Vous êtes sur la page 1sur 37

Basic Analysis of Shell Structures

OBJECTIVE/SCOPE:

To describe the basic characteristics of pre- and post-buckling shell behaviour and to explain and
compare the differences in behaviour with that of plates and bars.

PREREQUISITES

Lecture 8.6: Introduction to Shell Structures

RELATED LECTURES:

Lecture 6.1: Concepts of Stable and Unstable Elastic Equilibrium

Lecture 8.1: Introduction to Plate Behaviour and Design

Lecture 8.4.1: Plate Girder Behaviour & Design I

Lectures 8.5.1: Introduction to Design of Box Girders

SUMMARY:

The combined bending and stretching behaviour of shell structures in resisting load is discussed; their
buckling behaviour is also explained and compared with that of struts and plates. The effect of
imperfections is examined and ECCS curves, which can be used in design, are given. Reference is also
made to available computer programs that can be used for shell analysis.

1. INTRODUCTION

Lecture 8.6 introduced several aspects of the structural behaviour of shells in an essentially qualitative
way. Before moving on to consider design procedures for specific applications, it is necessary to gain
some understanding of the possible approaches to the analysis of shell response. It should then be
possible to appreciate the reasoning behind the actual design procedures covered in Lectures 8.8 and
8.9.

This Lecture, therefore, presents the main principles of shell theory that underpin the ECCS design
methods for unstiffened and stiffened cylinders. Comparisons are drawn with the behaviour of columns
and plates previously discussed in Lectures 6.6.1, 6.6.2 and 8.1.

2. BENDING AND STRETCHING OF THIN SHELLS

The deformation of an element of a thin shell consists of the curvatures and normal displacements
associated with out-of-surface bending and the stretching and shearing of the middle surface. Bending
deformation without stretching of the middle surface, as assumed in the small deflection theory for flat
plates, is not possible, and so both bending and stretching strains must be considered.

If the shape and the boundary conditions of a shell and the applied loads are such that the loads can be
resisted by membrane forces alone, then these forces may be found from the three equilibrium conditions
for an infinitely small element of the shell. The equilibrium equations may be obtained from the equilibrium
of forces in three directions; that is, in the two principal directions of curvature and in the direction normal
to the middle surface. As a result, the three membrane forces can be obtained easily in the absence of
bending and twisting moments and shear forces perpendicular to the surface. An example is an
unsupported cylindrical shell subjected to uniform radial pressure over its entire area (Figure 1).
Obviously, the only stress generated by the external pressure is a circumferential membrane stress. The
assumption does not hold if the cylinder is subjected to two uniform line loads acting along two
diametrically opposed generators (Figure 2). In this case, bending theory is required to evaluate the
stress distribution, because an element of the shell cannot be in equilibrium without circumferential
bending stresses. Circumferential bending stresses are essential to resist the external loads, and
because the wall is thin and has very little flexural resistance, they greatly affect its load carrying
resistance [1].
Significant bending stresses usually only occur close to the boundaries, or in the zone affected by other
disturbances, such as local loads or local imperfections. Locally, the resulting stresses may be quite high,
but they generally diminish at a small distance from the local disturbance. Bending stresses may,
however, cause local yielding which can be very dangerous in the presence of repeated loadings, since it
can result in a fatigue fracture.

It is normally more structurally efficient if a shell structure can be configured in such a way that it carries
load primarily by membrane action. Simpler design calculations will usually also result.

3. BUCKLING OF SHELLS - LINEAR AND NON-LINEAR BUCKLING THEORY

Buckling may be regarded as a phenomenon in which a structure undergoes local or overall change in
configuration. For example, an originally straight axially loaded column will buckle by bowing laterally;
similarly a cylinder may buckle when its surface crumples under the action of external loads. Buckling is
particularly important in shell structures since it may well occur without any warning and with catastrophic
consequences [2-4].

The equations for determining the load at which buckling is initiated, through bifurcation on the main
equilibrium path of a cylindrical shell, may be derived by means of the adjacent equilibrium criterion, or,
alternatively, by use of the minimum potential energy criterion. In the first case, small increments (u1, v1,
w1) are imposed on the pre-buckling displacements (u0, v0, w0)

u = u0 + u1

v = v0 + v1 (1)
w = w0 + w1

The two adjacent configurations, represented by the displacements before (u0, v0, w0) and after the
increment (u, v, w) are analysed. No increment is given to the load parameter. The function represented
by (u1, v1, w1) is called the buckling mode. As an alternative, the minimum potential energy criterion can
be adopted to derive the linear stability equations. The expression for the second variation of the potential
energy of the shell in terms of displacements is calculated. The linear differential equations for loss of
stability are then obtained by means of the Trefftz criterion. Readers requiring a more detailed coverage
of shell buckling are advised to consult [4].

