase the critical buckling load.
If the ends of the column are clamped, that is, prevented from rotating, the
buckling curve changes from a half-wavelength sine to a full wavelength
cosine, effectively halving the length of the column, and increasing the
critical load by a factor of 4. A brace in the middle of the column that
prevents it from moving to one side also halves the buckling length, and
increases the critical load by a factor of 4.

For a short column, P' will become large, and at some point the compressive
strength of the material will be exceeded. Let P" = Aσ be the maximum
load for failure by compression. Rankine combined these results
empirically to give a maximum load P by the formula 1/P = 1/P' + 1/P",
which is quite conservative, and seems to give good results for columns
between

• When a structure ( subjected usually to compression ) undergoes visibly large


displacements transverse to the load then it is said to buckle. Buckling may
be demonstrated by pressing the opposite edges of a flat sheet of cardboard
towards one another. For small loads the process is elastic since buckling
displacements disappear when the load is removed.
Local buckling of plates or shells is indicated by the growth of bulges, waves
or ripples, and is commonly encountered in the component plates of thin
structural members.

Buckling proceeds in manner which may be either :

stable - in which case displacements increase in a controlled fashion as loads


are increased, ie. the structure's ability to sustain loads is maintained,
or
unstable - in which case deformations increase instantaneously, the load carrying
capacity nose- dives and the structure collapses catastrophically.

Neutral equilibrium is also a theoretical possibility during buckling - this is


characterised by deformation increase without change in load.
Buckling and bending are similar in that they both involve bending moments. In
bending these moments are substantially independent of the resulting deflections,
whereas in buckling the moments and deflections are mutually inter-dependent - so
moments, deflections and stresses are not proportional to loads.
If buckling deflections become too large then the structure fails - this is
a geometric consideration, completely divorced from any
material strength consideration. If a component or part thereof is prone to buckling
then its design must satisfy both strength and buckling safety constraints - that is why
we now examine the subject of buckling.

Buckling has become more of a problem in recent years since the use of high strength
material requires less material for load support - structures and components have
become generally more slender and buckle- prone. This trend has continued
throughout technological history, as is demonstrated by bridges in the following
sequence :

The Pont du Gard in Provence was completed by the Romans


in the first century AD as part of a 50km aqueduct to convey
water from a spring at Uzès to the garrison town of Nemausus
(Nimes). The bridge is constructed from limestone blocks fitted
together without mortar and secured with iron clamps. The
three tiered structure avoids the need for long compressive members.

The Royal Border Bridge, Berwick upon Tweed, was built by


Robert Stephenson whose father George built the Stockton and
Darlington Railway (the first public railway) in 1825. Opened
in 1850, the bridge continues today as an important link in the
busy King's Cross (London) - Edinburgh line. The increased
slenderness of the columns compared to the Pont du Gard reflect technological
improvements over many centuries.

The Crymlyn Viaduct over the Ebbw Alley opened in 1857 as


Welsh coal mining expanded. It was constructed of wrought
and cast iron, and remained the highest railway viaduct in the
UK until its closure in 1964 due to increased locomotive
weights (1908 photo). The advance from masonary to the
slender metal compressive members which make up each column requires substantial
bracing to prevent buckling

The Humber road bridge, opened in 1981, comprises a


continuously welded closed box road deck suspended from
catenary cables supported on reinforced concrete towers.
Suspension bridges eliminate the need for struts other than the
two towers, however avoiding buckles in other slender
components becomes an issue

The dangers associated with over-slender build were tragically driven home by the
collapse of the Tacoma Narrows road bridge over the Puget Sound in 1940. Although
this failure was apparently due to wind- structure aerodynamic coupling rather than
buckling as such, this film clip demonstrates graphically the ability of large structures
to undergo significant elastic deflections.

Buckling of thin-walled structures

A structure may fail to support its load when a connection snaps, or it bends until it is
useless, or a member in tension either pulls apart or a crack forms that divides it, or a
member in compression crushes and crumbles, or, finally, if a member in
compression buckles, that is, moves laterally and shortens under a load it can no
longer support. Of all of these modes of failure, buckling is probably the most
common and most catastrophic.

Leonhard Euler long ago showed that there was a critical load for buckling of a
slender column. A column, of course, is simply a common case of a compression
member. With any smaller load, the column would remain straight and support it.
With any larger load, the least disturbance would cause the column to bend sideways
with an indefinitely large displacement; that is, it would buckle. The simplest case is
that of a column pinned at both ends, that is, free to rotate, under a load P. Let x be the
distance along the column, and y its displacement to the side. The bending moment at
any point x is Py, which increases with the displacement, it should be noticed. This
means that any buckling merely encourages further buckling, explaining why such
failure is catastrophic.

