You are on page 1of 17

GOODRICH PROPRIETARY

February 2004






ELASTIC BUCKLING and POSTBUCKLING

of

SHELL STRUCTURES


REQUIREMENTS and GUIDE

for

FINITE ELEMENT MODELING



GOODRICH PROPRIETARY
ELASTIC BUCKLING AND POSTBUCKLING OF SHELL STRUCTURES

REQUIREMENTS AND GUIDE

FOR

FINITE ELEMENT MODELING




TABLE OF CONTENTS



1) GENERAL REQUIREMENTS FOR STRUCTURAL STABILITY ANALYSES...................... 3
2) OUTLINE TO FEM APPROACH FOR POSTBUCKLING ANALYSES................................... 4
3) MESH REFINEMENT GUIDELINES.......................................................................................... 6
4) EXAMPLES OF MESH REFINEMENT EFFECTS..................................................................... 8
4.1 Flat Panel Mesh Size Sensitivity Results .................................................................................. 8
4.2 Cylindrical Shell Mesh Size Sensitivity Results Part 1............................................................. 9
4.3 Cylindrical Shell Mesh Size Sensitivity Results Part 2........................................................... 10
5) ANALYSIS FACTORS, ISSUES AND COMPLICATIONS..................................................... 12
5.1 Finite Element Incompatibility Leads To Phantom Modes..................................................... 12
5.2 FEM Analysis Factors and Complications .............................................................................. 13
5.3 Post-Buckling Initial Imperfections......................................................................................... 15
6) NASTRAN FEM ANALYSIS LIMITATIONS........................................................................... 16
7) REFERENCES............................................................................................................................. 17




BY: Rich Alloway
PHONE: 619.498.7448
REVISION DATE: Feb 23, 2004
GOODRICH PROPRIETARY

1) GENERAL REQUIREMENTS FOR STRUCTURAL STABILITY ANALYSES

Buckling shall not cause components that are subject to instability to collapse when ultimate loads
are applied, nor shall buckling deformation from limit loads degrade the functioning of any system
or produce changes in loading (or deflections) that are not accounted for, Ref. 1. Buckling is the
instability phenomenon in a beam, plate or shell structure where disproportionately large nonlinear
structural deformations occur from a relatively small increase in the external loading. Buckling
deformations can develop in a gradual or sudden manner.

Buckling shall not be permitted at or below the maximum fatigue loads on the structure. For
fatigue critical structure with buckling allowed below limit load, the effect of small imperfections
on the fatigue stress levels shall be evaluated. In addition, structures that buckle below limit loads
shall be assessed for material yielding at limit load levels according to the program requirements.

Evaluation of buckling strength shall consider the combined action of primary and secondary
stresses and their effects on (1) general instability resulting in the collapse of the component, (2)
local or panel instability, (3) reductions in stiffness due to plastic deformation and (4) reductions in
stiffness due to creep deformation.

All structural components that are subject to compressive in-plane stresses under any combination
of ground loads, flight loads, or loads resulting from temperature changes shall be analyzed or
tested for buckling failure. Design loads for buckling (collapse of the component) shall be ultimate
loads, except that any load component that tends to alleviate buckling shall not be increased by the
ultimate factor of safety. Thus, in the case of a cylinder under destabilizing crush pressure and
torsion loads, both limit loads shall be increased by the ultimate factor of safety. However, when a
stabilizing burst pressure is combined with a destabilizing torsion load, the destabilizing torsion
limit load is increased by the factor of safety and then combined with the limit burst pressure (no
factor of safety is applied to the pressure since it stabilizes the cylinder).

The analysis of all structural members shall include consideration of the relative deflection of
adjacent members and the resulting loads imposed on all members. In addition, deflections of all
major load-carrying structure and equipment-support structure shall be calculated to ensure that the
structure does not interfere with an adjacent system and to ensure that equipment is not adversely
affected by motion of the equipment package support points.

When the design thickness is not specified in the program specification, the following definition
will be used. The design thickness, t
D
, for each structural member other than mechanically or
chemically milled pressure vessels shall be determined by the following steps:
1) Determine the thickness variations in the as received material and those due to forming and
processing
2) Determine the maximum thickness, t
Max
, and minimum thickness, t
Min

3) Calculate the mean thickness based on equal plus and minus tolerances: t
D1
= (t
Max
+ t
Min
)/2
4) Calculate a lower bound thickness based on strength or stability calculations:
For strength design use: t
D2
= 1.10* t
Min

GOODRICH PROPRIETARY
For stability design use: t
D2
= 1.05* t
Min

5) The design thickness, t
D
, is the minimum of t
D1
and t
D2
i.e. t
D
= min( t
D1
, t
D2
)
The mean and minimum design thicknesses shall include allowances for cumulative material
damage or material loss resulting from repeated exposure to the design environment (wear, fretting,
ect.). The design thickness for mechanically or chemically milled pressure vessels shall be the
minimum thickness (i.e., mean thickness minus the lower tolerance). The above design thickness
discussion was taken from Ref. 2.