In practice, for some problems, the results obtained by these analyses are adequate and in accordance
with experiment. In other cases, such as the buckling of an axially compressed cylinder, the results can
be positively misleading as they may substantially overestimate the actual carrying resistance of the shell.
The use of these methods leads to the following value for the axial buckling load of a perfect thin elastic
cylinder of medium length:

(2) Assuming = 0,3 for steel gives cr = 0,605 E

This buckling load is derived on the assumption that the pre-buckling increase of the radius due to the
Poisson effect is unrestrained and that the two edges are held against translational movement in the
radial and circumferential directions during buckling, but are able to rotate about the local circumferential
axis. These edge restraints are usually called "classical boundary conditions".

Equation (2) is of little use to the designer because test results yield only 15-60% of this value. The
reason for the big discrepancy between theory and experimental results was not understood for a long
time and has been the subject of many studies; it can be explained as follows:

The boundary conditions of the shells have a significant effect and can, if modified, give rise to lower
critical loads. Many authors have investigated the effects of the boundary conditions on the buckling load
of cylindrical shells. The value given by Equation (2) refers to a real cylinder only if the edges are
prevented from moving in the circumferential direction, i.e. v = 0 (Figure 3). If this last condition is
removed and replaced by the condition nxy = 0 (i.e. free displacement but no membrane stress in the
circumferential direction) a critical value of approximately 50% of the classical buckling load is obtained.
This boundary condition is quite difficult to obtain in practice and cylinders with such edge restraints are
much less sensitive to imperfection than cylinders with classical boundary conditions; they are, therefore,
not of primary interest to the designer. If, instead, the top edge of the cylinder is assumed to be free, the
critical buckling load drops down to 38% of the critical value given by Equation (2). In general, it can be
stated that if a shell initially fails with several small local buckles, the critical load does not depend to any
great extent on the boundary conditions, but, if the buckles involve the whole shell, the boundary
conditions can significantly affect the buckling load.
The critical buckling load may also be reduced by pre-buckling deformations. To take these deformations
into account, the same boundary conditions in both the pre- and post-buckling range must be included.
The consequence is that during the compression prior to buckling, the top and bottom edges cannot
move radially (Poisson's ratio not being zero) and, therefore, the originally straight generators become
curved. The post-buckling deformations are not infinitely small and the critical stress is reduced.

The complete understanding of the reason for the large discrepancies between theoretical and
experimental results in the buckling of shells has caused much controversy and discussion, but now the
explanation that initial imperfections are the principal cause of the phenomenon is generally accepted.
4. POST-BUCKLING BEHAVIOUR OF THIN SHELLS

The starting point for this illustrative study of the post-buckling behaviour of a perfect cylinder, under axial
compression (Figure 3), is Donnell's classical equations [2]. A suitable function for w (trigonometric) may
be assumed and introduced into the compatibility equation, expressed in terms of w and of an adopted
stress function F. The quadratic expressions can be transformed to linear ones by means of well known
trigonometric relations. Then the stress function F, and as a consequence the internal membrane
stresses, may be computed. The expression for the total potential energy can then be written, and
minimized, to replace the equilibrium equation. The solution is improved by taking more terms for w.

In Figure 4, the results obtained by using only two buckling modes are shown and compared with the
curves obtained later, i.e. with a greater number of modes. The results show that the type of curve does
not change by increasing the number of modes, but the lowest point of the post-buckling path decreases
and can attain a value of about 10% of the linear buckling load. In the limiting case, i.e. where the number
of terms increases to infinity, the lowest value of the post-buckling path tends to zero, while the buckling
shape tends to assume the shape of the Yoshimura pattern (Figure 5). It is the limiting case of the
diamond buckling shape that can be described by the following combination of axi-symmetric and
chessboard modes.

(3)
It is worth noting that the buckling load associated with either the combination or the two single modes is
the same and is given by Equation (2).

