Vous êtes sur la page 1sur 33

COLUMN

A column or pillar in architecture and structural engineering is an structural element that transmits,
through compression, the weight of the structure above to other structural elements below. For the purpose
of wind or earthquake engineering, columns may be designed to resist lateral forces. Other compression
members are often termed "columns" because of the similar stress conditions. Columns are frequently used to
support beams or arches on which the upper parts of walls or ceilings rest. In architecture, "column" refers to such a
structural element that also has certain proportional and decorative features. A column might also be a decorative
element not needed for structural purposes; many columns are "engaged with", that is to say form part of a wall.

Contents

[hide]

1 History

2 Structure

o 2.1 Nomenclature

o 2.2 Equilibrium, instability, and loads

o 2.3 Extensions

o 2.4 Foundations

3 Classical orders

o 3.1 Doric order

o 3.2 Tuscan order

o 3.3 Ionic order

o 3.4 Corinthian order

o 3.5 Composite order

4 Solomonic

5 Notable columns

6 See also

7 References

[edit]History

Equilibrium, instability, and loads


Main article: Buckling#columns
These are composed of stacked segments and finished in the Corinthian style (Temple of Bel, Syria)

[hide]

Materials failure modes

Buckling Corrosion Creep Fatigue

Fouling Fracture Hydrogen embrittlement

Impact Mechanical overload

Stress corrosion cracking Thermal

shock Wear Yielding

As the axial load on a perfectly straight slender column with elastic material properties is increased in magnitude, this
ideal column passes through three states: stable equilibrium, neutral equilibrium, and instability. The straight column
under load is in stable equilibrium if a lateral force, applied between the two ends of the column, produces a small
lateral deflection which disappears and the column returns to its straight form when the lateral force is removed. If the
column load is gradually increased, a condition is reached in which the straight form of equilibrium becomes so-called
neutral equilibrium, and a small lateral force will produce a deflection that does not disappear and the column remains
in this slightly bent form when the lateral force is removed. The load at which neutral equilibrium of a column is
reached is called the critical or buckling load. The state of instability is reached when a slight increase of the column
load causes uncontrollably growing lateral deflections leading to complete collapse.

For an axially loaded straight column with any end support conditions, the equation of static equilibrium, in the form of
a differential equation, can be solved for the deflected shape and critical load of the column. With hinged, fixed or free
end support conditions the deflected shape in neutral equilibrium of an initially straight column with uniform cross
section throughout its length always follows a partial or composite sinusoidal curve shape, and the critical load is
given by

where E = elastic modulus of the material, Imin = the minimal moment of inertia of the cross section, and L = actual
length of the column between its two end supports. A variant of (1) is given by

Table showing values of K for structural columns of various end conditions (adapted from Manual of Steel Construction, 8th edition,

American Institute of Steel Construction, Table C1.8.1)

where r = radius of gyration of [column]cross-section which is equal to the square root of (I/A), K = ratio of the longest
half sine wave to the actual column length, and KL = effective length (length of an equivalent hinged-hinged column).
From Equation (2) it can be noted that the buckling strength of a column is inversely proportional to the square of its
length.
When the critical stress, Fcr (Fcr =Pcr/A, where A = cross-sectional area of the column), is greater than the proportional
limit of the material, the column is experiencing inelastic buckling. Since at this stress the slope of the material's
stress-strain curve, Et (called the tangent modulus), is smaller than that below the proportional limit, the critical load at
inelastic buckling is reduced. More complex formulas and procedures apply for such cases, but in its simplest form
the critical buckling load formula is given as Equation (3),

where Et = tangent modulus at the stress Fcr

A column with a cross section that lacks symmetry may suffer torsional buckling (sudden twisting) before, or in
combination with, lateral buckling. The presence of the twisting deformations renders both theoretical analyses and
practical designs rather complex.

