Meek 1989

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 19

COMPUTER METHODS IN APPLIED MECHANICS AND ENGINEERING 72 (1989) 57-75

NORTH-HOLLAND

LARGE DISPLACEMENT ANALYSIS OF SPACE-FRAME STRUCTURES

J.L. MEEK and S. LOGANATHAN


Department of Civil Engineering, University of Queensland, St. Lucia, Queensland 4067, Australia

Received 21 September 1987


Revised manuscript received 3 May 1988

The current research is mainly concerned with the geometrically nonlinear static analysis of
three-dimensional space-frame structures. The elastic analysis of frame structures by means of the finite
element method in the post-buckling range inevitably involves the solution of large systems of
nonlinear equations. The most satisfactory way of solving such problems is to combine the arc-length
method (Riks and Crisfield) within each increment with the Newton-Raphson method (NR method)
as the iteration strategy. For large joint rotations, the joint orientation matrix suggested by Oran has
been used to update the rotational displacement of a joint. The present study deals with the 'imperfect'
approach to trace the secondary paths of three-dimensional frame structures, the particular examples
studied being a two-hinged deep arch and a shallow geodesic dome. Eigenvectors are calculated at
bifurcation points to force the structure on to the secondary path by introducing a small perturbation
either in load or in geometry,

1. Introduction

The elastic stability analysis of space-frame structures, involving the solution of large
systems of nonlinear equations, has been extensively studied by various authors and is based
on either the finite element method [1-4] or on the beam-column approach [5-10]. In the
finite element formulation, the deformation of a given structure is described by sets of (N)
deformation parameters which are known as generalized coordinates. The load deformation
history of a structure represents a curve in an (N + 1)-dimensional space spanned by the
deformation parameters and the externally applied load. Such a curve is usually referred to as
an equilibrium path or a deformation path.
In describing the motion of the element, a total Lagrangian or an Eulerian description can
be employed; the former in terms of the initial position; the latter in terms of the final
deformed state. Computationally, an Eulerian formulation is strictly an updated Lagrangian
approach [11]. The total Lagrangian approach offers advantages since the initial configuration
remains constant. This simplifies the formulation and computational algorithm. However, this
approach is limited by the magnitude of joint rotations. To overcome this problem of large
rotations, an updated Lagrangian approach [12-14] can be employed, which separates the
effect of pure member deformations from the joint displacements.
For large joint rotations, the rotational displacements of a joint are accumulated using a
joint orientation matrix as presented by Oran [10]. The incremental joint orientation matrix is
calculated in each load increment and the total joint orientation matrix updated for the next

0045-7825/89/$3.50 © 1989, Elsevier Science Publishers B.V. (North-Holland)


5~ J.L. Meek, S. Loganathan, Large displacements of space-frame structures

increment. Due to the variation of the principal directions from one cross-section to another,
end section orientation matrices are required to obtain an average value of the orientation
matrix of the member cross-section axes.

2. Finite element formulations

In the present study, the finite element method is employed to derive the nonlinear
equations governing the behaviour of spatial frames. The derivation is based on the assump-
tions that:
- t h e material is homogeneous, isotropic, and linearly elastic;
- the cross-sections are doubly symmetric and constant throughout the length, thus excluding
coupling of the torsional stiffness to that of the bending and axial stiffness;
- t h e external loads are conservative and applied at the joints only;
-transverse shear deformations and warping effects are neglected; and
-lateral deflection curves are cubic and axial displacement is linear.
There are two types of coordinate systems which are employed in the derivation, namely a
fixed global set of coordinates and a local convective system which rotates and translates with
the member. As such, the joint displacements are separated from member deformations. The
axial shortening due to the 'bowing' of the member is included in the analysis.
The secant stiffness relations between the moments and rotations of the three-dimensional
frame element (Fig. 1) may be obtained by the first variation of the total potential energy, and
are expresed as

- 7 - + 36J °. + L[ 2et'
M~z = r[ 4EIz 4Pl] Pl
(:)
7 N]~,'

-=t t Y6 o,.+ -T-+-~ffj~,, (2)


[ --7--
4Ely +
4PI] + [t 2Elf
-gg-]o,, Pl
M~y (3)
7 N]°J, '

= 7 3"0 0/y+L I +-~ff'JOjy, (4)


and the torsional moment and axial force are given by
GJ
Mx=lO ~ , ,,
t51

(e )
P = EA -[ - 8 b , (6)

8 b -~60 (202 - O~,Oj,+ 20/,2 +20~y_ Oeeiy + 20~) (7)


where 5b is the axial lengthening due to bowing, EA is the axial rigidity; GJ is the torsional
rigidity, El), and E1 z are the flexural rigidities about the y-axis and z-axis, respectively, and I is
the initial member length.
I,L. Meek, S. Loganathan, Large displacements of space-frame structures 59

