Sunteți pe pagina 1din 13

In science, buckling is a mathematical instability, leading to a failure mode.

Theoretically, buckling is
caused by abifurcation in the solution to the equations of static equilibrium. At a certain stage under
an increasing load, further load is able to be sustained in one of two states of equilibrium: a purely
compressed state (with no lateral deviation) or a laterally-deformed state.
Buckling is characterized by a sudden sideways failure of a structural member subjected to
high compressive stress, where the compressive stress at the point of failure is less than the
ultimate compressive stress that the material is capable of withstanding. Mathematical analysis of
buckling often makes use of an "artificial" axial load eccentricity that introduces a secondary bending
moment that is not a part of the primary applied forces being studied. As an applied load is increased
on a member, such as a column, it will ultimately become large enough to cause the member to
become unstable and is said to have buckled. Further load will cause significant and somewhat
unpredictable deformations, possibly leading to complete loss of the member's load-carrying
capacity. If the deformations that follow buckling are not catastrophic the member will continue to
carry the load that caused it to buckle. If the buckled member is part of a larger assemblage of
components such as a building, any load applied to the structure beyond that which caused the
member to buckle will be redistributed within the structure.
Contents
[hide]

1 Columns
o

1.1 Self-buckling

2 Buckling under tensile dead loading

3 Constraints, curvature and multiple buckling

4 Flutter instability

5 Various forms of buckling

6 Bicycle wheels

7 Surface materials

8 Cause

9 Accidents

10 Energy method

11 Flexural-torsional buckling

12 Lateral-torsional buckling
o

12.1 The modification factor (Cb)

13 Plastic buckling

14 Dynamic buckling

15 Buckling of thin cylindrical shells subject to axial loads

16 Buckling of pipes and pressure vessels subject to external overpressure

17 See also

18 References

19 External links

Columns[edit]

A column under a concentric axial load exhibiting the characteristic deformation of buckling

The eccentricity of the axial force results in a bending moment acting on the beam element.

The ratio of the effective length of a column to the least radius of gyration of its cross section is
called the slenderness ratio (sometimes expressed with the Greek letter lambda, ). This ratio
affords a means of classifying columns. Slenderness ratio is important for design considerations. All
the following are approximate values used for convenience.

A short steel column is one whose slenderness ratio does not exceed 50; an intermediate
length steel column has a slenderness ratio ranging from about 50 to 200, and are dominated by
the strength limit of the material, while a long steel column may be assumed to have a
slenderness ratio greater than 200 and its behavior is dominated by the modulus of elasticity of
the material.

A short concrete column is one having a ratio of unsupported length to least dimension of the
cross section equal to or less than 10. If the ratio is greater than 10, it is considered a long
column (sometimes referred to as a slender column).

Timber columns may be classified as short columns if the ratio of the length to least
dimension of the cross section is equal to or less than 10. The dividing line between intermediate
and long timber columns cannot be readily evaluated. One way of defining the lower limit of long
timber columns would be to set it as the smallest value of the ratio of length to least cross
sectional area that would just exceed a certain constant K of the material. Since K depends on
the modulus of elasticity and the allowable compressive stress parallel to the grain, it can be
seen that this arbitrary limit would vary with the species of the timber. The value of K is given in
most structural handbooks.

If the load on a column is applied through the center of gravity (centroid) of its cross section, it is
called an axial load. A load at any other point in the cross section is known as an eccentric load. A
short column under the action of an axial load will fail by direct compression before it buckles, but a
long column loaded in the same manner will fail by buckling (bending), the buckling effect being so
large that the effect of the axial load may be neglected. The intermediate-length column will fail by a
combination of direct compressive stress and bending.
In 1757, mathematician Leonhard Euler derived a formula that gives the maximum axial load that a
long, slender, ideal column can carry without buckling. An ideal column is one that is perfectly
straight, homogeneous, and free from initial stress. The maximum load, sometimes called the critical
load, causes the column to be in a state of unstable equilibrium; that is, the introduction of the
slightest lateral force will cause the column to fail by buckling. The formula derived by Euler for
columns with no consideration for lateral forces is given below. However, if lateral forces are taken
into consideration the value of critical load remains approximately the same.

where
= maximum or critical force (vertical load on column),
= modulus of elasticity,
= area moment of inertia,

= unsupported length of column,


= column effective length factor, whose value depends on the conditions of end support
of the column, as follows.
For both ends pinned (hinged, free to rotate),
= 1.0.
For both ends fixed,
= 0.50.
For one end fixed and the other end pinned,
= 0.699....
For one end fixed and the other end free to move laterally,
is the effective length of the column.

