0% found this document useful (0 votes)
49 views

345-Finite Rotation Analysis and Consistent Linearization Using Projectors

Uploaded by

jinshuaixu
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
49 views

345-Finite Rotation Analysis and Consistent Linearization Using Projectors

Uploaded by

jinshuaixu
Copyright
© © All Rights Reserved
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 32

Computer Methods in Applied Mechanics and Engineering 93 (1991) 353-384

North-Holland

Finite rotation analysis and consistent


linearization using projectors
B. Nour-Omid and C.C. Rankin
Computational Mechanics Section, Lockheed Palo Alto Research Laboratory, Palo Alto, CA 94304,
USA

Received February 1990

A systematic procedure is presented that may be used to extend the capabilities of an existing linear
finite element to accommodate finite rotation analysis. This procedure is a generalization of the work
presented in Computers and Structures, Volume 30, pp. 257-267. The basis of our approach is the
element-independent co-rotational algorithm, where the element rigid body motion (translations and
rotations) is separated from the deformationai part of its total motion. The variation of this
co-rotational relation results in a projector matrix, with the property that a consistent internal force
vector is invariant under its action. The consistent tangent stiffness matrix is shown to depend on this
invariance condition through the derivative of the projector. This results in an unsymmetric tangent
matrix whose anti-symmetric part depends on the out-of-balance force vector. In this paper we prove
that using the symmetric part of the tangent matrix, the Newton iteration retains its quadratic rate of
convergence. This approach has been used to solve a number of large rotation test example problems.
The results demonstrate that it is possible to analyze structures undergoing large rotations within a
general co-rotational framework, using simple and economical finite elements. The resulting improve-
ments in the performance of these simple elements are brought about through the use of convenient
software utilities as pre- and post-processors to the element routines.

1. Introduction

Optimum design of aerospace components requires the simulation of structural response


well into the nonlinear postbuckling regime. Such simulations, in turn, require robust analysis
tools that perform as well for large rotations as they do when the response is infinitesimal and
linear. There are a number of finite elements that perform reliably for response analysis in the
moderate rotation regime. Rather than to introduce new element formulations, our focus is to
construct a systematic and efficient procedure for upgrading existing finite elements to allow
arbitrarily large rotations, with moderate strains. The basis of our approach is the element-
independent co-rotational algorithm [1-3]. This procedure separates the element rigid body
motion from its smaller deformational response, thus allowing small or moderate rotation
strain-displacement relations to be used in the context of arbitrarily large total rotations. It
belongs to a family of co-rotational procedures, first introduced in 1969 by Wempner [3] and in
1973 by Belytschko and co-workers [1, 4]. The concept of a rotating frame attached to the
element was also used by Horrigmoe and Bergen [5] and Argyris et al. [6]. De Veubeke [7]
used a single rotating frame, attached to the structure, for dynamic analysis of flexible bodies.

0045-7825/91/$03.50 © 1991 Elsevier Science Publishers B.V. All rights reserved


354 B. Nour-Oraid, C.C. Rankin, Finite rotation analysis and consistent linearization

Typically, the directions of the axes of this frame are chosen based on a subset of the
displacement degrees of freedom of the element, and thus the element need only be
formulated for linear or near linear response with respect to this frame. In this way, the
desired nonlinearities are introduced via the rotation of the local frame. In 1986, Rankin and
Brogan [2] introduced the element-independent co-rotation as a general framework for large
rotation analysis in three dimensions. For an overview of co-rotation for large rotation analysis
see [8]. In the co-rotation approach, the deformational part of the displacement is extracted by
purging the rigid body components before any element computation is performed. This
pre-processing of the displacements may be performed outside the standard element routines
and thus is independent of the elemen,~ type (except for slight distinctions between beam,
triangular and quadrilateral elements).
The element-independent approach presented in [2] performs well for many situations. For
certain analyses, however, difficulties were observed with the co-rotational formulation when
u~ed in conjunction with some shell elements. What was common among these elements is
that they did not possess correct invariance to even infinitesimal rigid body motion; that is, the
internal force field derived from their strain energy was not in se|t'-equilibrium. Adverse
stiffening of the element was especially evident with curved surfaces modeled using flat
elements, a phenomenon commonly known as warping-sensitivity. In a previous paper [9], we
addressed this question where the self-equilibrium of the internal force vector, normally
satisfied at the element level, was enforced through .a projector operator acting on the
computed internal forces. Projectors are matrices with special properties that may be used to
purge certain unwanted components of a vector by operating on them. Projectors have been
used previously in finite element development [10]. In [9], a projector arises naturally from the
derivative of the local co-rotational element frame with respect to the total displacement
degrees of freedom. This derivative is evaluated as part of the variation of the strain energy
used to obtain the element nodal forces. The action of the projector operator on a
non-equilibrated element force vector brings it into equilibrium. Moreover, the rigid body
components of an incremental displacement vector are eliminated when multiplied by the
projector. Finally, it transforms a linear element stiffness matrix to one with correct rigid body
properties. In a recent paper, Crisfield [11] successfully used a co-rotational formulation for
large rotation analysis of three-dimensional beams. His approach, although specific to beams,
incorporates many of the concepts presented in [9].
In the work presented here, we assume that the projected internal forces may be obtained
through the consistent variation of the strain energy with respect to the global freedoms.
Similarly, the element tangent stiffness is obtained as the consistent linearization of the
projected internal forces. This linearization involves the projector developed in [9] and its
derivative, the derivative of the orthogonal transformation from global to local coordinate
system (i.e., the co-rotating basis), and the original element tangent stiffness matrix. In
addition, by omitting the small rotation assumption for the deformational degrees of freedom
and choosing a suitable definition of the local element rotational freedoms, we obtain a
modified element stiffness matrix that possesses correct rigid-body invariance even for large
rotations. As a consequence of this invariance, we show that a consistent tangent stiffness is
related in a special and explicit manner to the internal force vector. This relation leads to a
simple test that may be easily used to validate existing element formulations for rotational
invariance.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 355

In Section 2, we present the basic concept of this paper in terms of infinitesimal motion.
Section 3 contains the main body of the paper where the consistent linearization of the
co-rotational formulation is presented. This lincarization results in the projector operator, its
derivative and the consistent element tangent stiffness matrix. A proof of the quadratic
convergence resulting from the use of the symmetric part of the tangent matrix in a
Newton-Raphson iteration is also given in this section. For formulations that lead to stiffness
matrices not satisfying the correct rigid-body invariance, the extensions presented in this
section provide a means for modifying the tangent matrix to meet this invariance requirement.
The computational procedure used in this approac.h is summarized in Section 4o This
procedure may be implemented through a set of element-independent utilities that is used to
manipulate the element displacement and force vectors and the element tangent stiffness
matrix. This procedure is used to solve a number of example problems and the results are
presented in Section 5. An important practical result is illustrated using a Bernoulli beam
element that has no bending-torsion coupling. Using the element-independent procedure
presented here, the proper coupling is introduced through the rotational invariance condition.
Finally, we establish that the performance improvements demonstrated in [9] for linear
analysis may be extended to problems displaying highly nonlinear behavior involving very
large rotations.

2. Invariance properties under infinitesimal rigid body motion

We begin with a body under an internal stress state undergoing an infinitesimal rigid body
translation and rotation. The strain energy associated with this state is uneffected by this
motion and thus must satisfy certain invariance conditions under rigid body motions. In turn,
this invariance property can be translated into a set of conditions on the internal force vector
and the tangent stiffness matrix. These conditions were derived in [9] and for completeness are
included in this section. Here we focus on the small motion case, since it allows us to
demonstrate the basic concepts in this paper by using simple linear algebra operations. In a
later section, we will extend the formulation to include the large rotation case.
We define • as the linearly independent set of rigid body modes associated with a single
finite element. In three dimensions, ~ has six column vectors that describe rigid body
translations and the rotations about the three coordinate axis. Given a nodal displacement
vector 8d, one can obtain a modified displacement 8d derived from 8d that contains all the
information describing the same deformation as 8d. 8d is obtained by subtracting the
components of the rigid body modes (columns of ~ ) from 8d. This can be formally written as

~l = Sd + IPb , (1)

where b is a vector that contains the components of 8d along each rigid body mode. Equation
(1) defines the N components of 8d in terms of the N components of 8d and three new
translation and three new rotation parameters in b. For 8d to be the pure deformational part
of 8d it must be orthogonal to the columns of * . This orthogonality condition results in the
following relation for b:
356 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

b= (2)