For recommended practices, refer to NASA SP-8007, SP-8019, SP-8032, SP-8068, and NACA TN-
3781 through TN-3786 for closed form solutions to plate and shell collapse loads.


2) OUTLINE TO FEM APPROACH FOR POSTBUCKLING ANALYSES

The corresponding Finite Element Method (FEM) numerical approach for evaluating buckling is to:
1) Develop a finite element model of the structure
2) Ensure that the finite element mesh size is fine enough to capture the critical buckling modes.
Consider the loading conditions, boundary conditions and shell geometry when selecting
element size. Guidelines are provided in Section 3 for to determine reasonable starting element
sizes.
3) Utilize design thickness defined in section 1.
4) Calculate the elastic buckling eigenvalue and eigenvector (buckling mode shape)
5) Verify the mesh size is adequate by running an additional eigenvalue analysis on the refined
mesh. The eigenvalue should not change by more than 5% and the critical eigenvector should
not change either.
6) Select eigenvectors that are known to be imperfection sensitive, like high circumferential wave
numbers for axially loaded cylinders in addition to the lowest values
7) Determine the initial imperfection of the structure due to manufacturing tolerances on surface
waviness allowable dents etc.
8) Scale the maximum displacement of the eigenvector to the initial imperfection size and add it to
the nominal geometry of the structure. This stabilizes the numerical solution and allows one to
investigate multiple modes for imperfection sensitivity. The lowest linear eigenvalue may not
be the most critical mode in a post-buckling analysis.
9) Analyze the structure with imperfections for large deflection
10) Perform limit load permanent deformation/detrimental deformation checks
11) Perform collapse load and detrimental deformation checks
12) Collapse loads are generally determined by the loss of static stability, which numerically
manifests itself as an inability to converge on a solution during a large deflection analysis.
However, loss of static stability (solution divergence) may infer collapse, but this is not an
absolute indicator. There may be dynamic shifts to shorter wavelength modes (higher buckling
modes in the case of a cylinder under collapse pressure). Here, the shift in wavelength is
dynamic but the structure may well become statically stable again once the shift has occurred.
13) Eigenvalue extractions during the nonlinear large deflection analysis may be useful to verify
that the loss of convergence is due to a buckling. These extractions can increase the accuracy
of the eigenvalue, but it is still a linear solution based on the current stiffness of the part during
GOODRICH PROPRIETARY
the solution process. It may over predict the buckling load for imperfection sensitive structures
and cannot anticipate softening due to plasticity. It is best to run the analysis until it fails to
converge. The resulting eigenvectors (mode shapes) may be useful in providing better
imperfection shapes than the simple linear eigenvector analysis thus providing a more stable
solution.
14) If performing a creep buckling analysis, refer to reference 10 for loading simplifications and
modeling guidelines.

A design value for the initial imperfection will be used in postbuckling analyses. This design
imperfection shall consider acceptable handling damage, the surface profile tolerance, surface
waviness and the low end bound of half the shell thickness in the case of axially loaded cylinders.
The surface waviness, d/L, is specified on the component drawing where d is the half wave height
and L is the half wave length. This half wave-length corresponds to the distance between buckling
mode nodes i.e. zero deflection points. The design imperfection height, , shall be equal to the
minimum value determined from the following three cases:
1) The maximum profile tolerance of the surface
2) The maximum surface waviness, d/L, times the half wavelength (in the direction of the
applied load) of the buckling mode being analyzed
3) The formula: = 0.00164 r + 0.051 t , where r is the radius of the shell and t is the
thickness of the shell
This imperfection height may be used to set the allowable dent depth from handling damage. If
higher damage levels must be accommodated, then this larger dent depth will be used for the design
imperfection height, .

The resulting design imperfection height shall be used to scale the buckling mode under
investigation to be added to the nominal geometry of the structure. If case1 or 2 above is used to
determine the design imperfection, the structure will need to be inspected to show compliance to
the drawing profile and waviness requirements.

A post-buckling analysis example for a frame stiffened composite fan cowl is provided in reference
14. Both static and dynamic post-buckling analyses were used to evaluate the structures collapse
load. A general overview document along with NASTRAN non-linear models are attached to this
analysis record.
GOODRICH PROPRIETARY
3) MESH REFINEMENT GUIDELINES

Finite element solutions are a function of mesh refinement. In general, the finer the mesh provides
a better answer than the courser mesh. As with any numerical method, accuracy comes at price
(more storage, model complexity, model turn around times, ect). A common way to assess the
accuracy of a model is to build an initial model and then refine the entire model and compare the
results for convergence. This method may take many models to come up with a good mesh.