A comprehensive overview of post-buckling theory is given in [5]. As will be discussed later, a realistic
theory for shell buckling has to take into account the unavoidable imperfections that appear in real
structures. Figure 6 shows the influence of imperfections on the strength of a cylinder subject to
compressive loading and Figure 7 shows typical imperfections.
5. NUMERICAL ANALYSIS OF SHELL BUCKLING
Simple types of shells and loading are amenable to treatment by analytical methods. The buckling load of
complex shell structures can, however, be assessed only be means of computer programs, many of
which use finite elements and have a stability option. CASTEM, STAGS, NASTRAN, ADINA, NISA,
FINELG, ABAQUS, ANSYS, BOSOR and FO4BO8 are some of the general and special purpose
programs available. Correct use of a complicated program requires the analyst to be well acquainted with
the basis of the approach adopted in the program.

The stability options and the reliability of the numerical results depend on the method of analysis
underlying each specific program, and on the buckling modes considered. Analysis of various types may
be performed:

1. Geometrical changes in the pre-buckling range are ignored, the pre-buckling behaviour of the
structure is thus assumed to be linear, and the buckling stress corresponds to that at the
bifurcation point B which is found by means of an eigenvalue analysis (Figure 8a). Applied to a
simple shell, this procedure yields the classical critical load. w denotes the lateral deflection of the
shell wall at some representative point.
2. Non-linear collapse analysis enables successive points on the non-linear primary equilibrium path
to be determined until the tangent to the path becomes horizontal at the limit point (Figure 8b). At
that stage, assuming weight loading, as is normally the case for engineering structures, non-
linear collapse ("snap-through") occurs.
3. Investigating bifurcation buckling from a non-linear pre-buckling state involves a search for
secondary equilibrium paths (corresponding to different buckling modes, e.g. different numbers of
buckling waves along the circumference of an axi-symmetric shell) that may branch off from the
non-linear primary path at bifurcation points located below the limit point (Figure 8c). The lowest
bifurcation point provides an estimate of the buckling load.
4. General non-linear collapse analysis of an imperfect structure consists of determining the non-
linear equilibrium path and the limit point L for a structure whose initial imperfections and plastic
deformations are taken into account (Figure 8d). The limit load, which is the ordinate of L, causes
the structure to "snap-through".
The four load-deflection diagrams given in Figure 8 may relate, for example, to a spherical cap subjected
to uniform radial pressure acting towards the centre of the sphere; in this case the critical failure mode
depends on the degree of shallowness of the cap.

6. BUCKLING AND POST-BUCKLING BEHAVIOUR OF STRUTS, PLATES AND SHELLS

Equilibrium paths, for a perfectly straight column, a perfectly flat plate supported along its four edges, and
a perfectly cylindrical shell, presented in the preceding Lecture 8.6, are repeated here (Figure 9) for
completeness.
In each diagram, represents the uniformly applied compressive stress, cr its critical value given by
classical stability theory, and U the decrease in distance between the ends of the members.
Each point on the solid or dashed lines represents an equilibrium configuration which is at least
theoretically possible, in the sense that the conditions for equilibrium between external and internal forces
are met.

Simple elastic shortening, according to Hooke's law, is reflected by the three straight lines OA. They
represent the pre-buckling, primary, or fundamental state of equilibrium, in which the column, the plate
and the shell remain perfectly straight, flat and cylindrical, respectively.

As long as < cr, the primary equilibrium is stable, i.e. if a minute accidental disturbance (a very small
lateral force, for example) causes a slight transverse deformation of the member, the deformation
disappears when its cause is removed, and the member returns of its own accord to its previous
configuration. Any point of the line OA, which is located above B, represents, however, unstable
equilibrium, i.e. the effect of a disturbance, even an infinitely small one, does not disappear with its cause,
but instantaneously increases and the member is set in (violent) motion, deviating further and irreversibly
from its previous equilibrium configuration. Some minor cause of disturbance always exists, for example,
in the form of an initial shape imperfection or of an eccentricity of loading. A state of unstable equilibrium,
therefore, although theoretically possible, cannot occur in real structures.

When the stress reaches its critical value, cr, a new equilibrium configuration appears at point B. This
configuration is quite different from the primary one and features lateral deflections and bending of the
strut, the plate, or the wall of the shell.

If the new configuration is characterised by displacements with respect to the primary state of equilibrium
which increase gradually from zero to high (theoretically infinite) values, the post-buckling states of
equilibrium are represented by points on a secondary equilibrium path which intersects with the primary
path at the bifurcation point B.

In fact, B is the lowest of an infinite number of bifurcation points, but the paths branching off from all the
others represent highly unstable equilibrium and have no practical significance.