Yl
M~, P M,=
. ~ .¢"0,z ~ljz .. P M,

H= P Mi~ ~ 1 . . ~ H~I~.. P Mx
~ ~ ( ~ . . . . ~_~ )) =X
k® ---t:e- ~ 7 ~
,1 (a)

,1. A ~

'~ ~ k'dIJE~j,A~:lj l"6j,a~6j

_..,~r,
~,"u4i ~ "F6i,Au6~
"
(b)
Fss'Ausj

F . ~, F2'j,Au2j TF~J,Bul~ FI,,,AU4~


.~" 4M

y )) ;.~
~(J..) ~" F6i,&u6.
"o," J

Z
(c)
Fig. 1. Member deformations and associated forces for a three-dimensional frame element. (a) Member basic
displacements and forces. (b) Member intermediate displacements and forces. (c) Member nodal forces and
displacements in global coordinates.

Partial differentiation of (1)-(6) with respect to each of the displacement variables yields
the incremental relationship

{AS} = [td{AV}, (8)


where
(AS} = (AM,=, AMiz,AM,,,AMjy,AM,,AP)',
{av} = <a0,z, a0jz, aO,~, a0~, a0x, Ae)',
and [k] is the incremental stiffness matrix based on basic member coordinate system.
60 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

By considering geometry and equilibrium,

(F) =[AI{S}, (9)

where {F} are the nodal forces in local coordinates, {S} are the basic member forces, and [A]
is the transformation matrix from member basic force-displacements to member nodal
force-displacements in local coordinates.
Using the contragredient principle,

{AV} = [ A ] ' { A u } , (10)

where {Au} are the incremental nodal displacements in the local coordinate system.
The incremental form of (9) is given by

( A F } = [ A ] ( A S } + [AA]{S}, (11)
where AA is the change in [A] resulting from {Au}.
Hence from (8) and (10),

(At} = ([al[kl[a]' + [DI)(Au}, (12)

where matrix [D] is obtained from [AA] {S} by rearranging the nodal displacements {~u}
and the basic member forces {S}.
Let [r] be the direction cosine matrix of the local member axes with respect to the global
axes. Then, the transformation matrix is given by

[R]- [[r], It], [r], [r]j. (13)


Thence,

{:'F~ = [e](A,~), (14)


and by contragredience

(~u} = [e]'(Au). (15)


Substituting (12) and (15) into (14) results in

{AF} = [el([Al[kl[al' + [DI)[R]'{Au} ,


i.e.
(~F) = [k~](au). (16)

The matrices [k], [A], [AA], and [D] are explicitly given in [17]. For complete details of the
derivation, the reader should consult [17, 19].
J.L. Meek, S. Loganathan, Large displacements of space-frame structures 61

3. Oran's joint orientation matrix

The corotational approach has been used by Meek and Tan [17] to study the post-buckling
behaviour of space-frame structures. Initially, the joint orientation matrix [a] is assumed to be
parallel to the global axes,

Ofll Of12 at3]


[a] = / O~21 Cg22 l~g23l " (17)
LO/31 0f32 0f33j
The incremental rotations are still small and the theorem of superposition still applies. Hence,
the three rotations are applied independently as shown in Fig. 2.
Hence, the incremental rotation matrix for a joint is
0 --A~'~3 A~'~2 ]
[Aa]= aa~ o -aa, (18)
--A~'~2 A~'~1 0
where
Aa,/ fAu,/
for joint i
IAu+J

for joint j .
AU12)
Hence, the incremental joint orientation matrix is given by
[Aa] ffi [ A a ] [ a ] . (19)

Thus the updated joint orientation matrix [a] + [Aa] may be evaluated from (17) and (19).
Due to the variation of principal directions from one cross-section to another, end-section
orientation matrices are required to obtain an average value of the orientation matrix of the
member cross-section axes,

[p<,.~] = [,~,>][r(O)]. [p,,] = [o~,,l[,.(m]. (20)

2 2

//~ -&ft2
...........L. I

-~Q'i
(a) (b). (c} +
Fi~. 2. Incremental rotation of a joint. (a) .,x~,, (b) A~'~2, (C) A~'~3.
62 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

where [p(i)] and [p(J)] are the 'end-section orientation matrices' for end-sections (i) and ( j ) ,
respectively, and [r (°)] is the member orientation matrix in the undeformed configuration.
In the case of small relative rotations of the member, the simple coordinate transformation
yields

[e(')]=-0,~ , [e(~)]-0jz 1 0 (21)


Oiy 0 Ojy 0 1

The r.ew end-section orientation matrices will be [p(~)][e °)] and [p(J)][e(J)]. The member y-
and z-axes are then defined as the principal directions of an average cross-section. Hence,

[r] = ½([p(')][e c')] + [p(')][e(J)]) . (22)

4. Program logic

A brief outline of the program logic is given below.