Eccentricity of the load, or imperfections such as initial crookedness, decreases column strength. If the axial load on
the column is not concentric, that is, its line of action is not precisely coincident with the centroidal axis of the column,
the column is characterized as eccentrically loaded. The eccentricity of the load, or an initial curvature, subjects the
column to immediate bending. The increased stresses due to the combined axial-plus-flexural stresses result in a
reduced load-carrying ability.

Column elements are considered to be massive if minimal side dimension is equal or more than 400 mm. Massive
columns have ability to increase concrete strength during long time period (even during exploitation period). Taking
into account possible loads onto structure increase in future (and even threat of progressive failure - terroristic
attacks, explosions etc.) - massive columns have advantage comparing with not ones. A little economy today has no
sense as usual for future. Moreover relatively small sections are not technological for reinforced structures during
their production. Balance between economy, mass of structures and so called "sustainable" construction is
necessary.

[edit]Extensions

Construction of Sigismund's Column inWarsaw, detail of the 1646 engraving.

When a column is too long to be built or transported in one piece, it has to be extended or spliced at the construction
site. A reinforced concrete column is extended by having the steel reinforcing bars protrude a few inches or feet
above the top of the concrete, then placing the next level of reinforcing bars to overlap, and pouring the concrete of
the next level. A steel column is extended by welding or bolting splice plates on the flanges and webs or walls of the
columns to provide a few inches or feet of load transfer from the upper to the lower column section. A timber column
is usually extended by the use of a steel tube or wrapped-around sheet-metal plate bolted onto the two connecting
timber sections.
when e<12 then it is known as short column

[edit]Foundations

A column that carries the load down to a foundation must have means to transfer the load without overstressing the
foundation material. Reinforced concrete and masonry columns are generally built directly on top of concrete
foundations. A steel column, when seated on a concrete foundation, must have a base plate to spread the load over
a larger area and thereby reduce the bearing pressure. The base plate is a thick rectangular steel plate usually
welded to the bottom end of the column.

[edit]Classical orders
Main article: Classical order

Church of San Prospero, Reggio Emilia,Italy

The Roman author Vitruvius, relying on the writings (now lost) of Greek authors, tells us that the
ancient Greeks believed that their Doric order developed from techniques for building in wood in which the earlier
smoothed tree trunk was replaced by a stone cylinder.

[edit]Doric order
Main article: Doric order

The Doric order is the oldest and simplest of the classical orders. It is composed of a vertical cylinder that is wider at
the bottom. It generally has neither a base nor a detailed capital. It is instead often topped with an inverted frustum of
a shallow cone or a cylindrical band of carvings. It is often referred to as the masculine order because it is
represented in the bottom level of the Colosseum and the Parthenon, and was therefore considered to be able to hold
more weight. The height-to-thickness ratio is about 8:1. The shaft of a Doric Column is always fluted.

The Greek Doric, developed in the western Dorian region of Greece, is the heaviest and most massive of the orders.
It rises from the stylobate without any base; it is from four to six times as tall as its diameter; it has twenty broad
flutes; the capital consists simply of a banded necking swelling out into a smooth echinus, which carries a flat square
abacus; the Doric entablature is also the heaviest, being about one-fourth the height column. The Greek Doric order
was not used after c. 100 B.C. until its rediscovery in the mid-eighteenth century.

[edit]Tuscan order
Main article: Tuscan order

The Tuscan order, also known as Roman Doric, is also a simple design, the base and capital both being series of
cylindrical disks of alternating diameter. The shaft is almost never fluted. The proportions vary, but are generally
similar to Doric columns. Height to width ratio is about 7:1.

[edit]Ionic order
Main article: Ionic order

The Ionic column is considerably more complex than the Doric or Tuscan. It usually has a base and the shaft is often
fluted (it has grooves carved up its length). On the top is a capital in the characteristic shape of a scroll, called
a volute, or scroll, at the four corners. The height-to-thickness ratio is around 9:1. Due to the more refined proportions
and scroll capitals, the Ionic column is sometimes associated with academic buildings. Ionic style columns were used
on the second level of the Colosseum.