This section provides some mesh refinement guidelines based on some typical types of structures
and loading conditions one will encounter. Flat plate and cylindrical shell are used as an example
of how mesh refinement can affect linear eigenvalue analysis results is provided in the section 4.
Additional considerations, including knockdown factors to be applied to the FEM eigenvalues, are
provided in section 5.

Mesh size guidelines for eigenvalue and postbuckling analyses are as follows:
1. For an axial loaded cylinder in compression, the maximum finite element edge length,
L
e
, is equal to rt L
e
35 . 0 = in both directions. This wave estimate is plays out well in
the axial direction. The lowest eigenvalue modes may have relatively small wave
numbers in the hoop direction, which may be adequate for eigenvalue analysis in some
cases. But in post-buckling analysis, the higher wave number modes are more sensitive
to imperfections and thus more critical than the lower hoop wave number modes that
may have a slightly smaller eigenvalue. Thus, it is safest to use the same length in both
directions.
2. For flat panels, the maximum finite element length is equal to one twelfth of the critical
buckling modes full wavelength. As a starting point, a square plate loaded on one edge
with fixed rotation edges would have a full wave buckle across the panel. As a result,
the model would require a 10x10-element mesh density to capture the lowest buckling
mode. For the simple support case, the lowest buckling mode has a half-wave length
requiring a 5x5-element mesh density. To accurately capture the shear buckling-mode
requires a 12x12 mesh, which drives the mesh refinement requirement.
3. For a cylinder under external uniform pressure with supported ends may have a different
element length in the axial direction than around its circumference. The maximum
element size along the axis of the cylinder should be one tenth of the axial length of the
cylinder. For a cylinder with a radius, r, length, L, and thickness, t, the number of
waves around the circumference, n, is approximately:
4
4
58 16 Z (r/L) n + = ,
where Z is the Batdorf parameter
2
2
1 v
rt
L
Z = . The maximum circumferential
element edge length, Le, is:
4
4
) / ( 58 16 / 2 . 0 L r Z r Le + = .

4. Cylinders, of honeycomb panel construction, may be evaluated by utilizing the Batdorf
parameter in its more general form:
D
C
r
L
Z
12
2
= , where C is the extensional stiffness
per unit width (C=Et for a skin) and D is the bending stiffness per unit width. A
GOODRICH PROPRIETARY
honeycomb panel with skins of equal thickness t has an extensional stiffness C=2Et and
a bending stiffness D=Eh
2
t / [(2(1-v
2
)] , where h is the distance between the centroid of
the inner and outer skins. Thus, the Batdorf parameter for cylinders of honeycomb
construction becomes:
h
v
r
L
Z

=
3
1
2 2
.

5. For a cylinder in torsion, the starting element edge length is the minimum of: the
cylinder radius divided by four, or the length of the shell divided by twelve.
6. Edge affects can also contribute significant reductions in cylindrical shell collapse by as
much as 20% due to just the Poissons effect in the gauge area transitioning to restrained
ends. The required starting finite element edge length, L
e
, is equal to rt L
e
75 . 0 =
refinement in the axial direction of the shell, with a minimum of two elements in from
the edge. Smaller edge lengths are required to capture stress effects in post-buckling
analysis. In the direction of the circumference, refinement will be governed by the
allowable aspect ratio of the elements of less than 10 to 1.
7. In general, if higher wave number modes are critical due to the shape of the plate and
the loading conditions, the mesh should be refined until the critical mode has at least
twelve elements along each critical buckling modes full wavelength.
8. When using beam elements one needs to be sure that the implementation of the element
is capable of capturing torsional and lateral instability of the beam under bending as
well as axial loading. Note that FE beam elements do not account for the crippling of
thin sections in a column, which will reduce the overall column buckling load capacity.
9. Membrane/axial load carrying structures that have a discontinuity in the membrane load
path need to consider the resulting offsets effect on the stiffness and bending loads that
occur in at the discontinuity. Examples of this kind of discontinuity are a single lap and
a bent up flange type joint.
10. When post processing the results, it is often handy to fringe plot the rotations on the
mode shape. This can help to determine the level of fixity a stiffener adds along the
plates edge.

These above analysis guidelines should provide a reasonable initial mesh size for buckling and
postbuckling analysis. The wavelength of the critical mode shape should be checked for
recommended number of elements along its length and the eigenvalue checked for convergence on
a finer mesh.