Step 1. Initialization and input parameters such as geometric and material properties,
connectivity, boundary conditions, and reference load.
Step 2. Determination of the skyline profile of the stiffness matrix.
Step 3. Restart option, if necessary.
Step 4. For each load cycle:
4.1. For each iteration:
4.1.1. Computation of tangent stiffness matrix [17] (provision for NR and
mNR methods).
4.1.2. Factorization of stiffness matrix by using modified Cholesky's [15]
method (k r = L D L t).
4.1.3. Check for negative pivots.
4.1.4. Eigenvalue analysis and eigenvectors for imperfect approach [22].
4.1.5. Solve for displacements due to reference load vector.
4.1.6. Determination of load factor at beginning of each iteration (skip 4.1.7
and 4.1.8 for first iteration of each load cycle).
4.1.7. Computation of displacement increment due to residual forces.
4.1.8. Load factor determination for subsequent iteration by using modified
constant arc-length strategy by Riks and Wempner [16]. Selection of
the appropriate root [21] and restart option for complex roots.
4.1.9. Updating joint orientation matrix [10], geometry, translational dis-
placement, and external load vectors.
4.1.10. Computation of internal member forces for each element.
4.1.11. Calculation of the nodal resisting forces of the element in the global
coordinate system and assembly to form the actual resisting forces of
the structure.
4.1.12. Determination of the unbalance or residual force vector.
J.L. Meek, S. Lcganathan, Large displacements of space-frame structures 63

4.1.13. Equilibrium check and test for convergency by considering force norm
[201.
4.1.14. If Step 4.1.13 is not satisfied, go to Step 4.1.
4.1.15. Check for number of iterations. If the allowable number of iterations
is exceeded, go to Step 5.
4.2. Print the results and save relevant data for restart option, if necessary.
4.3. For next load cycle, go to Step 4.
Step 5. Stop.

$. Solution techniques

The most satisfactory method of solving the nonlinear problem is to combine incremental
methods with an iteration technique. The well-known, efficient constant arc-length method
suggested by Ramm [16] and Crisfield [21] has been used within each increment in the present
analysis. All the displacements are controlled in the arc-length method, unlike the displace-
ment control method in [25]. The iteration strategy involved in the analysis is the Newton-
Raphson method (NR method). It is necessary to reform and refactorize the tangent stiffness
matrix in the NR method for each iteration. On the other hand, in the modified Newton-
Raphson method (mNR method) the tangent matrix has to be assembled and factorized only
at the beginning of each increment. There are advantages and disadvantages in each of these
two iteration strategies. The computational cost of each iteration in an increment is less in
mNR method. Unfortunately, the convergency rate is often poor in this method. In some
structural analysis problems, the NR method consumes less (cpu) time than the mNR method.
The iterative process is continued until the convergence, determined by the ratio of Euclidean
norm of the residual forces to that of the external applied load to be less than a specified
amount, is satisfied [20] and a check is kept on the stability of solutions by monitoring the sign
of the determinant of the reduced stiffness matrix. The modified Cholesky's method is used to
factorize the stiffness matrix in the skyline solver proposed by Felippa [15]. The determinant
criterion during factorization could be used to determine the sign of the incremental load
factor. However in some cases, this criterion is insufficient to trace the load-deflection path
and an alternate criterion based on incremental work was found to be superior [23].
Subsequent experience has shown that both the sign of the determinant and the sign of the
incremental work must be checked from one increment to the next for the structural systems
which exhibit snap back characteristics.
There are two possible techniques termed the 'perfect' and the 'imperfect' approach to
follow the secondary path passing through the bifurcation point. Eigenvectors have to be
computed at the bifurcation point for 'imperfect' approach, however, a small perturbation in
load or in geometry may be imposed to trace the secondary path or post-bifurcation path
without requiring to compute the eigenvectors at the bifurcation point. In this approach a
difficulty arises in determining the location of the small perturbation prior to commencement
of the analysis. To overcome this problem, the current program has the option which allows
the user to restart the solution process from the last converged solution point. An algorithm
based on the Jacobi method has beem employed [22] in the eigenvalue analysis to compute the
eigenvectors corresponding to ~,t~e smallest eigenvalue. Then a small force imperfection is
added to the structure to activate the corresponding eigen mode.
64 J, L. Meek, S. Loganataan, Large displacements of space-frame structures

6. Constant arc-length strategy of Crisfield

A simple algorithm presented by Batoz and Dhatt [24, 25] was found by them to be a better
procedure to compute the displacements due to external load increment {P} and residual
forces {R},

{AUe} = [ k l - l { P } , (23)

{AUr}=[kl-l{R} , (24)
where [k] is the tangent stiffness matrix, {Aue} is the incremental displacement due to
external load increment, and {Au r} is the incremental displacement due to residual forces.
The actual external load increment applied is AA{P} and the loading factor, AA, is obtained
by specifying that the displacement increment at the degree of freedom, n, satisfies a
displacement, ~, given by
~= a,~{au~")} + {au'/'}. (25)
From (25),
~ - (Au'/)}
AA = (26)
{Au~~,)