The great difference between the strut, the plate and the cylinder is embodied in their post-buckling
behaviour. In the case of the column (Figure 9a), the secondary path, BC, is very nearly horizontal, but in
reality it curves imperceptibly upwards; the equilibrium along BC is almost neutral (it is, strictly speaking,
weakly stable). For the plate (Figure 9b) the secondary path, BC, climbs above B, although less steeply
than before; the plate deflects laterally, more and more under a gradually increasing load, but the
equilibrium at points on BC is stable.

After bifurcation, the point representing the state of equilibrium of an axially loaded cylinder (Figure 9c), in
theory, can travel along the secondary path BDC. The equilibrium at points located below B on the solid
curve is, however, highly unstable and, hence, cannot really exist. What would happen after point B is
reached, if it were possible to manufacture a perfect cylinder from material of unlimited linear elasticity
and to support and load or deform it in the theoretically correct manner, depends on the loading method.

When displacements, u, of one plate of a supposedly rigid testing machine with respect to the other plate
are imposed in a controlled manner, buckles suddenly appear in the wall of the cylinder. The compressive
stress drops at once from cr to the ordinate of point E (only a fraction of cr), while the shortening of the
cylinder remains equal to ucr, the abscissa of B. In contrast with bifurcation, finite displacements are
involved in the transition between the equilibrium configurations represented by points B and E; such an
occurrence is called snap-through. The buckling process is further complicated by the existence of
different intersecting equilibrium paths, which correspond to different numbers of circumferential buckling
waves and which have the same general shape as BDC. Some parts of these paths represent stable
equilibrium, while other parts represent unstable equilibrium; after the initial snap-through from state B to
state E, the shell can jump repeatedly from one buckling configuration to another.
When the load, rather than the displacement, u, is controlled a different effect occurs; if, for example, a
load = 2 rt cr is imposed, the overall shortening of the cylinder almost instantly increases from ucr to the
abscissa of point F, and its wall suddenly exhibits deep buckles, while the average compressive stress
remains equal to cr. It should be noted that this "snap-through" has dynamic characteristics which are
not considered in this description.

Esslinger and Geier [5] explain the fundamentally different behaviour of columns, plates and shells by the
following argument illustrated by Figures 10, 11 and 12.
The differential equation

expresses the lateral equilibrium of any element of an axially loaded strut when bifurcation occurs (Figure
10a) by stating that the deflecting force per unit length, due to the external loads, Fcr, given by the first
term, cancels out the restoring force per unit length, due to the internal bending stresses, given by the
second term. Both the deflecting forces (Figure 10b) and the restoring forces (Figure 10c) are
proportional to the lateral deflection. Consequently, equilibrium of the column is independent of the
magnitude of the transverse deformation and of u, for a given constant axial force Fcr.

The restoring forces which balance the lateral forces (Figure 11b) deflecting a buckling plate (Figure 11a),
are due not only to longitudinal and transverse bending moments (Figure 11c), but also to transverse
membrane forces (Figure 11d). The restoring forces due to membrane action are zero, as long as the
plate is flat, but they then increase proportionately to the square of its lateral deflection. As a result, the
compressive external load required for equilibrium increases together with the lateral deformation and
with the plate shortening u.

Figure 12a shows the radial component of the buckling pattern of a compressed cylinder at the bifurcation
point. Outward displacements of a curved surface cause tensile membrane forces, while inward
displacements generate compressive membrane forces. Figure 12b gives a more accurate picture of an
inward buckle of very small amplitude; it is seen that the original sign of the circumferential curvature of
the shell wall is not reversed at the start of buckling. The radial forces arising from the combination of the
membrane forces with the curvature of the deformed cylinder, which still has its initial sign, are shown in
Figure 12c. These radial forces all tend to counteract buckling. Hence the high resistance of a perfect
cylinder to the initiation of buckling, given by Equation (2). Increasing inward displacements cause the
change of circumferential curvature to exceed the magnitude, 1/r, of the original curvature of the cylinder,
as shown in an exaggerated manner in Figure 12d, and more realistically in Figure 12e. In the region of
the inward buckles, the wall of the cylinder is now curved inwards and, as a result, the compressive
membrane forces in these areas no longer resist the appearance of dents, but precipitate them (Figure
12f). Hence, the total restoring effect of the membrane forces has now weakened substantially compared
with the state prevailing at the bifurcation point. The upshot is that, once buckling has started, equilibrium
is conceivable only under decreasing axial load.