GOODRICH PROPRIETARY
4) EXAMPLES OF MESH REFINEMENT EFFECTS

To start with, it suffices to say that finite element solutions are a function of mesh refinement. In
general, the finer the mesh provides a better answer than the courser mesh, but at a cost. A mesh
refinement study was performed using a flat plate and a short cylinder as examples to illustrate the
issue of mesh convergence.

4.1 Flat Panel Mesh Size Sensitivity Results

The first example is an eigenvalue analysis of square flat plate, with a 10.00 edge length, a 0.100
thickness, a 10.0 msi elastic modulus and a Poissons ratio of 0.30. The analysis was carried out
with NASTRAN version 2001.0.7, using the QUAD4 shell element. FEM results for different level
of mesh refinement are compared to theoretical hand calculation values from the SMART tools
program. Flat plates have a low sensitivity to imperfections. As a result, the theoretical eigen-
value buckling strength correlates well enough with test data to be used directly as allowable
values. The raw results for buckling load in lbs/in are presented in the following table:

Square Plate MxM Mesh Refinement Study From
NASTRAN-2001.0.7 QUAD4 Elements
Critical Nxx Critical Nxy Elements Per
Edge
Simple Clamped Simple Clamped
Theory 361.5 921.0 838.0 1,346.2
3 381.9 1,733.7 2,175.8 16,719.0
5 365.9 1,042.7 1,096.4 2,228.2
8 360.4 933.6 911.1 1,525.3
12 359.0 912.6 861.9 1,375.7
100 358.6 908.5 836.5 1,318.4

The percent difference between the FEM and hand calculations are presented in the next table:

Square Plate Mesh Refinement Study % Error From
NASTRAN-2001.0.7 QUAD4 Elements VS. Theory
Axial Nxx Shear Nxy Elements
Per Edge
Simple Clamped Simple Clamped
3 5.6 88.2 159.6 1142.0
5 1.2 13.2 30.8 65.5
8 -0.3 1.4 8.7 13.3
12 -0.7 -0.9 2.8 2.2
100 -0.8 -1.4 -0.2 -2.1

The driving load case for mesh refinement is the shear buckling case, which requires a mesh
refinement of 10 elements per edge to yield results that are within 10% of the actual critical load
level. The preferred mesh size is 12 elements per edge.

GOODRICH PROPRIETARY
4.2 Cylindrical Shell Mesh Size Sensitivity Results Part 1

A second example illustrates the effect of mesh refinement on a short circular cylinder typical of a
skin bay between stiffeners. Cylinder parameters used in this example are a radius of 50.0, 10.0
axial length, 0.063 thickness, a 10.0 msi elastic modulus and a Poissons ratio of 0.30. Two of the
load conditions (hoop and shear) are considered along with two end boundary conditions (simple
and clamped).

The FE displacements in the radial, hoop and axial directions are Ur, Ut, and Uz respectively. The
FE rotations in the radial, hoop and axial directions arer, t and z, respectively. Simple support
displacement boundary conditions at one end of the cylinder were Ur=Uz=z=0 with two point
pinned in the Ut=0, while displacements at the other end were Ur=z=0. Clamped edge conditions
added no rotation about the hoop direction, i.e. t=0 to the simple support conditions. The hoop
loads, Ntt were applied to the model by a lateral pressure, while the shear load Nzt was applied to
both ends as forces in the hoop direction producing a torsion load. The simple end support
boundary condition simulates the effect of frames with open sections, while the fixed end condition
is similar to that of a closed section frame.

The analysis was carried out with NASTRAN version 2001.0.7, using the QUAD4 shell element.
FEM results for different levels of mesh refinement are compared to theoretical hand calculation
values from the SMART tools program as well as the allowable values. The allowable values are
slightly lower than the theoretical values due to a slight imperfection sensitivity of cylindrical shells
to this loading condition and were derived empirically. The FEM eigenvalue analysis does not
account for this reduction, although a post-buckling analysis has this capability. The raw results for
buckling load in lbs/in are presented in the following table:

Cylinder Mesh Refinement Study Using NASTRAN-2001.0.7 QUAD4
Elements With Theoretical Values
Critical Ntt Critical Nzt
Hoop
Elements
Axial
Elements Simple Clamped Simple Clamped
Theory 166.2 204.0 275.8 331.2
Allowable 132.0 185.0 205.3 266.0
72 5 363.3 452.0 815.1 1002.2
72 12 332.9 433.2 704.4 873.2
144 5 190.7 241.7 405.8 520.2
144 12 183.8 236.5 349.2 420.5
288 5 167.6 213.6 336.4 433.8
288 12 164.4 212.4 291.0 348.7