The bending moment is related to the curvature by Py = M = -YI d2y/dx2, since for
small y the second derivative is the reciprocal of the radius of curvature. Y is Young's
modulus, and I is the moment of inertia of area of the cross-section of the column. I is
the sum of elements of area times the square of their distances from the neutral axis,
as in the bending of a beam. The differential equation is easy to solve, with the result
that y = A sin ax + B cos ax, where a2 = P/YI. B = 0, since y(0) = 0, and the condition
A sin aL = 0 must also be satisfied, where L is the length of the column. This means
that aL = π, 2π, ... , or the smallest value of P is given by P' = π2YI/L2. The constant A
itself remains indeterminate.

The interpretation of this result is that for P < P' the column remains straight and A =
0. For P > P', the column is unstable and buckles. P' is the critical load for buckling. It
is found from practice that this theory gives good results for columns that are more
than 30 times longer than wide. P' is proportional to 1/L2, so the supporting capacity
of a column decreases rapidly with an increase in length. We also notice that an
eccentric load will cause a decrease in strength, since there will be a bending moment
and a curvature from the start. The column will be stable under sideways impacts and
will return to straightness when such load is removed. The moment of inertia I will be
different for different directions of buckling if the cross-section is not axially
symmetrical. The direction with the minimum I will be the critical one.

There is a good incentive to look for ways to increase the critical buckling load. If the
ends of the column are clamped, that is, prevented from rotating, the buckling curve
changes from a half-wavelength sine to a full wavelength cosine, effectively halving
the length of the column, and increasing the critical load by a factor of 4. A brace in
the middle of the column that prevents it from moving to one side also halves the
buckling length, and increases the critical load by a factor of 4.

For a short column, P' will become large, and at some point the compressive strength
of the material will be exceeded. Let P" = Aσ be the maximum load for failure by
compression. Rankine combined these results empirically to give a maximum load P
by the formula 1/P = 1/P' + 1/P", which is quite conservative, and seems to give good
results for columns betweenequilibrium path bifurcates into two symmetric
secondary paths as illustrated. Clearly the critical Euler load limits the column's safe
load capacity.

Local buckling of an edge-supported thin plate does not necessarily lead to total
collapse as in the case of columns, since plates can generally withstand loads greater
than critical. However the P-q curve illustrates plates' greatly reduced stiffness after
buckling, so plates cannot be used in the post- buckling region unless the behaviour in
that region is known with confidence.
It should be emphasised that the knee in the P-q curve is unrelated to any elastic-
plastic yield transition; the systems being discussed are totally elastic. The knee is an
effect of overall geometric rather than material instability.

This photograph illustrates local buckling of a model box girder constructed from thin
plates, not unlike the road deck of the Humber bridge above.

Inclined striations are caused by shear loading in the web of a beam or in a torqued
tube giving rise to compressive buckling stresses at 45o to the longitudinal direction as
predicted by Mohr's circle.
The behaviour of a compressed shell after buckling is quite different to that of a
plate; in this case an unstable ( negative ) stiffness is accompanied by a sudden
reduction of load capacity.

Since the displacements are uncontrolled in most practical systems, shells behave in a
snap- buckling mode - ie. as an increasing load reaches the bifurcation point, the
cylinder must undergo an instantaneous increase in deflection ( "snap" ) to the
point 1 in order to accomodate the increasing load. A subsequent decrease in load is
accomodated by a corresponding decrease in buckling deflection until the point 2 is
reached whereupon the structure again snaps instantaneously - this time back to the
point 3 on the primary path.
Clearly this behaviour makes it imperative in design to apply large safety factors to
the theoretical buckling loads of compressed cylinders.

It has been noted that a pressure


vessel head is subjected to
a compressive hoop stress at its
junction with the cylinder.

The two photographs here (from


Ramm op cit) show both inward
and outward buckles arising from
this compression in the
torispherical heads of internally
pressurised 3 m diameter stainless
steel vessels.

Longitudinal stresses in a
vertical cylinder may also
promote buckling as these two
photographs illustrate (from
Rhodes & Walker op cit).
Warning of impending failure of
the 7.3 m diameter vitreous
enamelled silo on the left is
provided by the visible buckles.
Grain pours out of the buckled
bin on the right - the ladder gives
an idea of the bin size.
Torsional buckling of columns can arise when a section under compression is very
weak in torsion, and leads
to the column rotating
about the force axis.

More commonly, where the section does not possess two axes of symmetry as in the
case of an angle section, this rotation is accompanied by bending and is known as
flexural torsional buckling.
Lateral buckling of beams is possible when a beam is stiff in the bending plane but
weak in the transverse plane and in torsion, as is the I-beam of the sketch.