The actual incremental displacement is given by

{Au} = Aa(Au~} + (Au,}. (27)

Equation (26) is applicable to the first iteration of the load cycle. Subsequent iterations are
performed in the same manner with 8 specified to be zero. This iteration strategy described by
Powell and Simons [26] will fail to converge when the controlling displacement snaps back
from one load level to another. To circumvent this problem, Riks [27] and Wempner [28]
developed the following constraint equatior which couples all the displacements and external
incremental loads through fixing the length of the increment;

{au}'{au) + a ~ / ( a e } ' { a p } = a i ~ , (28)

where Al is the incremental arc-length.


The above constraint equation was originally added to the incremental stiffness matrix.
Unfortunately, the inclusion of an additional constraint equation destroys the symmetry and
bandness of the stiffness matrix. Batoz and Dhatt [25] suggested the two-step technique
described above to circumvent this problem. The constraint equation has been modified by
Ramm [16], where both iterations in a plane normal to the tangent and iteration in a sphere
are described and are illustrated in Fig. 3.
The constraint equation was reformulated by Crisfield [21] for the purpose of numerical
analysis and its superiority has been numerically proven by many researchers. The modified
J.L. Meek, S. Loganathan, Large displacements of space-frame structures 65

mthat solution
/ normalto tangent
I / / final sotuho.

~1 I Ily \_

"1/J
°/
DISPLAEEMENT, u

Fig. 3. Constraint surfaces for the arc-length method.

constraint equation is as follows:

{ A u } t { A u } -- Al 2 " (29)

At the start, a unit loading parameter (i.e., AA(n = 1) is chosen as the scalar multiplier of
an arbitrary external load and the incremental displacement is determined by the following
equation:

{Au (n} ~[k] -1AA(n{P}. (30)


Hence, the prescribed generalized arc-length
(31)
In the paper of Meek and Tan [19], it was suggested that the sign of AA(1) must be based on
the sign change of the incremental work done, AW, where

awffi {au"~I'{e}. (32)

If the sign of AW has changed from that of the previous increment, the sign of AA(t) is
reversed from that of its previous value.
At the ith iteration, after the initial increment,

{Au~n} = [kl-'{P} (33)


and
{A u~'~} - [kl-' {R¢'-'~}, (34)

where {R (i-1)} is the residual force at the end of the (i - 1)th iteration.
66 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

The iterative displacement {Au t°} is

{ a u " ' ) = aa "(au' + {aU r°} • (35)

The incremental displacement up to the ith iteration (8u °)} is given by


{aU (/)} = {aU (1-1)} -3L {Au(i)} , (36)

According to (31),
{ auO )} t { au t, )} = Al 2 " (37)

Substituting (35) and (36) into (37) results in

a(AA(')) 2 + b(AA Ci)) + c = 0 , (38)


where
O) t (O
a = {Au e } {au e },

b = 2[{8u °-')} + {Au,(i) }] t {Au eO) },


C ~''~[{aU (i-I)} "Ji"{AU(rl)}]t[{aU(i-I)} + {Au~i))]- Al 2 .

The real roots are possible if and only if

b e - 4ac >~O. (39)

Selection of an appropriate root was suggested by Crisfield [21] to avoid doubling back on
the solution path. This will be such that the angle between the previous incremental
displacement {Su ('- t) } and the present incremental displacement {Su t° } should be positive. If
both choices of AA°~ result in positive values, then the appropriate root is the one nearest to
the linear solution,

AA(i) = - c / b . (40)

Meek and Tan [17, 18, 19] suggested that (38) yields imaginary roots when the structure
exhibits multiple instability directions at a point. From the authors' experience, these
problems are encountered when analyzing space structures such as the geodesic dome and the
two-hinged deep arch. For the strategy proposed herein, better control is achieved by reducing
the arc-length [M] and the arbitrary external load [P] proportionally. Thus, let the imaginary
roots occur at the ]th increment, and the tangent stiffness matrix is recomputed at ( ] - 1)th
increment from the stored relevant parameters. Both the arc-length and the external load are
reduced by half and the iteration is continued. Once the solution is determined, the arc-length
and the external load revert to their initial values. From the authors' eYperienee with the
above strategy, this modification has proved to be a stable, efficient, and automatic strategy
[31] for snap-through problems.
J.L. Meek, S. Loganathan, Large displacements of space-frame structures 67