Ionic capital

[edit]Corinthian order
Main article: Corinthian order

The Corinthian order is named for the Greek city-state of Corinth, to which it was connected in the period. However,
according to the architectural historianVitruvius, the column was created by the sculptor Callimachus, probably
an Athenian, who drew acanthus leaves growing around a votive basket. In fact, the oldest known Corinthian capital
was found in Bassae, dated at 427 BC. It is sometimes called the feminine order because it is on the top level of the
Colosseum and holding up the least weight, and also has the slenderest ratio of thickness to height. Height to width
ratio is about 10:1.

[edit]Composite order
The Composite order draws its name from the capital being a composite of the Ionic and Corinthian capitals. The
acanthus of the Corinthian column already has a scroll-like element, so the distinction is sometimes subtle. Generally
the Composite is similar to the Corinthian in proportion and employment, often in the upper tiers of colonnades.
Height to width ratio is about 11:1 or 12:1.

[edit]Solomonic
Two of Constantine's Solomonic columns in their present day location on a pier in St Peter's, Rome. Part of Bernini's Baldachin, inspired by

the original columns, is in the foreground.

A Solomonic column, sometimes called "barley sugar", begins on a base and ends in a capital, which may be of any
order, but the shaft twists in a tight spiral, producing a dramatic, serpentine effect of movement. Solomonic columns
were developed in the ancient world, but remained rare there. A famous marble set, probably 2nd century, was
brought to St Peter's, Rome by Constantine I, and placed round the saint's shrine, and was thus familiar throughout
the Middle Ages, by which time they were thought to have been removed from the Temple of Jerusalem.[1] The style

Buckling
From Wikipedia, the free encyclopedia

For other uses, see Buckling (disambiguation).

[hide]

E
Materials failure modes

Buckling Corrosion Creep Fatigue

Fouling Fracture Hydrogen embrittlement

Impact Mechanical overload

Stress corrosion cracking Thermal

shock Wear Yielding

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

a bifurcation 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: an undeformed state or a laterally-deformed state.

In practice, buckling is characterized by a sudden failure of a structural member subjected to high compressive

stress, where the actual compressive stress at the point of failure is less than the ultimate compressive stresses that

the material is capable of withstanding. For example, during earthquakes, reinforced concrete members may

experience lateral deformation of the longitudinal reinforcing bars. This mode of failure is also described as failure

due toelastic instability. Mathematical analysis of buckling makes use of an axial load eccentricity that introduces a

moment, which does not form part of the primary forces to which the member is subjected. When load is constantly

being applied on a member, such as column, it will ultimately become large enough to cause the member to become

unstable. Further load will cause significant and somewhat unpredictable deformations, possibly leading to complete

loss of load-carrying capacity. The member is said to have buckled, to have deformed.

Contents

[hide]

1 Columns

o 1.1 Self-buckling

2 Buckling under tensile dead loading

3 Flutter instability

4 Limit point vs bifurcation buckling

5 Bicycle wheels

6 Surface materials

7 Energy method

8 Flexural-torsional buckling

9 Lateral-torsional buckling

10 Plastic buckling
11 Dynamic buckling

12 Buckling of thin cylindrical shells subject to axial loads

13 Buckling of pipes and pressure vessels subject to external overpressure

14 See also

15 References

16 External links

[edit]Columns

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.

A short concrete column is one having a ratio of unsupported length to least dimension of the cross section not
greater than 10. If the ratio is greater than 10, it is 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 thespecies 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 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 direct 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, = 2.0.

is the effective length of the column.

Examination of this formula reveals the following interesting facts with regard to the

load-bearing ability of slender columns.

1. Elasticity and not 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 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 (hinged at the ends).

Since the moment of inertia of a surface is its area multiplied by the square of a

length called the radius of gyration, 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 strut 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 of the critical buckling

load based on an initial (small) curvature. The Rankine Gordon formula is also

based on experimental results and suggests that a strut 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.