Using this result in (1), the pure deformationai part of 8d may be written as

8d=PBd, where P=I- ~Tt(l//'tl//)-ll//'t ; (3),(4)

P has the property P = p2 and therefore by definition is a projector matrix. The particular P
given in (4) acts as a filter to purge the rigid body components when it operates on 8d. Note
that P is not invertible since its eigen-values are either zero or unity. This indicates that 8d
does not contain all the information in 8d. The action of the projector P retains the
deformational part of the motion while discarding the rigid body motion. While the P given in
(4) is symmetric for infinitesimal motion, this is not the case in general. In the following
sections we will derive P for large rigid body rotations and show that it is unsymmetric.
The invariance of the strain energy under rigid body motion can then be stated as

6( d) = (5)
The internal force vector is the derivative of the strain energy with respect to the displacement
parameters. Since d and a are in the same coordinate system, a consistent finite element
discretization should produce the same internal force vector when using either d or d as the
displacement parameters. Thus

O~b = Oq~ (since ~b )

_ o,¢,

Odi oai '


(6)

where f/, d i and a i are the i-th components off, d and d, respectively. Here we use Einstein's
convention where a summation is implied over repeated indices. The result in (6) for
consistent finite elements where f = dc~/Od= adp/Od can be put in matrix form as

f= PT. (7)

The above equation is obtained by substituting the differential form of (3) given by

e = oJ/Od (8)
into the last part of (6). Equation (7) requires that the internal force vector computed using a
consistent finite element formulation be in self-equilibrium. This result is independent of the
element formulation. It is simply a condition due to kinematics and a statement of the
invariance of the stress state under rigid body motion, which is of fundamental importance.
Consider next a finite element formulation resulting in an internal force vector fwhich does
not satisfy (7). In other words, f is not a self-equilibrating set of forces. Define f through
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 357

f=e~. (9)

The result of this transformation is to produce a force vector that is in self-equilibrium. This
can be shown by noting that f given in (9) satisfies (7) as a consequence of the fact that P is a
projector matrix (i.e., P = p2). Thus, the internal force vector computed by an element
formulation which does not correctly account for the invariance properties under rigid body
motion can be restored through the transformations in (9).
The tangent stiffness matrix is computed as the second derivative of the strain energy with
respect to the displacement and rotation parameters. Taking the derivative of (9), the (i, j)
component of the tangent stiffness matrix is obtained as

k,j - odi odj odj t Odi Odm


0din 02, Or,, 02dm OPmi
od, Oa,noan od, Od i Odi Odm- Pmif~,nnPnj + ~ f,n "

To derive the last equation we used the definition of P given in (8) where Pii are the
components of this matrix. In matrix form this equation becomes
Opt,,, Opt
K = e'Ke + ~ -07 i m = e ' K e + - ~ : ]; (10)

Pm is the m-th row of P and the operation denoted by ':' represents a contraction and implies
forming a linear combination of a set of matrices with each component of f as weights. These
matrices are the derivatives of the columns of P with respect ot d. The result of this
contraction is the so-called 'geometric' stiffness matrix. Under small deformation assumption
the derivative of the projector matrix becomes constant and the internal forces become small
compared to the material stiffness. As a result, the second term in (10) vanishes and the
tangent stiffness matrix reduces to pt~p derived in [9].
For large rotation analysis, it is possible to express the incremental equilibrium equations in
terms of a parameterization of the orthogonal rotation matrices at the nodes defining the
configuration. When the stresses are derivable from a potential energy functional, the set of
incremental equilibrium equations has the symmetric tangent in (10) as its matrix coefficient.
The solution to this symmetric system results in a set of three incremental rotation parameters
at every node. The use of these parameters yields equivalent generalized forces that cannot be
interpreted easily for general purpose use.
Our approach is to introduce a change of variables to instantaneous spin variables (direct
counterparts ot angular velocities) which yield the commonly accepted interpretation for
moments and forces and results in simple expressions for the projector and the tangent
stiffness matrix. Since the potential energy cannot be expressed in terms of these instanta-
neous variables, the resulting tangent matrix is not symmetric in general. Fortunately, using
the symmetric part of this matrix yields the same favorable rate of convergence as will be
shown in the sequel.
358 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

3. Including large rotatic,ns effects

Extending the results in the previous section to large rotations requires a change of
variables for the global displacements and rotations. The global displacements are written in
terms of three quantities:
(a) the displacements of a point, attached to the finite element, and chosen to be the origin of
a local coordinate system,
(b) the rotation of these local axes,
(c) a local displacement in these axes.
Similarly, the global rotation is written in terms of the rotation of the local axes, and a local
rotation relative to these axes. This change of variable is referred to as the co-rotation method.
An important property of the co-rotation frame is that while it translates with the element,
unlike convected coordinates it remains an orthogonal frame.
Consider a body undergoing a motion from its initial configuration to the current one (see
Fig. 1). The displacement vector of a typical point P with position vector X ~ in the global
coordinate is denoted by u g and its rotation which is defined in terms of an orthogon,"l matrix
denoted by T g. The superscripts g and e denote quantities in global and local element
coordinate system. T g transforms any vector attached to P from the initial configuration to the
current one. A point O is chosen in the body, with position vector Xog in the global
coordinates, to be the origin of a local element frame. The initial orientation of this local
element frame is defined by the orthogonal transformation E o. The same triad E o is also
attached to point P in the initial configuration. As the body undergoes its motion, the origin of
the local frame translates and rotates. This translation is denoted by u o and the corresponding
rotation is defined by the orthogonal matrix E with respect to the initial global coordinates. In
the current configuration, the position vector of P is given by x ~ = X ~ + u ~, while that for the
origin is x~ = X~ + u,~,. The position vector of P in the local element coordinate system is
simply x ~ = E ' ( x ~ - x g,). The last equation is derived by transforming a global quantity to the
local element frame. The displacement of P in the local element frame is then given by

INITIAL \ ~i~.~ I \\-u • " ~

Fig, 1. Motion and deformation of a typical body.


B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 359

a ° = e'0: + x - - XD- x . (11)

A superposed bar over any kinematic quantity (i.e. ~ ) denotes a deformational quantity that
is small in magnitude compared to its corresponding global quantity.
The transformation from the point P in the current configuration to the global coordinate
system can be obtained in two different ways. One is the product of T g and Eo (since the
transformation from point P to o in the initial configuration i s the identity). The other is the
product of the rotation tensor equivalent to T ~ defined by T g and E. T g is given by

7"~ = E f ~ E ' . (12)

This equation simply transforms T¢ from the local coordinate system to the global one to
obtain T ~. Noting that in the initial configuration the axis of the triad at point P are parallel to
those of the local element frame, we have

TgE = TgE,,. (13)

Substituting for Tg from (12) into the above we obtain

T~ = E ' T g E o . (14)

Note that T~ and 1~ are small rotations describing deformations and thus are close to identity.
Here, we introduce a notation that was used in [9] and represents a 3 x 3 skew-symmetric
matrix (a spin tensor) as a function of the corresponding axial vector (the spin axis). This
matrix function may be described in terms of the cross-product operation and takes the form

0 -% % ]
g}=spin(oJ) = o J x = ~3 0 --~1 , (15)
-- ~2 ~1 0

where toi, i = 1 , . . . , 3, are the components of the axial vector ~o. This equation shows the
correspondence between the product of a skew-symmetric tensor and the cross product
operation, specifically,

spin(to)r=to× r = -r x t o = -spin(r)to (16)

for any pair of three-dimensional vectors r and o~.


Moreover, associated with every skew-symmetric matrix, g~, there is an orthogonal matrix,
Q, that is related to g~ through the matrix exponential (see Appendix A)

Q = en " (17)

Q is orthogonal since

Q~ = e nt = e - n = Q - : . (18)
360 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

io ..ooo., o :o = , I - - _ _ o- . - . o , ,, s . . , , , , - - o . j
log. o-I I,,, = AXIAL (~2} I
Fig. 2. Equivalence of vectors, skew-symmetric matrices and orthogonal matrices in 3-D.

This result (exhibited in Fig. 2) holds for any size matrix, whereas the equivalence of a vector
and a skew-symmetric matrix is only valid in three dimensions.
The above results can be used to introduce the equivalent rotation angles for T g and T~
used in (14). Accordingly

0 g = log~(T g) and Og = axial(Og), 0 ~ = iog~(f ~) and 0~ = a x i a l ( O ' ) .