A decreasing determinant gives an indication that a bifurcation or limit p~int region is being
approached. It was suggested by Ramm [16] and Crisfield [21] that the sign of AA°) should
follow that of the previous increment, unless the sign of the determinant has changed. If not, a
sign reversal is applied to the load increment. It is, however, reported by Kani et al. [29] and
also by the authors in [19, 32, 33] that in spatial structures with multiple negative diagonal
terms in the factorized tangent stiffness matrix before the onset of the limit point this criterion
will not work. An alternative criterion based on the incremental work done suggested by
Meek and Tan [17] has been found to overcome this difficulty. This is equivalent to using the
current stiffness parameter as proposed by Kani et al. [29] and Crisfield [30]. However, the
incremental work or current stiffness parameter criterion on its own will break down if a
structure exhibits snap-back characteristics. This problem was encountered by the authors
when analyzing a shallow geodesic dome structure. To circumvent this problem, both the
determinant and incremental work criteria have been employed in the present investigation.
Even with all the above-mentioned strategies to improve convergence, it is still not an easy
task to trace the complete load-deflection path. Oscillating solutions still occur at points either
just after or just befme limit points. When the pivots of factorized tangent stiffness matrix are
examined, it is found that multiple negatives occur, indicating the presence of an eigenvalue
cluster. When a point where the solution oscillates occurs it is necessary to revert to the
imperfect approach mentioned above. The previously converged load point is used and the
eigenvalues and the eigenvectors are calculated from this point. These correspond to the
multiple pivots on the diagonal of the reduced equations. It is then necessary to activate all the
relevant eigenmodes. This is achieved by applying small forces at the appropriate nodes and
directions. In the example of the shallow geodesic dome, one such set of eigenmodes has been
given (Fig. 11). It is crucial to the analysis that these imperfections are employed. If, for
example, a symmetric set of perturbation forces is applied, the eigenmode may simply lock up
and the solution path may still oscillate between two load values. Alternatively it also has been
found in the present studies, that if the inequality in (39) is turned into an equality, for the
proposes of obtaining the limiting values of the arc length, that the structure can be forced to
snap through to a stable equilibrium point.

7. Numerical examples

EXAMPLE 7.1. Two-hinged deep arch. The hinged arch depicted in Fig. 4 is loaded at its
apex by a concentrated load. The present analysis modelled the arch with eight beam
elements. Geometric symmetry of the structure has not been considered because of the
obvious occurrence of an asymmetric buckling mode. In this example, both the 'perfect'
approach and the 'imperfect' approach have been employed, as mentioned above, to trace the
primary path and the secondary or bifurcation paths.
The results obtained from the 'perfect' approach are shown in Fig. 5 and denoted as 1. On
tracing the primary bifurcation path, the number of negative diagonals in the factorized
tangent stiffness matrix is changing from zero to two up to the limit point. It is also noted in
the analysis that if the determinant criterion is employed in determining the direction of the
arc-length increment, the solution strategy will fail to converge and will not be able to
continue the rising fundamental path. Incremental work concept [19], found to be superior to
68 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

P, v

~ 0 ' ~

L 160 .I
Radtus, R = 100 in
Thtckness, t = lm
Area, A = 1In2
E = 10|bl;n 2

Fig. 4. Geometry of a two-hinged deep arch.

overcome this problem, was employed in the analysis. After the occurrence of the symmetric
snap-through point or the limit point, the number of negative pivots in the reduced stiffness
matrix is three. The two negative pivots in the rising fundamental path show that two branches
of secondary paths are possible. Firstly, a small geometric imperfection is added to the original
structure and the resulting solution path is denoted as 2 in Fig. 5.
Eigenvalue analysis is performed at the load point where the first negative pivot occurred.
The appropriate eigenmode is found from this analysis. This indicates that an out-of-plane
force is necessary to force the solution onto the secondary path. Three such forces of different

1 4. No Imperfection- perfect eppro•ch


2 x Geometric Imperfection at crown
1200 3 o Z force st crown : lO.OIb,
4 • : 5.OIb.
5 • : 1.0lb.
1000 6 o Moment at crown : e.41b.in.
T& = 64.01b,in.
8: =640,OIb.in.

800 \
a \
c~
a
4[
600 \
o
,.J

400
\
200

, • , • ,
DEFLEGTIONv, in.

-200
÷

Fig. 5. Load-deflectioncurves for a two-hinged deep arch.


].L. Meek, S. Loganathan, Large displacements of space-frame structures 69

magnitude were applied individually at the crown and the results obtained numbered 3, 4, and
5 in Fig. 5, respectively. Similarly for the second negative pivot, an out-of-plane moment at
the crown is found to be the necessary perturbation force. The structure is reanalyzed for
three different moment imperfections and the results numbered as 6, 7, and 8 on the same
figure, respectively.
This problem has been extensively treated by Huddlesto.n [34] and by Da Deppo and
Schmidt [35]. The nonlinear asymmetric buckling load compares favourably with Da Deppo
and Schmidt. The resulting secondary equilibrium path by employing moment imperfection
compare very well with the published curves. Unfortunately, the authors could not find the
solution in the literature for the first branch to compare with the present results.

