Вы находитесь на странице: 1из 51

CU-CAS-05-02

CENTER FOR AEROSPACE STRUCTURES

Unied Formulation of Small-Strain Corotational Finite Elements: I. Theory

by C. A. Felippa and B. Haugen

January 2005

COLLEGE OF ENGINEERING UNIVERSITY OF COLORADO CAMPUS BOX 429 BOULDER, COLORADO 80309

Unied Formulation of Small-Strain Corotational Finite Elements: I. Theory

C. A. Felippa and B. Haugen

Department of Aerospace Engineering Sciences and Center for Aerospace Structures University of Colorado Boulder, Colorado 80309-0429, USA Email: carlos.felippa@colorado.edu FEDEM Technology Inc 2933 5th St., Boulder, CO 80304, USA Email: bjorn.haugen@fedem.co

Report No. CU-CAS-05-02 April 2004, revised January 2005

Contributed to Computer Methods in Applied Mechanics and Engineering for the Special Issue on Shells, edited by E. Ramm, M. Papadrakakis and W. A. Wall. To appear 2005.

A Unied Formulation of Small-Strain Corotational Finite Elements: I. Theory


Carlos A. Felippa and Bjorn Haugen
Department of Aerospace Engineering Sciences and Center for Aerospace Structures University of Colorado, CB 429 Boulder, CO 80309-0429, USA Email: carlos.felippa@colorado.edu Web page: http://titan.colorado.edu/Felippa.d/FelippaHome.d/Home.html FEDEM Technology Inc 2933 5th St. Boulder, CO 80304, USA Email: bjorn.haugen@fedem.co

Abstract This article presents a unied theoretical framework for the corotational (CR) formulation of nite elements in geometrically nonlinear structural analysis. The key assumptions behind CR are: (i) strains from a corotated conguration are small while (ii) the magnitude of rotations from a base conguration is not restricted. Following a historical outline the basic steps of the Element Independent CR formulation are presented. The element internal force and consistent tangent stiffness matrix are derived by taking variations of the internal energy with respect to nodal freedoms. It is shown that this framework permits the derivation of a set of CR variants through selective simplications. This set includes some previously used by other investigators. The different variants are compared with respect to a set of desirable qualities, including self-equilibrium in the deformed conguration, tangent stiffness consistency, invariance, symmetrizability, and element independence. We discuss the main benets of the CR formulation as well as its modeling limitations. Key words: geometrically nonlinear structural analysis, corotational description, shell nite elements

TABLE OF CONTENTS
Page

1. 2.

3.

4.

5.

6.

7. A. A.1. A.2. A.3. A.3. A.4. A.5. A.6.

Introduction The Emergence of CR 2.1. Continuum Mechanics Sources . . . . . . . . . . 2.2. FEM Sources . . . . . . . . . . . . . . . . 2.3. Shadows of the Past . . . . . . . . . . . . . . 2.4. Linking FEM and CR . . . . . . . . . . . . . 2.5. Element Independent CR . . . . . . . . . . . . Corotational Kinematics 3.1. Congurations . . . . . . . . . . . . . . . 3.2. Coordinate Systems . . . . . . . . . . . . . . 3.3. Coordinate Transformations . . . . . . . . . . . 3.4. Rigid Displacements . . . . . . . . . . . . . . 3.5. Rotator Formulas . . . . . . . . . . . . . . 3.6. Degrees of Freedom . . . . . . . . . . . . . . 3.7. EICR Matrices . . . . . . . . . . . . . . . 3.8. Deformational Translations . . . . . . . . . . . . 3.9. Deformational Rotations . . . . . . . . . . . . Internal Forces 4.1. Force Transformations . . . . . . . . . . . . . 4.2. Projector Properties . . . . . . . . . . . . . . Tangent Stiffness 5.1. Denition . . . . . . . . . . . . . . . . . 5.2. Material Stiffness . . . . . . . . . . . . . . 5.3. Geometric Stiffness . . . . . . . . . . . . . . 5.4. Consistency Verication . . . . . . . . . . . . Three Consistent CR Formulations 6.1. Consistent CR formulation (C) . . . . . . . . . . 6.2. Consistent Equilibrated CR Formulation (CE) . . . . . 6.3. Consistent Symmetrizable Equilibrated CR Formulation (CSE) 6.4. Formulation Requirements . . . . . . . . . . . 6.5. Limitations of the EICR Formulation . . . . . . . . Conclusions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . . . . . . . . . . .

. . . .

. . . . . . . . . . . . . . . . . . . . . .

1 2 2 2 2 4 4 5 5 7 8 9 9 10 11 13 14 15 15 16 17 17 17 18 19 19 19 20 20 20 21 22 24 24 24 25 25 26 27 27 27 28 28

THE MATHEMATICS OF FINITE ROTATIONS Spatial Rotations . . . . . . . . . . . . . . . Spinors . . . . . . . . . . . . . . Spin Tensor and .Axial.Vector . . . . . . . . . . . . . Spinor Normalizations . . . . . . . . . . . . . . Spectral Properties . . . . . . . . . . . . . . . From Spinors To Rotators . . . . . . . . . . . . . . Rotator Parametrizations . . . . . . . . . . . . . . . A.6.1. Rotator From Algebra . . . . . . A.6.2. Rotator from Geometry . . . . . . A.7. Rotators for .All Seasons . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . .

A.8. A.9. A.10. A.11. A.12. A.13. A.14.