Using these results in (14), one obtains the following relation:

O~ = log~(E eO~Eo) . (19)

The change of variables defined by (11) and (14), or equivalently (19), describes the
primary variables u and e in terms of the new variables a and e in a local coordinate frame.
~(~, R) is assumed to be unique and differentiable. Thus, using_the chain rule, the internal
force vector is obtained in terms of the derivatives of ~ and 0 with respect to the global
displacements and rotations at the nodes. Accordingly,

r where J,~t, =
d,i,;
dU.o
dat;
d6,;]
=-~|
OUa
d0t,| (20), (21)
de,, Lde,,j do,: do,,j
Through a further application of the chain rule, the derivatives with respect to 0 and 0 in
the Jacobian transformation J given above can be written in terms of the exponential of 0 and
0; namely T and T. Equivalently these derivatives can be obtained in terms of to and ~. The
derivative of to represent instantaneous rotations (see Appendix A). As a result 8to = ~T T t
and 8 ~ = BT T'. These relations yield

L~,, = tt(e,~) -o ,
'P,,hH(Ob)
' (22)
where
r 0~" / .
(0,,)= 0 A(O,~,) ' H(0,;)= 0 A(0,;) and P,,,, /0,~ z 0-z/' (23)

Otoa J
A(#) represents the derivative of a with respect to to, namely ##/eto and is a function of O
(see (A. 14) in Appendix A), where the projector relates the total increment of motion in the
element frame to its deformational part through
R Nour-Omid, C.C. Rankin, Finite rotation analysis aud c¢ngistent linearization 361

8d~= P a d ~. (24)

An explicit form for H may be obtained directly from (A.14).


A modified form for the internal force vector that is equivalent to that givci~ in (20) may be
defined as

[ o4, / o,,: ]
f:(u,~)=La6/a~:j.
Using the definition for H a n d (23), the above modified internal force vector can be related to
that in (21) through
Nnodes Nnodes
f , ,°( u , to) = H ( O : ) t? ~ ( u , O) = P',,t,H(O;)'j;(a,O)= ~ e'j;(a,,~). (25)
b b

It is important to note that the above form of the internal force vector is based on a change of
variables from 0 to a~. Once an incremental change, 8oJ, is computed, the new T is evaluated
by multiplying the exponential of spin(~ito) by the old T. 0 is calculated using the natural
logarithm of T.

3.1. Derivation of the projector

We begin by taking the variation of (11) and (14). Thus, we obtain

,~a '~ = 8 E ' ( u g + X g - ugo - Xgo) + E t ( $ u g - ~ u ~ ) and ~'~ = aE' TgEo + E' 8T ~ E o ,
(26), (27)

where 8a denotes an infinitesimal change in a. Note that X ~ and X g are constant undeformed
coordinates in local element and global coordinate systems, respectively• The variation of an
orthogonal matrix is required in the above two equations. We introduce ~ . as the instanta-
neous axis of rotation for E in the global coordinate system• oJ~ satisfies

~E = spin(Sco~)E. (28)
This relation can be used to simplify (26) as follows:

aa e = E'l-spiJl(a~,~)(u ~ +x ~ - Uo~-X~o)+(au~-au~o)]
= - E t s p i n ( a c o g e ) E E ' ( u g + X g - u ° - X~o) + E ' ( a u ~ ~.~o)
• e
= -spm(8oJE)E ;
(u g + X g
- uog - X og) + ( 5u~ - ~u~o) (29)
= - s p i n ( ~ o ~ ) ( a e + X e) + (Sue - 5U~o),

where the last equation is obtained by substituting for E ~(ug + X g - Uog -- Xog ) f r o m ( 11 ). Note
that in the first of these equations the expression in the square brackets is simply the
362 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

deformation part of the displacements in the local element coordinates. Similarly for (27),

~ = [ - E t spin(~tog)T g + E t spin(stog)TglE,,
= [spin(~to ~) - spin(~to~ )]EtTgEo = [spin(Sto e) "spin(~to ~-)] # e . (30)

In the above we used the fact that g transforms a vector and a tensor from the global
coordinate system into the local element coordinates according to

r e = Err g and spin(r e) = Et spin(rg)E.

Equation (30) can be written in terms of the instantaneous rotation axis for T~ to get
e
8~ e = 8to ~ - 8to E . (3!)
At this point we assume that the finite elements are invariant under pure translation, a
requirement that is normally satisfied if the element passes the patch test. Hence, u~ may be
omitted from (29) to obtain

~tie = _spin(~to~).r. • + ~u ~ = spin(x ~) Bto~ + Su ~ , (32)

where x e = X e + ti e. The last equation is obtained using the cross product property of the spin
operator, spin(x e) is a skew-symmetric matrix constructed from the components of x e
according to (15) and is simply the m o m e n t a r m tensor.
With the above assumptions, the P given in (23) may be described in terms of its 6 x 6
blocks:

Oqto E • ~ 0toE]
I8,,h + spin(x,~)
spin(x,,) ~ ~,,F~
J
= I8.b _
Pah
0to E (33)
18"h 0toe
(guh
where

~rt= [ - s p i l ( x : ) ] and Fh= #toe/#tohj,


] (34),(35)

where the subscripts a and b refer to quantities at two typical nodes a and b in the finite
element discretization. In (35), #toe/0ub and #toE]Otob and thus F represent the variation of
the base vectors of E (the element frame) with respect to the displacements at all the nodes.
We include explicit expressions for F in Appendix B. A detailed derivation of these results is
presented in [9] where the bi-orthogonality condition between

F= F~ and ~=

n de node
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent iinearization 363

is established for the choice of local elemep.t irames described in Appendix B. This
bi-orthogonality condition takes the form

U'~=l. (36)

3.2. Tangent matrix and derivative o f the projector


The element tangent stiffness matrix may be obtained by differentiating zhe interna)l force
vector with respect to the element nodal parameter. Then

K,a, = Of~ / Od g , ( 37 )

where d g is the set of parameters associated with node a in the global coordinate system, d~g
g The internal
has six components; three translations, u,g and three instantaneous rotations, toa.
force vector in local frame given by (25) can be transformed to the global frame using an
orthogonal matrix. This result and its dual in the virtual work relation yields

fg = G P ~ " : , 8d ~ = PG' 8d g , (38)


where

G = diag[E] = I ]
.
""
E
is a block diagonal matrix with the 3 x 3 transformation matrix E as its diagonal blocks.
Taking the variation of (38) we obtain

6fg = GP' 8 f ~ + G 8P t f " + 8G ptfc . (39)

The first term on the right-hand side of (39) simplifies as

G P t 8 ] '~ = G P ' - ~ 6d '~ = Gpt,K'~ta '`, ~'~) "~'~ 5 d ~

= Get/~(~ ~, ~ ' ) p ~-~ 8d g = G p t / ~ ( ~ ~, o ~ ) e G t 8d g . (40)

This term is due to the symmetric linear stiffness term transformed by the projector to possess
the correct rigid body modes and then transformed to the global coordinate system. The
second term in (39) depends on the derivative of the projector operator, P. This derivative
may be obtained directly from (33) re~ulting in the following relation:

8P = - ~ ~U' - 8 ~ U t . (41)

The derivation of ~T requires the definition of the local element frame. In Appendix B, we
have shown that for the given choices of the c!em:ent frame U can be written as

r = Y + ¥~', (42)
364 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

~" and Y are constant matrices with entries that are 0, 1 or - 1 and _ is a 3 x 3 square matrix
that varies with the element deformation. The derivative o f / " depends on that of ~, through

8F = Y 8~'. (43)

The bi-orthogonality condition between/" and qv (eq. (36)) and the dependence o f / " on the
element deformation (Eq. (42)) yields

~ t ~/~ ._}_~ t t y , ~ = I . (44)

In this equation both ~ and ~ t y are invertible for the choices of the element frame
considered in Appendix B. Thus

~= [qtty]-'(I- qrtY). (45)

This equation establishes that the components of the derivative o f / ' depend only on those for
W. Differentiating (44) and simplifying the result using (42) yields an expression for 8 ~ in
terms o f / ' , Y and 8W. Thus

8 ~ , = - ( q t ' Y ) -I 8 ~ P ' / ' . (46)

Substituting this relation in (43), we obtain 8/" directly in terms of 8qf as

8/" = - Y(rFtY) -~ 8q vt F . (47)