[edit]Self-buckling

A free-standing, vertical column, with density , Young's

modulus , and radius , 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 1.86635...

[edit]Buckling under tensile dead loading

Fig. 1: Single-degree-of-freedom structure showing buckling under tensile dead

loading.

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. 1, 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. 2. 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.

[edit]Flutter instability

Structures subject to a follower (nonconservative) load may suffer

instabilities which are not of the buckling type and therefore are not

detectable with a static approach. For instance, the so-called 'Ziegler

column' is shown in Fig.3.

Fig.3: A sketch of the 'Ziegler column', a two-degree-of-freedom 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 ( ) and

divergence ( ),

where and .

Fig.4: A sequence of deformed shapes at consecutive times

intervals of the structure sketched in Fig.3 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.4 (upper part) together with the divergence

instability (lower part) consisting in an exponential

growth.

Recently, Bigoni and Noselli (2011)[5] have

experimentally shown that flutter and divergence

instabilities can be directly related to dry friction,

watch themovie for more details.

[edit]Limit point vs bifurcation


buckling
Bifurcation buckling[6][7] is sometimes called Euler

buckling even when applied to structures other

than Euler columns. As 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.

In structures experiencing limit point instability, if

the load is increased infinitesimally beyond the


critical load, the structure undergoes a large

deformation into a different stable configuration

which is not adjacent to the original configuration.

An example of this type of buckling is a toggle

frame (pictured) which 'snaps' into its buckled

configuration.
[edit]Bicycle wheels

A conventional bicycle wheel consists of a thin

rim kept under high compressive stress by the

(roughly normal) inward pull of a large number of

spokes. It can be considered as a loaded column

that has been bent into a circle. As such, if spoke

tension is increased beyond a safe level, the

wheel spontaneously fails into a characteristic

saddle shape (sometimes called a "taco" or a

"pringle") like a three-dimensional Euler column.

This is normally a purely elastic deformation and

the rim will resume its proper plane shape if


spoke tension is reduced slightly.

[edit]Surface materials
Sun kink in rail tracks

Buckling is also a failure mode

in pavement materials, primarily with concrete,

since asphalt is more flexible. Radiant heat from

the sun is absorbed in the road surface, causing it

to expand, forcing adjacent pieces to push

against each other. If the stress is great enough,

the pavement can lift up and crack without

warning. Going over a buckled section can be

very jarring to automobile drivers, described as

running over a speed hump at highway speeds.

Similarly, rail tracks also expand when heated,

and can fail by buckling, a phenomenon

called sun kink. It is more common for rails to

move laterally, often pulling the underlain

railroad ties (sleepers) along .

[edit]Energy method

Often it is very difficult to determine the exact

buckling load in complex structures using the

Euler formula, due to the difficulty in deciding the

constant K. Therefore, maximum buckling load

often is approximated using energy conservation.

This way of deciding maximum buckling load is

often referred to as the energy method in

structural analysis.

The first step in this method is to suggest a

displacement function. This function must satisfy

the most important boundary conditions, such as

displacement and rotation. The more accurate the

displacement function, the more accurate the

result.
In this method, there are two equations used (for

small deformations) to approximate the "inner"

energy (the potential energy stored in elastic

deformation of the structure) and "outer" energy

(the work done on the system by external forces).

where is the displacement

function and the subscripts

and refer to the first and second

derivatives of the displacement. Energy

conservation yields:

[edit]Flexural-torsional

buckling

Occurs in compression members

only and it can be described as a

combination of bending and

twisting of a member. And it must

be considered for design purposes,


since the shape and cross sections

are very critical. This mostly occurs

in channels, structural tees,

double-angle shapes, and equal-

leg single angles.

[edit]Lateral-torsional

buckling

When a simple beam is loaded

in flexure, the top side is


in compression, and the bottom

side is in tension. When a slender


member is subjected to an axial

force, failure takes place due to

bending or torsion rather than

direct compression of the material.

