Anda di halaman 1dari 24

Experimental Perspective

on the Buckling of Pressure


Vessel Components
J. Bachut
Institute of Physics,
Cracow University of Technology,
ul. Podchoraz_ych 1,
Krakow, 30-085 Poland

This review aims to complement a milestone monograph by Singer et al. (2002, Buckling
ExperimentsExperimental Methods in Buckling of Thin-Walled Structures, Wiley, New
York). Practical aspects of load bearing capacity are discussed under the general umbrella of buckling. Plastic loads and burst pressures are included in addition to bifurcation and snap-through/collapse. The review concentrates on single and combined
static stability of conical shells, cylinders, and their bowed out counterpart (axial compression and/or external pressure). Closed toroidal shells and domed ends onto pressure
vessels subjected to internal and/or external pressures are also discussed. Domed ends
include: torispheres, toricones, spherical caps, hemispheres, and ellipsoids. Most experiments have been carried in metals (mild steel, stainless steel, aluminum); however,
details about hybrids (copper-steel-copper) and shells manufactured from carbon/glass
fibers are included in the review. The existing concerns about geometric imperfections,
uneven wall thickness, and influence of boundary conditions feature in reviewed
research. They are supplemented by topics like imperfections in axial length of cylinders,
imperfect load application, or erosion of the wall thickness. The latter topic tends to be
more and more relevant due to ageing of vessels. While most experimentation has taken
place on laboratory models, a small number of tests on full-scale models are also referenced. [DOI: 10.1115/1.4026067]
Keywords: cones, ellipsoids, hemispheres, toricones, torispheres, toroids, external, internal pressure, buckling, plastic load, burst pressure

Introduction

Buckling of thin-walled shell-like components, frequently


found in pressure vessels, has been the subject of numerous studies over many decades. A substantial wealth of accumulated
knowledge exists in the form of books, conference proceedings,
reviews, and other published material. To this end, books covering
experimental aspects of buckling of pressure vessels or their components include: [16]. Reference [7] can be regarded as a milestone piece of work entirely devoted to experimental
methodologies associated with, per se, buckling. It contains a
comprehensive list of references. Conference proceedings dealing
entirely with buckling include Refs. [813]. A large number of
conference papers in the proceedings are explicitly devoted to
buckling experiments. Summary of the current set of design recommendations on buckling prone, thin-walled shell components,
e.g., cylinders, cones, and/or doubly-curved shells can be found in
Ref. [14]. The authors list some existing shell stability technical
issuesmany of which are not addressed in the NASA recommendations (Refs. [1517]). Initial geometric imperfections, nonlinear prebuckling deformations, boundary conditions, load
introduction effects, combined loads, and variation in material
properties are just a sample of topics still awaiting further investigations. The reliance on arbitrarily chosen knock-down factors is
also echoed in Refs. [18,19]. There have also been a number of
published reviews of research on buckling, e.g., Refs. [2022]. In
many instances, experimental data was used to develop design
standards. A good example would be the predecessor of the current code [23], where design of externally pressurized hemispherical/torispherical domes is entirely empirical. It is based on the
lower bound to known experimental results being further reduced
by a safety factor [24]. Authors of a review Ref. [25] poses a quesManuscript received August 18, 2013; final manuscript received November 14,
2013; published online December 30, 2013. Editor: Harry Dankowicz.

Applied Mechanics Reviews

tion: why, despite a great research effort, has the scientific community failed to develop procedures for shell design that are not
based on empirical data, i.e., on lower-bound followed by a
knock-down factor? They do not provide a definitive answer to
this dilemma which still exists more than ten years on. Instead
they propose a number of procedures aiming at improvement of
over-conservativeness of the lower-bound design philosophy for
the case of axially compressed cylinders (including FRP cylinders). In a review of the buckling resistance of thin and slender
structures typically found in nuclear industry, Ref. [26], back in
1984, categorized vessel related components, prone to buckling,
as: (i) stiff (buckling in plastic range), (ii) medium (elastic/plastic
buckling), and (iii) soft (elastic buckling). The criterion used here
was the ratio, REY, of elastic bifurcation-to-first yield load. For
REY  5 the component was deemed to be stiff. For REY  0:2,
structure was classified as soft. In this context, the results of 42
experimental buckling tests have been compared with computed
predictions of buckling. The tests were on metallic torispherical
and elliptical heads, cylinders under: shear, axial compression or
external pressure, spherical caps, spheres, and stiffened/unstiffened baffles. Scatter of results is provided here as Table 1 (from
Ref. [26]). Table 1 shows the difference between tests and numerical predictions of buckling together with the number of tests carried out. It is seen here that the range of errors varies from 30%
to 50%. The discrepancies between experimental and computed
values have been attributed to: (i) specimen geometry, (ii) boundary conditions, and (iii) material data. It was also noted that in
some cases the buckling was difficult to observe experimentally
and it was subjective (internally pressurized domes; for example).
Progress has been made in assessing the issues leading to big discrepancies between test data and theory since the publication of
Ref. [26]. Testing methods, for example, have improved and they
have been more focused [2730]. It is true to say that buckling
experiments are now better instrumented than in the past; for
example, in Refs. [27,28,3033]. Equally, more rigorous

C 2014 by ASME
Copyright V

JANUARY 2014, Vol. 66 / 010803-1

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Table 1 Comparison of 42 experimental and computed results


from Ref. [26] (Pexp tl is experimental buckling load, and Pc is
computed buckling load)
Pexp tl  Pc
Pexp tl
30% to 20%
20% to 10%
10% to 0%
0% to 10%
10% to 20%
20% to 30%
30% to 40%
40% to 50%

Number of tests
3
5
5
7
6
6
5
5

computational models have been developed, e.g., Ref. [7],


[3439]. References [7] and [30], for example, provide a thorough
review of experimental techniques and results obtained in a wider
context of structures prone to buckling. Recent review papers,
Refs. [20,21], address structural behavior of domed ends and add
to the accumulated know-how base. Design methodology is available in the form of various codes [1517,23,4042], and are also
available in a specialized stability handbook [43] (see also Ref.
[44]). As mentioned earlier, one of unresolved issues within structures prone to buckling is the effect of initial shape imperfections
on the magnitude of buckling load. Despite efforts aiming at finding a universal answer to the dilemma, it still remains a subject of
active research. It is true to say that the detrimental effect of initial
shape imperfections has been quantified for a range of structural
components and loading conditions. However, it is, by-and-large,
still based on a component by component basis. An attempt of a
unified approach to the design of imperfection sensitive structural
components can be found in Ref. [42], where a better manufacturing quality of a member results in the attainment of higher buckling loads. However, research into the derivation of less restrictive
knock-down factors still continues, e.g., Ref. [45]. A different
way of reducing imperfection sensitivity of axially compressed
cylinders is explored in Ref. [46]. Numerical results suggest that
cylinders filled with, and/or surrounded by a compliant core, can
be less imperfection sensitive. Finally, it is worth mentioning a
better access to the specialized buckling-related, and web based
resources, e.g., Refs. [47,48].

Cylindrical Shells

Cylindrical shells constitute a backbone of pressure vessels and


of other load bearing components in variety of on land, in the sea,
and in the air applications. As mentioned earlier, the comprehensive review of cylindrical shells under buckling conditions is
available in Ref. [7]. Historical background into early shell buckling tests is available in Ref. [49]. The earliest shell buckling tests
on thin-walled tubes under axial compression and bending were in
18451850. These tests were followed by experimental buckling
tests on tubes under external and/or internal pressure. Reference
[49] lists milestones in experimental buckling tests, and motivation behind them, for cylindrical shells for up to the 1970s. Tubular bridges motivated the very first buckling tests on tubes in the
mid-19th century. The next demand for experimental data came
from shipbuilders at the end of the 19th century. The needs arising
from the design of tunnel linings and submarines dominated
experimentation early in the 20th century. The latter designs specifically stimulated experimentation on stiffened cylinders. Aircraft structures were associated with tests on much thinner
cylinders. By then, the range covered the radius-to-wall thickness,
R/t, between 35 and 1440. The concept of knock down factor
appears to have been founded at that time. The next big impetus
in buckling of cylindrical shells took place in the 1960s and it was
associated with careful studies of buckling and post-buckling
patterns via high speed photography and the use of photoelasticity.
010803-2 / Vol. 66, JANUARY 2014

Further details about these milestones together with information


about sources are given in Ref. [49]. Review of research into buckling of cylinders and domed closures onto cylindrical vessels is
available in Ref. [50]. Plain cylinders, cylinders reinforced by: (i)
rings, (ii) stringers, and (iii) rings and stringers are included. Loading includes: (i) axial compression, (ii) external pressure, and (iii)
axial compression and external pressure. Nonaxial compression is
also mentioned. Some of more recent research into static stability
of cylindrical shells is reviewed in what follows.
2.1 External Pressure. Corrosion damage to the wall thickness can affect the buckling strength of externally pressurized vessels. These effects have recently been researched for cylindrical
shells and domed ends. Buckling strength of cylindrical pressure
hulls with artificial damage introduced to the wall has been
researched in Refs. [5154]. Motivation for this comprehensive
experimental program was linked to naval submarines, which, if
not adequately protected, can suffer from corrosion damage. Once
the surface corrosion is identified in the pressure hull, a range of
repair possibilities exist. This includes anything from weld buildup of the lost wall thickness to the full replacement of affected
shell plates. However, all of these routes carry side effects, e.g.,
appearance of residual stresses, geometrical distortions, or change
of material properties in heat affected zones. An alternative
approach would be to allow operation of a corrosion affected vessel, but within a modified envelope of operations. References
[51,52] detail buckling tests on laboratory models of approximately 220 mm diameter, 2.5 mm nominal wall thickness, and
manufactured from 6082-T6 aluminum alloy. Buckling tests on
20 ring stiffened shells subjected quasi-static external hydrostatic
pressure were carried out. Eight models were near-perfect and
the rest had localized wall thinning in the form of rectangular
patches positioned at half length. The wall thickness loss of up to
25% was introduced on the outer surface, only. Corrosion to Ttype rings was also considered. Here, the breadth of the localized
flange damage was up to a maximum of 50% (with the web being
intact). Comparison of test data with submarine design formulae
prediction of collapse pressure is provided for both intact and corroded models. It is seen here that the experimental collapse values
are higher than those given by the existing design manual. However, at the same time, the reduction of collapse and yield pressures caused by the wall thinning was 20 and 40%, respectively,
when compared to intact models. Tests on additional eight models, three intact and five with simulated corrosion, are described in
Ref. [54]. This time a different aluminum alloy, AA-6082-F28,
was used. The principal idea behind choosing this alloy was its
elongation at break, which is comparable to steel plate used in
construction of submarines. Uniaxial tests have shown anisotropy
in material properties and the elongation at break ranging from 9
to 20% was recorded. However, the profile of stressstrain curves
resembled that of true material used in submarine pressure hulls.
Tests showed that collapse strength was related not only to the
yield point of the material, but also to the plastic reserve of the
material. In consequence, the full stressstrain curve would be
needed in the numerical predictions of collapse. Results based on
elastic perfectly plastic modeling would lead to conservative estimates of buckling strength. Although the tests were carried out on
aluminum models, their pattern of failure resembled that of hulls
made from high-strength steel found in real submarine pressure
hulls. Several proprietary finite element (FE) codes were
employed in Ref. [53] to utilize the vast amount of experimental
data collected (in Ref. [52]) in order to numerically estimate collapse pressures. Accuracy of the FE results is about 11%, while
the conventional design manuals are accurate to within 20% for
intact models, and 26% for corroded models. It has transpired that
the fine details had to be modeled and Ref. [53] provides these
details. The sensitivity of buckling pressures to manufacturing
defects in pipes with the ratio of diameter-to-wall-thickness, D/t,
within 10 to 25 has been investigated in Ref. [55].
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 1

Cylinders with uneven length at the top end

2.2 Axial Compression. Buckling strength of axially compressed cylindrical shells still attracts sizeable amount of research.
One specific topic has been devoted to buckling of variable length
cylinders under axial compression, as sketched in Fig. 1. When
two or more cylindrical segments form a prime load bearing structure then the interaction between two neighboring segments
become critical when the load is axial compression. The possibility of buckling of either one or both segments complicates the segment-to-segment interaction. Typical application exists in
aerospace where the gap between segments is filled by shimming
[56]. Once axial compression is applied to two segments where
there is a variable gap between them, then the uneven load results.
Diminishing axial gaps result in a variable length of hoop contact
between two cylinders and in localized plastic deformations.
These local effects at the imperfect end of the cylinder can propagate along the shells length. They in turn can trigger asymmetric
bifurcation buckling or collapse, and as such they can pose design
limitations. One aspect of this problem was studied in Ref. [57].
Cylinders with sinusoidal waviness of axial length were subjected
to axial compression by a rigid disk moving vertically. Numerical
results have highlighted a complicated nature of the interaction
between the plate and imperfect cylinder. A big drop in buckling
strength has been obtained for relatively small amplitude of waviness in length. In Ref. [58], 18 mild steel cylinders with the
length-to-radius ratio, L/R  2.4 and with the radius-to-wall thickness ratio, R/t  185 were collapsed by axial compression. Cylinders had variable length at one end of the sinusoidal profile. The
amplitude-of-axial-waviness-to-wall thickness ratio, 2A/t, was
varied between 0.05 and 1.0. Experimental results show that buckling strength strongly depends on the axial amplitude of imperfection (see Fig. 2). Average imperfect cylinders, with 2A/t 1.0, are
able to support 49% of experimental buckling load obtained for
geometrically perfect model. The largest sensitivity of buckling
strength was associated with small amplitudes of axial length. For
example, for axial length imperfection amounting to 25% of wall
thickness the buckling strength was reduced by 40%. It appears
that the number of sinusoidal waves in the imperfection profile
plays a secondary role, i.e., its role in reducing the buckling
strength is not a dominant factor. The paper provides experimental
details and comparisons with numerical results based on the FE
analyses.
Applied Mechanics Reviews

2.3 Combined Loading


2.3.1 Straight Cylinders. Experimental program aiming at
reassessing NASA Space Vehicle Design Criteria guide SP-8007
(Ref. [15]) is reported in Refs. [5961]. Metallic cylinders with
the radius-wall-thickness ratio, R/t, varying between 250 and
1500 were buckled under static loading. Experimental models
were from copper, aluminum or stainless steel. Their diameter
was 135 mm. The whole program consisted from 150 carefully
conducted buckling tests. The role of geometrical imperfections
was of particular interest and it was closely monitored during
tests. It was noted, for example, that internal pressure reduced the
influence of imperfections on buckling, resulting in higher buckling strength. A stable post-buckling behavior of copper model (R/
t 1350) was obtained for the case of simultaneous action of internal pressure and bending. But for other cases, internal pressure
triggered local yielding and this, in turn, accelerated elasticplastic buckling. Some models were retested under a different set of

Fig. 2 Buckling response for axially compressed cylinders


with various waviness of length

JANUARY 2014, Vol. 66 / 010803-3

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 4 Load carrying capacity of equivalent barrels as a function of barrelling, D/Ro. Photographs of tested models at a, b, c,
and d.