It often happens that a system is prone to buckling in various modes. These usually
interact to reduce the load capacity of the system compared to that under the buckling
modes individually. An example of mode interaction is the thin box section which
develops local buckles at an early stage of loading, as shown greatly exaggerated
here.
The behaviour of the column is influenced by these local buckles, and gross column
buckle will occur at a load much less than the ideal Euler load. The Steel Structures
Code, AS 1250 op cit. sets out rules for the avoidance of mode interaction in large
components, and its guidelines should be followed in design.

Buckling has mixed blessings in automotive


applications.
The photograph on the left illustrates how
local buckling of a car's thin-walled A-pillar
dramatically reduces passenger cell integrity
in the event of roll-over.
Conversely, the energy absorbed by plastic
buckling can reduce significantly the injuries
suffered by a vehicle's occupants in the event
of a crash. The energy absorption capability of
thin- walled sections is demonstrated clearly
by the experiment photographed on the right.
(from Murray op cit)
The detailed analysis of most practical buckle-prone structures is too complex
mathematically to attempt here. We therefore examine instead some mechanisms
which demonstrate (un)stable behaviour similar to that of structures. The mechanisms
allow us to appreciate buckling behaviour and the tools used to analyse it, and to
introduce the concept of imperfections which must occur in practical components and
which have a relatively large effect on buckling behaviour and safety.
This work leads to the derivation of a design equation for practical columns, in which
the twin failure modes of strength and geometric instability invariably interact. This
interaction is apparent also in the behaviour of cracks.

Compression Members

Compression members, such as columns, are mainly subjected to axial forces. The
principal stress in a compression member is therefore the normal stress,

The failure of a short compression member resulting from the compression axial
force looks like,

EULER BUCKLING

However, when a compression member becomes longer, the role of the geometry
and stiffness (Young's modulus) becomes more and more important. For a long
(slender) column, buckling occurs way before the normal stress reaches the strength
of the column material. For example, pushing on the ends of a business card or
bookmark can easily reproduce the buckling.

For an intermediate length compression member, kneeling occurs when some areas
yield before buckling, as shown in the figure below.

In summary, the failure of a compression member has to do with the strength and
stiffness of the material and the geometry (slenderness ratio) of the member.
Whether a compression member is considered short, intermediate, or long depends
on these factors. More quantitative discussion on these factors can be found in the
next section.

Design Considerations

In practice, for a given material, the allowable stress in a compression member


depends on the slenderness ratio Leff / r and can be divided into three regions: short,
intermediate, and long.

Short columns are dominated by the strength limit of the material. Intermediate
columns are bounded by the inelastic limit of the member. Finally, long columns are
bounded by the elastic limit (i.e. Euler's formula). These three regions are depicted
on the stress/slenderness graph below,

Consider a long simply-supported column under an


external axial load F, as shown in the figure to the left.
The critical buckling load is given by Euler's formula,

where E is the Young's modulus of the column material, I is


the area moment of inertia of the cross-section, and L is the
length of the column.

Note that the critical buckling load decreases with the square
of the column length.

Extended Euler's Formula

In general, columns do not always terminate with simply-supported ends. Therefore,


the formula for the critical buckling load must be generalized.

The generalized equation takes the form of Euler's formula,


where the effective length of the column Leff depends on the boundary conditions. Some
common boundary conditions are shown in the schematics below:

The following table lists the effective lengths for columns terminating with a variety of
boundary condition combinations. Also listed is a mathematical representation of the
buckled mode shape.

Theoretical Engineering
Boundary Effective Effective
Buckling Mode Shape
Conditions Length Length
LeffT LeffE

Free-Free L (1.2·L)

Hinged-Free L (1.2·L)

Hinged-Hinged
(Simply- L L
Supported)

Guided-Free 2·L (2.1·L)

Guided-Hinged 2·L 2·L

Guided-Guided L 1.2·L

Clamped-Free
2·L 2.1·L
(Cantilever)

Clamped-Hinged 0.7·L 0.8·L

Clamped-Guided L 1.2·L

Clamped-Clamped 0.5·L 0.65·L


In the table, L represents the actual length of the column. The effective length is often
used in column design by design engineers.

Governing Equation for Elastic Buckling

Consider a buckled simply-supported column of length L under an external axial


compression force F, as shown in the left schematic below. The transverse
displacement of the buckled column is represented by w.

The right schematic shows the forces and moments acting on a cross-section in the
buckled column. Moment equilibrium on the lower free body yields a solution for the
internal bending moment M,

Recall the relationship between the moment M and the transverse


displacement w for an Euler-Bernoulli beam,

Eliminating M from the above two equations results in the governing equation for the
buckled slender column,

Buckling Solutions

The governing equation is a second order homogeneous ordinary differential


equation with constant coefficients and can be solved by the method of characteristic
equations. The solution is found to be,

where . The coefficients A and B can be determined by the two boundary


conditions , which yields,

The coefficient B is always zero, and for most values of m*L the coefficient A is
required to be zero. However, for special cases of m*L, A can be nonzero and the
column can be buckled. The restriction on m*L is also a restriction on the values for
the loading F; these special values are mathematically called eigenvalues. All other
values of F lead to trivial solutions (i.e. zero deformation).