If the beam is not supported in the

lateral direction (i.e., perpendicular

to the plane of bending), and the

flexural load increases to a critical

limit, the beam will fail due to

lateral buckling of the compression

flange. In wide-flange sections, if

the compression flange buckles

laterally, the cross section will also

twist in torsion, resulting in a failure


mode known as lateral-torsional

buckling.

[edit]Plastic buckling

Buckling will generally occur

slightly before the theoretical

buckling strength of a structure,

due to plasticity of the material.

When the compressive load is near

buckling, the structure will bow

significantly and approach yield.

The stress-strain behaviour of

materials is not strictly linear even

below yield, and the modulus of

elasticity decreases as stress

increases, with more rapid change

near yield. This lower rigidity

reduces the buckling strength of

the structure and causes

premature buckling. This is the


opposite effect of the plastic

bending in beams, which causes


late failure relative to theEuler-

Bernoulli beam equation.

[edit]Dynamic buckling

If the load on the column is applied

suddenly and then released, the

column can sustain a load much

higher than its static (slowly

applied) buckling load. This can

happen in a long, unsupported

column (rod) used as a drop

hammer. The duration of

compression at the impact end is

the time required for a stress wave

to travel up the rod to the other

(free) end and back down as a

relief wave. Maximum buckling

occurs near the impact end at a

wavelength much shorter than the

length of the rod, at a stress many

times the buckling stress if the rod

were a statically-loaded column.

The critical condition for buckling

amplitude to remain less than

about 25 times the effective rod

straightness imperfection at the

buckle wavelength is

where is the impact

stress, is the length of

the rod, is the elastic

wave speed, and is the

smaller lateral dimension of a


rectangular rod. Because the

buckle wavelength depends

only on and , this

same formula holds for thin

cylindrical shells of

thickness .[8]

[edit]Bucklingof thin
cylindrical shells
subject to axial loads

Solutions of Donnel's eight

order differential equation

gives the various buckling

modes of a thin cylinder under

compression. But this

analysis, which is in

accordance with the small

deflection theory gives much

higher values than shown

from experiments. So it is

customary to find the critical

buckling load for various

structures which are

cylindrical in shape from pre-

existing design curves where

critical buckling load Fcr is

plotted against the ratio R/t,

where R is the radius and t is

the thickness of the cylinder

for various values of L/R, L the

length of the cylinder. If cut-

outs are present in the

cylinder, critical buckling loads


as well as pre-buckling modes

will be affected. Presence or


absence of reinforcements of

cut-outs will also affect the

buckling load.

[edit]Buckling
of pipes
and pressure
vessels subject to
external
overpressure

Pipes and pressure vessels

subject to external

overpressure, caused for

example by steam cooling

down and condensating into

water with subsequent

massive pressure drop, risk

buckling due to

compressive hoop stresses.

Design rules for calculation of

the required wall thickness or

reinforcement rings are given

in various piping and pressure

vessel codes.

[edit]See also

Perry Robertson formula

Wood method

[edit]References

1. ^ Kato, K. (1915).

"Mathematical

Investigation on the

Mechanical Problems

of Transmission

Line". Journal of the


Japan Society of

Mechanical

Engineers 19: 41.

2. ^ Ratzersdorfer, Julius

(1936). Die

Knickfestigkeit von

Stben und

Stabwerken. Wein,

Austria: J. Springer.

pp. 107109.

3. ^ Cox, Steven J.; C.

Maeve McCarthy

(1998). "The Shape of

the Tallest

Column". Society for

Industrial and Applied

Mathematics 29: 547

554.

4. ^ D. Zaccaria, D.

Bigoni, G. Noselli and

D. Misseroni Structures

buckling under tensile

dead load.

Proceedings of the

Royal Society A, 2011,

467, 1686-1700.

5. ^ D. Bigoni and G.

Noselli, Experimental

evidence of flutter and

divergence instabilities

induced by dry friction.