Fig. 3 Combined stability plot for cylindrical shell (axial compression, F, versus external pressure, p) [67]

loads. While these tests were within elastic domain, the concerns
were raised about the retest data. In view of this, Ref. [61] provides two sets of test data, i.e., single test and retest buckling
results. In a separate study, the influence of the thermal insulation
layer onto buckling performance of cantilevered cylindrical shell
is reported in Ref. [62]. This is both an experimental and numerical study. Vacuum induced buckling tests of small steel cylinders
are reported in Refs. [63,64]. Models were mass manufactured
industrial containers for storage of paint. Initial geometry of cylinders generators was carefully measured and the loading
amounted to hydrostatic external pressure. The content of Ref.
[65] is in many respects unique. It provides insights into load carrying capacity of on land vertical storage tanks, usually found in
the petroleum industry, under buckling conditions. Four steel
tanks with volume capacity ranging from 1000 m3 to 65,000 m3
(with radii between 5 m and 35 m) were in situ measured for geometry (geometrical imperfections), and then buckled by external
pressure (through the application of internal vacuum). The paper
provides a wealth of practical information including: design code
estimates, implications of quality of manufacturing, knock down
factors, and FE analyses (including axial compression generated
by roof-loading).
2.3.2 Bowed-out Cylinders. Buckling strength of axially compressed bowed out cylindrical shell can be larger than buckling
load of mass equivalent cylinder, as shown in Fig. 3 (with details
in Refs [66,67]). At the same time, barrels are more efficient in
supporting external pressure. This idea has been explored in Refs.
[6873] for underwater applications. Earlier background to the
above idea can be found in Refs [74,75]. It is seen in Fig. 4 that
external hydrostatic pressure increases with the amount of outward barrelling; it reaches maximum and then it drops when the
shell becomes the outer half of a toroid. A number of mild steel
models have been machined to test numerical predictions. All
models had the same mass as the reference cylinder and they had
integral flanges at both ends. Each shell after over machining was
stress relieved. Wall thickness and shape have been measured
prior to testing. The diameter of the tested shells was about
200 mm, their length varied from 75 to 100 mm in order to secure
the constant mass while the wall thickness was nominally 3.0 mm.
Thick flat plates were attached to tested models, which were then
filled with oil and vented to the atmosphere. Models were
010803-4 / Vol. 66, JANUARY 2014

immersed in 350 mm  1000 mm vessel and single incremental


quasi-static pressure was applied while the amount of expelled oil
was measured. All shells failed suddenly with a loud bang and
rapid outflow of oil from inside. Hence, there has been no difficulty in identifying buckling pressure. Experimental buckling
pressures varied from 8 to 22 MPa. Figure 4 indicates four configurations a, b, c, and d, which were subject of experimentation. Bifurcation buckling was predicted for points a and b.
Photographs in the figure show the lobar mode of failure at a,
and b. At point c, both the machined barrel and tested model
are seen in the insert. Photographs of collapsed barrels at points
c and d are also depicted in Fig. 4. Good agreement has been
obtained between experimental results and numerical predictions.
The ratio of pnum =pexp tl varied from 0.98 to 0.99 for reference cylinders, and 0:90  pnum =pexp tl  1:02 for mass equivalent barrels.
The sensitivity of buckling/collapse pressures to the initial, eigenmode type geometric imperfections has been assessed for both cylindrical and bowed out geometries. It appears that barrelling does
not necessarily increase the sensitivity of buckling pressure to
shape deviations from perfect geometry. Reference [68] provides
further details. Buckling tests carried out in Ref. [68] proved that,
on a like for like basis, barrels were able to support pressure of
nearly 90% higher than mass equivalent cylinders. A more practical configuration of bowed out shell would be associated with a
barrel having the same length and top/bottom radii as the master
cylinder. This would require the readjustment of the wall thickness in the barrel in order to have both shells of the same mass. In
addition, a shells generator could be searched in a different class
of profiles than circular arcs. This approach was adopted in Ref.
[76]. The shape of the generator was assumed to be defined by a
generalized ellipse [77]
n2

n1 
x
y

1
(1)
Ro D
0:5 Lo n3
where Ro is a barrels radius at top/bottom ends; D is the amount
of barrelling at the equator.
Parameters n1, n2, and n3, which strongly influencing the barrels meridian, were chosen as optimization variables. Simulated
annealing, SA, was utilized in the search for the optimal shape
leading to the maximum of buckling pressure under the equality
condition imposed on masses, m, of the cylinder and barrel, i.e.,
mcyl mbarrel.
A section through the design space is plotted in Fig. 5. It is seen
here that a significant increase in buckling pressure above the cylinders one was predicted. Detailed calculations were performed
for the reference cylindrical geometry given by Lo/Ro 1.0 and
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 5 The magnitude of failure pressure for a shell generator


described by generalized ellipse [76]. View of barrels shape at
bifurcation corresponding to pmax.

Ro/to 33.33. The reference cylinder under consideration, fails


through the bifurcation buckling at pbif 10.5 MPa. The eigenmode corresponds to n 7 circumferential waves. The SA algorithm has found the global optimum at popt 14.71 MPa, and it
corresponded to the optimal design vector (n1, n2, 2n3/Lo)opt
(2.2, 2.0, 1.0). The predicted failure mechanism at the optimum
was through the axisymmetric collapse. A section through the
design space is shown in Fig. 6, where it is seen that feasible domain is not convex. Two nominally identical barrels were collapsed experimentally and they failed at 16.97 MPa and
16.83 MPa, respectively. After removal from the test tank both
barrels were photographed and these are seen as inserts in Fig. 6.
It is worth noting that the experimental buckling pressure for two
equivalent cylinders was 11.66 MPa and 11.58 MPa, respectively.
For barrels, this gives an increase of 45% in magnitude of external
hydrostatic pressure above the master cylinders buckling
strength. Depending on numerical modeling the ratio, pnum/pexp tl,
varied from 0.90 to 1.02. Additional calculations have shown that
the shape optimization has not created the solution which would
be dismissed from a practical point of view because of greatly
enhanced imperfection sensitivity of buckling load to initial geometric imperfections. It is seen from results given in Ref. [76] that
both the initial (cylinder) and optimal (barrel) designs have a comparable sensitivity to initial shape deviations from perfect
geometry.

Fig. 6 Plot of the cost function versus design vector components n1 5 n2. Also, photograph of two tested barrels E1 and
E1a [76].

Applied Mechanics Reviews

Fig. 7 Buckling strength of two-segment vessel versus the


flange thickness [67]

The above results prompted the examination of multisegment


vessels made from bowed out cylinders (see Ref. [78]). It was
decided to investigate bowed out cylindrical shells which had the
same wall thickness as the reference cylinder. Although, per se,
these were not equivalent models since the mass was not the
same. Nevertheless, large increases in external buckling pressures
were predicted by numerical calculations. One of the outstanding
issues was the joining of two neighboring segments. One possibility was to have the integral flanges of adjoining segments bolted
together. Here, dimensions of flanges heavily influenced the load
carrying capacity of the whole assembly as illustrated in Fig. 7.
Two configurations were chosen for experimentation and they are
denoted in Fig. 7 as models DB1 and DB2. Numerical calculations
indicated that in both cases, the excessive yielding at the joints
controlled the failure. Post-collapse pictures of models DB1 and
DB2 are also seen in Fig. 7. One 4-segment vessel has also been
studied. Its collapsed view of global nature is shown in Fig. 8 (see
Ref. [79] for discussion of inter-stiffener collapse and overall collapse of cylinders). Connecting segments by bolting external
flanges performed satisfactorily for given arrangements. However,
not all possibilities, e.g., external versus internal flanges were
explored. In addition, the wall thickness of barrels was kept constant. It might be beneficial if variation of the wall thickness is
allowed in order to mitigate the edge effects. Numerical results
obtained in Ref. [70] indicate that in barrels made from fiber reinforced plastics (FRP), it is possible to reduce the edge effects
through the appropriate lamination stacking. In particular, free radial displacements at the top and bottom edges do not automatically lead to an inferior performance when compared with an

Fig. 8 Collapsed four-segment vessel [67]

JANUARY 2014, Vol. 66 / 010803-5

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

equivalent cylinder. However, no experimental verification of this


has been carried out.

Conical Shells

Buckling of cones has closely been associated with buckling


performance of cylindrical counterparts. Strong motivation for experimental and theoretical research into buckling of cones had
been, and still is [80,81], rooted into their substantial role in missiles and in space launchers. Account of these efforts is provided
in Ref. [82]. Past experiments on unstiffened conical shells are
briefly reviewed in Ref. [83]. Between 1958 and 2008, there have
been 484 buckling tests on unstiffened cones. Reference [83]
shows the number of tests per year, the type of applied loading, together with the source of data and the type of material from which
the tested shells were made. Apart from plain, i.e., not reinforced
cones, a range of stiffened cones has also been tested in the past:
111 in the same period. Most tests were on cones reinforced by
rings, and in most cases cones were made from steel. By far the
most frequent buckling tests have been carried out on cones subjected to a single load. However, buckling tests were also carried
out when two or more loads were applied at the same time. Details
about these tests can be found in Refs. [8385]. The above tests
have been carried out in elastic range. The next sections summarize recent experimental research effort for cones subjected to
external or internal pressure, axial compression, and cones subjected to combined loading. Most of these papers are related to
elasticplastic buckling with very few experimental data available
here.
3.1 Externally Pressurized Cones. It has to be mentioned
that studies into elasticplastic buckling of cones have not been a
subject of extensive research effort. Nevertheless some fresh experimental results have been reported. For example, details about
buckling tests on ten mild steel cones subjected to quasi-static
external pressure are reported in Ref. [86]. Shells were manufactured by rolling flat sheets followed by longitudinal welding. The
base diameter was about 300 mm and the wall thickness varied
between 0.5 mm and 0.8 mm. The collapse was a gradual process
but it is not reported what was the capacity the vacuum source
which served as means of loading. Measured geometry was utilized in subsequent FE analyses. The ratio of pexp tl/pnumerical, varied between 0.40 and 0.92. Tests on 19 steel, seamless cones
obtained by using the metal spinning, are described in Ref. [87].
The base diameter of all models was 500 mm and the wall thickness was 0.635 mm. Four models were unstiffened while the
remaining had external rings at various positions on the slant.
Models were subjected to quasi-static external hydrostatic pressure. The experimental buckling pressures were established visually once any bulge was visible on the cone surface. In this case,
the ratio of pexp tl/pnumerical varied between 0.62 and 0.80. In a separate research, with scattered results in Refs [8890], a series of
15 mild steel cones reinforced by external rings have been tested
under quasi-static external hydrostatic pressure. It is reported that
the failures were sudden with a loud bang accompanying failure.
Once a cone failed the pressure instantly dropped (for example,
from 14.46 MPa to 7.32 MPa for cone no. 10 in Ref. [88]). Strain
gauges on the inner surface, primarily used to indicate the number
of lobes in the failure mode, recorded large plastic strains just
prior to buckling. Thus, subjective identification of buckling load
appears to be removed during the latter tests. Comparisons
between experimental and numerical predictions measured by the
ratio, pexp tl/pnumerical, varied from 0.24 to 0.84 (Tables 4 and 5 of
Ref. [88]; see also Table 6 of Ref. [89], and Table 3 of Ref. [90]).
In an earlier series of buckling experiments on ring-reinforced
steel and aluminum cones summarized in Ref. [91], the same ratio
varies from 0.56 to 0.87. Over the years, a number of attempts
have been made to find a simplified design equation which could
provide an estimate of the buckling strength of cones subjected to
hydrostatic pressure. One such equation has been proposed in Ref.
010803-6 / Vol. 66, JANUARY 2014

[92] as a result of extensive parametric studies. Reference [93]


compares estimates given in Ref. [92] with numerical predictions
given by BOSOR5 code, Ref. [94] (see Fig. 9). It is seen here
that all Bosor5 predictions are higher than those given in
Ref. [92]. Comparisons have also been made between 17 test data
on steel/aluminum cones and the proposed equation. The ratio,
pexp tl/pnumerical, was found to be between 0.57 and 1.41. Estimates
of buckling pressures for thicker cones were on a safe side while
for thinner cones the above ratio was on unsafe side (between
0.57 and 0.91).
3.2 Internally Pressurized Cones. Results of an experimental study into the buckling of six scaled-down models of the inner
vessel relevant to a liquid metal fast breeder nuclear reactor are
reported in Ref. [95]. The models consisted of two cylindrical
shells connected by cone and toroid (all from stainless steel).
These compound shells were manufactured by rolling, forming,
and welding, and they were subjected to internal pressure and/or
concentrated loads applied to the stand-pipes. Diameters of cylinders were 420 mm and 810 mm with the wall thickness being
0.8 mm. The ratios of experimental buckling load to the FE estimates were between 0.95 and 2.02. Conical shells are frequently
used in metallic silos and storage tanks where internal pressure is
often an important loading condition. References [9698] discuss
internally pressurized cone-cylinder intersections under uniform
internal pressure. Details about a single, mild steel test conecylinder fabricated by rolling and welding steel sheets of 1 mm
thickness are given in Ref. [96]. Radius and length of the cylinder
were both 500 mm. Cone apex half angle was 60 degrees. The FE
results indicated that buckling load of the cone-cylinder intersections subjected to internal pressure appears not to be sensitive to
initial geometric imperfections. Test data on a further three conecylinder intersections is provided in Ref. [97]. Cones were made
by rolling and seam welding of 1 mm thick mild steel sheets.

Fig. 9 Comparison of buckling loads with Bosor5 predictions


[93]

Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Their apex half-angle was 40 deg. Under single incremental loading, they failed by bifurcation with a number of hoop waves at the
conecylinder junction. Additional conecylinder model, this
time with a horizontal ring of 20 mm depth and placed at the junction, is tested in Ref. [98]. It appears that in all three references,
an exact bifurcation buckling was difficult to establish experimentally and remained a fairly subjective process. The above three
references provide detailed discussions around this subject.
3.3 Buckling Under Axial Compression. Recent tests are
reported in Refs. [99,100]. Five steel models fabricated from steel
sheets by cold rolling and longitudinal seam-welding were collapsed by quasi-static axial compression. The base diameter of all
models was 450 mm and the wall thickness varied between
0.7 mm and 0.9 mm. Heavy plates were attached at both ends to
simulate clamped boundary conditions. Models failed through the
formation of axisymmetric bulge at the small-radius end. The test
ultimate loads were normalized by Rankine limit load, and the ratio varied between 0.98 and 1.11. In addition, 40 aluminum cones
fabricated by spinning [100] were subjected to axial squashing
between platens. The base diameter varied between 256 mm and
304 mm with the wall thickness being between 0.625 mm and
1.45 mm. Although the main objective was to evaluate the energy
absorption and folding mechanisms on the post-collapse path, the
collapse loads are also reported. Results of the FE analyses for
some models are given, and the ratio of experimental collapse force
to the FE estimate, Fexp tl/Fnumerical, was between 0.17 and 0.75.
3.4 Combined Loading. It appears that there has been very
limited research into elasticplastic buckling performance of conical shells, under the action of combined loads. The previous
research effort has primarily been limited to buckling by external
hydrostatic pressure which in fact corresponds to combined loading due to axial compression resulting from pressure action on
top/bottom flanges, e.g., Refs. [86,88,89,91].
In order to gain a better understanding of buckling of cones
under combined loading, two series of buckling tests on laboratory size steel cones have recently been carried out in Liverpool.
Models were CNC-machined and all of them had integral top and
bottom flanges. They were subjected to: (i) axial compression, (ii)
external radial pressure, or (iii) any combination of pressure and
axial compression. Nominal dimensions of the first set of models,
13 in total with designated names C1, C13, were as follows: the
cone semiangle, b 26 deg; the ratio of the larger radius, r2, to
wall thickness, t, was r2/t 34.0; the wall thickness, t 3.0 mm.
Details about experimentation and related FE computations can
be found in Refs. [83,93,101106]. The second set of ten models
had the following geometry: b 14 deg; r2/t 54.0, and the wall
thickness, t 2.0 mm. Here, details about the experiments and the
FE results can be found in Refs. [107111]. Figure 10 shows two,
as manufactured models (b 26 deg in Fig. 10(a), and b 14 deg
in Fig. 10(b)). None of the models were stress relieved prior to
testing. Combined loading was applied to them using arrangement
shown in Fig. 10(c). Heavy top and bottom plates were attached
to each cone. Flanges were partially embedded into the plates in
order to secure clamped-clamped boundary conditions as realistic
as possible. Axial compression was applied through a ram
attached to the bottom plate and connected, through an internal
bar and pivoted coupler, to the top flange (see Refs. [83,101,102).
The whole arrangement seen in Fig. 10(c) was immersed in
350 mm  1000 mm pressure tank. External pressure in the tank
was controlled manually. At the same time, through a separate
pressure line, compressive force was applied via the ram. Various
loading scenarios were explored during experiments. Both sets of
models developed plastic straining before buckling. Hence different loading paths were explored first, and this included: (i) pressure preloading followed by incremental axial force loading, (ii)
axial force preloading followed by incremental radial pressure,
and (iii) proportional loading. References [102,104,106] discuss
Applied Mechanics Reviews

Fig. 10 View of machined cones: (a) b 5 26 deg, (b) b 5 14 deg,


and (c) arrangement for combined loading. Adapted from Ref.
[103].

these in detail. It is shown in them that the magnitudes of buckling


load remain nearly the same for different loading paths; although
there were small differences in values of plastic strains at buckling. Figure 11(a) depicts loading paths for the first set of cones
(models C1,, C13). All models were filled with oil and vented
to the atmosphere. During loading, the amount of expelled oil was
measured. Failure of all models was sudden with large outflow of
oil and accompanied by a loud bang. As mentioned earlier, both
sets of cones were machined from two different billets of steel
and tests were carried out in order to establish the exact material
properties. Several round tensile specimens (10 mm diameter,
200 mm long) were cut from billets in different directions and
they have not been stress relieved prior to uniaxial tensile tests.
Full details about the material properties are available in
Refs. [102,104] for the first set of cones (b 26 deg). Tensile
stressstrain curves confirmed mild steel characteristics of material, i.e., well defined upper/lower yield followed by horizontal
plateau. Details about the evaluation of material properties for the
second set of cones are available in Refs. [108,111]. Under uniaxial tension the second material exhibits continuous strain hardening without a clearly defined yield point. Hence, 0.2% proof stress
was assumed for the yield point. The FE computing related to the
above experimentation identified the first yield envelope in addition to the collapse envelope. These are depicted in Fig. 11 for
b 26 deg and in Fig. 12 for b 14 deg models. It is seen here
that the first yield envelope is of by-and-large bilinear format.
For lower values of axial force, the plastic strains begin to grow
from larger-radius end of the cone. For higher values of axial
force, the plastic strains start to develop at the smaller-radius end
of the cone. The hatched area in Fig. 11 indicates the elasticplastic domain. Experimental results shown in Fig. 11 are cast in two
different formats. Experimental data points in Fig. 11(a) are nortl
malized: by average experimental collapse force, Fexp
avg , and by
tl
(note
that
there
average experimental buckling pressure, pexp
avg
were two nominally identical models tested under pure axial compression and two models tested under pure radial pressure). Experimental data points shown in Fig. 11(a) are normalized by the FE
predicted collapse force, Fcoll
o , and by the FE predicted pure radial
collapse pressure, pcoll . Both interactive plots shown in Figs. 11(a)
and 11(b) were obtained for elastic perfectly plastic modeling of
steel, experimentally measured average geometry and constant,
average wall thickness. Design guidance for the cones subjected
to interactive loads, as discussed above, is provided by the ASME
code [41]. Reference [42], on the other hand, provides the design
guidance for the cases of pure axial compression and lateral pressure only. Magnitudes of design loads were obtained for all tested
cones by using both codes in order to compare them with the test
JANUARY 2014, Vol. 66 / 010803-7

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 11 Combined stability plots for b 5 26 deg cones [106]. Loading paths shown in Fig. 11(a) while configurations given by
the current Design Codes are superimposed in Fig. 11(b).