7. IMPERFECTION SENSITIVITY

The behaviour of actual imperfect components differs from the theoretical behaviour described above and
is represented by the dotted curves in Figure 9. They show that true bifurcation of equilibrium does not
actually occur in the case of real structural members. However, the solid lines provide an approximate
picture - the smaller the initial imperfections, the truer the picture is - of the behaviour of the component,
and therein lies the significance of the bifurcation buckling concept.

The dotted lines in Figures 9a and 9b, have been drawn for a column and a plate with slight initial
curvature. It can be seen that the carrying resistance of the strut is not much lower than the theoretical
buckling load, provided that the imperfection is not too great. One can conclude from Figure 9b that the
equilibrium path of an imperfect plate may not exhibit any discontinuity when the compressive stress
increases beyond cr, and also that the plate may possess a considerable post-buckling strength reserve.
If it is thin, this reserve may be considerably greater than the bifurcation buckling load. Raising the stress
beyond cr for the perfect plate does not bring about immediate ultimate failure. Both the column and the
plate finally fail by yielding caused by excessive bending.
Owing to the imperfection of a real cylinder, the dotted equilibrium path does not display the very sharp
high peak B which is a feature of the theoretical equilibrium path OBDC. The culminating point G or H
(Figure 9c) of the dotted line, called a limit point, is at a much lower level than the bifurcation point, even
when the amplitude of the initial deviations from the perfect cylindrical shape is minute. The lower dotted
curve is the equilibrium path for a cylinder with somewhat larger imperfections. When the loading is due
to weight and happens to correspond to the limit point, the curve must jump horizontally from G or H
towards the right hand branch of the curve. The concomitant shortening, u, of the steel shell is so large,
and due to buckles which are so deep, that normally part of the wall material is strained into the plastic
range and so the buckling phenomenon, in this case a snap-through or non-linear collapse, is almost
always catastrophic.

One should not infer from the description in the preceding paragraph that only imperfect structural
components display behaviour characterized by a limit point. Due to gradual changes in the geometry of a
perfect structure, its primary equilibrium path may be non-linear from the outset of loading and, indeed,
feature a limit point.

As a summary two points can be established:

1. The real collapse stress, uG or uH (Figure 9c), is much lower than the theoretical critical stress,
cr, for the perfect shell, even though the imperfections may be hardly perceptible.
2. Nominally identical shells collapse under markedly different loads because the unintentional
actual imperfections of such shells, as erected, are different in magnitude and in distribution, and
because an appreciable decrease in ultimate load may result from slightly larger imperfections.

A sweeping generalization to the effect that all shells are always very sensitive to deviations from the
perfect shape would be unwarranted. The imperfection sensitivity depends on the type of shell and
loading. It may vary from slight to extreme, even for the same kind of shell under different loading
conditions. For example, the imperfection sensitivity of cylindrical shells under uniform external pressure
is quite low, whilst the same shells are highly imperfection sensitive when they are compressed in the
meridional direction. The difference relates to the buckling mode; under axial load, the buckling modes
are characterised by waves which, compared to the diameter, are short in both the longitudinal and the
circumferential direction. Small initial imperfections, which may occur anywhere on the surface of the
cylinder and which are likely to have roughly the same shape as some of the critical buckles, tend to
deepen under increasing load and to trigger off a snap-through at an early loading stage. The buckling
pattern under external pressure, however, consists of buckles which are long in the meridional direction,
and less numerous in the hoop direction, and therefore probably of considerably larger size than the
principal initial dents and bulges.

Another factor that should be mentioned as contributing to the imperfection sensitivity of axially loaded
cylinders is the multiplicity of different buckling modes associated with the same bifurcation load. Any
realistic theoretical treatment of the buckling problem is complicated further by the existence of residual
stresses due to cold or hot forming and/or due to welding. Behaviour is also affected by the appearance
of plastic deformations in the steel and, in some cases, by the presence of stiffeners. The non-linear
structural behaviour of the shell may be due to the latter, as well as to changes in the geometry resulting
from the deformation of the shell.

In conclusion, imperfections are the main cause of the large difference between the ultimate load
obtained in tests and the theoretical buckling load. A wide scatter of results for nominally identical shells
can be seen in Figure 13a where the ratio of experimental buckling loads, Fu, against the theoretical
values, Fcr, for axially loaded cylinders are given for different r/t ratios. Figure 13b gives the factors
proposed by ECCS to reduce the theoretical buckling load to values appropriate for design.
8. CONCLUDING SUMMARY

• Bending and stretching are the modes by which shell structures carry loads.
• For shell structures, in industrial applications, buckling may be the critical limit state due to
slenderness effects.
• Imperfections are the main cause of the very significant difference between the theoretical and
the experimental buckling load.
• There are fundamental differences in initial buckling behaviour between shells and plates.
• In practice, shell buckling analysis can be applied only to special structures which have been
manufactured/constructed using strict quality control procedures that minimise imperfections.