This relation can now be used to eliminate 8/" from (41) to obtain

8P ffi ~/"t 8qt ( r t q v ) - t y t - 8 q v F t . (48)

This is the most_general form for 8P. For many elements such as triangular and quadrilaterml
shell elements, Y = 0. This result can be used to simplify the expression for iv, and ~, reduces
to

'-~ -- (t//tY) - l . (49)

Thus 8/" simplifies as follows:

8P = W / " 8~' ( r t ~ ) - , y , _ 8 ' P / ' t = ( ~ / ' t _ I) 8 W / " ' = - P 8W F t . (50)

The second term in (39) involves the product of the variation of pt and j~. Then using (50)
we have

G 8P' ] ° = - c . r 8qr' 8V.' re(u, to), (51)


where the last of (25) was used to eliminate the deformational part of the displacement and
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent iinearization 365
rotation quantities, ii and o~. Using the explicit form for ~' in (35) and partitioning.f~ as

f~(u, to)= [ n: ] (52)


Lm:J'
(51) can be simplified further to obtain
Nnodes n:] Nnodes
~a I ~a-e ]
GSPtf~(~,~)=-GF ~ [spin(Sx:) 0] m : j = G r [spin(n:) 0] 8o3:

= -GFFtp ~d e = -61"Ftpo t ~O g ' (53)

where (24) and the second ef (38) were used to replace deformational quantities (;~.e., 8~ ¢ and
8a5 ~) with total displacement and rotation increments in the element frame and

~a= [Spl n a">]. (54)

Note that F is independent of the moments in if(u, to).


The third term in (39) depends on the variation of G. Using the definition of G and (28) in
local form we have

8G P ~ = 8G f~ = G diag[spin(Sto~)lf e
= -6P 8to~ = - ~ P r ' ~d ° = - 6 ~ r ' 6 8d ~ , (55)
where
[ spin(n:) ] (56)
/~" = L spin(m:) 1"
The result in this equation can be combined with that in (53) and (40) to obtain the
following expression for the tangent stiffness matrix:

K ~ = a(e'~,(a ~, ~ ) e - r~'e - Pr')6'. (57)


ke
The first term in this equation depends on the tangent stiffness matrix/~(i, d ) that is
obtained based on small displacement assumption and therefore is symmetric. However, the
second and third terms are unsymmetric, and hence the tangent matrix is unsymmetric. It is
useful to note that the unsymmetric part of the element tangent matrix has a rank of six. In
[12, 13], Simo and Vu-Quoc obtained a similar unsymmetry of the tangent matrix. Fortunate-
ly, the unsymmetric term becomes zero when the system is in equilibrium. That is,

IIK-Ktll ~ 0 ~ II~u,~)ll ~ 0 . (58)

This is a sufficient condition for a quadratic rate of convergence in a Newton-Raphson type


iteration, as shown in the following theorem.
366 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

CG~_,sider the Newton-Raphson type algorithm where at each step one solves the linear
system of equations given by

K~i, Ad(i) = -f(d,i)) , (59)

where K s and K A are the symmetric and antisymmetric part of the tangent matrix defined by

K s = ½(K + K ' ) , K A = ½(K- K ' ) . (60)

T H E O R E M . Ignoring the antisymmetric part of the tangent matrix when solving the linearized
system of equations in a Newton-Raphson type algorithm results in a quadratic rate of
convergence if

IIK ,II ~< allf(d.OII (61)

where a is a constant.

P R O O F . The Taylor series expansion at d(i ) is

f(d(i)) = f(d(i_l)) + K(,_I) Ad(i_t ) + r(Adti_,)) , (62)

where r is the remainder which depends on Adti_~ and

IIr(ad._,))ll flllad._, l! 2 . (63)

with/3 some other constant. Writing K,_ t~ in terms of its symmetric and skew-symmetric parts,
we have
s
f(d<;)) = f(d(,_l) ) + Kt,_l) Aden_I) + K ,a_ l ) Adti_t ) + r(Ad._l~) .

By using (59) for the ( i - 1)st iteration, the first two terms on the right-hand side vanish
resulting in the unbalanced force vector in terms of the anti-symmetric part of K and the
remainder vector, r, to get

f(d(i)) = K A
(,- t~ Adt,- ~ + r(Ado- ~)). (64)

Then, substituting this equation back into (59), we obtain

Ad.) = - ( K ( so ) - I { K (,-l)
A Ad(i-l) + r(Adt,-~))} (65)

and taking norms

Ilad.,ll Ilad._,,ll + IIr(ad._,>)ll} • (66)


Using the original assumption in (61) and taking norms of (59) for the ( i - 1)st iteration, we
arrive at the following inequalities:
B. Nour-Omid, C.C. Rankin, Finite rotatwn analysis and consistent iinearization 367

A
IIg.-,,ll ~< a[[f(d,+_l,[I <<- ~ll( K ,,-,~11
s IIAd,,-,,ll . (67)

Substituting the last result in (66) to eliminate the antisymmetric part of K we obtain

IIAd,,,ll ~ II(K~+,)-'II {all K .-,,ll


~ IIAd.-,,ll ÷ ~llAd,+_,,ll 2} (68)
or
lid,,=,,- d.,ll ~< ~',lld,,, - d,+-,>ll 2 . (69)
where
= II(g~,,)-' II(~lIg,%-,~ll +/3). (70)

Thus if the antisymmetric part of the stiffness vanishes at the same rate as the residual,
convergence is quadratic whether the symmetrized or unsymmetrized stiffness is used. This is
precisely the situation with the tangent matrix given in (57) where the nonsymmetric part of
the stiffness is proportional to the residual. []

3.3. Properties of K ~
'The consistent stiffness K e given in (57) has some unusual properties that come about as a
result of large rotations. It is well known [6] that moments defined in either body-centered or
space-centered frames are not conserving. The unsymmetry in the tangent matrix away from
equilibrium is a direct consequence of this fact.
Given that the projector annihilates the rigid body modes 1/" (i.e., p~t = 0) the product of
the tangent matrix and these set of rigid body modes reduces to

K~ : -/~. (71)

This result is equivalent to the statement that the forces and moments respond to an
infinitesimal rotation, 8to, the same way as vectors according to 8f= 8oJ x f.
Next, consider the product of ~ ' and K~ Using (57) together with the orthogonality
property in (36) we have

¢'K ~ = - f ' P - ¢':r' = -f' + (f'¢- ¢':)r'. (72)

where the last equation is obtained by substituting for P from (33). The quantity in the
parenthesis consists of sums of products of spin tensors, and when expanded out becomes
Nnodes

spin(n:)spin(x:)-spin(x:)spin(n:)-spin(m:)
Q

Nnodes

= ~ spin(n: x x : - m : ) , (73)

where we used the identity s p i n ( a ) s p i n ( b ) - spin(b)spin(a)= spin(a x b) to obtain the last


equation. The result in (73) is idei~tically zero since n: and m: have been modified by the
action of the projector operator (see (25)) to force self-equilibrium on re. Thus, (72) reduces
to
368 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

+,'/+o = - P ' . (74)

Moreover, (73) can be used to obtain an alternative form for the stiffness matrix given in (57).
Thus
/++ = e ' / + e - rP'e- fir'= pt/~e_/..,pt -I--/'pt'I/fFt - fir'

= e'/~e- I'.i~' + r~'fr' - fir'= e'ge - FF t- etfr'. (75)

To obtain the last result, we first expand P in terms of ~t and F. Next, we substitute the
relation/~t~ t = ~ t f obtained from (73), and finally combine terms to get pt.
Using (75), the symmetric part of the tangent matrix takes the form
e
Ksymm = P'K.P - F F t p - F F t = P'KP - I F ' - ptFFt , (76)
where
F = ½(F + F ) . (77)

The blocks in F associated with a typical node a are of the form


= [ spin(n~! ]
Fa L½ spin(m,,)J" (78)

It appears that for the symmetric part of the tangent matrix, the moments rotate by half the
amount that the forces do. This fact can be verified directly for a variety of 'elements' that are
defined to be objective to large rotations in the first place. These moments are equivalent to
the 'semi-tangential' moments introduced by Argyris and discussed in detail in [6].
The product of the symmetric part of the tangent matrix given in (76) and gt becomes
e
K~ymmgt= -F. (79)

This equation provides the bases for a very simple and general test for the invariance of a
generalized tangent stiffness operator to infinitesimal rotations. One only has to compute two
quantities;
(a) ~ , which depends only on the current coordinates of the nodes,
(b) F, which depends on the internal force.
If (76) is not satisfied, one may conclude that the element stiffness matrix is not consistent
with the internal force vector. Elements that do not pass this test may produce erroneous
results when used in large deformation analysis or when the elements are distorted. This test is
valid for any conserving system that yields a symmetric stiffness, however such stiffness is
derived. The above is not a valid test for non-conserving systems, such as follower force
problems where 'load stiffness' terms are included in the tangent matrix or cases where
element boundary conditions themselves are non-conserving.