Fig. 12 Combined stability plot for ten, b 5 14 deg, cones [110]

data. Estimates of implicitly safe values are shown in Fig. 11(b)


for tested models C1,, C13. The ASME predictions are marked
by open circles while the ECCS predictions are marked by asterisks. It is seen here that for some load configurations the recommended values would generate plastic straining under single
incremental loading. This happens, for example, for cones C8, C7,
C12, C5/C6, C1/C2 subjected to combined loading as well as to
C11 subjected to pure lateral pressure. The ECCS based design
010803-8 / Vol. 66, JANUARY 2014

pressure for C11 also falls into the elasticplastic domain. The
same applies to cones C3/C4 subjected to pure axial compression.
While there is a substantial margin of safety against buckling
under a single incremental loading, the predictions of failures
inside of the elasticplastic domain can be of concern under
repeating loads. In the latter case, cones can fail through growth
of plastic strains if there is no shakedown. As mentioned earlier
the second set of experiments was carried out in order to verify
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 13 Photographs of collapsed cones by external hydrostatic pressure


(b 5 26 deg for C6 model, b 5 14 deg for CS6 model)

these unexpected results. It is seen in Fig. 12 that some of recommended configurations would buckle with plastic strains being
developed. This applies to cones CS1, CS6, CS7 (ASME, Ref.
[41]), and CS3/CS4 (ECCS, [42]). Buckling under pure lateral
pressure predicted by both codes remains elastic (CS2/CS5).
Hence the same observation about safety under repeating loads
remains true here for cones with code-based design loads falling
inside of the elasticplastic domain. Finally, views of buckled
cones C6 and CS1 are seen in Fig. 13.
Initial geometric imperfection in conical shells can lower the
buckling strength and there have been numerous studies addressing this topic. However, only a handful of tests have been carried
out on cones with deliberately built-in shape deviation from perfect geometry. It has been generally believed that inward dimpletype axisymmetric shape deviations from perfect geometry are the
most dangerous imperfections, i.e., leading to the largest reduction
of the buckling load. Reference [112], for example, examined the
influence of axisymmetric inward-bulge type shape imperfection
on the magnitude of buckling load (axial compression). The
imperfections were deliberately introduced during manufacturing
(by electro-deposition of copper). Thirty cones with the top end
radius-to-the wall-thickness ratio ranging from 181 to 1115 were
tested. The extreme sensitivity of the buckling load to initial
imperfections of the order of a fraction of the wall thickness was
confirmed experimentally. Imperfections found in real structures
are likely to have neither axisymmetric nor have the shape of
buckling mode but they rather occur locally. One needs to exert
high degree of skills when assessing the load carrying capacity of
complex structures prone to buckling. A three stage approach is
advocated in Ref. [113] for the assessment of imperfection sensitive real structure. The effect of localized dimple-type imperfections on the buckling strength of axially compressed cones was
addressed in Ref. [114]. The FE results showed that buckling load
of the cone with inward axisymmetric imperfection was nearly
equal to the buckling load of local imperfections which extended
60 deg or more around the circumference. Twenty high quality epoxy conical shells were buckled in Ref. [115] by axial compression. The prime objective here was to develop experimental
knock-down factor against buckling. The paper also contains data
on axially compressed imperfect cylinders (built-in inward dimple). Recent numerical results given in Ref. [110] show that outward dimple-type shape distortion can be as bad as the
corresponding inward dimple. This directly contradicts the long
standing view that the inward shape imperfections constitute the
worst case. A subsequent study [116] considered geometrically
imperfect conical shells subjected to axial compression, external
pressure, or simultaneous action of both loads. Axisymmetric
shape imperfection was assumed to be an inward dimple, outward
dimple, or coexisting inward and outward dimples. The profile of
inward bulge was described by (see Fig. 14)

dz

8
>
>
<0



p
>
>
: di cos3
z  zi
bA

Applied Mechanics Reviews

9
bA >
>
2 =
bA >
;
jz  zi j  >
2

where di is the amplitude of the inward imperfection, zi is position


of its center along the generator, and bA is the extension of the
imperfection along the cones slant. The shape of outward axisymmetric bulge was also considered. Its form was given by Eq.
(2) in which zi was substituted by zo, and the slant extension was
assumed to be characterized by, bB (see Ref. [116] for details).
From a design point of view, it would be desirable to know in
advance what would be the lower bound response to any possible
shape deviations one could encounter in reality (inner and/or
outer). Typical parameters which influence the buckling strength
were assumed, and the worst scenario was sought using the Tabu
Search optimization method. Results were obtained for mild steel
cones for which earlier test results on perfect models had been
carried out [103,110]. Interactive diagrams obtained for inward
bulge, outward bulge, and coexisting inward/outward bulges are
shown in Fig. 15. It is seen from Fig. 15 that both inward and outward imperfections can significantly reduce the load carrying
capacity. However, the largest shrinkage of the interactive diagram is obtained for the case of coexisting inward and outward
initial shape imperfections in cones generator. Illustration of
imperfect initial geometry together with shape of the generator at
the collapse is plotted in Figs. 15(b)15(d) for selected points. In
Ref. [118], initial geometric imperfections were taken in the form
of the eigenmode, a single wave extracted from the eigenmode
and localized smooth dimple modeled analytically. Load carrying

jz  zi j 

(2)
Fig. 14 Geometry of inward, axisymmetric dimple imperfection
[116]

JANUARY 2014, Vol. 66 / 010803-9

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 15 The worst interactive stability plots for different imperfection profiles (a). Collapsed shapes at points a, b, and c [116].

capacity of imperfect models was computed along the combined


stability domain using the finite element proprietary code. The FE
results showed that buckling strength of axially compressed and
imperfect cone was only 55% of geometrically perfect model.
Buckling strength of a cone subjected to lateral pressure; on the
other hand, amounted to 43% of the corresponding value of perfect model. However, it was the shrinkage of stability plot of
imperfect cone which was found to be significant. For imperfect
cones subjected to combined axial compression and external pressure, the collapse envelope shrunk by 48% with the elastic sub-set
being reduced by 51%. Numerical study into imperfection sensitivity of buckling loads for cones with the semiangle, b 30, 55,
60, 65, and 70 deg for the case of axial compression, was only carried out in Ref. [119]. Imperfections were taken in the form of
eigenshapes and the effect of different boundary conditions on
buckling was examined. Results are comparable to those obtained
in Ref. [118].
It appears that, so far, there has been no experimental verification of imperfection sensitivity of buckling loads for cones subjected to simultaneously acting axial compression and external
pressure.
Finally, Ref. [117] details development of a test rig for buckling
tests of shell components subjected to pressure and/or axial (centric/eccentric) loading. Test results were obtained for conical
shells made from Mylar. These delivered interactive stability diagrams. The effect of off-axis axial compression on the buckling
strength of pressurized cone was also investigated. Shape measurements gave information on the quality of tested models.

Domed EndsExternal Pressure

4.1 Hemispheres, Torispheres, and Toricones. End closures onto cylindrical vessels can take different shape forms ranging from flat plates to domed ends. The latter can include
spherical caps, hemispherical, ellipsoidal, torispherical or toriconical shapes. These shells also find way into other specialized
applications, e.g., large outer space mirrors. When the loading is
such that the internal stress is dominated by membrane stress
resultants and their associated, relatively high stretching stiffness
then the load carrying capacity becomes very efficient. In doubly
curved ends, the curvature of a shells mid-surface, together with
the high ratio of stretching to bending stiffness, generally leads to
a nonlinear interaction of membrane and bending effects. In cases
like that, the load carrying capacity of domed ends strongly
depends on their geometry, boundary conditions, material
010803-10 / Vol. 66, JANUARY 2014

behavior, type of applied load and the presence, or absence, of initial geometric imperfections. Static stability of domed ends has
been researched for decades both theoretically and experimentally, e.g., Ref. [120]. Review of past efforts in this area can be
found, for example, in Refs [7,121]. It is worth noting here that
the first tests on externally pressurized spherical caps were at the
end of the 19th century. Buckling tests on domed ends of other
shapes have continued until today and they were driven by variety
of reasons. For example, in view of raised concerns that aluminum
caps tested in Ref. [122] had the base diameter only between
20 mm and 50 mm, it was decided to re-examine the sudden drop
in buckling strength around the shallowness parameter k, k 4.0,
1=4
where k  231  t2  H=t1=2 (see Fig. 16 for notation). Six
mild steel caps with 200 mm diameter were carefully machined
from a billet. Each model had an integral heavy base ring, and
joining arrangement of the shell with base ring is shown in
Fig. 16. Three rings were designed to fail elastically and the
remaining three to fail within the elasticplastic range. Six test
points are superimposed in Fig. 16 on the original 1963 test data
by Krenzke and Kiernan. It is seen that indeed there is a minimum
of the load carrying capacity for k  4.0, irrespective whether it is
in the elastic or elasticplastic regime. Further details are available in Ref. [123]. An interesting piece of experimentation is
reported in Ref. [124]. Spherical caps manufactured from brass
with the wall thickness, t 0.4 mm, were subjected to outward radial load applied at the base of the cap. Experimental buckling
loads compared favorably with the theory provided.
An empirical approach to design of externally pressurized
hemispheres, adopted by British Standards Institutions BSI 5500
(now PD 5500), has been reviewed in Ref. [24]. The paper
addresses elastic and elasticplastic buckling. It points to the most
consistent safety factor over the experimental data, together with
a definition of allowable shape that includes both overall shape
allowance and a local defect parameter. Extension of PD 5500 to
externally pressurized torispheres triggered concerns about the
safety factor for sharp knuckle torispheres. Experimental results
of Ref. [125] demonstrate the kind of issues (see Fig. 17). It is
seen here that several test points plot below the lower bound curve
to all previously known experimental points, i.e., domes 1A, 1C,
4A, 4C, 7A, and 7C (see Table 2 for model descriptions). While
domes 1A and 1C are outside the stipulated lower limit on
r/D 0.06, it is the remaining heads tested at Brown University,
which surprisingly fell below the PD 5500 design curve (after
being multiplied by the 1.5 safety factor) which raised the concern
about the universality of the proposed experimental-lower-bound.
In view of these findings, it has transpired there was not much
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 16 Plot of buckling load versus shallowness parameter, k. Also, view of collapsed cap and joining arrangements between the cap and integral base flange
(adapted from Ref. [123]).

Fig. 17 Safe and unsafe domains in PD 5500 code. Models 1A and 1C fall outside admissible geometry stipulated by PD 5500
(a). Test data for ten machined and two spun torispheres is plotted in (b) (pe 1:21Et 2 =Rs2 and pyss 2ryp t=Rs ) adapted from
Ref. [138].
Table 2 Geometry, material properties, collapse pressures, and parameters K and D for externally pressurized torispheres
E
Dome

r/D

Rs/D

Rs/t

L/D

(GPa)

1
1A
1C
4A
4C
7A
7C
P2/1
P2/2
P2/3
P2/4
P4A/1
P4A/2

0.059
0.0427
0.0427
0.0492
0.0492
0.0501
0.0501
0.060
0.058
0.058
0.059
0.060
0.062

1.01
0.749
0.749
1.235
1.235
0.989
0.989
1.05
1.05
1.03
1.04
0.775
0.764

563.64
128.99
126.09
61.47
61.47
45.189
45.189
143.99
148.60
146.00
148.60
83.0
83.0

0.37
0.20
0.20
0.33
0.33
0.37
0.37
0.05
0.05
0.05
0.05
0.09
0.09

207.0
192.4
207.0
192.4
207.0
192.4
207.0
208.0
208.0
208.0
208.0
212.0
212.0

information in the literature on torispherical heads with knuckle


radii, r, in the region of lower limit of r/D 0.06. In response to
this, two research programs were carried out in Liverpool. The
first program concentrated on 24 torispherical heads, some hot
Applied Mechanics Reviews

ryp

pexp tl
(MPa)

370.0
220.4
435.2
243.6
365.4
243.6
365.4
430.0
430.0
430.0
430.0
426.0
426.0

0.128
0.662
1.172
3.862
5.655
5.793
7.586
1.71
1.78
1.78
1.67
4.65
4.62

Ref.

0.643
4.387
2.445
8.329
5.974
11.33
8.126
2.178
2.110
2.148
2.110
3.887
3.887

0.105
0.173
0.182
0.522
0.510
0.576
0.503
0.307
0.330
0.324
0.309
0.485
0.482

[120]
[125]
[125]
[125]
[125]
[125]
[125]
[128]
[128]
[128]
[128]
[127]
[127]

pressed and some cold spun, to acquire information in this range


[126128]. The heads had R/t ratios between 85 and 330 and they
were about 0.75 m in diameter. A few of the test results fell below
the recommended PD 5500 design curve [23] (after they had been
JANUARY 2014, Vol. 66 / 010803-11

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 18 Various stages of manufacturing of a hemisphere to be externally pressurised ((a) and


(b)). View of collapsed hemisphere (c).

multiplied by the 1.5 safety factor, see domes P2 and P4 in


Fig. 17). This led to the second set of tests on 16 torispherical and
2 hemispherical end-closures. Nominal dimensions of these
domes were as follows: diameter of approximately D  800 mm,
and wall thickness, t  6 mm. The ratio of the radius of spherical
portion, Rs, to diameter D, varied from Rs/D 0.75 to Rs/D 1.0.
The ratio of the knuckle radius, r, to diameter, D, varied from r/
D 0.064 to r/D 0.18. Each dome had L 50 mm cylindrical
flange. Some of tested heads were petal-welded simulating a procedure commonly used for fabrication of large end-closures. This
is illustrated in Fig. 18. Eight pressed petals and one spherical cap
are shown in Fig. 18(a). The wooden rig seen in Fig. 18(b) was
used to assemble the segments and then weld them. The shells
investigated were expected to be sensitive to deviations from the
perfect shape (see, for example, Refs [129135]). It was, therefore, decided to carefully monitor shells shape prior to testing,
i.e., after pressing, cutting, and welding. The shape of each head
was measured along 72 meridians at 2.5 deg intervals within the
spherical cap and 5 deg intervals within the knuckle. The cylindrical portion was scanned at 5 mm intervals. The shape measurements were made on the shells inside surface using an LVDT
transducer and a computer based data acquisition system. On average, there were about 2000 measured points per dome. The
same grid was used to measure the domes thickness using an ultrasonic probe. Material properties were obtained from test plates
which had accompanied the heads through their heating and heattreatment cycles. Results were obtained from two specimens per
test plate and, in modeling the stressstrain curve for the use in
subsequent numerical calculations, a multisegment technique was
used. Experimental value of the Youngs modulus was 189.0 GPa
with 0.2% proof stress being, ryp 665 MPa. The above data was
incorporated in several different ways into the numerical analyses.
Detailed description of manufacturing, pretest measurements, and
testing is given in Ref. [136]. Additional information is given in
Ref. [137]. It has been found that there was no definitive pattern
to the relative collapse pressures of the welded and nonwelded
heads. Furthermore, the collapse pressure of all tested domes,
were higher than the predictions of PD 5500 multiplied by the
safety factor of 1.5 (although some of the heads, as delivered, did
not pass the PD 5500 allowable tolerances on shape). Calculations
showed that the loss of strength, in almost all cases, was due to
formation of a single, localized, dimple which gradually deepened

on the post-collapse path. This resembled closely the experimental


collapse mode (see Figs. 18 and 19). The ratios of the experimental to the predicted collapse pressures by the FE were in the range
from 0.85 to 1.14 (for torispheres) and (1.02, 0.91) for two welded
hemispheres. Hemispheres described in Ref. [136] were relatively
thick, i.e., with R/t  62 and manufacturing the whole head would
be difficult due to the possibility of wrinkling. Thinner hemispheres on the other hand could be spun. Here, a series of 0.58 m
diameter with R/t varying between 190 and 780 were examined in
Ref. [131] both experimentally and numerically. One of the main
conclusions made was that one should use the minimum shell
thickness for design purposes and not rely on the average wall
thickness since three test results plotted below the design curve
when the average wall thickness was used. Additional computations have indicated that torispheres with sharp knuckle, r/D 
0.06, might collapse well below the recommended design curve
[138]. A series of 12 collapse tests were carried out on torispheres
with r/D 0.6 in order to verify these findings. There were two
models having nominally the same geometry. Ten heads were
machined laboratory models and further two heads were industrially spun. Experimental results are shown in Fig. 17. It is seen in
Fig. 17(b) that all experimental data falls below the recommended
design curve. The implication of this is that the radius of the
knuckle in the torisphere needs to be considered in the calculations/design. Four titanium alloy spheres made by welding two
halves were collapsed by external pressure as a part of
research into deep sea vehicle [139,140]. Spheres had the
internal diameter of 500 mm, the wall thickness of about 9.0 mm,
and they disintegrated under implosion-type failure, at pressures
55.058.0 MPa. Experimental results compared satisfactorily with
the design equations which took into account shape and wall
thickness imperfections.
It is worth noting here the use of multilayer materials for the
construction of domed ends. These could be entirely layered metal
constructions, multiply constructions assembled from fiber reinforced plastics or hybrids. The first case can be illustrated by
recent buckling tests on spun hemispheres and torispheres from
copper-steel-copper flat sheets, as described in Refs. [78,141,142].
Flat sheets of three layer copper-steel-copper hybrid material
were manufactured by rolling diffusion. The resulting bonding of
layers was strong enough to withstand manufacturing of torispherical and hemispherical domes by spinning. No debonding was

Fig. 19 Photographs of collapsed petal-welded and plain, nominally identical, torispheres

010803-12 / Vol. 66, JANUARY 2014

Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 20 Female molding tool after the first ply being draped. Also, view of hemisphere after the
collapse test.