Journal of the

Mechanics and Physics

of Solids, 2011, 59,

22082226.
6. ^ "Buckling of Bars,

Plates, and Shells" By

Robert M. Jones

7. ^ "Observations on

eigenvalue buckling

analysis within a finite

element context" by

Christopher J. Earls

8. ^ Lindberg, H. E., and

Florence, A.

L., Dynamic Pulse

Buckling, Martinus

Nijhoff Publishers,

1987, pp. 1156, 297

298.

Timoshenko, S. P., and

Gere, J. M., Theory of

Elastic Stability, 2 ed.,

McGraw-Hill, 1961.

Nenezich,
M., Thermoplastic

Continuum Mechanics,

Journal of Aerospace
Structures, Vol. 4, 2004.

The Stability of Elastic

Equilibrium by W. T.
Koiter, PhD Thesis, 1945.

Dhakal Rajesh and Koichi

Maekawa (October

2002). "Reinforcement

Stability and Fracture of

Cover Concrete in

Reinforced Concrete
Members. [1]
Willian T. Segui (2007).

Steel Design Fourth

Edition. United States.


Chris Carson.

Analysis and design of

flight vehicle structures-


E.F.Brune

[edit]External links

The complete theory and

example experimental

results for long columns

are available as a 39-

page PDF document

at http://lindberglce.com/t
ech/buklbook.htm

Video on Buckling under


tensile dead loading

Laboratory for Physical

Modeling of Structures

and Photoelasticity

(University of Trento,
Italy)

http://www.midasuser.co

m.tw/t_support/tech_pds/f

iles/Tech%20Note-

Lateral%20Torsional%20
Buckling.pdf

View page ratings


Rate this page

What's this?

Trustworthy

Objective
Complete

Well-written

I am highly

knowledgeable about this

topic (optional)

Submit ratings

Categories:

Elasticity (physics)
Materials science
Mechanical failure modes
Structural analysis

Log in / create account


Article
Talk
Read
Edit
View history

Main page
Contents
Featured content
Current events
Random article
Donate to Wikipedia
Interaction
Help
About Wikipedia
Community portal
Recent changes
Contact Wikipedia
Toolbox
Print/export
Languages

Deutsch

Espaol

Franais
Italiano

Magyar
Nederlands

Polski
Portugus

Suomi
Svenska

This page was last modified on 24

January 2012 at 04:44.

Text is available under the Creative

Commons Attribution-ShareAlike

License; additional terms may apply.

See Terms of use for details.

Wikipedia is a registered

trademark of the Wikimedia

Foundation, Inc., a non-profit

organization.

Contact us

Privacy policy

About Wikipedia

Disclaimers

Mobile view


The slenderness ratio is the ratio between the height or length of a structural element (such as
a column, or strut) and the width or thickness of the element.

For example, if a rectangular column is 6m high, and 400mm by 600mm in cross-section, then
its slenderness is 6000/600 = 10 in one direction and 6000/400 = 15 in the other direction.

The higher the slenderness ratio, the more slender the structural element is. How slender a
structural element is allowed to be depends upon the material it is made from. Steel can be
more slender than concrete, for example.

In structural engineering calculations, the slenderness is often denoted as the element's


"effective" length divided by something called the radius of gyration. The radius of gyration is
a measure of the average distance of the material from the centroid (centre of gravity) of the
element's cross section. This can be calculated as r = (I/A)0.5, where I is the second moment
of area, or second moment or inertia, of the cross section and A is the area of the cross
section. The second moment of area is described here:
http://en.wikipedia.org/wiki/Second_moment_of_area.

The effective length of an element is determined by how it is fixed at its ends. The effective
length is the length of the column that will form half a sine wave if it buckles. If it is "pinned",
or has hinged ends, the effective length is the true length of the element. If it is a cantilever
(fixed at one end but free at the other), the effective length is twice the true length. If it is
fully fixed at both ends the effective length is 0.7 times the true length, but this is in reality
very difficult to achieve, so often a real structural element is considered to be only nominally
fixed and the effective length is taken to be 0.85 times the true length. See
http://www.corusconstruction.com/en/reference/teaching_resources/architectural_studio_refe
rence/elements/design_of_columns_and_struts/column_end_fixity/ for some photos and
further explanation.