For simple support condition, the Ntt hoop load has 29 circumferential waves and a axial wave,
while the Ntz shear load has 27 circumferential waves and 1 axial wave. For clamped edge support
condition, the Ntt hoop load has 32 circumferential waves and 1 axial wave, while the Ntz shear
load has 30 circumferential waves and 1 axial wave. The approximation from Section 3 estimates
the hoop load to have 32 waves, which is in good agreement with the FEM results. Percent error
between the NASTRAN and theoretical results are shown in the following table:

GOODRICH PROPRIETARY
Cylinder Mesh Refinement % Error Using NASTRAN-2001.0.7 QUAD4
Elements VS. Theory (Add 25% for error from Allowable)
Ntt % Error Nzt % Error
Hoop
Elements
Axial
Elements Simple Clamped Simple Clamped
72 5 118.6 121.6 195.6 202.6
72 12 100.3 112.4 155.4 163.6
144 5 14.8 18.5 47.2 57.0
144 12 10.6 15.9 26.6 27.0
288 5 0.8 4.7 22.0 31.0
288 12 -1.1 4.1 5.5 5.3

There is a general rule of thumb that at least 5 elements are required per half wavelength to provide
accurate results. When this rule is applied to the simple supported shear load case, 300 elements
are required in the circumferential direction along with 10 in the axial direction results in an over
prediction of less than 8%. If one were to have judged a wavelength in this case the error would
increase to 20%.

Note that allowable values are determined by empirical methods and applied to the FEM
eigenvalue analysis results as a knockdown factor.

4.3 Cylindrical Shell Mesh Size Sensitivity Results Part 2

Axial loading of cylinders has been a difficult problem to solve. The results are highly influenced
by shell imperfections and end boundary conditions. This section shows the effects of end
conditions on the buckling eigenvalues. Data from three end displacement boundary conditions
(BC) are investigated. Those from Section 4.2 are the first to be used. The second BC adds Ut=0
to only the first end, while the third condition adds Ut=0 to both ends. The axial forces are applied
to the ends of the cylinder to create axial running loads of 1 lb/in (Nzz=1) for the eigenvalue
analysis. The raw results are provided in the following table:

Axially Loaded Cylinder Mesh Refinement Study Using NASTRAN-2001.0.7 QUAD4 Elements
With Slightly Different End Conditions Results In Large Differences
Edge 1 Ur=Uz=0
Edge 2 Ur=0, Nz=1
Edge 1 Ur=Uz=Ut=0
Edge 2 Ur=0, Nz=1
Edge 1 Ur=Uz=Ut=0
Edge 2 Ur=Ut=0, Nz=1
Hoop
Elements
Axial
Elements Simple Clamped Simple Clamped Simple Clamped
Theory 480.4 480.6 480.4 480.6 480.4 480.6
Allowable 122.0 130.3 122.0 130.3 122.0 130.3
72 5 300.8 634.0
72 12 250.8 541.9
144 5 300.7 580.3 482.7 640.2
144 12 250.7 539.1 483.5 573.7
288 5 300.9 524.5 381.6 559.6 482.6 588.6
288 12 250.6 494.6 340.7 483.4 539.9
288 24 244.3 493.7

GOODRICH PROPRIETARY
Again, the theoretical and allowable values are provided for comparison purposes. Note that the
simple edge support results can vary by a factor of two, just by changing the circumferential
displacement boundary condition. These changes in eigenvalue do not reflect imperfection effects,
which further reduce the load carrying capacity of the shell.

Allowable values are significantly lower than the theoretical values due to the high imperfection
sensitivity of cylindrical shells to axial loading. These allowable values were derived empirically
as a knockdown factor applied to the theoretical value. The FEM eigenvalue analysis does not
account for this additional 75% reduction, although a post-buckling analysis has this capability.


GOODRICH PROPRIETARY
5) ANALYSIS FACTORS, ISSUES AND COMPLICATIONS

5.1 Finite Element Incompatibility Leads To Phantom Modes

Finite element compatibility can become a significant issue when mixing different element types.
These incompatibility issues show up routinely when modeling honeycomb panels with solid and
shell elements, as well as, modeling skin-stiffened structures with shell and beam elements.
Element compatibility is when elements have matching displacement fields between nodes. This is
a fundamental requirement for element selection when using the finite element method.