Fig. 21 Pattern of distorted fibers after draping (a). One quarter of draped fabric superimposed
on the FE grid (b) [151].

found in post-collapsed heads. Filament winding has been widely


used for manufacturing pressure vessels to be loaded by internal
pressure. Wrapping of extruded cylindrical barrels by woven FR
is also known. The use of FRP as a material for manufacturing
domed ends to be subjected to external pressure has a much
shorter history. The use of patches of woven CFRP for reinforcement of deliberately made imperfect, steel torispheres has been
studied in Ref. [143]. Mild steel heads were machined with the
increased-radius flat patch at the apex. They were then reinforced
on inside with the aim of restoring the initial buckling strength
when subjected to external pressure. Details about damaged by
buckling, laboratory size hand laid-up domes, which were
repaired and then retested are available in Ref. [144]. Various
aspects of manufacturing and buckling/collapse of externally pressurised domed ends can be found in Refs. [145154]. These references consider shells of about 200 mm and 800 mm diameter,
their manufacturing, testing and theoretical analyses. The larger
heads were either filament wound or draped from woven fabric. In
both cases, the prepreg material was used (predominantly CFRP),
and there were no internal liners when loaded by external pressure. Some heads, both torispherical and hemispherical, were
laminated using petalled segments as illustrated in Fig. 20. These
segments of woven cloth were butt-jointed and it is seen in Fig.
20 that failure is due to large crack running in the hoop direction
while the butt-joining was not affected. Although the heads were
axisymmetric, their material properties were not. Fiber distortion
and the wall thickness were found from the draping algorithm.
Fiber orientation was measured along a number of meridians
using specially constructed jig. Measured distribution of the fiberangle compared well with predictions given by the draping algorithm [151]. Figure 21 shows how the woven fabric distorts when
draped over doubly curved surfaces. Draping process could be
enhanced by moving the focal point away from the apex in order
to mitigate the wall thickness build-up around the edges
[151,152,155].
The inclusion of toroidal (knuckle) segment between cylinder
and conical vessel end closure is a natural way leading to
Applied Mechanics Reviews

diffusion of the stress jump at the junction (see, for example,


Refs. [38,156]). Results based on parametric studies carried out in
Ref. [156] show that the inclusion of the knuckle can significantly
increase elastic buckling strength. Larger the (r/D)-ratio larger is
the buckling strength. For a wide range of the apex semiangle,
30 deg  b  75 deg, toricones are stronger than the corresponding
cones alone. However, there is very little experimental data in
support of the knuckles role and its influence. With this in mind,
eight steel toriconical shells have been buckled by quasi-static
external pressure in order to measure this problem [157]. The diameter of all models was 200 mm at the base and their wall thickness was 2 mm. The apex semiangle was b 45 deg for all shells.
A summary of experimental results, together with numerical estimates of bifurcation buckling pressures, are provided in Table 3.
It is seen here that experimental buckling pressures vary only
between 3.9 and 4.4 MPa despite large variation in shape. The ratio of experimental to numerically predicted values of buckling
loads varied between 1.03 and 1.18. Numerical estimates of buckling loads given in Table 3 are based on overall average geometry,
axisymmetric model and the use of elasticplastic modeling of
Table 3 Comparison of test data with computed results for
externally pressurised steel cones/toricones
pfailure
exp tl
Model

r/D

T1
T1a
T2
T2a
T3
T3a
T4
T4a

0.2
0.2
0.1
0.1
0.05
0.05
0.0
0.0

pfailure
numerical

 failure 
pdesign

(MPa)
4.38
4.35
4.14
4.10
3.86
3.86
4.14
4.14

3.95
3.97
3.52
3.50
3.68
3.42
4.03
3.99

pfailure
exp tl
pfailure
numerical

6.39
6.39
5.20
5.20
4.55
4.55
5.16
5.16

1.11
1.10
1.18
1.17
1.05
1.13
1.03
1.04

JANUARY 2014, Vol. 66 / 010803-13

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 22 View of cone T4 as machined (a), and after collapse (b). Buckled toricones T2 and T2a
are seen in (c) and (d). External pressure in all cases.

Fig. 23 View of collapsed ellipsoidal shells (adapted from Ref. [159]). External pressure.

material. Views of buckled toricones are provided in Fig. 22. Reference [156] provides design equations for elastic buckling pressures derived from parametric studies, and these are extended to
plastic region as well. According to Reference [156], plastic buckling pressure of a toricone, ppl
toricone , can be approximated by the
following expression:
el
pfailure
design 0:5po 0:4ptoricone

(3)

where elastic buckling pressure, pel


toricone , is approximated by
 t 2:5


2 0:75
E sinbcosb1:5
pel
toricone 13:68 1  t
D

(4)

and D is given by
p
D D 2r cosb  1 2 rt sinb

(5)

The quantity, po is referred to as the yield pressure, and it corresponds to the pressure which causes the spread of plastic strains
anywhere in the shell reaching half of the wall thickness. The values of, po, for different values of (r/D), the yield point of material,
ryp, Youngs modulus, E, and the (D/t)-ratio can be read from
design diagram. When read from Fig. 6 in Ref. [156] these values
are: 4.95 MPa, 2.81 MPa, 2.25 MPa, 2.03 MPa for models T1/T1a,
T2/T2a, T3/T3a, T4/T4a, respectively. The resulting failure pressures are given in column 5 of Table 3. All estimated values are
significantly higher, ranging from 18 to 47%, than the test data.
However, it needs to be said that design the equations, Eqs.
(3)(4), have been obtained for geometrically perfect shells, elastic perfectly plastic modeling of material, and for cone/toricones
being supported by cylindrical shell unlike here [157] where the
heavy base ring simulated clamped boundary conditions. Furthermore, reading values of, po, from a nomogram was not accurate.
However, the tendency of buckling pressure variation with the
knuckle size was similar between the Eqs. (3)(4) and the current
test data.
4.2 Ellipsoids. Elliptical shells of revolution can be used in
specialized applications, e.g., in pressure hulls for rescue-type
submersibles. Early test data is available in Refs. [158,159] where
two prolate ellipsoids, machined from 7075-T6 aluminum, were
collapsed under incremental external pressure (see Fig. 23) for
their view after collapse. The semiaxes were B 75 mm and
A 25 mm, while the wall thickness was 0.76 mm. Despite
010803-14 / Vol. 66, JANUARY 2014

different arrangements for boundary conditions around the equator, the collapse pressures differ only by about 10%. Results of the
buckling tests on an additional 33 machined epoxy resin models
together with the underlining theory are available in Ref. [160].
Elliptical shells of revolution can also be used to close the ends of
externally pressurized vessels, with typical applications in the
submersibles and space vehicle industry. Details about a recent
numerical and experimental study into buckling of steel ellipsoidal domes loaded by static external pressure can be found in Refs.
[77,161,162]. A range of geometries and thicknesses of domes
was examined, as was the influence of different boundary conditions. Shells were examined on the basis of having the same mass.
This meant that all shells were analyzed on a like for like basis,
and as such, each domes performance was easily quantified. Numerical analyses of both perfect and imperfect shells were carried
out. Two kinds of imperfections were considered: deterministic
imperfections derived from measured dimensions and eigenmode
imperfections. The main focus was on prolate domes, i.e., those
taller than a hemisphere of the same radius (Fig. 24). Numerical
predictions were confirmed by pressurizing six laboratory scale
prolate domes to destruction. Three current design codes summarized in these studies included: ASME VIII, PD5500, and the
ECCS. At present, prolate domes are not included in the three
codes. The method of calculating design pressures was outlined
and recommendations were made for the possible inclusion of
prolate ellipsoids into the codes. Currently allowed ellipsoids
(right of hemisphere) and suggested inclusion of prolate domes
(left of hemisphere) are depicted in Fig. 24. All domes in this figure have the same mass. Points (t1, t2, t3) are experimental points
(two tests per point). View of three nominally identical pairs of
prolate ellipsoids after testing is shown in Fig. 24. The ratio A/B
was 0.8, 0.65 and 0.5 for ellipsoids t1, t2, and t3, respectively. It is
also worth mentioning that the new European design rules for
externally pressurized vessels, EN 1993-1-6/Eurocode3, Part 1.6,
Refs. [163,164], do not contain prolate ellipsoidal shells. Oblate
ellipsoids, especially with (B/A)-ratio of two, have been frequently used as closures on cylindrical vessels. Oblate ellipsoids
with other (A/B)-ratios have not been widely used. The same
applies to domed ends whose generators were subject of structural
optimization. References [165,166] are two examples where the
shape of the meridian was searched for the best performance and
the optimum was subsequently subject of experimental verification. The latter reference examined the performance of externally
pressurized ellipsoids through the formal optimization process. In
this study, the highest buckling load was to be found within a
range of geometries of generalized ellipsoidal domes loaded by
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 24 Currently allowed and proposed ellipsoids (a), prolate and oblate geometries sketched in (b), (c), and view of three
pairs of elliptical domes after tests (d)

external pressure. Generalized ellipsoids are a variation of standard ellipsoids in that the exponents used to size the dome are variables. A generalized ellipse is of the form
 x v1  y v2

1
(6)
A
B

where

For a given dome geometry A and B, one has a two-dimensional


design space of variables v1 and v2. In terms of geometry of dome,
when v1 v2 1 the dome is conical. When v1 v2 2 the dome
is a standard 1:2 ellipsoid. Additionally if A B the dome
becomes hemispherical. As v1 and v2 are further increased, the
dome tends towards a closed cylinder. Details about the material
properties can be found in Ref. [166], and they were modeled as
being elasticperfectly plastic. The boundary conditions at the
base of the dome were set to fully clamped.
A hemisphere of the same material properties, boundary conditions, and thickness ratio D/tH 100 had a failure pressure of
pH 11.01 MPa. Mode of failure for this shell was axisymmetric
collapse occurring at the base of the shell. This hemisphere
(hereon termed the reference hemisphere) was the benchmark for
assessing the pressure resistance of generalized ellipsoids.
The condition of constant mass was imposed on all shells,
meaning that thickness of the ellipsoids is determined from its
shape. Where mass is calculated from:

The adaptive Tabu search method was employed to determine values of v1 and v2, which give the maximum pressure resistance of
ellipsoidal domed ends. The results of the Tabu search are shown
in Fig. 25. Also shown in Fig. 25 are the failure pressures for
standard ellipsoids (i.e., v1 v2 2) of the same material, boundary conditions and mass. The optimum shell corresponds to point
a in Fig. 25 and has geometry of v* (1.92857, 1.57937, 1.4).
The optimal dome is 20% stronger than the reference hemisphere,
and has a failure pressure of pmax 12.84 MPa, mode of failure is
axisymmetric collapse.
To confirm numerical results obtained, four geometries of the
dome were machined from a mild steel billet (global optimum
plus further three geometries). The geometries of dome machined
correspond to points a, b, c, and d in Fig. 25, and their
nominal dimensions are given in Ref. [166]. All ellipsoids had the
same mass. Domes were machined in pairs, to demonstrate repeatability of the experiment and also act as a safeguard should one of
the pair be damaged, e.g., during manufacture. The domes were
machined with integral flanges in order to attach them to a base
plate for testing, and also to make sure that no radial movement of
the base was allowed during testing. Before testing, domes were
carefully measured for any variations in shape and thickness
details are in Refs. [166,167].
The experimental failure pressures of all the domes tested are
listed in Table 4. Also given are numerical predictions made using
BOSOR5 based on average thickness. The numerical results are
normalized by the experimental pressures and shown in parentheses. The numerical predictions are all within eight percentage of
the experimental values.
It is worth noting here that structural optimization of pressure
vessel components subject to buckling constraints poses multifaceted challenge. At the structural analysis level, significant insight
is needed into the effects of, for example, the effects of initial geometric imperfections, nonlinear preloading, boundary conditions,
or follower-type loading on the type and magnitude of buckling.
At the optimization level, there is usually very little knowledge
about the nature of design space. For the two cases discussed in
the case of ellipsoids the design space proved to be nonconvex. In
situations like these the use of zero-order approach appears to be a
very efficient approach (Tabu search in this case). A reliable

mshell mref ; with m Stq

(7)

where S is surface area and was calculated by integrating numerically. This condition means that taller, prolate shells will have a
thinner wall due to their larger surface area. Conversely, shallow,
oblate domes will be thicker than the reference hemisphere. In all
cases, the shell wall thickness was kept constant as one moved
along the meridian.
The modes of failure considered here were bifurcation buckling, pbif, and axisymmetric collapse, pcoll. The lower of these two
failure loads was taken to be the critical failure load, pcr.
The optimization problem can be formally expressed as
popt v max pcr v1 ; v2 ; v3

(8)

subject to the constraints


1:5  v1  2:5

(9)

1:5  v2  2:5

(10)

0:3  v3  4:0

(11)

Applied Mechanics Reviews

pcr pcr =pH

(12)

v3 A=B

(13)

v v 1 ; v 2 ; v 3

(14)

JANUARY 2014, Vol. 66 / 010803-15

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 25 Failure pressures of optimized generalized ellipsoids. Also shown is the failure of standard ellipsoids (m1 5 m 2 5 2.0).
Points a, b, c, and d denote experiments (two tests per point) [166].

Table 4 Experimental and numerical values of buckling loads


of CNC machined mild steel domes (numerical values are normalised by experimental values and given in parentheses)
Model
a
c
b
d

a1
a2
c1
c2
b1
b2
d1
d2

pexptl (MPa)

pnumerical (MPa)

13.24
13.46
7.98
8.14
10.21
10.14
5.07
5.07

12.88 (0.97)
12.90 (0.96)
8.18 (1.03)
7.98 (0.98)
9.77 (0.96)
9.69 (0.96)
5.47 (1.08)
5.39 (1.06)

reanalysis could be adopted (usually with a substantial number of


man-years of development and quality assurances). The above
illustrates how this approach provided results for components
with complicated buckling behavior. Ellipsoidal shells have
recently been considered for tanks in rocket launchers. For example, an overall reduction in space vehicle mass has motivated
parametric studies of buckling performance of oblate bulkheads
for a propellant tank [168]. The paper considers both the oblate ellipsoidal shell and the combination of cylinder and bulkhead with
buckling being one of prime design considerations. The buckling
of the three ring-reinforced prolate ellipsoids subjected to external
pressure is reported in Ref. [169]. Three shells, of 200 mm base
diameter, were cast using thermoset polyurethane plastic. They
were then reinforced by equally spaced and externally attached
rings (ten in two domes and 13 in the third model). All three
domes failed through elastic bifurcation buckling. The reinforced
shells were at least four times stronger than their plain counterparts (based on the FE estimates).