3.4. Change ~ f l~Triables f r o m O~ to ~ e

In the prewou,:', section, the internal force and the tangent stiffness were derived using the
instantaneous rotations ~e. In this section we derive these quantities directly in terms of 0~.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent iinearization 369

For some elements with a complete set of rotational degrees of freedom (i.e., elements
derived from C ~ shell theory that include drilling stiffness) the current configuration is given in
terms of rotation angles O. The interrelationship between the large rotation 0 and orientation
is given by (A. 1) for any rotation matrix T. In our case where elements are imbedded in the
corotation framework, these quantities are the deformational part, re, of the total rotation.
From (25), the change of variable in the internal force from 0 to to is carried out according
to

--¢2 --C --~ t~¢ --C


f.(u , ~) = H ( O o ) 7 o ( u , 0~), (80)
where the superposed tilde denotes explicit dependence on t~~ for the native element internal
force and subsequently for the tangent stiffness matrix. The element tangent stiffness matrix is
obtained for any two nodes a and b by differentiating the above equation to obtain

oi: owo , -o
/~,, _ Of: = H'~ ~ + 6"b - - : f : = H~KahHb + 6'~b Oto,, " f : (81)

where the derivative of H is given in Appendix A. For moderate deformational rotations the
contribution to the stiffness from the derivative of H vanishes, a condition that is always
satisfied in the limit of a fine grid.

4. Computational procedure

The element-independent co-rotation algorithm derived in terms of the projector operator


and its derivatives is summarized in Table I. All the computations relating to element
quantities are performed in Step 1 of this table. This step can be broken down into three basic
stages. The first stage, labeled (a) in Table 1, filters ihe displacements and rotations in the
global coordinates to obtain local deformational quantities. Stage (b) simply computes the
internal force vector and tangent stiffness matrix with existing element software using the
deformational quantities evaluated in stage (a) and applying the transformations in (80) and
(81). When the element deformation is sufficiently small (i.e., for a fine mesh) or for elements
that do not use rotations as primary degrees of freedom, the operations involving H in (80)
and (81) may be omitted. Stage (c) involves filtering the internal force vector to ensure
self-equilibrium and modifies the tangent stiffness matrix to satisfy the condition in (79).

5. Numerical examples

In this section we present the results obtained from the analyses of some example problems
to demonstrate the characteristic of the procedure described in this paper. These examples
require only a few degrees of freedom, are simple to describe, but nevertheless are capable of
very complicated nonlinear response in the large rotation regime. In the first two examples,
we used simple rod elements with cubic interpolation for the bending shape functions, and
linear interpolation for axial and torsional freedoms, and two Gauss integration points. These
370 B. Nour-Omid, C.C. Rankin, Finite r o t a t i o n analysis and consistent linearization

Table 1
Element ~ndependent, projector based algorithm for finite rotation analysis
I. For each element perform the following:
(a) Compute E, Uog from ug.
~ = E~(u g + X ~ - uo _ x D _ x o
~ = E ' T e , Eo
6 ~ = log~(T ~)
(b) Evaluate f f ( t i ~,/)~) a n d / ~ ( t i ~, 6~).
Compute H and its derivative.
Compute fe(li~, t~ ¢) and/~e(li c, o3~) (eq_s. (80) and (81)).
(c) Compute • (eq. (34)) and U from ti ~, T ~ and X ~
f~ = p~f~
Compute F from f f (eq. (78)).
K ~ = PtK~P - UF t - P'FF'
Transfer to global coordinates:
/ ~ = GKCG ~ a n d f f = Gff

2. Assemble and solve:

KAd=f

3. Update:

U~ := Ug + AUg

Tg : ~-. espirL(,,~¢a'g}Tg

4. Terminate when converged

traditional Bernoulli elements have been in use in the Structural Analysis of General Shells
(STAGS) [14] computer code for many years, and without the above element-independent
methods, are known to underestimate the displacement field for even moderate rotations. We
used this element without changes. Only the element-independent modifications to the
displacement field, the nodal force vector and the stiffness matrix described here were carried
out. The response to the first example involves large twist of a slender cross section. We found
it was more effective to orient the element frame using a midpoint interpolation from the
triads at the two ends of the beam (rather than only one mode, as described here). This
interpolation is straightforward and does not change the content of this paper. The result is a
slight change to the first column of/1.
Our first example is that of a hinged right-angle frame first introduced by Agryris [15], and
later analyzed by Simo and Vu-Quoc [12, 13]. The geometry and boundary conditions are
illustrated in Fig. 3. Using the symmetry condition about the y-z plane through the apex, a
finite element model for half of the frame is sufficient to capture the nonlinear response of the
frame. At the support, translation along x and rotation about z is permitted. The loading is a
moment about z at this end. The problem is made more difficult by the slenderness of the
cross section (Fig. 3) with a thickness/height ratio of 1/50. Hence, the orientation of the cross
section (i.e., as it twists out of plane) along a side of the frame is critical, and will severely test
our formulation. This problem involves truly large three-dimensional rotations, with the frame
B. Nour-Omid, C.C. Rankin, b~nite rotation analysis and consistent linearization 371

LENGTH L = 255.0mm
y
WIDTH w = 30.0mm
THICKNESS t = 0.6 mm ~,~
YOUNG'S MODULUS E = 71240 N/ram 2 f -
POISSON'S RATIO v :: 0.31y///~. L CROSS-SECTION

/J/
Z

Fig. 3. Right angle frame under applied end moment.

rotating a full 360 ° out of plane as the supported end rotates (in plane) a full circle. The
response has two paths. A primary path represents two-dimensional response in the x - y
plane, and a secondary path exhibits the full three-dimensional response. Separating these two
paths are bifurcation points at ---626.7 N mm. The finit,~ element model used for this analysis
consists of ten equally-spaced rod elements. This bifurcation point was also obtained from a
linear eigen-analysis. Previously, it was impossible to evaluate the bifurcation point from such
an analysis using the unmodified elements since there was no coup,ling between torsion and
bending in the original element formulation and thus in the stability matrix.
From these runs, it is clear that our new formulation properly accounts for torsional-
bending coupling that arises out of the requirement of rotational invariance. Our nonlinear
analysis proceeded in three steps. First, we marched up the primary path (vertical segment,
Fig. 4) until very near the critical point. Second, we initiated a new bifurcation analysis from
the nonlinear state. Last, we used the lowest (out-of-plane) mode to continue the solution on

2 -2°°V
_4ooV
-600~

-800 | I I I I I I J
-200 -150 -100 -50 0 50 100 150 200
Z C O O R D I N A T E OF APEX
Fig. 4. Applied moment versus Z-displacement of the apex of the hinged right-angle frame.
372 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

200

150

100
x. i
Q. 50
Q.
0
ILl 0

z
-50
rr
0
0 -100
0

-150

-200

-250 I i . I I I I I
-200 -150 -100 -50 0 50 100 150 200
Z COORDINATE OF APEX

Fig. 5. Projection of the trajectory of the apex on the y - z plane.

the secondary path. This last step required the use of the Equivalence Transformation
bifurcation processor [16] recently introduced into STAGS. No perturbation loads or displace-
ments were ever necessary to switch from one solution branch to the other. The plot shown in
Fig. 5 gives the trajectory of the apex of the frame in the y - z plane. Note that the apex is
constrained from moving in the x-direction. In this figure, we show two complete revolutions
of the frame, each cycle of which requires about 130 solution points. In the first cycle, the
frame rotates full-circle, only to return to its initial configuration with the opposite moment
load. Our arc-length solution algorithm permitted the analysis to continue right past the
second (negative) critical point, to repeat the trace in reverse. That is, rotation and translation
at the left end traced the s~:~mecurve backwards to the initial, positively-loaded state, but with
the lateral displacement of opposite sign. In fact, we were able to continue the analysis ad
infinitum, repeating the path cycle for cycle as long as desired. There is complete symmetry of
the load path about the two critical points, each of which is identical in magnitude, a point
discussed fully in [12, 13] It should be noted that full Newton iteration was required for this
tricky problem. Full quadratic convergence was observed, indicating that our linearization was
complete and consistent. Note also the size of the displacements, which are of the same order
as the structure itself.