When honeycomb panels are modeled with solid elements for core and shell elements for the thin
skins, it must be recognized that the shell rotations are not compatible with the solid elements.
Shell element rotations implicitly result in a non-linear displacement field between nodes resulting
in an incompatibility with the assumed displacement field of the solid element. This can lead to
phantom FEM buckling modes because the thin shell elements are essentially pinned at each node
and large in size. This is essentially equivalent to intra-cell skin buckling (dimpling) where the
core cell size is the finite element edge length. Intra-cell buckling occurs with cell (element edge
length) sizes on the order of less than a half of an inch. Buckling modes of this type are principally
dominated by rotations with little or no displacement, resulting in a random looking displacement
mode shape in PATRAN. Please note that these (large) rotations are not read into PATRAN with
the default setting. In general, the skin should be modeled as a membrane element to maintain
compatibility with the solid elements. This is achieved in NASTRAN by only entering a material in
the membrane stiffness location in the PSHELL card. The use of PCOMPS in NASTRAN gives a
shell element with a bending stiffness making it incompatible with solid elements.

In honeycomb panel to land transition areas, where the solids elements transition to lands modeled
with shell elements, there is a need to couple the rotations of the solid element edge to the
rotational degree of freedom used by the shell element. This can be achieved with RBE elements to
couple the solid element edge rotation to the rotational DOF. Care must be taken to make sure that
the nodal coordinate systems are consistent and to exclude the core though thickness displacement
in the RBE element. The resulting interface moments and shear loads can be used with a beam on
an elastic foundation hand calculation to determine the skin/doubler to core peel stresses. Small
element sizes are required to account for the coupling between the core flat-wise stiffness and skin
bending stiffness. In order to get the correct coupling, the shell elements must be smaller than the
decay length for a beam on an elastic foundation. This decay length may be calculated from the
following formula:

4
2
3
) 1 ( 3
d c
d d
d
E
ht E
L

=

Where E
d
is the skin modulus, E
c
is the core flat-wise compression modulus, h is the core height
and t
d
is the combined thickness of the skin and doubler. For a 3 lb/cu-ft aluminum core, an
aluminum skin and douber combined thickness of 0.06 and a 1.0 core height, the required
GOODRICH PROPRIETARY
element size would need to be around 0.26. This is quite small compared to typical element sizes
used to model nacelle honeycomb structures. Another reason this should not be done is previous
discussion on shell buckling over solid elements.

Complications can also arise when modeling skin-stiffener type structures. These structures are
often modeled with shell and beam elements that have incompatible rotational stiffness. The shell
element only has only two rotational stiffness while the beam element has all three. This can
result in a similar buckling issue as found when mixing shell and solid elements. The shell normal
rotational stiffness is zero. Any rotation about the shell normal is modeled by the in-plane linear
displacement field along the element edge and not by this nodal rotation. Because beam elements
are cylindrical bending elements that have a curved displacement field between nodes (defined by
the nodal rotations), its displacement field does not match the linear shell element edge
displacements. This is another example of an incompatible set of elements.

5.2 FEM Analysis Factors and Complications

Complications can arise in post-buckling analyses aimed at determining the collapse load of a
structure. These complications are best described with specific examples.

The first example is of a two-part L frame structure consisting of an end-loaded column that is
supported on the top by a rigid connection to the end of cantilevered beam and a pinned support on
the bottom. If one performs a post-buckling analysis by adding the initial imperfection as a scaled
buckled mode shape, the response of the structure will be different from that if one adds or
subtracts the imperfection shape onto the perfect structure. In one case the postbuckling behavior
will be stable and not sensitive to initial imperfections. In the other case the postbuckling behavior
will be unstable and sensitive to the size of the initial imperfection. This example shows that how
one induces the initial imperfection shape on the perfect structure can result in substantially
different failure loads.

The second example is that of an end loaded cylinder. There are many closely spaced modes that
have very different post-buckling behavior. One set of these modes is completely stable while
another mode with a similar eigenvalue but different mode shape is unstable and with its collapse
load being strongly influenced by the size of the initial imperfection. The unstable buckling mode
has its collapse load reduce to 42% of the perfect structure collapse load with an initial
imperfection equal to one half of the thickness of the shell. Thus it may be necessary to check
more than one buckling mode to determine the collapse load of a structure Ref. 4. Koiters
analyses shows that an initial imperfection equal to the thickness of the shell reduces the collapse
load by a factor of five to 20% of that of a perfect structure Ref. 5. The mesh refinement
requirements are also rather substantial because the critical wavelength (due to imperfection
sensitivity not just the lowest eigenvalue) is generally short in both the axial and hoop directions.
The half-wavelength equation rt L
w
72 . 1 = was obtained from Reference 9, where r is the radius
of the cylinder and t is its thickness. Because the critical mode shape tends to be a diamond pattern,
this wave-length (in hoop and axial directions) is fairly small compared to the size of shell. In
addition, the finite element size required to accurately resolve this small wavelength requires five
elements along the length of the half wavelength for most element types, Ref. 4. This makes the
GOODRICH PROPRIETARY
maximum finite element edge length, L
e
, equal to rt L
e
344 . 0 = . End and edge affects have also
been cited as contributing to lowering the collapse load of cylindrical shell under axial compression
by as much as 20% Ref. 6.