4.3 Toroids. Toroidal shells of closed cross-section have


been used commonly for liquid or gas storage in vehicles, aerospace structures and in specialized underwater applications. Due
to its compacted shape, a toroid is also suitable for breathing apparatus. It has proven to be a very desirable storage vessel in
space-limited applications since it permits various systems structures to be routed through its central opening [170]. Thin-walled
toroids, when subjected to external pressure, or vacuum, can
buckle. Early buckling tests are described in Ref. [171], and the
known experiments are listed in Ref. [172]. Results of extensive
parametric studies into stability of externally pressurized toroids
(with circular and noncircular cross-sections; with perfect and
imperfect geometries; metallic as well as from composites), can
010803-16 / Vol. 66, JANUARY 2014

Fig. 26 Bifurcation and collapse pressures for toroidal shell


with circular cross-section and r/t 5 18.74 [177]

be found in Refs. [173178]. Consider a closed toroidal shell, of


uniform wall thickness, t, and being subjected to uniform external
pressure, p. Assume that the mid-surface radius of the tube is, r,
and the distance from the shells axis of rotation to the center of
the cross section is, R, as sketched in Fig. 26. It is not immediately
obvious what boundary conditions should be applied to toroidal
shell shown in Fig. 26. This problem has been addressed in a number of previous studies. One particular approach, adopted in Ref.
[176], was based on a series of numerical runs in which various
combinations of restraints were imposed. The boundary conditions (BCs), which gave the lowest buckling pressure, were therefore identified numerically. Results shown in Fig. 26 have been
obtained for BCs applied at the inner and outer perimeters of the
equatorial plane. The prebuckling and buckling boundary conditions were different. At the prebuckling phase of the analysis:
u 0.0, v  w  b  free. They were applied at the inner equator
only. The outer equator was left unconstrained. The BCs at the
buckling phase of computing were: u  v  w  b  free at both
the inner and outer equators. Typical results can be seen in Fig. 26
where in the insert, Fig. 26(a), prebuckling shape is shown, while
in 26(b) the buckling mode is plotted. Further details and other
results are given in Ref. [176]. Results seen in Fig. 26 show that
bifurcation buckling takes place for toroids with bigger (R/r)ratios. More compact toroids fail by collapse. The transition from
bifurcation to collapse depends not only on the (r/t)-ratio but also
on the yield point of material, ryp. The transition value, (R/r)tr,
between bifurcation and collapse can be calculated from the following equation:
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 27 Two spun halves prior to welding into TS1. Toroids TS1 and TS2 after collapse [177].

the inside of shells was open to the atmosphere during the application of external pressure. The aim here was to avoid an
implosion type collapse of shells. The external pressure was
applied at Dp 0.04 MPa. The end of the load bearing capacity
was associated with a sudden loud bang, large outflow of oil
through the vent, and big drop in pressure. Photographs of shells
after removal from the pressure tank are depicted in Fig. 27 and
Fig. 28. Experimental collapse pressures are given in Table 5 together with numerical estimates based various modeling assumptions. It is seen that a reasonable agreement has been obtained
between experimental failure pressures and numerically predicted
values. The ratio of (pexp tl/pnum) varies between 0.82 and 1.0
while results, which have so far been published in the literature,
oscillate between 0.40 and 1.40.

5
Fig. 28 View of stainless steel toroidal shell, TE1 being lowered to pressure tank for testing (a), and the model after collapse (b) [177]
Table 5 Externally pressurised steel toroids: comparison
of experimental and numerical pressures for different
numerical models (1 5 nominal, 2 5 average, 3 5 min, 4 5 max,
5 5 variable)

Domed EndsInternal Pressure

According to a recent report by the US Department of Energy


[180], there are eight failure modes to be found in internally pressurized vessels, and they are: (i) excessive elastic deformation,
including elastic instability, (ii) excessive plastic deformation,
(iii) brittle fracture, (iv) stress rupture/creep deformation, (v) plastic instabilityincremental collapse, (vi) high strainlow cycle fatigue, (vii) stress corrosion, and (viii) corrosion fatigue.
The current section will review recent research effort related to
elastic/elasticplastic buckling, excessive plastic deformation (plastic
loads), and plastic instabilityincremental collapse (burst pressure).

pnum (MPa)

TS1
TS2
TE1

pexpt(MPa)

9.44(0)
7.10(0)
8.68(0)

8.10(0)
5.66(0)
8.68(0)

5.68(0)
3.68(0)
7.18(0)

9.40(0)
6.64(0)
9.64(0)

8.40(c)
5.20(c)
7.48(0)

8.40
4.28
7.24

 

1:71
R
1:13 0:0486 ryp=E
r=t2:34
r tr

(15)

Equation (15) has been derived in Ref. [176] by curve-fitting of


numerical data generated by wide parametric studies.
The available literature suggests that experimental results on
externally pressurized closed toroidal shells are rare (see for
example Refs. [172,179]). Details about manufacturing, preexperiment measurements, and tests on three steel toroids subjected to external pressure are reported in Refs. [177,179]. Two
models, TS1 and TS2, were manufactured from mild steel by spinning two halves and then welding them around the inner and outer
equatorial perimeters (see Fig. 27). The third model, TE1, has
been assembled by welding four 90-deg stainless 316/316L steel
elbows (see Fig. 28). Prior to tests, shells were filled with oil and
Applied Mechanics Reviews

5.1 Buckling. Domed ends onto internally pressurized cylinders appear in much wider engineering applications than externally pressurized counterparts. Their safe and efficient design has
attracted a significantly larger amount of research effort since, on
one hand, the thinner heads are prone to buckling, and the transition region between bifurcation buckling and axisymmetric yielding is not well defined; however, it is of practical relevance, on
the other hand. In addition, studies to develop inelastic and limit
analyses have also been carried out. Previous work in this area
has been well documented and regularly reviewed. A good
source of relevant information can be found in Refs.
[2,7,20,22,27,181188].
An approximate expression for the limit pressure corresponding
to appreciable plastic deformations in internally pressurized torispheres can be found in Ref. [189]. Experimental verification of
plastic limit analysis for ten torispherical and for three toriconical
shells was discussed in Ref. [190]. Experimental work into the
failure of 12 torispherical ends subjected to internal pressure is
reported in Ref. [191]. All models were manufactured from aluminum alloy and some of them failed through asymmetric bifurcation buckling while the thicker models failed by plastic
deformation. Results of a parametric study into elastic buckling,
and first yielding, for torispherical shells were studied in Ref.
[181]. Comparison of test and theory for asymmetric elasticplastic bifurcation buckling of torispheres is available in Ref. [185].
JANUARY 2014, Vol. 66 / 010803-17

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Experimental bifurcation buckling loads for spun torispherical


and ellipsoidal dished ends are compared with the predictions of
numerically computed values in Ref. [192]. Buckling tests on two
steel torispherical heads are reported in Ref. [186]. These models
were large diameter, approximately 4.8 m shells, and they were
assembled from a number of welded petals. A set of 16 buckling
tests on spun steel torispheres was reported in Ref. [193]. Diameter of these heads was 0.5 m and the ratio of the knuckle radius to
diameter varied from 4 to 10%. The effect of initial shape deviations from perfect geometry on the bifurcation buckling was also
discussed. The results of 190 experimental tests on internally pressurized torispherical heads were collated and analyzed in Refs.
[187,188]. Most of the analyzed torispheres failed by asymmetric
elastic or inelastic bifurcation buckling. In the paper, attention
was also paid to the plastic collapse and burst pressures. Approximate equations for elastic and elasticplastic buckling pressures
(torispherical and ellipsoidal heads) can be found in Ref. [20] together with the details about sources of past studies. They contain
information on more than 200 buckling experiments on internally
pressurized headsincluding buckling tests on large industrially
manufactured domes, e.g., Ref. [186]. The equations would allow
a quick estimate of failure pressure. It is worth noting here the
fact that computed bifurcation buckling can depend on the choice
of plasticity model, i.e., deformation or J2 flow theory. Discussion
of this topic with experimentation on small machined steel heads
is available in Ref. [183]. Buckling tests have confirmed the existence of asymmetric bifurcation as predicted by the deformation
theory while computations based on J2 flow theory concluded that
there should not be bifurcation buckling. The existing paradox
between J2 flow theory and deformation theory of plasticity manifests itself in a wider class of structural problems where buckling
is a possibility (see Refs [7,194,195]). The suggestion to use both
approaches, in order to establish the sensitivity of the predictions
to both models of plasticity, appears appropriate [196]. It is worth
noting here that the unique definition of buckling load can sometimes be very subjective; see for example Ref. [183]. One way
forward here was the adoption of the Southwell plot despite substantial plastic straining. In this way, the subjectivity could be
removed from experimentation (see Refs. [183,197,198]). Elastic
buckling can sometimes go unnoticed and, as a result, it can lead
to serious consequences. Reference [199] illustrates this for the
case of a rear pressure bulkhead in a wide body civilian aircraft.
Due to pressure differences between inside of the cabin and outside of the fuselage, the rear pressure bulkhead can undergo buckling when in the air (i.e., skin within a pocket of reinforcements
by stringers and rings). Upon landing affected skin pops back to
the initial shape. Investigations reported in Ref. [199] concluded
that multiple buckling of the skin within a single skin-pocket was
a cause of in-flight rupture of the rear pressure bulkhead. This led
to the cabin pressure loss and an emergency landing. An interesting picture of multiple pocket-type buckling of a skin in a wing of
flying glider can be seen on the opening page of Ref. [48] (n.b., a
very valuable and multifaceted internet resource on buckling of
shells). Spars and rings provide a load bearing structure for the
gliders wing but under flight conditions the skin can buckle
within a reinforced pocket, and in a number of pockets. While
there is no immediate loss of structural integrity, multiple in-out
buckles can potentially lead to a skins rupture/tearing.
Asymmetric bifurcation is not the only form of design limitation for internally pressurized heads. Under single incremental
loading, plastic load has been introduced as a measure of structural integrity [200] and this concept has been applied to a number
of pressure vessel geometries and piping components. The next
section discusses this concept for internally pressurized torispherical ends and for closed toroidal shells.
5.2 Plastic Loads. There is a continuing discussion aimed at
identification of the best failure criterion for torispherical
heads which do not buckle. The use of plastic collapse loads, as
010803-18 / Vol. 66, JANUARY 2014

Fig. 29 Comparison of computed plastic load pC1 with experimental values. Also, plot of experimental burst pressures. View
of burst models K1, K2, and K6, adapted from Ref. [201].

recommended in Ref. [200] is a frequently adopted approach (see


also Refs. [201203]). The above papers use a common approach
for evaluation of a plastic load, using a graphical relationship
between internal pressure and apex deflection. In Ref. [201], five
pairs of mild steel torispheres were used for experimental extraction of plastic loads. The load-deflection curves were used to
obtain plastic load, pC1 (twice-the-yield-point-deflection) and pC2
(twice-elastic-slope). All models had the same wall thickness and
the same mass, WT. The latter was related to the mass of a hemisphere, WH, having the same diameter. The ratio WT/WH 0.693
was assumed for all tested torispheres and this resulted in a range
of heads having different geometry. Figure 29 shows the comparison of experimental and computed values of pC1 (with pC2 being
only marginally higher and not plotted). Calculations have shown
that calculated plastic collapse pressures differ by no more than
12% from those obtained in experiments. In addition, single measurements of the apex deflection proved to be an effective way of
establishing the twice-the-end-of-proportionality plastic loads.
Incremental loading continued until burst, and most of tested
domes ruptured at the apex with a single long crack (as depicted
in Fig. 29). It is seen here that a substantial reserve of strength
beyond the collapse loads existed. The ratio of (pburst/pC1) varies
here from 1.93 to 2.81 (or from 1.67 to 2.81 if a premature leak is
taken into account). Reference [200] also points out that the evaluation of the plastic load should employ a more objective and justified by physics criterion, i.e., for pressure tests, the relevant
quantities would be pressure and change of internal volume. This
approach was adopted in Refs. [69,204208] in order to obtain
plastic loads for torispherical, ellipsoidal and toroidal models
(mild steel, stainless steel and AA6061-T1 aluminum alloy). Reference [204] studied plastic loads based on (i) load versus deflection and on (ii) a work criterion associated with volumetric
changes in internally pressurized steel torispheres. In both cases,
end of linearity was adopted as a reference point for calculation of
the plastic loads. Plastic loads based on the apex deflection were
found to be comparable to those calculated by using volumetric
changes. This was not true for the knuckle deformations. The corresponding plastic loads were consistently smaller than the apex
or volumetric ones. In addition, it was shown that apex deflections
cannot be used as indicators of the plastic loads for some shapes,
e.g., for prolate ellipsoidal heads. As part of the experimentation,
three machined torispherical models were tested under monotonically increasing pressure up to burst. The ratio of experimental to
numerical values of plastic loads was between 0.92 and 0.99. As
Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Fig. 30 History of plastic strains growth versus number of pressure cycles in steel
torisphere T8 (a). Also views of burst oblate ellipsoidal models: (b) mild steel, and
(c) aluminum (adapted from Ref. [206,207]).

in the previous case, a substantial reserve of strength existed


beyond plastic loads. The ratio of the burst pressure to experimental plastic load varied from 1.9 to 2.3. Generally, a reasonable
comparison of the experimental and numerical results was
obtained. But if these domes were to operate in the post-yield
range where cyclic load could be present, the role of the plastic
load was less clear than the shakedown load. This aspect of pressure loading was explored in Ref. [206] for mild steel heads where
the concept of a first-cycle shakedown pressure was adopted to
compare the relative values of plastic and shakedown loads.
Experimentally studied geometries included torispheres (four
heads) and ellipsoidal shells (two prolate and two oblate models).
The sequence of loading consisted of the following steps: (i)
incremental loading to volumetric plastic load then unload, (ii) cycling loading until shakedown achieved then unload, and (iii)
incremental loading until burst. Torispherical and elliptical end
closures made from mild steel have not lost their serviceability
when pressurized to the level of plastic loads. Experimental measurements of strains at the most stressed point has always led to
shake down to purely elastic state at the plastic load, pV1. Experimental burst pressures were from 1.8 to 3.4 times the plastic load,
pV1. Figure 30 illustrates the accumulation of plastic strains for
torisphere T8 (Rs/D 1.0, r/D 0.10, D/t 25) for which experimental loads were as follows: plastic load, pV1 18.0 MPa, shakedown load, pshk 21.92, and burst pressure, pburst 60.69 MPa.
This particular head lost its integrity by a longitudinal crack passing through the apex (similar to seen in Fig. 29). View of oblate
ellipsoidal head, E1, after burst is also depicted in Fig. 30. This
study was followed by a search for optimal domes with piece wise
distribution of wall thickness subject to shakedown constraints
[209]. In a separate study, the problem of finding the minimum
mass of domed closures on internally pressurised cylinder was
sought [210]. This particular motivation came from light weight
on-board tanks to be used in natural gas vehicles. The plastic load
based on global/volumetric criterion was used and selective experimental benchmarking followed using eight mild steel heads. Plastic and shakedown loads for domes made from strain-hardening
material were studied in Ref. [207]. Results of numerical calculations were benchmarked experimentally using four torispherical
and one oblate ellipsoidal heads machined from AA6061-T1
Applied Mechanics Reviews

aluminum alloy. Shakedown was based here on kinematic strainhardening, and ellipsoidal model, EAL1, after the burst is depicted
in Fig. 30. It appears appropriate here to mention the availability
of 1904 seminal paper by M. T. Huber, which has been translated
into English in 2004 as Ref. [211]. This work has formed the basis
for the HuberMises yield criterion.
While the role of plastic loads for internally pressurized heads
is still being researched, it is the burst pressure which is of greater
value from a practical point of view as it gives an indication of the
margin of safety for a single incremental loading. This is an important quantity, especially, at a design stage or at an emergency
situation.
5.3 Burst Pressure. A number of studies which are relevant
to the issue of burst pressure are available in the literature, and a
brief summary can be found in Ref. [212]. A procedure for numerical calculation of pressure at tensile plastic instability for internally pressurized axisymmetric pressure vessels has been
developed in Refs. [213215]. The proposed pressure is to be an
upper bound to the burst pressure that could be achieved in real
vessels. While the failure mode caused by plastic instability has
been studied analytically and experimentally, it is excessive plastic deformation which is a more probable mode of failure than
bursting due to plastic instability (see Ref. [216]). A recent theoretical and experimental study into burst of internally pressurized
domes can be found in Ref. [212]. The burst pressure is based on
excessive plastic deformation rather than on plastic instability. It
is postulated to use the true plastic strain, eup , corresponding to the
ultimate tensile strength, UTS, for computing the magnitude of
burst pressure. One needs not only the magnitude of plastic strain
but also a place where this strain is to be attained. Based on the
above criterion for admissible magnitude of plastic strain, it is further postulated to define the burst pressure as follows: pburst is the
pressure at which the equivalent plastic strain, PEQQ, reaches the
ultimate plastic strain, eup , anywhere at the mid-surface of a structure. A series of calculations have been performed for mild steel
shallow spherical caps, tested previously for buckling, Ref. [123],
and for torispherical heads tested previously for plastic loads, Ref.
[201]. The latter were from mild steel. In addition, four
JANUARY 2014, Vol. 66 / 010803-19

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

torispheres were freshly machined from aluminum alloy AA6061T1 to complement the series of tests. Hence, burst tests were carried out on: six steel and four aluminum torispheres and six spherical caps. Computed burst pressures for steel spherical caps were
between 21.1 and 4.9% below experimental values. At the
same time, the errors between computed plastic instability pressures and the experiments were between 35.1 and 51.8% (overestimated experimental results). For steel torispheres the
respective ranges of errors were: [12.1%, 32.4%] and
[17.6%, 42.8%]. In aluminum torispheres the ranges were:
[0.4%, 14.0] and [12.7%, 20.8%]. It is seen here that the
magnitudes of computed plastic instability pressures were above
experimental values for all tested models. Derivation of plastic
instability load for internally pressurized stainless steel toroid,
followed by burst tests of two models, can be found in
Refs. [217,218]. Seventeen 45 liter carbon steel toroidal tanks
were burst-tested during qualifying procedure for the anticipated
LPG usage [219,220]. Water was the pressurization media. Tanks
burst when volume increased by about 11%, and the standard
deviation on average burst pressure of 8.56 MPa was about
0.220 MPa. The FE assessment of structural integrity was based
on monitoring the maximum deflections at the apex and at outer
equatorial plane. Other relevant work in this area can be found in
Refs. [221,222].