LENGTH L : 240.0 mm
POLAR MOMENT Ix = 2.16 mm4 ]~
SECOND MOMENTS ly : Iz = 0,0833 mm4
YOUNG'S MODULUS E = 71240 N/mm 2 ~ I~
POISSON'S RATIO v = 0.31 Oy -- 0 ~
SHEAR MODULUS G = 27190 N / m m 2 M ~

"qj ;
Fig. 6. Hockling problem of a cable under torsie':l.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 373

Our second example is an idealization of a cable segment, shown in Fig. 6. This analysis is
an example of hockling [17-19], in which a cable is twisted until its capacity to store energy in
torsion is exceeded. At this point the cable bifurcates into a lower energy bending mode
relieving some of the energy stored in torsion, as can be seen from twisting a garden hose. A
complete segment consisting of 20 equal rod elements was constructed to model the cable in
Fig. 6. The left end is allowed to translate in and rotate about the x axis. The other end is
fixed. The deformed shapes of the cable is shown in Fig. 7. The result in Fig. 8 shows the
variation of the twist angle with the applied load. Initially, the rod twists without any lateral
displacement. A critical point is reached, beyond which the rod becomes unstable. Again, the
equivalence transformation [16] is used to switch to the unstable branch. Here, the rod
exhibits large lateral displacement as the load decreases. The deformed shape of the rod takes
a helical form with the free end pulling in until the cable becomes a perfect circle. The load is
negative at this second critical point. The equivalence transformation is used once again to

H O R I Z O N ~
Fig. 7. Deformed shapes of the cable as it hockles.

250
200
E~, 150
100

o,

~ -50 -

-100

-150
0 1 2 3 4 5 6 7
TWIST ANGLE (rad)
Fig. 8. Variation of the applied torque with the twist angle for the cable hockling problem.
374 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

l- L -I
I '1 I I I I I I 1 I I.I .I I.I I.I . . . . .i i.i i.r l ii ii
~,,1,,

l
LiD i LENGTH L = 255.0mm
]~ i WIDTH w = 30.0ram
IG F THICKNESS t = 0.6mm
~ I YOUNG'S MODULUS E = 71240 N/rnm 2
FINITE-ELEMENTMODELS ~/r~./~/ POISSON'SRATIO v = 0.31
Fig. 9. Clamped right angle frame problem and its FE discretizations.

jump onto a third branch that involves simple untwisting of the cable until the cable turns into
a circular configuration free of torsion. The energy stored in torsion has been transferred into
bending. As before, this analysis depended only on the element-independent modifications
that were used to convert a rod element not designed for this type of analysis to one fully
capable handling the complicated three-dimensional response.
The last example problem is the clamped right angle frame problem of Fig. 9 subjected to
an in-plane point load at the end. The properties and dimensions of this ,~rame are identical to
the first problem. However, the loading and boundary conditions are different and conse-
quently the problem has no symmetry. The existing 410 shell element in the computer
program STAGS was used to construct two different finite element models for the complete
frame (Fig. 9). The coarse mesh model has 17 elements while the fine mesh has 64 elements
(Fig. 10). The stability limits for these models are given in Table 2 and compared to those in
[13, 15, 20]. These results show the improvements that are obtained using the projector
operator.
Once past the stability limit, the structure exhibits a secondary out-of-plane deflection
response. No perturbation load is required to switch to this secondary mode when the

2.5

2.0

"~I.S !

Q
uJ
~. 1.0
o.

0.5 FINE MESH


..... COARSE MESH

0 I I I I I
0 10 20 30 40 50 SO
LATERAL DISPLACEMENT AT LOADED TIP (ram)
Fig. 10. Post-buckling response of the clamped frame for 17 and 64 element mesh.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 375

Table 2
Critical buckling load of the right angle frame
Reference Element # Elements Critical load
[15] Beam 20 1.088
[13] Beam 20 1.090
[15] Triangle shell 86 1.145
[20] Quad. shell 68 1.137
[20] - Cony. 1.128
Present Quad. shell w/o proj. 17 0.881
Present Quad. shell with proj. 17 1.138
Present Quad. shell w/o pro i. 64 1.089
Present Quad. shell with proj. 64 I. 130

equivalence transformation is used. The structure loses most of its stiffness immediately after
it buckles. The stiffness remains small while increasing slightly until the loaded tip undergoes a
deflection of about 50.0 mm. Past this point, there is a stiffening effect due to membrane
contributions that come about as a result of stretching the entire assemblage.

6. Conclusions

In this paper, a systematic "procedure is presented for extending the capabilities of an


existing linear finite element of accommodate finite rotation analysis. The basis of our
approach is the element-independent co-rotational algorithm where the element rigid body
motion (translation and rotations) are separated from the deformational part of its total
motion. This is achieved in the conventional manner by attaching a local frame to the element,
such that the motion of the element with respect to this local frame fully describes the
deformation. The internal strain energy of an element depends only on this deformation and is
independent of the rigid body part of the motion. We used this fact to obtain invariance
conditions that must be satisfied by the element internal force vector and tangent stiffness
matrix. These conditions may be applied in a single element test to validate the element
formulation for large rotation analysis, using a simple projector matrix. When the internal
force vector is in self-equilibrium, the action of the projector has no effect on the internal
forces and moments. However, when self-equilibrium is not satisfied by the basic element
formulation the product of the projector by the internal force vector produces a new set of
forces and moments that are in equilibrium.
The tangent stiffness matrix obtained from a consistent linearization of a co-rotationally
based element is shown to be unsymmetric, where the unsymmetric part depends only on the
unbalanced forces. This fact is shown to result in quadratic rate of convergence when the
symmetric part of the tangent matrix is used in the Newton-Raphson iteration scheme.
The above procedure is used to solve three different example problems whose solutions
exhibit large rotation behavior. Two of these problems were previously analyzed in
[13, 15, 20]. The results of our studies demonstrate that it is possible to analyze structures
undergoing large rotations within an efficient and general framework, using simple and
376 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

economical finite elements. These improvements in the performance of the elements are
brought about through the use of convenient software utilities as pre- and post-processors to
the element routines thus requiring no modification to existing finite element software.

Acknowledgment

We are grateful to Dr. G.M. Stanley for many enlightening technical discussions and
suggestions that helped improve this paper and to Mr. F.A. Brogan for the effective
implementation of this procedure into the STAGS program. This research was supported by
the Independent Research Program of LMSC.

Appendix A. Exponential of spin tensors and its derivative

The exponential of a 3 x 3 skew-symmetric matrix, O =spin(0), was first given by


Rodriguez and takes the form

T=e°=l+~O+
sin O 1 - cos O 02
02 , (A.1)

where 0 = II011, This identity may be obtained using the series definition of the matrix
exponential and a repeated substitution of

(~3 = __,02 0 .
(A.2)

The last equation is the result of the Cayley-Hamilton theorem and is obtained by substituting
O in its characteristic equation. It is interesting to note that the results of Rodriguez in (A. 1)
is a special case of a more general theorem that states that the continuous function of any
n x n matrix may be expressed as a matrix polynomial of order n - 1 in that matrix [21, p. 96].
The coefficients of this polynomial will depend on the spectrum of the matrix. It is interesting
to note that the approximation to the exponential of O used by Hughes and Winget [22] (i.e.,
( I - ½0)-J(l + ½0)) is simply the (1, 1) Pade approximation to the exponential. This is one
of the few approximations to. (A.1) that yields an orthogonal matrix.
The inverse of (A.1) (i.e., log~ T) can be found by forming the anti-symmetric part of T.
Accordingly,

O = log~ T = arcsin(r)
2r axial(T- T t) (A.3)

where ~"is given by

r = ] ] a x i a l ( T - T')[[. (A.4)

A robust algorithm for computing the log of a 3 x 3 orthogonal matrix may be found in [23].
To obtain the derivative of T, we rewrite (A.1) in terms of the unit vector O = 0/O and
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 377

8 = spin 0 as

T = e °° = I + sin 0 0 + (1 - cos 0 ) 8 2 . (A.5)

Then O and # can be treated as independent parameters; one representing the length of 0 and
the other its direction. As a result
dT OT
8T = ~-~ 80 + ~-~" 8 8 . (A.6)