Due to the grossly unconservative results given by the theoretical buckling load prediction for
axially loaded cylinders, a correction factor was developed in Reference 12. The design-buckling
load is calculated by multiplying the theoretical buckling load by the correction factor. This
approach can be applied directly to FEM results for axially loaded cylinders. The equation for the
correction factor, , is:

) 1 ( 901 . 0 1
/ ) 16 / 1 ( t r
e

= ,

where: r is the radius of the shell and t is the thickness of shell. The equation is applicable for
r/t<1500.

Additional reduction factors used to correlate theoretical to the test data were obtained from
Reference 13. They are: 0.75 for cones/cylinders under external pressure; 0.69 for cones/cylinders
under torsion, 0.56 for shallow toroidal segments under external pressure, and 0.35 for bowed out
toroidal segments under axial tension (curved inlet forward bulkhead with radial tension loads).
As can be seen from the latter factors, the geometry of the shell greatly influences the knockdown
factor. Thus illustrating the need to use appropriate knockdown factors when using the results from
a linear buckling analysis weather it was derived from an FE analysis or a hand analysis.

For short cylinders/cones (Z<40), reductions can be less than stated in the latter paragraph.
Reference 10 provides the theoretical buckling coefficients as a function of Z, where is the
correction factor that correlates the theoretical to the observed buckling loads and Z is the Batdorf
parameter. The FE buckling correction factor is calculated in the following three steps. First
calculate the buckling coefficient, k as a function of Z, using the recommended correction
coefficient . Second, calculate the theoretical buckling coefficient by assuming is equal to one.
The FE buckling correction factor is equal to the corrected value of the buckling coefficient, k(Z),
divided by the theoretical buckling coefficient, k(Z). As an example, Reference 12 shows a graph
of the buckling coefficient for a cylinder under lateral pressure with simply supported edges that
can be approximated by the following equation:

Z k + = 08 . 1 98 . 3
2


where =0.5625 is the theory to allowable correction factor. For a cylinder with Z=20 and
=0.5625 then k=5.291 while the theoretical value corresponding to =1.0 yields k=6.119. The
resulting reduction factor is 0.86 verses the 0.75 value for large Z values.

GOODRICH PROPRIETARY
5.3 Post-Buckling Initial Imperfections

FEM post-buckling analyses require initial imperfections for accurate collapse load estimates,
improving the stability of solution and individual mode sensitivity to imperfections. This is
especially true for axially loaded cylinders where the theoretical values can be many times higher
than observed collapse values. References 4 and 5 relate the imperfection to thickness ratio to a
buckling correction factor. Reference 12 relates the design correction factor to the radius to
thickness ratio. This allows one to obtain the imperfection to thickness ratio as a function of radius
to thickness ratio. When the results are plotted, there is a linear relation between the two ratios.
The relationship derived with the data from Reference 4 results in the more conservative of the two
equations. The resulting equation for the imperfection height, , is:
= 0.00164 r + 0.051 t
where r is the radius of the shell and t is the thickness of the shell.

Due to the set up and computation time required for postbuckling analyses, there is a need to reduce
the amount of runs required to analyze structures that have multiple buckling modes with closely
spaced eigenvalues. One such approach given in Reference 7 has been to apply up to six
imperfection distributions at a time to a single structural analysis. This approach can be viewed
loosely as a basis for an imperfection space similar to a basis of a vector space. The total
imperfection height was set equal to the thickness of the buckled component. Other imperfection
heights and modes were studied in addition to the imperfection space approach in References 7
and 8.

Careful thought is required when performing post-buckling analyses with a composite buildup of
different structural elements such as frames, plates, cylindrical shell and general shell structures.
This is definitely the case when plates and axially loaded cylinders are combined. This is because
the collapse load of an axially loaded cylinder is on the order of four times lower than the linear
eigenvalue predicts, while the plate collapse load can be higher than the eigenvalue predicts.
Hence, the selection of imperfections in the cylindrical shell becomes problematic due to the shear
number of eigenvectors that need to be looked at to find those that would influence the early
collapse of the axially loaded cylinder. The buckled mode shape indicates the critical locations for
imperfections to start the particular mode to buckle. Thus, a dent could well start the critical
collapse mode provided one has guessed the correct spot to put it. This is why imperfections are
applied in the shape of the buckling mode being investigated for postbuckling performance. When
using beam elements mixed with shell elements one needs to be sure that the implementation of the
element is capable of capturing torsional and lateral instability of the beam under bending as well
as axial loading Ref. 11.