Closure/Conclusions

The paper shows that experiments related to static buckling of


pressure vessel components have remained relevant to a wide
range of industries over the last decade or so. Skepticism about
the future role of experimentation into buckling, and raised two or
three decades ago, appears to move towards a two way fruitful
coexistence with computational mechanics. Nearly all tests quoted
in this paper had one sort or another theoretical/numerical part
mostly based on the FE approach. However, not all theoretical
predictions manifested themselves in close experimental outcomes, and part of the reason can be attributed to experimentation.
The known issues like material properties, boundary conditions,
loading mechanisms, and test repeatability have gained prominence in labs. Pretest measurements and online monitoring of tests
have seen a step change. The empirical knock-down factors are
being gradually reassessed. Buckling performance of epoxy-based
structural components, and corrosion affected members, is also
actively researched.

References
[1] Bushnell, D., 1985, Computerized Buckling Analysis of Shells, Martinus Nijhoff Publishers, Dordrecht, The Netherlands.
[2] Nash, W. A., 1995, Hydrostatically Loaded Structures, Pergamon Press, New
York.
[3] Ross, C. T. F., 1990, Pressure Vessels Under External Pressure: Statics and
Dynamics, Elsevier Applied Science, London.
[4] Ross, C. T. F., 2001, Pressure VesselExternal Pressure Technology, Horwood Publishing, Chichester, UK.
[5] Spence, J., and Tooth, A. S., eds., 1994, Pressure Vessel Design, Concepts
and Principles, E & FN Spon, London.
[6] Yamaki, N., 1984, Elastic Stability of Circular Cylindrical Shells, Series in
Applied Mathematics and Mechanics, Vol. 27, North-Holland, Amsterdam.
[7] Singer, J., Arbocz, J., and Weller, T., 2002, Buckling ExperimentsExperimental Methods in Buckling of Thin-Walled Structures, Vol. 2, John Wiley &
Sons Inc., New York.
[8] De Borst, R., Kyriakides, S., and Van Baten, T. J., eds., 2002, Stability and
Vibration in Thin-Walled Structures, Int. J. Nonlinear Mech., 37, pp.
5711002.
[9] Dubas, P., and Vandepitte, D., eds., 1987, Stability of Plate and Shell Structures, Ghent University, Belgium.
[10] Galletly, G. D., 1995, Buckling Strength of Imperfection-Sensitive Shells,
Thin-Walled Struct., 23, pp. vii-viii.
[11] Harding, J. E., Dowling, P. J., and Agelidis, N., eds., 1982, Buckling of Shells
in Offshore Structures, Granada Publishing Ltd., St Albans, UK.
[12] Jullien, J. F., ed., 1991, Buckling of Shell Structures, on Land, in the Sea, and
in the Air, Elsevier Science Publishers, London.
[13] Ramm, E., ed., 1982, Buckling of Shells, Springer-Verlag, Berlin.

010803-20 / Vol. 66, JANUARY 2014

[14] Nemeth, M. P., and Starnes, J. H., Jr., 1997, The NASA Monographs on
Shell Stability Design RecommendationsA Review and Suggested
Improvements, Proceedings of the 38th AIAA/ASME/ASCE/ASC Structures,
Structural Dynamics, and Materials Conference, A Collection of Technical
Papers, Part 4, April 710, Kissimmee, FL, Paper No. AIAA-97-1302.
[15] NASA, 1965, Buckling of Thin-Walled Circular Cylinders, NASA Space
Vehicle Design Criteria, Report No. NASA SP-8007.
[16] NASA, 1968, Buckling of Thin-Walled Truncated Cones, NASA Space Vehicle Design Criteria, Report No. NASA SP-8019.
[17] NASA, 1969, Buckling of Thin-Walled Doubly Curved Shells, NASA
Space Vehicle Design Criteria, Report No. NASA SP-8032.
[18] Birkemoe, P. C., 1996, Stability: Directions in Experimental Research, Eng.
Struct., 18, pp. 807811.
[19] Morris, N. F., 1996, Shell Stability: The Long Road from Theory to
Practice, Eng. Struct., 18, pp. 801806.
[20] Bachut, J., and Magnucki, K., 2008, Strength, Stability and Optimization of
Pressure Vessels: Review of Selected Problems, ASME Appl. Mech. Rev.,
61, p. 060801.
[21] Krivoshapko, S. N., 2007, Research on General and Axisymmetric Ellipsoidal Shells Used as Domes, Pressure Vessels, and Tanks, ASME Appl. Mech.
Rev., 60, pp. 336355.
[22] Teng, J. G., 1996, Buckling of Thin Shells: Recent Advances and Trends,
ASME Appl. Mech. Rev., 49, pp. 263274.
[23] PD 5500, 2005, Specification for Unfired Fusion Welded Pressure Vessels,
BSI, London.
[24] Newland, C. N., 1972, Collapse of Domes Under External Pressure, Vessels
Under Buckling Conditions, Paper No. C191/72, IMechE, London, W. Clowes
& Sons Ltd, London, pp. 4352.
[25] Arbocz, J., and Starness, J. H., Jr., 2002, Future Directions and Challenges in
Shell Stability Analysis, Thin-Walled Struct., 40, pp. 729754.
[26] Combescure, A., Hoffman, A., Devos, J., and Baylac, G., 1984, A Review of
ten Years of Theoretical and Experimental Work on Buckling, Recent Advances in Nuclear Component Testing and Theoretical Studies on Buckling, San
Antonio, TX, ASME PVP-1984, Vol. 89, ASME, New York, pp. 2332.
[27] De Paor, C., 2012, The Effect of Random Geometric Imperfections on The
Buckling of Thin Cylindrical Shells Due to External Pressure, Ph.D. thesis,
National University of Ireland, Cork, Ireland.
[28] Degenhardt, R., Bethge, A., Kling, A., Zimmermann, R., Rohwer, K., Klein,
H., Tessmer, J., and Calvi, A., 2007, Probabilistic Approach for Improved
Buckling Knock-Down Factors of CFRP Cylindrical Shells, Proceedings of
the 10th European Conference on Spacecraft Structures, Materials, and Mechanical Testing, DLR, Berlin, Paper No. CEAS-2007-434.
[29] Hilburger, M. W., Nemeth, M. P., and Starnes, J. H., Jr., 2004, Shell Buckling
Design Criteria Based on Manufacturing Imperfection Signatures, Report No.
NASA TM 212659.
[30] MacKay, J. R., and van Keulen, F., 2010, A Review of External Pressure
Testing Techniques for Shells Including a Novel Volume-Control Method,
Exp. Mech., 50, pp. 753772.
[31] Abramovich, H., Singer, J., and Weller, T., 2002, Repeated Buckling
and Its Influence on the Geometrical Imperfections of Stiffened
Cylindrical Shells Under Combined Loading, Int. J. Nonlinear Mech., 37, pp.
577588.
[32] Bisagni, C., and Cordisco, P., 2006, Post-Buckling and Collapse Experiments
of Stiffened Composite Cylindrical Shells Subjected to Axial Loading and
Torque, Compos. Struct., 73, pp. 138149.
[33] Teng, J. G., Zhao, Y., and Lam, L., 2001, Techniques for Buckling Experiments on Steel Silo Transition Junctions, Thin-Walled Struct., 39, pp.
685707.
[34] Hilburger, M. W., and Starnes, J. H., Jr., 2002, Effects of Imperfections on
the Buckling Response of Compression-Loaded Composite Shells, Int. J.
Nonlinear Mech., 37, pp. 623643.
[35] Legay, A., and Combescure, A., 2002, Efficient Algorithms for
Parametric Non-Linear Instability Analysis, Int. J. Nonlinear Mech., 37, pp.
709722.
[36] Reddy, J. N., Arciniega, R. A., and Wang, C. M., 2010, Recent Developments
in the Analysis of Carbon Nanotubes and Nonlinear Shell Theories, Shell
Structures: Theory and Applications, Vol. 2, W. Pietraszkiewicz and I. Kreja,
eds., CRC Press/Balkena, Leiden, The Netherlands, pp. 1118.
[37] Sliz, R., and Chang, M.-Y., 2011, Reliable and Accurate Prediction of the Experimental Buckling of Thin-Walled Cylindrical Shell Under an Axial Load,
Thin-Walled Struct., 49, pp. 409421.
[38] Wunderlich, W., Obrecht, H., Springer, H., and Lu, Z., 1989, A SemiAnalytical Approach to the Non-Linear Analysis of Shells of Revolution,
Analytical and Computational Models of Shells, Proceedings of The Winter
Annual Meeting of the ASME, San Francisco, CA, Dec. 1015, A. K. Noor, T.
Belytschko, and J. C. Simo, eds., CED-Vol. 3, ASME, New York, pp.
509536.
[39] Bachut, J., and Eschenauer, H. A., eds., 2001, Emerging Methods for Multidisciplinary Optimization, Springer, New York.
[40] ASME, 2004, Boiler & Pressure Vessel Code, ASME, New York.
[41] ASME, 2008, Code Case 2286-2, Alternative Rules for Determining Allowable External Pressure and Compressive Stresses for Cylinders, Cones, Sphere
and Formed Heads, Section VIII, Divisions 1 and 20 , Cases of the ASME
Boiler and Pressure Vessel Code, ASME, New York, pp. 113.
[42] ECCS TC8 TWG 8.4 Shells, 2008, Buckling of Steel ShellsEuropean
Design Recommendations, No. 125, 5th ed., ECCS, Multicomp Lda, Sintra,
Portugal.

Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

[43] Samuelson, L. A., and Eggwertz, S., 1992, Shell Stability Handbook, Elsevier
Applied Science, London.
[44] Eggwertz, S. F., and Samuelson, L. A., 1990, Buckling Strength of Spherical
Shells, J. Construct. Steel Res., 17, pp. 195216.
[45] MacKay, J. R., and van Keulen, F., 2013, Partial Safety Factor Approach to
the Design of Submarine Pressure Hulls Using Nonlinear Finite Element Analysis, Finite Elem. Des., 65, pp. 116.
[46] Obrecht, H., Rosenthal, B., Fuchs, P., Lange, S., and Marusczyk, C., 2006,
Buckling, Post Buckling and Imperfection-Sensitivity: Old Questions and
Some New Answers, Comput. Mech., 37, pp. 498506.
[47] Desicos, 2013, Desicos, www.desicos.eu/Publications.html
[48] Bushnell, B., and Bushnell, D., 2013, Shell Buckling, www.shellbuckling.
com
[49] Singer, J., 1997, Experimental Studies in Shell Buckling, Collection of
Technical Papers, Part 3, Proceedings of the 38th AIAA/ASME/ASCE/AHS/
ASC Structures, Structural Dynamics, and Materials Conference, Kissimmee,
FL, April 7-10, Paper No. AIAA-97-1075.
[50] Miller, C. D., 1984, Research Related to Buckling Design of Nuclear Containment, Nucl. Eng. Des., 79, pp. 217227.
[51] MacKay, J. R., van Keulen, F., Smith, M. J., and Pegg, N. G., 2010, Collapse
Mechanics in Externally Loaded Pressure Vessels With Simulated Corrosion
Damage, Shell Structures and Applications, W. Pietraszkiewicz and I. Kreja,
eds., CRC Press/Balkena, Leiden, The Netherlands, pp. 315318.
[52] MacKay, J. R., Smith, M. J., van Keulen, F., Bosman, T. N., and Pegg, N. G.,
2010, Experimental Investigation of the Strength and Stability of Submarine
Pressure Hulls With and Without Artificial Corrosion Damage, Mar. Struct.,
23, pp. 339359.
[53] MacKay, J. R. Jiang, L. Glas, A. H., 2011, Accuracy of Nonlinear Finite
Element Collapse Predictions for Submarine Pressure Hulls With and Without
Artificial Corrosion Damage, Mar. Struct., 24, pp. 292317.
[54] MacKay, J. R., van Keulen, F., 2012, The Sensitivity of Overall Collapse of
Damaged Submarine Pressure Hulls to Material Strength, ASME J. Offshore
Mech. Arct. Eng., 135, pp. 19.
[55] Lo Frano, R., and Forasassi, G., 2009, Experimental Evidence of Imperfection Influence on the Buckling of Thin Cylindrical Shell Under Uniform External Pressure, Nucl. Eng. Des., 239, pp. 193200.
[56] Albus, J., Gomez-Garcia, J., and Oery, H., 2001, Control of Assembly
Induced Stresses and Deformation Due to Waviness of the Interface Flanges
of the ESC-A Upper Stage, Proceedings of the 52nd International Astronautical Congress, Toulouse, France, pp. 19.
[57] Bachut, J., 2010, Buckling of Axially Compressed Cylinders With Imperfect
Length, Comput. Struct., 88, pp. 365374.
[58] Bachut, J., 2013, Buckling of Cylinders With Imperfect Length, Proceedings of the ASME 2013 Pressure Vessel and Piping Conference PVP2013,
Paris, Paper No. PVP2013-97316, ASME, New York, pp. 19.
[59] Da Silva, A., and Limam, A., 2010, Buckling of Thin Cylindrical
Shells Under Combined Loads: Bending, Compression and Internal
Pressure, Proceedings of the 16th US National Congress of Theoretical
and Applied Mechanics, USNCTAM2010-552, State College, PA, June
27July 2.
[60] Da Silva, A., 2011, Buckling of Thin Cylindrical Shells Under Combined
Loadings: Internal Pressure, Compression, Bending and Transverse Shear,
Ph.D. thesis, INSA Lyon, Lyon, France (in French).
[61] Mathon, C., and Limam, A., 2005, Experimental Collapse of Thin Cylindrical
Shells Submitted to Internal Pressure and Pure Bending, Thin-Walled Struct.,
44, pp. 3950.
[62] Da Silva, A., Limam, A., Lorioux, F., Radulovic, S., Taponier, V., and Leudiere,
V., 2012, Buckling of Thin Pressurized Cylinders Under Pure Bending or Axial
Compression: Rocket Launcher Applications, Proceedings of the 12th European Conference on Space Structures, Materials and Environmental Testing,
ESTEC, L. Ouwehand, ed., Noordwijk, The Netherlands, pp. 16.
[63] De Paor, C., Cronin, K., Gleeson, J. P., and Kelliher, D., 2012, Statistical
Characterisation and Modelling of Random Geometric Imperfections in Cylindrical Shells, Thin-Walled Struct., 58, pp. 917.
[64] De Paor, C., Kelliher, D., Cronin, K., Wright, W. M. D., and McSweeney, S.
G., 2012, Prediction of Vacuum-Induced Buckling Pressures of Thin-Walled
Cylinders, Thin-Walled Struct., 55, pp. 110.
[65] Hornung, U., and Saal, H., 2002, Buckling Loads of Tank Shells With
Imperfections, Int. J. Nonlinear Mech., 37, pp. 605621.
[66] Bachut, J., and Smith, P., 2003, Static Stability of Barrelled Shells Under
Hydrostatic Pressure, Proceedings of the ICPVT-10, J. L. Zeman, ed., Oesterreichische Gesellschaft fuer Schweisstechnik, Vienna, pp. 103109.
[67] Bachut, J., and Smith, P., 2008, Buckling of Multi-Segment Underwater
Pressure Hull, J. Ocean Eng., 35, pp. 247260.
[68] Bachut, J., 2002, Buckling of Externally Pressurised Barrelled Shells: A
Comparison of Experiment and Theory, Int. J. Pressure Vessels Piping, 79,
pp. 507517.
[69] Bachut, J., and Wang, P., 2001, Buckling of Barreled Shells Subjected to
External Hydrostatic Pressure, ASME J. Pressure Vessel Technol., 123, pp.
232239.
[70] Bachut, J., 2006, Strength and Bifurcation of Barrelled Composite Cylinders, Shell Structures: Theory and Applications, W. Pietraszkiewicz, and C.
Szymczak, eds., Taylor & Francis, London, pp. 203206.
[71] Bachut, J., 2010, Developments in Strength and Stability of Shell Components Used in Submersibles, Shell Structures: Theory and Applications, Vol.
2, W. Pietraszkiewicz and I. Kreja, eds., CRC Press/Balkena, Leiden, The
Netherlands, pp. 310.