= 2.0.

Examination of this formula reveals the following interesting facts with


regard to the load-bearing ability of slender columns.
1. Elasticity and not the compressive strength of the materials of
the column determines the critical load.
2. The critical load is directly proportional to the second moment
of area of the cross section.
3. The boundary conditions have a considerable effect on the
critical load of slender columns. The boundary conditions
determine the mode of bending and the distance between
inflection points on the deflected column. The inflection points
in the deflection shape of the column are the points at which
the curvature of the column change sign and are also the
points at which the internal bending moments are zero. The
closer together the inflection points are, the higher the resulting
capacity of the column.

A demonstration model illustrating the different "Euler" buckling modes. The


model shows how the boundary conditions affect the critical load of a slender
column. Notice that each of the columns are identical, apart from the boundary
conditions.

The strength of a column may therefore be increased by distributing the


material so as to increase the moment of inertia. This can be done
without increasing the weight of the column by distributing the material
as far from the principal axis of the cross section as possible, while
keeping the material thick enough to prevent local buckling. This bears
out the well-known fact that a tubular section is much more efficient
than a solid section for column service.
Another bit of information that may be gleaned from this equation is the
effect of length on critical load. For a given size column, doubling the
unsupported length quarters the allowable load. The restraint offered by
the end connections of a column also affects the critical load. If the
connections are perfectly rigid, the critical load will be four times that for
a similar column where there is no resistance to rotation (in which case
the column is idealized as having hinges at the ends).
Since the radius of gyration is defined as the square root of the ratio of
the column's moment of inertia about an axis to cross sectional area,
the above formula may be rearranged as follows. Using the Euler
formula for hinged ends, and substituting Ar2 for I, the following formula
results.

where

is the allowable stress of the column, and

is the

slenderness ratio.
Since structural columns are commonly of intermediate length, and
it is impossible to obtain an ideal column, the Euler formula on its
own has little practical application for ordinary design. Issues that
cause deviation from the pure Euler column behaviour include
imperfections in geometry in combination with plasticity/non-linear
stress strain behaviour of the column's material. Consequently, a
number of empirical column formulae have been developed to
agree with test data, all of which embody the slenderness ratio. For
design, appropriate safety factors are introduced into these
formulae. One such formula is the Perry Robertson formula which
estimates the critical buckling load based on an initial (small)
curvature. The Rankine Gordon formula (Named for William John
Macquorn Rankine and Perry Hugesworth Gordon (1899 1966))

is also based on experimental results and suggests that a column


will buckle at a load Fmax given by:

where Fe is the Euler maximum load and Fc is the maximum


compressive load. This formula typically produces a
conservative estimate of Fmax.

Self-buckling[edit]
A free-standing, vertical column, with density
modulus

, and cross-sectional area

, Young's

, will buckle under its

own weight if its height exceeds a certain critical height:[1][2][3]

where g is the acceleration due to gravity, I is the second


moment of area of the beam cross section, and B is the
first zero of the Bessel function of the first kind of order
-1/3, which is equal to 1.86635086...

Buckling under tensile dead


loading[edit]

Fig. 2: Elastic beam system showing buckling under tensile


dead loading.

Usually buckling and instability are associated to


compression, but recently Zaccaria, Bigoni, Noselli and
Misseroni (2011)[4] have shown that buckling and instability
can also occur in elastic structures subject to dead tensile
load. An example of a single-degree-of-freedom structure
is shown in Fig. 2,[where?] where the critical load is also
indicated. Another example involving flexure of a structure
made up of beam elements governed by the equation of

the Euler's elastica is shown in Fig.3. In both cases, there


are no elements subject to compression. The instability and
buckling in tension are related to the presence of the slider,
the junction between the two rods, allowing only relative
sliding between the connected pieces. Watch a movie for
more details.

Constraints, curvature and


multiple buckling[edit]

Fig. 3: A one-degree-of-freedom structure exhibiting a tensile


(compressive) buckling load as related to the fact that the
right end has to move along the circular profile labeled 'Ct'
(labelled 'Cc').