It is always strongly recommended to perform hand calculations to estimate structural performance
prior to a FEM eigenvalue or post-buckling analysis. This hand analysis provides an essential
check to the complicated numerical analysis used when employing the FEM to post-buckling
analyses.
GOODRICH PROPRIETARY
6) NASTRAN FEM ANALYSIS LIMITATIONS

Caution must be used when utilizing NASTRAN for post-buckling analyses of structures. One
main caution is that while NASTRAN has a large deflection capability it does not have a large
rotation capability. This can induce significant errors if large rotations occur in the structure prior
to collapse. The MARC code offers both large deflection and large rotation capability.
NASTRAN cbeam elements do not support beam offsets for large displacement or buckling
analysis, nor does it support shell element offsets.

There are additional constraints when using NASTRAN for post-buckling analyses. NASTRAN
does not correctly calculate thermal loads during thermal transients when either plasticity or creep
models are activated. The code will calculate significant thermal stresses even when the elastic
modulus of the material goes to zero. The plasticity algorithm does not allow for temperature
dependence of the material yield or work hardening of the material. NASTRANs creep algorithm
does allow for temperature dependence with a limited choice of material models to choose from,
but does not calculate the correct transient thermal stresses. NASTRAN does not allow different
temperature dependencies for primary and secondary creep. In addition, NASTRAN has no
provision for performing plasticity or creep analysis on honeycomb shell structures.

The MARC code correctly models the thermal load increments. It also allows for: primary and
steady creep rate temperature dependence, temperature dependant plasticity, has an automatic time-
step algorithm, allows plasticity and creep analysis with honeycomb shell elements, and provides a
critical time to creep buckle based on an eigenvalue analysis utilizing the creep tangent stiffness
matrix. Additional capabilities in thermo-plastic, contact, large displacement, large strain, thermal,
and coupled thermal/structural analysis are also available.

When composite materials are used, NASTRAN does not support ply-by-ply failure analysis.
MSC/MARC has this capability to automatically update individual ply stiffness as each fiber or
matrix failure occurs during the post-buckling simulation.

If any of the above limitations in the NASTRAN code limit the applicability or the accuracy of the
postbuckling analysis, it is recommended to use the MARC finite element code for postbuckling
analyses.

Post-buckling analyses on semi-stable structures can require substantial amounts of computation
time even for small to moderate finite element models. The numerical model needs to be as small
as possible yet be able to capture the critical buckling modes of the structure.

An approach to creep buckling analysis is presented in Ref. 10.


GOODRICH PROPRIETARY
7) REFERENCES

1) NASA SP-8057, "Structural Design Criteria Applicable to a Space Shuttle," March 1972, in
section 4.3.3 on page 4-3.

2) NASA SP-8057, "Structural Design Criteria Applicable to a Space Shuttle," March 1972, in
section 4.9.1 on page 4-15.

3) Don Brush and Bo Almroth, Buckling of Bars, Plates and Shells, McGraw-Hill, Inc 1975.
Page 251.

4) T. Hughes and T. Belytschko, "Recent Advances in Nonlinear Finite Element Analyses," July
24 1989, Palo Alto CA. Volume II, Shell-I, Figures 6.9-6.16

5) Don Brush and Bo Almroth, Buckling of Bars, Plates and Shells, McGraw-Hill, Inc 1975.
Page 234.

6) Don Brush and Bo Almroth, Buckling of Bars, Plates and Shells, McGraw-Hill, Inc 1975.
Page 331.

7) M Nemeth et al, "Nonlinear Analysis of the space shuttle Superlightweigh LO2 Tank: Part I
Behavior Under Booster Ascent Loads," AIAA Technical Paper AIAA-98-1838.

8) M Nemeth et al, "Nonlinear Analysis of the space shuttle Superlightweigh External Fuel Tank,"
NASA Technical Paper 3616, December 1996.

9) Raymond Roark, Formulas for Stress and Strain Third Edition, McGraw-Hill Book
Company, Inc., New York-Toronto-London, 1954, page 316.

10) R. Alloway, Creep Buckling Analysis Methods, BFGoodrich Report RHR 96-021, March
1996.

11) Don Brush and Bo Almroth, Buckling of Bars, Plates and Shells, McGraw-Hill, Inc 1975.
Pages 37-60.

12) NASA SP-8007, "Buckling of Thin Walled Circular Cylinders," August 1968, in section 4.2.1
on page 5.

13) NASA Technical Memorandum TM X-73305, "Astronautic Structures Manual," August 1975,
NASA Marshall Space Flight Center.

14) R. Alloway, CF6-80C2 RMS-060 Hollow Hat Fan Cowl FE Post Buckling Analysis,
Goodrich Analysis Record AR0100157, November 11, 2002.

You might also like