E X A M P L E 7.2. A shallow Geodesic Dome. The shallow geodesic dome, 1 : 10 rise-to-span


ratio, shown in Fig. 6 has been formerly analyzed by Noor and Peters [36] as a truss structure
for two independent loading systems. In the present analysis, the dome is idealized as 156
frame elements with a concentrated load acting at its apex. When analyzing this structure,
various numerical difficulties are encountered to trace the complete load-deflection path. The
solution strategies adopted in the analysis have been explained briefly in the previous section.
Oscillating solutions occurred either just before or just after the limit points.
The load-deflection curves for the vertical displacement of the nodes 1, 2, 8, and 20 are
depicted in Figs. 7, 8, 9, and 10, respectively. It should be noted that the buckling form is
symmetric about the apex. Multiple symmetrical snap-through behaviour is exhibited by this
structure. For the snap-through behaviour of this structure, the solution strategy fails to

x / ~ ~ 'W' ~'

v =0,3
G = 2,652x101° NIm 2
E = 6,895 x 101° NIm 2
A = 6,5 x lO'~'m2
xl
a = 6.Ore
' A/VVVXAFI f -- O,6m
kiiiiA ), t : 035 m
Equation of surface :
"~ ~ ~ '~ ~ I x~ + x~,{x~ +7,2) 2 : 60,84
I x3 I
I

Fig. 6. Geometry of the shallow geodesic dome.


70 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

r~
200

180

160

140

120
/s
.100
0
~ 8o 3
.1
6o

4o 1

~o

o
0.1 0.2 0.3 0.4 0,5 016~' 0.7 0.8 oig
-20 DISPLACEMENT V, m. ~ / ~
/

Fig. 7. Load-deflection curve for ~, shallow geodesic dome at node 1.

converge if the determinant criterion is employed in determining the direction of the


arc-length increment. Incremental work concept [19] is used to circumvent this problem. After
the first two snap throughs, the solution failed to converge with the above-mentioned strategy.
Oscillating solution occurred due to the snap-back characteristics of the structure. In the
present investigation, a sign reversal for the arc-length increment is implemented only if both
the determinant and the incremental work changes sign from the previous increment [23].
With this strategy, the solution path is determined for snap-back behaviour of the structure.

200

1SO

leo

140

•; 120
L"
o¢[ 100
g
8O

60-

40-

| I I ~1 I I I I
O,1 0,2 0,3 0.4 o,e 0.7 O.e o.g
-20 DISPLACEMENT V. m.

F~g, 8. Load-deflection curve for a shallow geodesic dome at node 2.


J.L. Meek, S. Loganathan, Large displacements of space-frame structures 71

200 [-2O0

180

100

140

i ; 120 -120 i ~
" I00
,<
o, oo

60 -60
40 -40
20 -20
I I I I I i I t I |
0.3 0.4 -3.2 -2.8~P~2.4 -2.0 -1 6 -1.2 -0.8 -0.4 0.4 0.8 1.2 1.0
~ DISPLACEMENT V x l O "~ in
-20

Fig. 9. Load-deflection curve for a shallow geodesic Fig. 10. Load-deflection curve for a shallow geodesic
dome at node 8. dome at node 20.

Unfortunately, with all the afore-mentioned strategies to improve convergence, it is not


sufficient to trace the complete load-deflection path. The 'perfect' approach will fail to
converge at this stage. When the pivots are examined, it is found that multiple negatives occur
showing the presence of an eigenvalue cluster. When this situation occurred, the relevant
parameters are stored at the previously converged load point for the restart option in the
current program. This is necessary to conserve the time and effort for the reanalysis, which
need not to start from the unloaded position. At this stage, it is necessary to revert to the
'imperfect' approach as mentioned earlier. Eigenvalues and eigenvectors are calculated from
the previously converged load point. These correspond to the multiple pivots on the diagonal
of the reduced stiffness matrix. Then, it is necessary to activate all the relevant eigenmodes.
One such eigenmode is shown in Fig. 11. Thus, it is crucial to the analysis that an appropriate
small load imperfection be applied to continue the path. In this example, a load imperfection
of 0.1 kN (0.004 x reference load at crown) is applied at nodes indicated by circles in Fig. 11.
This tmns the bifurcation problem into a limit point problem which is readily traversed.
In Fig. 11, the dominant downward and upward displacements are marked by circles and
black circles, respectively. One should observe that the upward displacements are relatively
larger than the downward displacements. Suppose, if the load imperfection be given at nodes
where the dominant displacement is in the upward direction, the solution path could not be
traversed completely, yielding oscillating solutions. Moreover, the upward loads are also
increasing enormously due to the proportional loading. It appears to be the best approach to
apply the load imperfection in the direction that is the same as that of the reference load
direction for the nodes which are moving downwards.
It is also noted in the current analysis that the occurrence of complex roots in (38) is
eliminated by applying a small force imperfection in the appropriate direction by means of
eigenvalue analysis. From the examples presented by previous researchers, it has been
72 J.L. Meek, S. Loganathan, Large displacements of space-frame structures

N~n~ EIGEN
~ " ~ VECTOR
X2 {x3 direction)

t 3
4
5
0.106
0.147
0.106
10 0.119
11 0.193
12 0.161
13 ~195
14 0.121

25 0.233
26 0.141

28 0.197
18 -0.056e

32 -0.0046
33 -0.0101

,~5543~ " 34 321


35 -0.0181
36 -0.0324
37 -0.019T

65~ sea 5r,~, 68~

Fig. 11. Plan view of the dominant eigenmode.