Applied Mechanics Reviews

[72] Jasion, P., and Magnucki, K., 2007, Elastic Buckling of Barrelled Shell
Under External Pressure, Thin-Walled Struct., 45, pp. 393399.
[73] Jasion, P., 2013, Stabilisation of a Post-Critical Behaviour of Sandwich
Cylindrical Shells, Shell Structures: Theory and Applications, Vol. 3,
W. Pietraszkiewicz and J. G
orski, eds., Taylor & Francis, London, pp.
195198.
[74] Bachut, J., 1987, Optimal Barrel-Shaped Shells Under Buckling Constraints, AIAA J., 25, pp. 186188.
[75] Bachut, J., 1987, Combined Axial and Pressure Buckling of Shells
Having Optimal Positive Gaussian Curvature, Comput. Struct., 26, pp.
513519.
[76] Bachut, J., 2003, Optimal Barrelling of Steel Shells Via Simulated
Annealing, Comput. Struct., 81, pp. 19411956.
[77] Eschenauer, H. A., 1989, Shape Optimization of Satellite Tanks for Minimum Weight and Maximum Storage Capacity, Struct. Optim., 1, pp.
171180.
[78] Smith, P., 2007, Optimisation of Shell Components Subject to Stability
Criteria, Ph.D. thesis, the University of Liverpool, Liverpool, UK.
[79] Kendrick, S. B., 1994, The Design of Externally Pressurised Vessels With
BS5500, Pressure Vessels Design, Concepts and Principles, J. Spence, and
A. S. Tooth, eds., pp. 291335.
[80] De Vries, J., 2009, The Imperfection Data Bank and Its Application, Ph.D.
thesis, Delft University, Delft, The Netherlands.
[81] Tyrrell, J., Cremers, J., and Wijker, J., 2005, Buckling Analysis and Qualification Static Load Testing of VEGA Interstage 1/2 Structure, Proceedings of
the 10th European Conference on Spacecraft Structures, Materials, and Mechanical Testing, DLR, Berlin, Paper No. CEAS-2007-433.
[82] Singer, J., 1999, On the Importance of Shell Buckling Experiments, ASME
Appl. Mech. Rev., 52, pp. R17R25.
[83] Bachut, J., and Ifayefunmi, O., 2010, Plastic Buckling of Conical Shells,
ASME J. Offshore Mech. Arct. Eng., 132, p. 041401.
[84] Berkovits, A., Singer, J., and Weller, T., 1967, Buckling of Unstiffened Conical Shells Under Combined Loading, Exp. Mech., 7, pp. 458467.
[85] Singer, J., 1965, On the Buckling of Unstiffened Orthotropic and Stiffened
Conical Shells, Proceedings of the 7th Congress of International Aeronautique, Paris, June 1416, pp. 122.
[86] Golzan, B. S., and Showkati, H., 2008, Buckling of Thin-Walled Conical Shells
Under Uniform External Pressure, Thin-Walled Struct., 46, pp. 516529.
[87] Barkey, M. E., Turgeon, M. C., and Varun Nare, T., 2008, Buckling of ThinWalled Truncated Cones Subjected to External Pressure, Exp. Mech., 48, pp.
281291.
[88] Ross, C. T. F., Little, A. P. F., and Adeniyi, K. A., 2005, Plastic Buckling of
Ring-Stiffened Conical Shells Under External Hydrostatic Pressure, Ocean
Eng., 32, pp. 2136.
[89] Ross, C. T. F., and Little, A. P. F., 2007, Design Charts for the General Instability of Ring-Stiffened Conical Shells Under External Hydrostatic Pressure,
Thin-Walled Struct., 45, pp. 199208.
[90] Ross, C. T. F., Andriosopoulos, G., and Little, A. P. F., 2008, Plastic General
Instability of Ring-Stiffened Conical Shells Under External Pressure, Appl.
Mech. Mater., 1314, pp. 213223.
[91] Ross, C. T. F., 2007, A Proposed Design Chart to Predict the Inelastic Buckling Pressures for Conical Shells Under Uniform External Pressure, Mar.
Technol., 44, pp. 7781.
[92] Esslinger, M., and Van Impe, R., 1987, Theoretical Buckling Loads of Conical Shells, Proceedings of ECCS Colloquium on Stability of Plate and Shell
Structures, P. Dubas and D. Vandepitte, eds., Ghent University, Belgium, pp.
387395.
[93] Bachut, J., 2011, On ElasticPlastic Buckling of Cones, Thin-Walled
Struct., 49, pp. 4552.
[94] Bushnell, D., 1976, Bosor5: Program for Buckling of ElasticPlastic Complex Shells of Revolution Including Large Deflections and Creep, Comput.
Struct., 6, pp. 221239.
[95] Bose, M. R. S. C., Thomas, G., Palaninathan, R., Damodaran, S. P., and Chellapandi, P., 2001, Buckling Investigation on Nuclear Reactor Inner Vessel
Model, Exp. Mech., 41, pp. 144150.
[96] Teng, J. G., and Zhao, Y., 2000, On the Buckling Failure of a Pressure Vessel
With a Conical End, Eng. Failure Anal., 7, pp. 261280.
[97] Zhao, Y., and Teng, J. G., 2001, Buckling Experiments on Cone-Cylinder
Intersections Under Internal Pressure, J. Eng. Mech., 127, pp. 12311239.
[98] Zhao, Y., 2005, Buckling Behaviour of Imperfect Ring-Stiffened Cone-Cylinder Intersections Under Internal Pressure, Int. J. Pressure Vessels Piping,
82, pp. 553564.
[99] Chryssanthopoulos, M. K., and Poggi, C., 2001, Collapse Strength of Unstiffened Conical Shells Under Axial Compression, J. Construct. Steel Res., 57,
pp. 165184.
[100] Gupta, N. K., Sheriff, N. M., and Velmurugan, R., 2006, A Study on Buckling of Thin Conical Frusta Under Axial Loads, Thin-Walled Struct., 44, pp.
986996.
[101] Bachut, J., and Ifayefunmi, O., 2009, Plastic Buckling of Conical Shells,
Proceedings of the ASME 2009 28th International Conference on Ocean, Offshore and Arctic Engineering, OMAE2009, Honolulu, HI, May 31June 5, Paper No. OMAE2009-79219.
[102] Bachut, J., and Ifayefunmi, O., 2010, Buckling of Unstiffened Steel Cones
Subjected to Axial Compression and External Pressure, Proceedings of the
ASME 2010 29th International Conference on Ocean, Offshore and Arctic Engineering, OMAE2010, Shanghai, P. R. C., June 611, Paper No.
OMAE2010-20518.

JANUARY 2014, Vol. 66 / 010803-21

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

[103] Bachut, J., and Ifayefunmi, O., 2012, Buckling of Unstiffened Steel Cones
Subjected to Axial Compression and External Pressure, ASME J. Offshore
Mech. Arct. Eng., 134, p. 031603.
[104] Ifayefunmi, O., 2011, Combined Stability of Conical Shells, Ph.D. thesis,
The University of Liverpool, Liverpool, UK.
[105] Ifayefunmi, O., and Bachut, J., 2011, The Effect of Shape, Thickness and
Boundary Imperfections on Plastic Buckling of Cones, Proceedings of the
ASME 2011 30th International Conference on Ocean, Offshore and Arctic Engineering, OMAE2011, Rotterdam, The Netherlands, June 1924, Paper No.
OMAE2011-49055.
[106] Ifayefunmi, O., and Bachut, J., 2012, Combined Stability of Unstiffened
ConesTheory, Experiments and Design Codes, Int. J. Pressure Vessels Piping, 9394, pp. 5768.
[107] Bachut, J., Ifayefunmi, O., and Corfa, M., 2011, Collapse and Buckling
of Conical Shells, Proceedings of the 21st International Offshore and Polar
Engineering Conference, ISOPE2011, Maui, HI, June 1924, H. W. Jin, B.
Newbury, M. Fujikubo, T. W. Nelson, and O. M. Akelson, eds., ISOPE,
Cupertino, CA, pp. 887893.
[108] Bachut, J., Muc, A., and Rys, J., 2011, Plastic Buckling of Cones Subjected
to Axial Compression and External Pressure, Proceedings of the ASME 2011
Pressure Vessel and Piping Conference, PVP2011, Baltimore, MD, July
1721, Paper No. PVP2011-57618.
[109] Bachut, J., 2012, Buckling of Truncated Cones With Localized
Imperfections, Proceedings of the ASME 2012 Pressure Vessel and
Piping Conference, PVP2012, Toronto, ON, July 1519, Paper No. PVP201278374.
[110] Bachut, J., 2012, Interactive Plastic Buckling of Cones Subjected to Axial
Compression and External Pressure, Ocean Eng., 48, pp. 1016.
[111] Bachut, J., Muc, A., Rys, J., 2013, Plastic Buckling of Cones Subjected to
Axial Compression and External Pressure, ASME J. Pressure Vessel Technology, 135, p. 011205.
[112] Arbocz, J., 1968, Buckling of Conical Shells Under Axial Compression,
Report No. NASA CR-1162.
[113] Jansen, E., Wijker, J., and Arbocz, J., 2007, A Hierarchical Approach for the
Buckling Analysisof the VEGA 1/2 Interstage, Proceedings of the 10th European Conference on Spacecraft Structures, Materials, and Mechanical Testing,
DLR, Berlin, Paper No. CEAS-2007-311.
[114] Cooper, P. A., and Dexter, C. B., 1974, Buckling of Conical Shells With
Local Imperfections, NASA Technical Memorandum, Report No. NASA TM
X-2991.
[115] Foster, C. G., 1987, Axial Compression Buckling of Conical and Cylindrical
Shells, Exp. Mech., 27, pp. 255267.
[116] Bachut, J., 2013, Combined Stability of Geometrically Imperfect Conical
Shells, Thin-Walled Struct., 67, pp. 121128.
[117] Bachut, J., and Stanier, D., 2012, Elastic Buckling of Conical Shells Under
Combined Loading of Axial Compression and External Pressure, Proceedings
of the Eleventh International Conference on Computational Structures Technology, B. H. V. Topping, ed., Civil-Comp Press, Stirlingshire, UK, Paper No.
163.
[118] Ifayefunmi, O., and Bachut, J., 2013, Instabilities in Imperfect Thick Cones
Subjected to Axial Compression and External Pressure, Mar. Struct., 33, pp.
297307.
[119] Wunderlich, W., and Albertin, U., 2000, Analysis and Load Carrying Behaviour of Imperfection Sensitive Shells, Int. J. Numer. Methods Eng., 47, pp.
255273.
[120] Jones, E. O., 1962, The Effect of External Pressure on Thin-Shell Pressure
Vessel Heads, ASME J. Eng. Ind., pp. 205219.
[121] Bachut, J., 1998, Pressure Vessel Components: Some Recent
Developments in Strength and Buckling, Prog. Struct. Eng. Mater., 1, pp.
415421.
[122] Krenzke, M. A., and Kiernan, T. J., 1963, Elastic Stability of Near-Perfect
Shallow Spherical Caps, AIAA J., 1, pp. 28552858.
[123] Bachut, J., 2005, Buckling of Shallow Spherical Caps Subjected to External
Pressure, ASME J. Appl. Mech., 72, pp. 803806.
[124] Chen, J.-S., and Huang, T.-M., 2006, Deformation and Reverse Snapping of
a Circular Shallow Shell Under Uniform Edge Tension, Int. J. Solids Struct.,
43, pp. 77767792.
[125] Washington, C. E., Clifton, R. J., and Costerus, B. W., 1977, Tests of Torispherical Pressure Vessel Head Convex to Pressure, WRC Bulletin, No. 227,
pp. 19.
[126] Galletly, G. D., Bachut, J., and Kru_zelecki, J., 1987, Plastic Buckling of
Imperfect Hemispherical Shells Subjected to External Pressure, Proc. Inst.
Mech. Eng., 201, pp. 153170.
[127] Galletly, G. D., Kru_zelecki, J., Moffat, D. G., and Warrington, B., 1987,
Buckling of Shallow Torispherical Domes Subjected to External Pressure
A Comparison of Experiment, Theory, and Design Codes, J. Strain Anal., 22,
pp. 163175.
[128] Warrington, B., 1984, The Buckling of Torispherical Shells Under
External Pressure, Ph.D. thesis, The University of Liverpool, Liverpool,
UK.
[129] Bachut, J., and Galletly, G. D., 1988, Clamped Torispherical Shells Under
External PressureSome New Results, J. Strain, 23, pp. 924.
[130] Bachut, J., and Galletly, G. D., 1993, Influence of Local Imperfections on
the Collapse Strength of Domed End Closures, Proc. Inst. Mech. Eng., 207,
pp. 197207.
[131] Bachut, J., and Galletly, G. D., 1995, Buckling Strength of Imperfect Steel
Hemispheres, Thin-Walled Struct., 23, pp. 120.

010803-22 / Vol. 66, JANUARY 2014

[132] Bachut, J., and Jaiswal, O. R., 1999, On the Choice of Initial Geometric
Imperfections in Externally Pressurized Shells, ASME J. Pressure Vessel
Technol., 121, pp. 7176.
[133] Galletly, G. D., Bachut, J., and Kru_zelecki, J., 1986, Plastic Buckling of
Externally Pressurized Dome Ends, Proceedings of the International Conference on Advances in Marine Structures, C. S. Smith and J. D. Clarke, eds.,
Elsevier Applied Science, London, pp. 238261.
[134] Galletly, G. D., and Bachut, J., 1991, Buckling Design of Imperfect Welded
Hemispherical Shells Subjected to External Pressure, Proc. Inst. Mech. Eng.,
Part C: J. Mech. Eng. Sci., 205, pp. 175188.

[135] Ory,
H., Reimerdes, H.-G., Schmid, T., Rittweger, A., and Gomez Garcia, J.,
2002, Imperfection Sensitivity of an Orthotropic Spherical Shell Under
External Pressure, Int. J. Nonlinear Mech., 37, pp. 669686.
[136] Moffat, D. G., Bachut, J., James, S., and Galletly, G. D., 1992, Collapse of
Externally Pressurised Petal-Welded Torispherical and Hemispherical
Pressure Vessel End-Closures, Pressure Vessel Technology, Report No.
ICPVT-7.
[137] Bachut, J., Galletly, G. D., and Moffat, D. G., 1991, An Experimental and
Numerical Study into the Collapse Strength of Steel Domes, Buckling of
Shell Structures on Land, in the Sea and in the Air, J. F. Jullien, ed., Elsevier
Applied Science, London, pp. 344358.
[138] Bachut, J., 1998, Buckling of Sharp Knuckle Torispheres Under External
Pressure, Thin-Walled Struct., 30, pp. 5577.
[139] Pan, B. B., Cui, W. C., Shen, Y. S., and Liu, T., 2010, Further Study on the
Ultimate Strength Analysis of Spherical Pressure Hulls, Mar. Struct., 23, pp.
444461.
[140] Pan, B. B., Cui, W. C., and Shen, Y. S., 2012, Experimental Verification of
the New Ultimate Strength Equation of Spherical Pressure Hulls, Mar.
Struct., 29, pp. 169176.
[141] Bachut, J., and Smith, P., 2007, Buckling of Multilayered Metal Composite
Domes, Proceedings of the 10th European Conference on Spacecraft Structures, Materials, and Mechanical Testing, DLR, Berlin, Paper No. CEAS2007-435.
[142] Bachut, J., 2009, Buckling of Multilayered Metal Domes, Thin-Walled
Struct., 47, pp. 14291438.
[143] Galletly, G. D., and Muc, A., 1988, Buckling of Fibre Reinforced Plastic
Steel Torispherical Shells Under External Pressure, Proc. Inst. Mech. Eng.,
Part C: J. Mech. Eng. Sci., 202, pp. 409420.
[144] Mistry, J., and Levy-Neto, F., 1992, The Behaviour of Repaired
Composite Domes Subjected to External Pressure, Composites, 23, pp.
271277.
[145] Bachut, J., Galletly, G. D., and Gibson, A. G., 1990, Collapse Behaviour of
CFRP Domes Under External Pressure, Proceedings of the IMechE Fourth
International Conference FRC-90, Fibre Reinforced Composites, Paper No.
C400/058.
[146] Bachut, J., Galletly, G. D., and Gibson, A. G., 1990, CFRP Domes Subjected
to External Pressure, Mar. Struct., 3, pp. 149173.
[147] Bachut, J., Galletly, G. D., and Levy-Neto, F., 1991, Towards Optimum Carbon-Fibre-Reinforced Plastic End Closures, Proc. Inst. Mech. Eng., 205, pp.
329342.
[148] Bachut, J., 1992, Influence of Meridional Shaping on the Collapse strength
of FRP Domes, Eng. Optim., 19, pp. 6580.
[149] Bachut, J., and Galletly, G. D., 1992, Externally Pressurized Hemispherical
Fibre-Reinforced Plastic Shells, Proc. Inst. Mech. Eng., Part C: J. Mech. Eng.
Sci., 206, pp. 179191.
[150] Bachut, J., 1993, Filament Wound Torispheres Under External Pressure,
Compos. Struct., 26, pp. 4754.
[151] Bachut, J., 2000, Modelling and Analysis of Multiply Torispheres Draped
from Woven Carbon Fabric, Comput. Struct., 76, pp. 19.
[152] Dong, L. Bachut, J., 1998, Analysis and Collapse of Thick Composite
Torispheres, Proc. Inst. Mech. Eng., Part E, 212, pp. 103117.
[153] Galletly, G. D., and Bachut, J., 1991, Collapse Strength of Composite
Domes Under External Pressure, Advances in Marine Structures, Vol. 2, C.
S. Smith and R. S. Dow, eds., Elsevier Applied Science, London, pp.
708732.
[154] Neto, F. L., 1991, The Behaviour of Externally Pressurised Composite
Domes, Ph.D. thesis, The University of Liverpool, Liverpool, UK.
[155] Bachut, J., and Dong, L., 1997, Use of Woven CFRP for Externally Pressurized Domes, Compos. Struct., 38, pp. 553563.
[156] Wunderlich, W., Obrecht, H., and Schnabel, F., 1987, Nonlinear Behaviour
of Externally Pressurized Toriconical ShellsAnalysis and Design Criteria,
Proceedings of the ECCS Colloquium on Stability of Plate and Shell Structures, P. Dubas and D. Vandepitte, eds., Ghent University, Belgium, pp.
373384.
[157] Bachut, J., 2013, Externally Pressurized ToriconesBuckling Tests, Shell
Structures: Theory and Applications, Vol. 3, W. Pietraszkiewicz and J. G
orski,
eds., Taylor & Francis, London, pp. 183186.
[158] Healey, J. J., 1965, Exploratory Tests of Two Prolate Spheroidal Shells
Under External Hydrostatic Pressure, David Taylor Model Basin Report No.
1868.
[159] Healey, J. J., 1965, Hydrostatic Tests of Two Prolate Spheroidal Shells, J.
Ship Res., 9, pp. 7778.
[160] Hyman, B. I., and Healey, J. J., 1967, Buckling of Prolate Spheroidal Shells
Under Hydrostatic Pressure, AIAA J., 5, pp. 14691477.
[161] Smith, P., and Bachut, J., 2006, Buckling of Externally Pressurised Prolate
Ellipsoidal Domes, Proceedings of PVP2006-ICPVT-11 2006 ASME Pressure Vessels and Piping Division Conference, Vancouver, BC, Canada, July

Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

[162]
[163]
[164]

[165]

[166]
[167]

[168]

[169]

[170]
[171]

[172]

[173]

[174]

[175]

[176]
[177]
[178]
[179]

[180]
[181]

[182]

[183]

[184]

[185]

[186]

[187]
[188]
[189]
[190]

2327, High-Pressure Technology, W. J. Bees and I. T. Kisisel, eds.,


ASME NDE Division, ASME Pipeline Systems Division, New York, pp.
329337.
Smith, P., and Bachut, J., 2008, Buckling of Externally Pressurised
Prolate Ellipsoidal Domes, ASME J. Pressure Vessel Technol., 130, pp. 19.
EN 1993-1-6/Eurocode3, 2004, Design of Steel Structures, Part 1-6: Stability
of Shell Structures, CEN, Belgium.
Wunderlich, W., 2004, Stability of Spherical Shells Under External Pressure, Progress in Structural Engineering, Mechanics and Computation, A.
Zignoni, ed., Taylor & Francis, London, pp. 155160.
Bachut, J., Galletly, G. D., and Moreton, D. N., 1990, Buckling of NearPerfect Steel Torispherical and Hemispherical Shells Subjected to External
Pressure, AIAA J., 28, pp. 19711975.
Bachut, J., and Smith, P., 2007, Tabu Search Optimization of Externally
Pressurized Barrels and Domes, Eng. Optim., 39, pp. 899918.
Smith, P., and Bachut, J., 2006, Optimization of Externally Pressurised
Domes Using Tabu Search, Proceedings of the 6th ASMO-UK/ISSMO Internatioanl Conference on Engineering Design Optimization, Oxford, UK, July
34, pp. 111119.
Smeltzer, S. S., and Bowman, L. M., 2002, Buckling Design Studies of
Inverted, Oblate Bulkheads for a Propellant Tank, Proceedings of the 43rd
AIAA/ASME/ASCE/AHS/ASC Structures, Structural Dynamics, and Materials Conference, April 2225, Paper No. AIAA-2002-1525.
Ross, C. T. F., Little, A. P. F., Chasapides, L., Banks, J., and Attanasio, D.,
2004, Buckling of Ring Stiffened Domes Under External Hydrostatic Pressure, Ocean Eng., 31, pp. 239252.
Bunkoczy, B., 1988, Methods for Manufacturing a Toroidal Pressure Vessel,
US Patent No. 4,790,472.
Nordell, W. J., and Crawford, J. E., 1971, Analysis of Behaviour of Unstiffened Toroidal Shells, Proceedings of the Pacific Symposium on Hydromechanically Loaded Shells, University of Hawaii, Honolulu, HI, IASS Paper
No. 4-4.
Galletly, G. D., and Bachut, J., 1995, Stability of Complete Circular and
Non-Circular Toroidal Shells, Proc. Inst. Mech. Eng., Part C: J. Mech. Eng.
Sci., 209, pp. 245255.
Bachut, J., and Jaiswal, O. R., 1998, Buckling of Imperfect Ellipsoids and
Closed Toroids Subjected to External Pressure PVP Vol. 368, Analysis and
Design of Composite, Process, and Power Piping and Vessels, D. K. Williams,
ed., ASME, New York, pp. 121128.
Bachut, J., and Jaiswal, O. R., 1998, Buckling Under External Pressure of
Closed Toroids With Circular and Elliptical Cross Sections, Advances in
Civil and Structural Engineering Computing Practice, B. H. V. Topping, ed.,
Civil-Comp Press, Edinburgh, UK, pp. 323334.
Bachut, J., and Jaiswal, O. R., 1999, Instabilities in Torispheres and Toroids
Under Suddenly Applied External Pressure, Int. J. Impact Eng., 22, pp.
511530.
Bachut, J., and Jaiswal, O. R., 2000, On Buckling of Toroidal Shells Under
External Pressure, Comput. Struct., 77, pp. 233251.
Bachut, J., 2003, Collapse Tests on Externally Pressurized Toroids, ASME
J. Pressure Vessel Technol., 125, pp. 9196.
Bachut, J., 2004, Buckling and First Ply Failure of Composite Toroidal Pressure Hull, Comput. Struct., 82, pp. 19811992.
Bachut, J., 2002, Collapse Tests on Externally Pressurised Toroids, PVP
Vol. 439, Pressure Vessel and Piping Codes and Standards, M. D. Rana, ed.,
ASME, New York, pp. 6572.
US Department of Energy, 2005, Design and Development Guide, US
Departmnet of Energy, Washington, DC.
Aylward, R. W., and Galletly, G. D., 1979, Elastic Buckling of, and, First
Yielding, in Thin Torispherical Shells Subjected to Internal Pressure, Int. J.
Pressure Vessels Piping, 7, pp. 321336.
Galletly, G. D., and Bachut, J., 1985, Torispherical Shells Under Internal
Pressure-Failure Due to Asymmetric Plastic Buckling or Axisymmetric
Yielding, Proc. Inst. Mech. Eng., 199, pp. 225238.
Galletly, G. D., Bachut, J., and Moreton, D. N., 1990, Internally Pressurized
Machined Domed EndsA Comparison of the Plastic Buckling Predictions of
the Deformation and Flow Theories, Proc. Inst. Mech. Eng., 204, pp.
169186.
Hamada, M., Morita, T., and Morisawa, Y., 1989, On the Buckling Problem
of a Pressure Vessel Torispherical Head Due to Internal Pressure, Int. J. Pressure Vessels Piping, 36, pp. 327340.
Lagae, G., and Bushnell, D., 1978, Elastic Plastic Buckling of Internally
Pressurized Torispherical Vessel Heads, Nucl. Eng. Des., 48, pp.
405414.
Miller, C. D., and Grove, R. B., 1986, Pressure Testing of Large-Scale Torispherical Heads Subject to Knuckle Buckling, Int. J. Pressure Vessels Piping,
22, pp. 147159.
Miller, C. D., 1999, Buckling Criteria for Torispherical Heads Under Internal
Pressure, WRC Bulletin, No. 444, pp. 199.
Miller, C. D., 2001, Buckling Criteria for Torispherical Heads Under
Internal Pressure, ASME J. Pressure Vessel Technol., 123, pp. 318323.
Shield, R. T., and Drucker, D. C., 1961, Design of Thin-Walled Torispherical
and Toriconical Pressure Vessel Heads, ASME J. Appl. Mech., 28, pp. 292297.
Save, M., 1972, Experimental Verification of Plastic Limit Analysis of Torispherical and Toriconical Heads, Pressure Vessels and Piping Design and
AnalysisA Decade of Progress, Vol. 1, G. J. Bohm, R. L. Cloud, L. C. Hsu,
and D. H. Pai, eds., ASME, New York, pp. 382416.

Applied Mechanics Reviews

[191] Kirk, A., and Gill, S. S., 1975, The Failure of Torispherical Ends of Pressure
Vessels Due to Instability and Plastic DeformationAn Experimental Investigation, Int. J. Mech. Sci., 17, pp. 525544.
[192] Galletly, G. D., 1981, Plastic Buckling of Torispherical and Ellipsoidal Shells
Subjected to Internal Pressure, Proc. Inst. Mech. Eng., 195, pp. 329345.
[193] Roche, R. L., and Autrusson, B., 1986, Experimental Tests on Buckling of
Torispherical Heads and Methods of Plastic Bifurcation Analysis, ASME J.
Pressure Vessel Technol., 108, pp. 138145.
[194] Bachut, J., Galletly, G. D., and James, S., 1996, On the Plastic Buckling Paradox for Cylindrical Shells, Proc. Inst. Mech. Eng., 210, pp. 477488.
[195] Giezen, J. J., Babcock, C. D., and Singer, J., 1991, Plastic Buckling of Cylindrical Shells Under Biaxial Loading, Exp. Mech., 31, pp. 337343.
[196] Bushnell, D., 1981, Buckling Od ShellsPitfall for Designers, AIAA J., 19,
pp. 11831226.
[197] Souza, M. A., Fok, W. C., and Walker, A. C., 1983, Review of Experimental
Techniques for Thin-Walled Structures Liable to Buckling, Exp. Tech., 7, pp.
2139.
[198] Singer, J., 1989, On the Applicability of the Southwell Plot to Plastic
Buckling, Exp. Mech., 29, pp. 205208.
[199] AAIB Bulletin No. 2/96, 1996, Rupture of Rear Pressure Bulkhead, Air
Accidents Investigation Branch, Aldershot, UK, pp. 3448.
[200] Gerdeen, J. C., 1979, A Critical Evaluation of Plastic Behaviour Data and a
United Definition of Plastic Loads for Pressure Components, WRC Bulletin,
No. 54, pp. 182.
[201] Bachut, J., 1995, Plastic Loads for Internally Pressurised Torispheres, Int. J.
Pressure Vessels Piping, 64, pp. 91100.
[202] Kalnins, A., and Updike, D. P., 1991, New design curves for torispherical
heads, WRC Bulletin, No. 364, pp. 161.
[203] Kalnins, A., and Rana, R. D., 1996, A New Design Criterion Based on Pressure Testing of Torispherical Heads, WRC Bulletin, No. 414, pp. 160.
[204] Bachut, J., and Ramachandra, L. S., 1997, The Failure of Internally Pressurised Vessel Heads Due to Yielding, Proceedings of the 8th International Conference on Pressure Vessel Technology, ICPVT-8, Montreal, QC, Canada,
July 2126, Design and Analysis, Vol. 2, A. Chaaban, ed., ASME, New York,
pp. 207216.
[205] Bachut, J., and Ramachandra, L. S., 1997, Shakedown and Plastic Loads for
Internally Pressurised Vessel End Closures, Proceedings of the International
Conference on Carrying Capacity of Steel Shell Structures, pp. 130137.
[206] Bachut, J., Ramachandra, L. S., and Krishnan, P. A., 1998, Experimental
and Numerical Investigation of Plastic Loads for Internally Pressurised Vessel
Heads, Pressure Vessel and Piping Codes and Standards, B. T. Lubin and T.
Tahara, eds., ASME, New York, pp. 345359.
[207] Bachut, J., Ramachandra, L. S., and Krishnan, P. A., 2000, Plastic and
Shakedown Loads for Internally Pressurised Domes Made from Strain Hardening Material, Proceedings of the ICPVT-9, Vol. 1, Sydney, Australia, pp.
5766.
[208] Bachut, J., 2000, Plastic Loads for Internally Pressurised Toroidal Shells,
ASME J. Pressure Vessel Technol., 127, pp. 151156.
[209] Bachut, J., and Ramachandra, L. S., 1997, Optimization of Internally Pressurized Torispheres Subject to Shakedown via GAs, Eng. Optim., 29, pp.
113129.
[210] Bachut, J., 1997, Minimum Weight of Internally Pressurised Domes Subject
to Plastic Load Failure, Thin-Walled Struct., 27, pp. 127146.
[211] Huber, M. T., 2004, Specific Work of Strain as a Measure of Material Effort,
Arch. Mech., 56, pp. 173190.
[212] Bachut, J., and Vu, V. T., 2007, Burst Pressures for Torispheres and Shallow
Spherical Caps, Strain, 43, pp. 2636.
[213] Updike, D. P., and Kalnins, A., 1994, Burst by Tensile Plastic Instability of
Vessels With Torispherical Heads, Recertification and Stress Classification
Issues, ASME PVP Vol. 277, J. N. Petrinec, ed., ASME, New York, pp.
8994.
[214] Updike, D. P., and Kalnins, A., 1997, Ultimate Load Analysis for Design of
Pressure Vessels Proceedings of the 1997 ASME Pressure Vessels and Piping
Conference, ASME PVP, Vol. 353, Pressure Vessel and Piping Codes and
Standards, ASME, New York, pp. 289293.
[215] Updike, D. P., and Kalnins, A., 1998, Tensile Plastic Instability of Axisymmetric Pressure Vessels, ASME J. Pressure Vessel Technol., 120, pp. 611.
[216] Langer, B. F., 1971, Design-Stress Basis for Pressure Vessels, Exp. Mech.,
11, pp. 111.
[217] Vu, V. T., 2008, Burst and Minimum Weight Design of Pressure Vessel
Components by Modern Optimisation Techniques, Ph.D. thesis, The University of Liverpool, Liverpool, UK.
[218] Vu, V. T., and Bachut, J., 2009, Plastic Instability Pressure of Toroidal
Shells, ASME J. Pressure Vessel Technol., 131, p. 051203.
[219] Kisioglu, Y., 2009, Burst Tests and Volume Expansions of Vehicle Toroidal
LPG Fuel Tanks, Turk. J. Eng. Environ. Sci., 33, pp. 117125.
[220] Kisioglu, Y., 2011, Burst Pressure Determination of Vehicle Toroidal Oval
Cross-Section LPG Fuel Tanks, ASME J. Pressure Vessel Technol., 133, p.
031202.
[221] Tong, R., and Wang, X., 1997, Simplified Method Based on the Deformation
Theory for Structural Limit AnalysisII. Numerical Application and Investigation on Mesh Density, Int. J. Pressure Vessels and Piping, 70, pp. 5158.
[222] Wasicek, M., Fischer, F. D., Sabirov, I., and Kolednik, O., The Burst Pressure
of a Cylindrical Vessel With a Conical Bottom and a Torispherical Head,
Proceedings of the ICPVT-10, J. L. Zeman, ed., Oesterreichische Gesellschaft
fuer Schweisstechnik, Vienna, pp. 95102.

JANUARY 2014, Vol. 66 / 010803-23

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Jan Bachut received his MS.c. in Physics in 1971 (Jagiellonian University, Krakow, Poland), Ph.D. in 1980,
and DS.c. in 1996. The latter two were in Mechanics and both from Cracow University of Technology
(CUT). For nearly a decade he worked at the Institute of Physics, CUT and then for more than 30 years he
taught and performed research at Mechanical Engineering, The University of Liverpool, UK. Currently he
holds the Chair in Mechanics at the Institute of Physics, CUT, and is Visiting Academic at the University of
Manchester Aerospace Research Institute, UK. He has published more than 150 papers in reviewed scientific
journals and proceedings and serves on Editorial Boards of Engineering Optimization and Technical Transactions journals. Recently he was co-chairman of Euromech Colloquium in Liverpool and CISM course in
Udineboth devoted to various aspects of structural optimization. He is a full member of the ECCS Technical Working Group 8.4 on Buckling of Shells, and has worked for Sub-Committee on Design Methods PVE/
5British Standards Institution. His research interests include buckling of shells, integrity of pressure vessels, and structural optimization of components aimed at on land, in the sea and in the air applications.

010803-24 / Vol. 66, JANUARY 2014

Transactions of the ASME

Downloaded From: http://appliedmechanicsreviews.asmedigitalcollection.asme.org/ on 07/16/2014 Terms of Use: http://asme.org/terms

Anda mungkin juga menyukai