The first term in this equation is simply the differentiation of the exponential with respect to a
scalar parameter where only the magnitude of the spin is allowed to vary. The derivative in
this term may be obtained directly from the exponential form of T. The second term is due to
the variation of the direction of the spin axis. Note that the second term involves the
differentiation with respect to two independent parameters. This term may be evaluated by
differentiating (A.5) directly. Thus

8T = 80 8 T + sin 0 8 8 + (1 - cos O ) ( 8 8 8 + 8 8 8 ) , (A.7)

80 and 8 0 may be obtained by differentiating the identities 02 = Ot0 and O = O 8 to get

0 t 80 88 = 1 0t .
8 0 = ~ , -~--~ ( O 2 8 0 - 800) (A.8)

In (30), the quantity that is of interest is 8 ~ = 8T T t. Differentiating the identity TT' = !


one obtains 8 ~ + [8~]' = 0. Hence, 8• is skew-symmetric. It denotes the instantaneous rate
of change of the rotation of the principle axis associated with T and when the differentiation is
carried out with respect to time, 8 ~ represents the angular velocity tensor. Multiplying 8T
from (A.7) by the transpose of T in (A.5) and using the identity 8 8 8 8 = 0, we obtain

8• =ST T' = 80 8 + sin 0 8 0 + (1 - c o s 0 ) ( 8 8 8 - 8 8 8 ) . (A.9)

This result is obtained directly from (A.5) and (A.7) and using the fact that 8 8 8 8 = 0.
Substituting for 80 and 8 0 from (A.8) in (A.9) one obtains the alternative form for 8,0 given
by

8 n = 0 - sin O Ot 80 O + sin O 1 - cos 0


03 ~80+ O2 (080-800) (A.10)

or, equivalently,
O - sin O sin 0 1 - cos 0
8t~ = axial(ST T t) = 03 O0' 80 + ~ 80 + 02 O 80, (A.1i)

where the identity O 80 = axial(O 8 0 - 8 0 O) is used to simplify (A.10). The results in


(A.11) were obtained at about the same time by Szwabowicz [24] and Simo [12].
378 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

The inverse of (A.11), namely the derivative of 0 in terms of 8to, is also required. To obtain
this relation, we first multiply (A.11) by 0 t resulting in

0 t 80 = 0 t ~ o J . (A.12)

This result is eventually substituted in the first term on the right-hand side of (A.11) to
eliminate its dependence on 80. Multiplying (A.11) by O and 0 2 results in two sets of
equations; one for O 80 and the other for O 2 80. Solving these equations for 8 0 0 results in

O sin O 192
~O 0 = 2(1 - cos 0 ) O 8~o - ½ 8oJ. (A.13)

Substituting the above in the third term on the right-hand side of (A.11) and using the
identity, f18t = 021 + 0 2 in the first term results in the following relation:

A = O0 = i - ½ 0 + r / O 2 (A.14)
0¢~O
where
2 sin 0 - 0(1 + cos 0 ) sin(½ O) - ~ O cos( ½0)
(Ao15)
r/= 202 sin O 02sin(½0)

In (61) we require the derivative of H contracted with the force vector. This in turn can be
computed through the derivative of A in (A. 14) contracted with a vector v. Differentiating the
product of A t and v, keeping v constant, we obtain

OA t 0
0-'-~ : v - ~ [Atv]A. (A.16)

Multiplyi::lg the right-hand side of (A. 14) by v and differentiating the result, the left-hand side
of the aoove equation becomes

OA'
OoJ : v = { - ½spin(v) + •[(O'v)l + Ov t - 2vO t] + I z O 2 v O t } A , (A.17)

where ~ depends on the derivative of r/and is expressed as

r/' 0 ( 0 + sin O ) - 8 sin2(½0)


/z = - ~ - = 404 sin2(~O) (A.18)

Where O is small, 17 and tz are computed from a truncated power series expansion. The axial
vector of the anti-symmetric part of the matrix in (A.17) is - ½ A v . In the case where v
corresponds to the internal moments at a node, one can observe that the assembled
contribution of this anti-symmetric matrix to the stiffness vanishes at equilibrium. This is the
same property demonstrated on the global level in Section 3.3.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent finearization 379

Appendix B. F for shells and beams

In this appendix we give an explicit expression for the infinitesimal motion of an element,
F, described in terms of the translations and rotations of the element nodes when a local
frame attached to the element undergoes an orthogonal transformation. This transformation
depends on the rotation of the basis vectors of E (i.e., the columns of the transformation from
the element local coordinates to global coordinates). A complete derivation may be found in
[9]. F depends on the current position of the nodes and the choice of the local element
coordinate system. Therefore, a separate expression must be obtained for each element (i.e.,
triangular, quadrilateral, etc.). An appropriate choice of the local frame can simplify the terms
in F. In the following the origin of the element local coordinates is chosen to be at node 1 of
the elements. The directions of the axis in the local frame was chosen separately for each
element. It should be noted that r is a function of the local deformational displacements. The
transformation to any other frame involves simple operations involving orthogonal matrices.

B. 1. Triangular shell elements


We choose the unit basis vector e 2 of E along the edge connecting nodes 1 and 3. That is,

= (B.1)

For details see Fig. B.1. Note that x ag is the global coordinates of node a in the deformed
configuration, and x~ is the same in local coordinates with the origin at node 1. Choosing the
third coordinate axis orthogonai to the plane through the three nodes of the triangle, we
obtain e~ as
e
e 3 = x 2 x x~,12A, (a.2)

where A is the area of the element given by A = ½Ix2 × x~l.


With the above choice of local element coordinate the infinitesimal rotation, F, becomes

e3

1 2

Fig, B.1. Local coordinate system for the triangular element.


380 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

1 1
0 0 2 0 0 0 0 0
/~3 ,'~'3

X2 1 ,~2
0 0 --~ 1- -~ 03. 0 0 1 0~ 0 0 i 2 03
X2 X3 X2 X2X3

1 1
--~ 0 0 0 0 0 20 0
_X3 X3 i

D
1 m
0 0 -0 0 -1 0 0 0 0 0 1
X3
2
X2 1
12 2 0 0 0 1 03 0 0 -1 03 0 0 1 03 (B.3)
X2X3 X3
1
0 0 --~ _10 0 0 0 0 -1 0 0
X3
where xi~ denotes the i-th component of the position vector of node a and 03 is the 3 x 3 matrix
of zeros corresponding to the rotational freedoms. Earlier, we assumed the elements to be
invariant under translations. The translational invariance is preserved in F and results in zero
column sums. Note that F in (B.3) and • in (32) satisfy the bi-orthonormality relation

~ t F = Ft~rt = 13 • (B.4)

The projector property of P is due to this bi-orthonormality condition.

B.2. Quadrilateral shell elements


In order to construct the local element frame for quadrilaterals, we choose the third
direction e 3 orthogonal to the two diagonals of the quadrilaterals (see Fig. B.2). The vectors
t~ e ~ e
along the diagonals are x 4 - x 2 and x 3 - x~. Since the origin of the frame is chosen to be at
node I, x~ is identically zero. Thus,
e
= x xD/2a , (a.5)
w h e r e a = ½[Xa(X4
1 2 - g~) + X3X2]
2 t is the normalizing factor and la] is the area of the element
projected on the plane normal to • 3. We also have

el = x4 x Ca//3, (a.6)
where/3 = X4z, and e 2 is obtained by completing a right-hand system and is given by

e2 = e3 x e 1 . (B.7)

With the above choice of local frame/-' takes the form


B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 381

e2

e3

Fig. B.2. Local coordinate system for the quadrilateral element.

F=Y.~, (B.8)
where
! 0 -1 0 0 0 0 0 1 0 0 0 ]
yt= 0 00300-10300003 00103
0 0 O0 0 000 -100
and
[X~i2a _(X 2 - X2)12a -X4(X4
3 2_ X2,)12a,O"
~-= X~ 2a X~12a 32
X4X3/2a~]
0

For elements with more than 4 nodes, the components of F associated with the first four
nodes are identical to those in (B.8). The contributions belonging to the additional nodes will
be identically zero. Again, F in (B.8) satisfies the bi-orthonormality relation (B.4), while
retaining its translational invariance.

B.3. Beam elements


The definition of the local element frame for a beam is more involved in three dimensions
(see Fig. B.2). The first coordinate axis is chosen to lie along the line joining the two beam

e2

Fig. B.3. Local coordinate system for the beam element.