9. REFERENCES

[1] Timoshenko, S. and Woinowsky-Krieger, S., "Theory of Plates and Shells", McGraw-Hill, New York
and Kogakusha, Tokyo, 1959.

[2] Flügge, W., "Stresses in Shells", Springer-Verlag, New York, 1967.

[3] Bushnell, D., "Computerised Buckling Analysis of Shells", Martinus Nijhoff Publishers, Dordrecht,
1985.

[4] Timoshenko, S. and Gere, J.M., "Theory of Elastic Stability", McGraw-Hill, New York and Kogakusha,
Tokyo, 1961.

[5] Esslinger, M. T., and Geier, B. M., "Buckling and Post Buckling Behaviour of Thin-Walled Circular
Cylinders", International Colloquium on Progress of Shell Structures in the last 10 years and its future
development, Madrid, 1969.
10. ADDITIONAL READING

ESDEP WG 8

PLATES AND SHELLS

Introduction to Shell Structures

OBJECTIVE/SCOPE

To describe in a qualitative way the main characteristics of shell structures and to discuss briefly the
typical problems, such as buckling, that are associated with them.

PREREQUISITES

None.

RELATED LECTURES

Lecture 6.1: Concepts of Stable and Unstable Elastic Equilibrium

Lecture 8.1: Introduction to Plate Behaviour and Design

Lecture 8.4.1: Plate Girder Behaviour and Design I

Lecture 8.5.1: Design of Box Girders

SUMMARY

Shell structures are very attractive light weight structures which are especially suited to building as well as
industrial applications. The lecture presents a qualitative interpretation of their main advantages; it also
discusses the difficulties frequently encountered with such structures, including their unusual buckling
behaviour, and briefly outlines the practical design approach taken by the codes.

1. INTRODUCTION

The shell structure is typically found in nature as well as in classical architecture [1]. Its efficiency is based
on its curvature (single or double), which allows a multiplicity of alternative stress paths and gives the
optimum form for transmission of many different load types. Various different types of steel shell
structures have been used for industrial purposes; singly curved shells, for example, can be found in oil
storage tanks, the central part of some pressure vessels, in storage structures such as silos, in industrial
chimneys and even in small structures like lighting columns (Figures 1a to 1e). The single curvature
allows a very simple construction process and is very efficient in resisting certain types of loads. In some
cases, it is better to take advantage of double curvature. Double curved shells are used to build spherical
gas reservoirs, roofs, vehicles, water towers and even hanging roofs (Figures 1f to 1i). An important part
of the design is the load transmission to the foundations. It must be remembered that shells are very
efficient in resisting distributed loads but are prone to difficulties with concentrated loads. Thus, in
general, a continuous support is preferred. If it is not possible to have a foundation bed, as shown in
Figure 1a, an intermediate structure such as a continuous ring (Figure 1f) can be used to distribute the
concentrated loads at the vertical supports. On occasions, architectural reasons or practical
considerations impose the use of discrete supports.
As mentioned above, distributed loads due to internal pressure in storage tanks, pressure vessels or silos
(Figures 2a to 2c), or to external pressure from wind, marine currents and hydrostatic pressures (Figures
2d and 2e) are very well resisted by the in-plane behaviour of shells. On the other hand, concentrated
loads introduce significant local bending stresses which have to be carefully considered in design. Such
loads can be due to vessel supports or in some cases, due to abnormal impact loads (Figure 2f). In
containment buildings of nuclear power plants, for example, codes of practice usually require the
possibility of missile impact or even sometimes airplane crashes to be considered in the design. In these
cases, the dynamic nature of the load increases the danger of concentrated effects. An everyday
example of the difference between distributed and discrete loads is the manner in which a cooked egg is
supported in the egg cup without problems and the way the shell is broken by the sudden impact of the
spoon (Figure 2g). Needless to say, in a real problem both types of loads will have to be dealt with either
in separate or combined states, with the conceptual differences in behaviour ever present in the
designer's mind.
Shell structures often need to be strengthened in certain problem areas by local reinforcement. A possible
location where reinforcement might be required is at the transition from one basic surface to another; for
instance, the connections between the spherical ends in Figure 1b and the main cylindrical vessel; or the
change from the cylinder to the cone of discharge in the silo in Figure 1c. In these cases, there is a
discontinuity in the direction of the in-plane forces (Figure 3a) that usually needs some kind of
reinforcement ring to reduce the concentrated bending moments that occur in that area.
Containment structures also need perforations to allow the stored product (oil, cement, grain, etc.) to be
put in, or extracted from, the deposit (Figure 3b). The same problem is found in lighting columns (Figure
3c), where it is general practice to put an opening in the lower part of the post in order to facilitate access
to the electrical works. In these cases, special reinforcement has to be added to avoid local buckling and
to minimise disturbance to the general distribution of stresses.