Buckling of an elastic structure strongly depends on the


curvature of the constraints against which the ends of the
structure are prescribed to move (see Bigoni, Misseroni,
Noselli and Zaccaria, 2012[5]). In fact, even a single-degreeof-freedom system (see Fig.3) may exhibit a tensile (or a
compressive) buckling load as related to the fact that one
end has to move along the circular profile labeled 'Ct'
(labelled 'Cc').

Fig. 4: A one-degree-of-freedom structure with a 'S'-shaped


bicircular profile exhibiting multiple bifurcations (both tensile
and compressive).

The two circular profiles can be arranged in a 'S'-shaped


profile, as shown in Fig.4; in that case a discontinuity of the
constraint's curvature is introduced, leading to multiple
bifurcations. Note that the single-degree-of-freedom
structure shown in Fig.4 has two buckling loads (one

tensile and one compressive). Watch a movie for more


details.

Flutter instability[edit]
Structures subject to a follower (nonconservative)
load[clarification needed] may suffer instabilities which are not of the
buckling type and therefore are not detectable with a static
approach.[6] For instance, the so-called 'Ziegler column' is
shown in Fig.5.

Fig. 5: A sketch of the 'Ziegler column', a two-degree-offreedom system subject to a follower load (the force P
remains always parallel to the rod BC), exhibiting flutter and
divergence instability. The two rods, of linear mass density ,
are rigid and connected through two rotational springs of
stiffness k1 and k2.

This two-degree-of-freedom system does not display a


quasi-static buckling, but becomes dynamically unstable.
To see this, we note that the equations of motion are

and their linearized version is

Assuming a time-harmonic solution in the form

we find the critical loads for flutter (


divergence (

) and

),

where

and

Fig. 6: A sequence of deformed shapes at


consecutive times intervals of the structure
sketched in Fig.5 and exhibiting flutter
(upper part) and divergence (lower part)
instability.

Flutter instability corresponds to a


vibrational motion of increasing amplitude
and is shown in Fig.6 (upper part) together
with the divergence instability (lower part)
consisting in an exponential growth.
Recently, Bigoni and Noselli (2011)[7] have
experimentally shown that flutter and
divergence instabilities can be directly
related to dry friction, watch the movie for
more details.

Various forms of
buckling[edit]
Buckling is a state which defines a point
where an equilibrium configuration

becomes unstable under a parametric


change of load and can manifest itself in
several different phenomena. All can be
classified as forms of bifurcation.
There are four basic forms of bifurcation
associated with loss of structural stability
or buckling in the case of structures with a
single degree of freedom. These comprise
two types of pitchfork bifurcation,
one saddle-node bifurcation (often referred
to as a limit point) and one transcritical
bifurcation. The pitchfork bifurcations are
the most commonly studied forms and
include the buckling of columns and struts,
sometimes known as Euler buckling; the
buckling of plates, sometimes known as
local buckling, which is well known to be
relatively safe (both are supercritical
phenomena) and the buckling of shells,
which is well-known to be a highly
dangerous (subcritical phenomenon).
[8]

Using the concept of potential energy,

equilibrium is defined as a stationary point


with respect to the degree(s) of freedom of
the structure. We can then determine
whether the equilibrium is stable, if the
stationary point is a local minimum; or
unstable, if it is a maximum, point of
inflection or saddle point (for multipledegree-of-freedom structures only) see
animations below.

Archetypal rigid link models with a single degree of f

basic buckling phenomena (see bifurcation diagrams

corresponding to q

Truss with spring tie (model shallow


tied arch).

Link-strut with rotational spring.

Link-strut with transverse


translational spring.

Animations of the variation of total potential energy (re

generic structural systems with the indicated bi

Two saddle-node bifurcations (limit


points).

Supercritical pitchfork bifurcation


(stable-symmetric buckling point).

Subcritical pitchfork bifurcation


(unstable-symmetric buckling
point).

Transcritical bifurcation (asymmetric


buckling point).

In Euler buckling,[9][10] the applied load is


increased by a small amount beyond the
critical load, the structure deforms into a
buckled configuration which is adjacent to
the original configuration. For example, the
Euler column pictured will start to bow
when loaded slightly above its critical load,
but will not suddenly collapse.

S-ar putea să vă placă și