(a)

(b)

(c)
Fig. 12. De[brined configurations of geodesic dome. (a) Original configuration. (b) Configuration at 1. (c) Con-
figuration at 2.
J.L. Meek, S. Loganathan, Large displacements of space-frame structures 73

(d)

(e)

(f)

(g)

(h)

(i)
Fig. 12(Cont.). Deformed configuration at geodesic dome. (d) Configuration at 3. (e) Configuration at 4. (f)
Deformation at 5. (g) Deformation at 6. (h) Deformation at 7. (i) Final deformed shape.
74 J.L. Meek, S. Loganathvn, Large displacements of space-frame structures

concluded that this type of problem is unlikely to occur. However, from the authors'
experience this will occur in structures exhibiting multiple inst~.~lity at a point. One such
problem is analyzed herein. The corresponding instability mode, at various positions num-
bered in Fig. 7, has been drawn in Fig. 12 as the perspective view.
Alternatively, when the compiex roots are obtained at the point where the multiple
negatives occurred, the limiting value of the arc-length could be obtained by utilizing equality
condition of (39). If this technique is employed, it has been found that the structure is forced
to snap through to a stable equilibrium point without showing any snap back characteristics.
More importantly, this strategy will not yield the complete solution of the previous strategy.

8. Conclusions

A solution strategy for the geometrically nonlinear analysis of elastic frame structures
subjected to static loads has been presented. The formulation has been adopted in an updated
Lagrangian approach to trace the load-deflection paths of space-frame structures undergoing
large deformations. For large joint rotations, Oran's approach has been employed in the
present study. The solution technique based on the constant arc-length method by Crisfield
was used and found to require considerable modification if the whole equilibrium path is to be
traced.
In the literature there has been little attention on overcoming imaginary roots at the limit
points. If the pure load incrementation is employed to circumvent these problem, either the
solution will drift away from the equilibrium path or will produce uncontrolled displacements
causing the solution to diverge. Whenever this type of problem occurs, even with small
increments, the displacements can still be too large. A simple method to prevent this is to
scale back both the load and the arc-length increment as mentioned earlier. However, this
strategy will break down at a point where an eigenvalue cluster is possible. As this stage, it is
necessary to revert to the 'imperfect' approach. The disadvantage of this method is the need to
compute the eigenvectors. On the other hand, it will solve the difficulty associated with the
practical implementation of th,," determination of the magnitude, type, and location of the
imperfections to be applied. The eigenvalue approach serves as a benchmark for the
formulation being proposed. Finally, the formulation and the solution strategies are validated
against the solution of highly geometrically nonlinear structures.

References

[1] R.D. Wood and d.C. Zienkiewicz, Oeo~i~trically nonlinear finite element analysis of beams, flames, arches
and axisyr~metric shells, Comput. & Structures 7 (1977) 725-735.
[21 S.N. Ren,~aa,, ,%nlinear static and dynamic aaalys~s of framed structures, Comput. & Structures 10 (1979~
879-897.
[3] G.H. Poweli, Theory of nonlinear elastic structures, ASCE J. Structural Div. 94 (1969) 2687-2701.
[4] R.H. Mailer and P.V. Marcal, Finite element analysis of nonlinear structures, ASCE J. Structural Div. 94
(1968) 2081-2105.
[5] F.W. Williams, An approach to the nonlinear behaviour of the members of a rigid jointed plane frame work
with finite deflections, Quart. J. Mech. Appl. Math. 7 (1964) 451-469.
[6] A. Jennings, Frame analysis including change of geometry, ASCE J. Structural Div. 94 (1968) 627-644.
{7] M. Papadrakakis, Post-buckling analysis of spatial structures by vector iteration methods, Comput. &
Structures 14 (1981) 393-402.
J.L. Meek, S. Loganathan, Large displacements of space-frame structures 75