Lecture 30:

Radius of Gyration & Buckling

The radius of gyration (r) describes the way in which the area of a cross-section is
distributed around its centroidal axis. If the area is concentrated far from the
centroidal axis it will have a greater value of r and a greater resistance to buckling. A
cross-section can have more than one radius of gyration and most sections have at
least two. If this is the case, the section tends to buckle around the axis with the
smallest value. The radius of gyration is defined as:

r = sqrt (I/A)
where

r = radius of gyration
I = moment of inertia
A = area of the cross section

All things being equal, a circular pipe is the most efficient column section to resist
buckling. This is because it has an equal radius of gyration in all directions and it has
the its area distributed as far away as possible from the centroid.

The steel columns shown below all have areas of 3-1/8 in2. The safe loads for an 8 ft
length are shown. The only difference between them is the way in which the cross-
sectional area is distributed about the centroid.

Buckling
Buckling is very similar to bending. Thus, the shape of the cross-section is very
important. The shape of the column also effects the way in which it will buckle.
Imagine for a moment a single sheet of paper (A4 or 8.5 x 11). If one would try to
simply stand it on edge it would be impossible unless the paper was folded. This
simple act of folding the paper actually increases the cross-sectional moment of
Inertia and thus the stiffness of the newly formed column. The stiffness of another
paper column could be futher increased by taking the paper and taping the long
edges together to create a tube. Now, the paper will be very stiff since the material
of the paper is distributed evenly as far away from the neutral axis as possible.

The load at which a column will begin to buckle is known as the Critical Buckling
Load (or critical load). A number of qualities of the column must be known in order to
determine this value. The Swiss mathematician Leonard Euler (1707 - 1783) derived a
formula in 1744 (known as the Euler Buckling Formula) to determine the load at
which a perfect column will buckle. It was a very important step in the history of
technology and remains important for column design today. The equation is only
accurate for columns which approach the perfect conditions for which he derrived the
equation.

Ncr = 2E I / (lk2)

In which the terms are defined as follows:

Ncr = Critical Buckling Load


E = Modulus of Elasticity
I = Moment of Inertia
lk = Effective Buckling Length

This equation can be modified by dividing both sides by the area of the column so
that the stress at which the column will buckle can be determined:

fcr = sigmacr = Ncr / A


= 2 E I /A(lk2)

now, knowing that r2 = I/A, this equation becomes:

= 2 E /(lk/r)2

Lambda, or the slenderness ratio is a value with which one can gage the relative
resistance of a column cross-section to buckling. Or, stated otherwise, the relative
ease in which a column WILL buckle. It is defined as

lambda = lk / r
Where lk is the buckling length and r is the radius of gyration.

Thus, the critical buckling stress can be expressed as,

fcr= 2 E / (lambda)2

The buckling length of a column depends on its physical length and its end
conditions. Euler discovered that if a column is hinged at both ends it will buckle in
the form of a sine curve with the inflection points at the hinges. This would be the
case in which the buckling length of a column is identical to its length. This is not the
case if the ends of the column are both fixed. The determination of the buckling
lengths for various column end conditions and frames is given below:

The magnitude of the internal forces is important to know in order to size a column. A
small amount of tensile stress has little effect on a wood or steel column, but could
cause problems in a concrete or masonry column.

Questions for Thought


What are the relationships between the various columnar elements within the
human skeleton? What are the various end conditions? What is the buckling length
of the columns in a fassade near you?
Additional Reading
Schodeck, Daniel. "Structures." Chapter 7.

Copyright 1995, 1996, 1997 by Chris H. Luebkeman and Donald Peting

Vous aimerez peut-être aussi