Local reinforcement is also often required at connections between shell structures, such as commonly
occur in general piping work and in the offshore industry. In these cases additional reinforcing plates are
used (Figure 3d), which help to resist the high stresses produced at the connections.

In contrast to local reinforcement, global reinforcement is generally used to improve the overall shell
behaviour. Because of the efficient way in which these structures carry load, it is possible to reduce the
wall thickness to relatively small values; the high value of the shell diameter to thickness ratios can,
therefore, increase the possibility of unstable configurations. To improve the buckling resistance, the shell
is usually reinforced with a set of stiffening members.

In axisymmetric shells, the obvious location for the stiffeners is along selected meridians and parallel
lines, creating in this way a true mesh which reinforces the pure shell structure (Figure 4a). On other
occasions, the longitudinal and ring stiffeners are replaced by a complicated lattice (Figure 4b), which
gives an aesthetically pleasing structure as well as mechanical improvements to the global shell
behaviour.
2. POSSIBLE FORMS OF BEHAVIOUR

There are two main mechanisms by which a shell can support loads. On the one hand, the structure can
react with only in-plane forces, in which case it is said to act as a membrane. This is a desirable situation,
especially if the stress is tensile (Figure 5a), because the material can be used to its full strength. In
practice, however, real structures have local areas where equilibrium or compatibility of displacements
and deformations is not possible without introducing bending. Figure 5b, for instance, shows a load acting
perpendicular to the shell which cannot be resisted by in-plane forces only, and which requires bending
moments, induced by transverse deflections, to be set up for equilibrium. Figure 5c, however, shows that
membrane forces only can be used to support a concentrated load if a corner is introduced in the shell.
It is worthwhile also to distinguish between global and local behaviour, because sometimes the shell can
be considered to act globally as a member. An obvious example is shown in Figure 6a, where a tubular
lighting column is loaded by wind and self-weight. The length AB is subjected to axial and shear forces,
as well as to bending and torsion, and the global behaviour can be approximated very accurately using
the member model. The same applies in Figure 6b where an offshore jacket, under various loading
conditions, can be modelled as a cantilever truss. In addition, for certain types of vault roofs where the
support is acting at the ends, the behaviour under vertical loads is similar to that of a beam.
Local behaviour, however, is often critical in determining structural adequacy. Dimpling in domes (Figure
7a), or the development of the so-called Yoshimura patterns (Figure 7b) in compressed cylinders, are
phenomena related to local buckling that introduce a new level of complexity into the study of shells. Non-
linear behaviour, both from large displacements and from plastic material behaviour, has to be taken into
account. Some extensions of the yield line theory can be used to analyse different possible modes of
failure.
To draw a comparison with the behaviour of stiffened plates, it can be said that the global action of shell
structures takes advantage of the load-diffusion capacity of the surface and the stiffeners help to avoid
local buckling by subdividing the surface into cells, resulting in a lower span to thickness ratio. A
longitudinally-stiffened cylinder, therefore, behaves like a system of struts-and-plates, in a way that is
analagous to a stiffened plate. On the other hand, transverse stiffeners behave in a similar manner to the
diaphragms in a box girder, i.e. they help to distribute the external loads and maintain the initial shape of
the cross-section, thus avoiding distortions that could eventually lead to local instabilities. As in box
girders, special precautions have to be taken in relation to the diaphragms transmitting bearing reactions;
in shells the reaction transmission is done through saddles that produce a distributed load.