[8] K.H. Chu and R.H. Rampetsreiter, Large deflection buckling of space frames, ASCE J. Structural Div. 98
(1972) 2701-2722.
[9] C. Oran, Tangent stiffness in plane frames, ASCE J. Structural Div. 99 (1973) 973-985.
[10] C. Oran, Tangent stiffness in space frames, ASCE J. Structural Div. 99 (1973) 987-1001.
[11] K.3. Bathe and H. Ozdemir, Elastic-plestic large deformation static and dynamic analysis, Comput. &
Structures 6 (1976) 81-92.
[12] D.W. Murray and E.L. Wilson, Finite element large deflection analysis of plates, ASCE J. Engrg. Mech. Div.
95 (1969) 143-165.
[13] T.Y. Yang, Matrix displacement solution to elastic problems of beams and frames, Internat. J. Solids and
Structures 9 (1973) 829-842.
[14] E Sharifi and E.E Popov, Nonh*aear buckfing of sandwich arches, ASCE J. Engrg. Mech. Div. 77 (1971)
1397-1412.
[15] C.A. Felippa, Solution of linear equations with skyline stored symmetric matrix, Comput. & Structures 5
(1975) 13-29.
[16] E. Ramm, Strategies for tracing the nonlinear response near limit points, in: W. Wunderlich, E. Stein and K.J.
Bathe, eds., Nonlinear Finite Element Analysis in Structural Mechanics (Springer, Berlin, 1981) 63-69.
[17] J.L. Meek and H.S. Tan, Large deflection and post buckling analysis of two and three dimensional elastic
spatial frames, Res. Rept. CE49, Department of Civil En;ineering, University of Queensland, Auslralia.
[18] J.L. Meek and H.S. Tan, A Stiffness matrix extrapolatio ~ strategy for nonlinear analysis, Comput. Meths.
Appl. Mech. Eagrg. 43 (1984) 181-194.
[19] J.L. Meek and H.S. Tan, Geometrically nonfinear analysis of space frames by an incremental iterative
technique, Comput. Meths. Appl. Mech. Engrg. 47 (1984) 261-282.
[20] P.G. Bergan and R.W. Clough, Convergence criteria for iterative process, AIAA J. 10 (1972) 1107-1108.
[21] M.A. Crisfield, A fast ineremental/iterative solution procedure that hr,ndles snap-through, Comput. &
Structures 13 (1981) 55-62.
[22] J. Greenstadt, The determination of the characteristic roots of a matrix by the .~acobi method, in: A. Ralston
and H.S. Wilf, eds., Mathematical Methods for Digital Computers (Wiley, Ne~,, York. 1966) 84-91.
[23] H.S, Tan, Finite element analysis of the elastic, nonlinear response of frames,~ plates and arbitrary shells to
static loads, Ph.D. Thesis, University of Queensland, Australia.
[24] W.E. Haisler, J.H. Stficklin and J.E. Key, Displacement incrementation in nonlinear structural analysis by the
self-correcting method, Internat. J. Numer. Meths. Engrg. 11 (1977) 3-10.
[25] J.L. Batoz and G. Dhatt, Incremental displacement algorithms for nonlinear l:roblems, Internat. J. Numer.
Meths. Engrg. 14 (1979) 1262-1267.
[26] G. PoweU and J. Simons, Improved iteration strategy for nonlinear structures, Internat. J. Numer. Meths.
Engrg. 17 (1981) 1455-146'7.
[27] E. Riks, An incremental approach to the solution of snapping and buckling problems, Internat. J. Solids and
Structures 15 (1979) 529-551.
[28] G.A. Wem~ner, Discrete approximations related to nonlinear theories of solids, Internat..1. Solids and
Structures 7 (19"Jl) 1581-1599.
[29] I.M. Kani, R.E. McConnel ~,~d T. See, rhe analysis and testing of a single layer shallow braced dome, in:
Proceedings of 3rd Internatiohal Conference on Space Structures, University of Surrey (1984) 613-61g.
[30] M.A. Crisfield, Accelerated solution techniques and concrete cracking, Comput. Meths. Appi. Mech. Engrg.
33 (1982) 585--607.
[31] J.L. Meek and S. Loganathan, Geometr, cally nonlinear analysis of space frame structures, in: Proceedings 5th
International Conference on Finite Element Methods, Melbourne, Australia (1987) 324-329.
[32] J.L. Meek and H.S. Tan, Large deflection analysis of space frames, ~m: Proceed~ng~ 9th Australasian
Conference on Mechanics of Structures and Mat¢~'ials, Sydney, Australia (1984) 142-1~16.
[33] J.L. Meek and H.S. 'ran, Geometrical nonlinear analysis of space frames by an incremental iterative
technique, in: Proceedings 4th World Congress and Exhibition on Finite Element Methods, Interlaken,
Switzerland (1984) 569-579.
[34] J.V. Huddleston, Finite deflection and snap through of high circular arches, J. Appl. Mech. 35 (1968) 763-760.
[35] D.A. Da Deppo and R. Schmidt, Side sway buckling of deep circular arches under a concentrated load, J.
Appl. Mech. 36 (1969) 325--327.
[36] A.K. Noor and J.M. Peter,~;, Instability analysis of space trusses, Comput. Meths. Appl. Mech. Engrg. 40
(1983) 199-218.

You might also like