The lowest load that causes buckling is called critical load (n = 1).

The above equation is usually called Euler's formula. Although Leonard Euler did
publish the governing equation in 1744, J. L. Lagrange is considered the first to
show that a non-trivial solution exists only when n is an integer. Thomas Young then
suggested the critical load (n = 1) and pointed out the solution was valid when the
column is slender in his 1807 book. The "slender" column idea was not quantitatively
developed until A. Considère performed a series of 32 tests in 1889.
The shape function for the buckled shape w(x) is mathematically called an
eigenfunction, and is given by,

Recall that this eigenfunction is strictly valid only for simply-supported columns.

Note: 1 Boundary conditions other than simply-supported will result in different


. critical loads and mode shapes.
2 The buckling mode shape is valid only for small deflections, where the
. material is still within its elastic limit.
3 The critical load will cause buckling for slender, long columns. In
. contrast, failure will occur in short columns when the strength of
material is exceeded. Between the long and short column limits, there is
a region where buckling occurs after the stress exceeds the proportional
limit but is still below the ultimate strength. These columns are classfied
asintermediate and their failure is called inelastic buckling.
4 Whether a column is short, intermediate, or long depends on its
. geometry as well as the stiffness and strength of its material.

intermediate Columns

The strength of a compression member (column) depends on its geometry


(slenderness ratio Leff / r) and its material properties (stiffness and strength).

The Euler formula describes the critical load


forelastic buckling and is valid only for long
columns. The ultimate compression strength of
the column material is not geometry-related and
is valid only for short columns.

In between, for a column


with intermediatelength, buckling occurs after the
stress in the column exceeds the proportional
limit of the column material and before the stress
reaches the ultimate strength. This kind of
situation is calledinelastic buckling.

This section discusses some commonly used


inelastic buckling theories that fill the gap
between short and long columns.

Tangent-Modulus Theory
Suppose that the critical stress σ t in an
intermediate column exceeds the proportional
limit of the material σ pl. The Young's modulus
at that particular stress-strain point is no
longerE. Instead, the Young's modulus
decreases to the local tangent value, Et.

Replacing the Young's modulus E in the Euler's


formula with the tangent modulus Et, the
critical load becomes,

The corresponding critical stress is,

Note: 1 The proportional limit σ pl, rather than the yield stress σ y, is used in the
. formula. Although these two are often arbitrarily interchangeable, the
yield stress is about equal to or slightly larger than the proportional limit
for common engineering materials. However, when the forming process
is taken into account, the residual stresses caused by processing can
not be neglected and the proportional limit may drop up to 50% with
respect to the yield stress in some wide-flange sections.
2 The tangent-modulus theory tends to underestimate the strength of the
. column, since it uses the tangent modulus once the stress on the
concave side exceeds the proportional limit while the convex side is still
below the elastic limit.
3 The tangent-modulus theory oversimplifies the inelastic buckling by
. using only one tangent modulus. In reality,the tangent modulus
depends on the stress, which is a function of the bending moment that
varies with the displacement w.

Reduced-Modulus Theory

The Reduced Modulus theory defines a reduced Young's modulus Er to compensate


for the underestimation given by the tangent-modulus theory.

For a column with rectangular cross section, the reduced modulus is defined by,

where E is the value of Young's modulus below the proportional limit.


Replacing E inEuler's formula with the reduced modulus Er, the critical load becomes,
The corresponding critical stress is,

Note: 1 The reduced-modulus theory tends to overestimate the strength of the


. column, since it is based on stiffness reversal on the convex side of the
column.
2 The reduced-modulus theory oversimplifies the inelastic buckling by
. using only one tangent modulus. In reality, the tangent modulus
depends on the stress which is a function of the bending moment that
varies with the displacement w.

Intermediate Column Design

Both tangent-modulus theory and reduced-


modulus theory were accepted theories of
inelastic buckling until F. R. Shanley published his
logically correct paper in 1946. The critical load of
inelastic buckling is in fact a function of the
transverse displacement w. According to
Shanley's theory, the critical load is located
between the critical load predicted by the
tangent-modulus theory (the lower bound) and
the reduced-modulus theory (the upper bound /
asymptotic limit).

However, the difference between Shanley's theory


and the tangent-modulus theory are not
significant enough to justify a much more complicated formula in practical
applications, especially when manufacturing defects in mass production and
geometric inaccuracies in assembly are taken into account.

Finally, if one must make and error in the design, engineers would much rather miss
on the safe side. This is the reason why many design formulas are based on the
overly-conservative tangent-modulus theory.

SOME PICTURES IN BUCKLING

Anda mungkin juga menyukai