3. IMPORTANCE OF IMPERFECTIONS

As was explained in previous lectures, the theoretical limits of bifurcation of equilibrium that can be
reached using mathematical models are upper limits to the behaviour of actual structures; as soon as any
initial displacement or shape imperfection is present, the curve is smoothed [2]. Figures 8a and 8b
present the load-displacement relationship that is expected for a bar and a plate respectively; the dashed
line OA represents the linear behaviour that suddenly changes at bifurcation point B (solid line). The plate
has an enhanced stiffness due to the membrane effect. The dashed lines represent the behaviour when
imperfections are included in the analysis.
As can be seen in Figure 8c, the post-buckling behaviour of a cylinder is completely different. After
bifurcation, the point representing the state of equilibrium can travel along the secondary path BDC.
Following B, the situation is highly dependent on the characteristics of the test, i.e. whether it is force-
controlled or displacement controlled. In the first case, after the buckling load is reached, a sudden
change from point B to point F occurs (Figure 8c) which is called the snap-through phenomenon, in which
the shell jumps suddenly between different buckling configurations.

The behaviour of an actual imperfect shell is represented by the dashed line. Compared with the
theoretically perfect shell, it is evident that true bifurcation of equilibrium will not occur in the real structure,
even though the dashed lines approach the solid line as the magnitude of the imperfection diminishes.
The high peak B is very sharp and the limit point G or H (relevant to different values of the imperfection)
refers to a more realistic lower load than the theoretical bifurcation load.

The difference in behaviour, compared with that of plates or bars, can be explained by examining the
pattern of local buckling as the loading increases. Initially, buckling starts at local imperfections with the
formation of outer and inner waves (Figure 9a); the latter represent a flattening rather than a change in
direction of the original curvature and set up compressive membrane forces which, along with the tensile
membrane forces set up by the outer waves, tend to resist the buckling effect. At the more advanced
stages, as these outer waves increase in size, the curvature in these regions changes direction and
becomes inward (Figure 9b). As a result, the compressive forces now precipitate buckling rather than
resist it, thus explaining why equilibrium, at this stage, can only be maintained by reducing the axial load.
The importance of imperfections is such that, when tests on actual structures are carried out, the
difference between theoretical and experimental values produces a wide scatter of results (see Figure
10). As the imperfections are unavoidable, and depend very much on the quality of construction, it is clear
that only a broad experimental series of tests on physical models can help in establishing the least lower-
bound that could be used for a practical application. Thus it is necessary to choose:

1. The structural type, e.g. a circular cylinder, and a fixed set of boundary conditions.
2. The type of loading, e.g. longitudinal compression.
3. A predefined pattern of reinforcement using stiffeners.
4. A strict limitation on imperfection values.

In consequence, the experimental results can only be used for a very narrow band of applications. In
addition, the quality control on the finished work must be such that the experimental values can be used
with confidence.

To allow for this, Codes of Practice [3] use the following procedure:

1. A critical stress, cr or cr, or a critical pressure, pcr, is calculated for the perfect elastic shell by
means of a classical formula or method in which the parameters defining the geometry of the
shell and the elastic constants of the steel are used.
2. cr, cr or pcr is then multiplied by a knockdown factor , which is the ratio of the lower bound of a
great many scattered experimental buckling stresses or buckling pressures (the buckling being
assumed to occur in the elastic range) to cr, cr or pcr, respectively. is supposed to account for
the detrimental effect of shape imperfections, residual stresses and edge disturbances. may be
a function of a geometrical parameter when a general trend in the set of available test points,
plotted with that parameter as abscissa, points to a correlation between the parameter and ;
such a trend is visible in Figure 10, where the parameter is the radius of cylinder, r, divided by the
wall thickness, t.
4. CONCLUDING SUMMARY

• The structural resistance of a shell structure is based on the curvature of its surface.
• Two modes of resistance are generally combined in shells: a membrane state in which the
developed forces are in-plane, and a bending state where out-of-plane forces are present.
• Bending is generally limited to zones where there are changes in boundary conditions, thickness,
or type of loads. It also develops where local instability occurs.
• Shells are most efficient when resisting distributed loads. Concentrated loads or geometrical
changes generally require local reinforcement.
• Imperfections play a substantial role in the behaviour of shells. Their unpredictable nature makes
the use of experimental methods essential.
• To simplify shell design, codes introduce a knock-down factor to be applied to the results of
mathematical models.

5. REFERENCES

[1] Tossoji, Ei., "Philosophy of Structures", Holden Day 1960.

[2] Brush, D.O., Almroth, B.O., "Buckling of Bars, Plates and Shells", McGraw Hill, 1975.

[3] European Convention for Constructional Steelwork, "Buckling of Steel Shells", European
Recommendation, ECCS, 1988.

Vous aimerez peut-être aussi