382 B. Nour-Omid, C.C. Ranldn, Finite rotation analysis and consistent iinearization

nodes. That is,


e, = Et(xg2 - x ~ ) / l x ~ - xgl = x /Ix l , (B.9)

In general, the remaining two axes may be chosen arbitrarily in the plane orthogonai to e t,
and are often chosen based on other criteria, such as the orientation of the cross-section. To
see how the element frame evolves, we introduce a vector, q, that moves with node one and
initially is coincident with e 2. q is rigidly attached to the nodal frame and changes its
orientation as the element deforms. With this coordinate system the resulting F take the form

r= + r.---, (B.10)
where
r 03t
i 0 0 0 -1 0 0 0 0 /ele l
Y'= 0 1 0 0 0 0 0 -1 03 , Y = / 03
-1 0 0 0 0 0 1 0 [. 03

and

,l li,] with r/= q~


q2

Here qj are the components of q. Note that if the beam is so deformed that q lies along the
first coordinate axis, then q2 becomes zero and rt--* inf. To avoid such a case the algorithm
redefines q to a vector that is not along e~. When this occurs, the expression for F remains the
same. Only r/changes and may become zero. Observe that the bi-orthonormality relations
between ~ and F hold. The principal difference here is that the variation of the element
frame is no longer independent of the rotations. This dependence occurs through the variation
o f the surface triad.

Appendix C. Nomenclature

a,b Node number indices.


axial ( ) The function that forms a vector from a spin tensor.
b Vector of components along rigid body modes.
d Vector of nodal unknowns (displacements and rotations).
di The i-th components of d.
Deformational part of displacements and rotations.
a, The i-th components of d.
e Superscript that denotes quantities in local element coordinates.
Eo Initial orthogonal transformation from the global to the local frame.
E Current orthogonal transformation from the global to the local frame.
oJE Instantaneous axis of rotation associated with E.
I_ The internal force vector.
f The unmodified internal force vector.
B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization 383

f, The i-th components off.


F,F,F Matrices that depend on internal force vector.
g Superscript that denotes quantities in global coordinates.
G The diagonal transformation matrix of E.
H The Jacobian transformation from instantaneous to total rotations.
J The Jacobian transformation from deformational to total displacements and rota-
tions.
J:b Sub-matrix of J relating nodes a and b.
K The tangent stiffness matrix.
The unmodified tangent stiffness matrix.
gs The symmetric part of the tangent stiffness matrix.
KA The antisymmetric part of the tangent stiffness matrix.
P Projector matrix.
Pm The m-th row of P.
P~j The components of P.
r The remainder vector in a Taylor series expansion.
spin( ) The function that forms a spin tensor from a vector.
T Orthogonal rotation matrix.
Deformational part of the orthogonal rotation matrix.
u Displacement vector.
uo Displacement vector of the origin of the local frame.
t~ Deformational part of the displacement vector.
X Initial position vector.
Xo Initial position vector of the origin of the local frame.
x Current position vector.
Xo Current position vector of the origin of the local frame.
F Rate of change in the instantaneous spin axis of the local frame with respect to the
nodal unknowns.
8a Infinitesimal change in a.
Strain energy functional.
Linearly independent set of rigid body modes.
0 Axial vector associated with O.
Axial vector associated with O.
Natural logarithm of T.
Natural logarithm of ]F.
to Axis of the instantaneous rotation associated with T.
03 Axis of the instantaneous rotation associated with T.
The spin tensor constructed from components of to.
The spin tensor constructed from components of o).

References

[1] T. Belytschko and B.J. Hsieh, Non-linear transient finite elements analysis with convected co-ordinates,
Internat. J. Numer. Methods Engrg. 7 (1973) 255-271.
384 B. Nour-Omid, C.C. Rankin, Finite rotation analysis and consistent linearization

[2] C.C. Rankin and F.A. Brogan, An element-independent co-rotational procedure for the treatment of large
rotations, ASME J. Pressure Vessel Techn. 108 (1986) 165-174.
[3] G. Wempner, Finite elements, finite rotations and small strains of flexible shells, Internat. J. Solids and
Structures 5 (1969) 117-153.
[4] T. Belytschko and L. Schwer, Large displacement transient analysis of space frames, Internat. J. Numer.
Methods Engrg. 11 (1977) 65-84.
[5] G. Horrigmoe and EG. Bergen, Instability analysis of free-form shells by flat finite elements, Comput.
Methods Appl. Mech. Engrg. 16 (1978) 11-35.
[6] J. Argyris, An excursion into large rotations, Comput. Methods Appl. Mech. Engrg. 32 (1982) 85-155.
[7] B.F. De Veubeke, The dynamics of flexible bodies, Internat. J. Engrg. Sci. 14 (1976) 895-913.
[8] M.K. Nygard and P.G. Bergan, Advances in treating large rotations for nonlinear problems, in: A.K. Noor
and J.T. Oden, eds., State of the Art Surveys on Computational Mechanics (ASME, New York, 1989).
[9] C.C. Rankin and B. Nour-Omid, The use of projectors to improve finite element performance, Comput. &
Structures 30 (1988) 257-267.
[10] T. Belytsehko, H. Stolarski and N. Crapenter, A C" triangular plate element with one-point quadrature,
Internat. J. Numer. Methods Engrg. 20 (1984) 787-802.
[11] M.A. Crisfield, A consistent co-rotational formulation for non-linear three-dimensional, beam elements,
Comput. Methods Appl. Mech. Engrg. 81 (1990) 131-150.
[12] J.C. Simo, A finite strain beam formulation. The three-dimensional dynamic problem. Part I, Comput.
Methods Appl. Mech. Engrg. 49 (1985) 55-70.
[13] J.C. Simo and L. Vu-Quoc, A three-dimensional finite strain rod model. Part If: Computational aspects,
Comput. Methods Appl. Mech. Engrg. 58 (1986) 79-116.
[14] B.O. Almroth, F.A. Brogan and G.M. Stanley, Structural analysis of general shells, Vol. II, User's
instructions for STAGSC-1, Lockheed Report LMSC-D633873, Lockheed Palo Alto Research Laboratory,
Palo Alto, CA, 1979.
[15] J.H. Argyris, H. Balmer, J.St. Doltsinis, EC. Dunne, M. Haase, M. Kleiber, G.A. Malejannakis, H.-E
Mlejenek, M. Miiller and D.W. Scharpf, Finite element method- the natural approach, Comput. Methods
Appl. Mech. Engrg. 17/18 (1979) 1-106.
[16] G.A. Thurston, F.A. Brogan and E Stehlin, Postbuckling analysis using a general purpose code, AIAA J. 24
(1986) 1013-1020.
[17] A.G. Greenhill, Proc. Inst. Mech. Engineers, London, 1883.
[18] F. Rosenthal, Greenhill's formula and the mechanics of cable hockling, NRL Report 7940, Naval Research
Laboratory, Washington, DC, 1975.
[19] R.V. Southwell, An Introduction to the Theory of Elasticity (Dover, New York, 1969).
[20] J.C. Simo, D.D. Fox and M.S. Rifai, On a stress resultant geometrically exact shell model. Part III:
Computational aspects of the nonlinear theory, Comput. Methods Appl. Mech. Engrg. 79 (1990) 21-70.
[21] Gantmacher, Matrix Theory, Vol. II (Chelsea, New York, 1977).
[22] T.J.R. Hughes and J. Winget, Finite rotation effects in numerical integration of rate constitutive equations
arising in largc ,eformation analysis, Internat. J. Numer. Methods Engrg. 15 (1980) 1413-1418.
[23] R.A. Spurrier, Comment on singularity-free extraction of a quaternion from a direction-cosine matrix, J.
Spacecraft 15 (4) (1978) 255.
[24] M.L. Szwabowicz, Variational formulation in the geometrically non-linear thin elastic shell theory, Internat. J.
Solids and Structures 22 (1986) 1161-1175.
[25] K.J. Bathe and S. Bolourchi, Large displacement analysis of three dimensional beam structures, Internat. J.
Numer. Methods Engrg. 14 (1979) 961-986.
[26] A. Cardona and M. Geradin, A beam element non-linear theory ith finite rotations, lnternat. J. Numer.
Methods Engrg. 26 (1988) 2403-2438.
[27] J.G. Simmonds and D.P. Danielson, Nonlinear shell theory with finite rotation and stress function vectors,
ASME J. Appl. Mech. 39 (94) (1972) 1085-1095.

You might also like