B. B.1. B.2.

C. C.1. C.2. C.3. C.4. C.5. D.

The Cayley. Transform. . . . . . . . . . . . . . . . . Exponential .Map . . . . . . . . . . . . . . . . Skew-Symmetric. Matrix . . . . . . . . . . . . . . . . . Relations From Rotators .to Spinors . . . . . . . . . . . . . . . . Spinor and.Rotator.Transformations. . . . . . . . . . . . . . . . Axial Vector.Jacobian . . . . . . . . . . . . . . . . Spinor and.Rotator.Differentiation . . . . . . . . . . . . . . . . A.14.1. Angular Velocities . . . . . . . . . . . A.14.2. Angular Accelerations . . . . . . . . . . A.14.3. Variations . . . . . . . . . . . . . CR MATRICES FOR TRIANGULAR SHELL ELEMENT Matrix S . . . . . . . . . . . . . . . . . . Matrix G . . . . . . . . . . . . . . . . . B.2.1. G by Side Alignment . . . . . . . . . . B.2.2. G by Least Square Angular Fit . . . . . . . B.2.3. Fit According to CST Rotation . . . . . . . B.2.4. Best Fit by Minimum LS Deformation . . . . BEST FIT CR FRAME Minimization .Conditions. . . . . . . . . . . . . . . . . Best Origin . . . . . . . . . . . . . . . . . Best Rotator . . . . . . . . . . . . . . . . . . Linear Triangle Best .Fit. . . . . . . . . . . . . . . . Rigid Body. Motion .Verication . . . . . . . . . . . . . . . . NOMENCLATURE References . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . .

. . . . . . .

. . . . . .

29 29 30 30 31 31 32 32 32 33 34 34 34 34 35 35 36 36 37 37 37 38 38 39 43

1. Introduction Three Lagrangian kinematic descriptions are in present use for nite element analysis of geometrically nonlinear structures: (1) Total Lagrangian (TL), (2) Updated Lagrangian (UL), and (3) Corotational (CR). The CR description is the most recent of the three and the least developed one. Unlike the others, its domain of application is limited by a priori kinematic assumptions: Displacements and rotations may be arbitrarily large, but deformations must be small. (1)

Because of this restriction, CR has not penetrated the major general-purpose FEM codes that cater to nonlinear analysis. A historical sketch of its development is provided in Section 2. As typical of Lagrangian kinematics, all descriptions: TL, UL and CR, follow the body (or element) as it moves. The deformed conguration is any one taken during the analysis process and need not be in equilibrium during a solution process. It is also known as the current, strained or spatial conguration in the literature, and is denoted here by C D . The new ingredient in the CR description is the splitting or decomposition of the motion tracking into two components, as illustrated in Figure 1. 1. The base conguration C0 serves as the origin of displacements. If this happens to be one actually taken by the body at the start of the analysis, it is also called initial or undeformed. The name material conguration is used primarily in the continuum mechanics literature. The corotated conguration C R varies from element to element (and also from node to node in some CR variants). For each individual element, its CR conguration is obtained through a rigid body motion of the element base conguration. The associated coordinate system is Cartesian and follows the element like a shadow or ghost, prompting names such as shadow and phantom in the Scandinavian literature. Element deformations are measured with respect to the corotated conguration.

Corotated C R Deformed (current, D spatial) C Motion splits into deformational and rigid Base (initial, undeformed, material) configuration C 0

2.

Deformational motion

Rigid body motion Global frame


Figure 1. The CR kinematic description. Deformation from corotated to deformed (current) conguration grossly exaggerated for visibility.

In static problems the base conguration usually remains xed throughout the analysis. In dynamic analysis the base and corotated congurations are sometimes called the inertial and dynamic reference congurations, respectively. In this case the base conguration may move at uniform velocity (a Galilean inertial system) following the mean trajectory of an airplane or satellite. From a mathematical standpoint the explicit presence of a corotated conguration as intermediary between base and current is unnecessary. The motion split may be exhibited in principle as a multiplicative decomposition of the displacement eld. The device is nonetheless useful to teach not only the physical meaning but to visualize the strengths and limitations of the CR description.

2. The Emergence of CR The CR formulation represents a conuence of developments in continuum mechanics, treatment of nite rotations, nonlinear nite element analysis and body-shadowing methods. 2.1. Continuum Mechanics Sources In continuum mechanics the term corotational (often spelled co-rotational) appears to be rst mentioned in Truesdell and Toupins inuential exposition of eld theories [81, Sec. 148]. It is used there to identify Jaumanns stress ux rate, introduced in 1903 by Zaremba. By 1955 this rate had been incorporated in hypoelasticity [82] along with other invariant ux measures. Analogous differential forms have been used to model endochronic plasticity [85]. Models labeled co-rotational have been used in rheology of non-Newtonian uids; cf. [17,78]. These continuum models place no major restrictions on strain magnitude. Constraints of that form, however, have been essential to make the idea practical in nonlinear structural FEA, as discussed below. The problem of handling three-dimensional nite rotations in continuum mechanics is important in all Lagrangian kinematic descriptions. The challenge has spawned numerous publications, for example [1,4,5,34,42,43,62,71,72]. For use of nite rotations in mathematical models, particularly shells, see [60,70,77]. There has been an Euromech Colloquium devoted entirely to that topic [61]. The term corotational in a FEM paper title was apparently rst used by Belytschko and Glaum [8]. The survey article by Belystchko [9] discusses the concept from the standpoint of continuum mechanics. 2.2. FEM Sources In the Introduction of a key contribution, Nour-Omid and Rankin [54] attribute the original concept of corotational procedures in FEM to Wempner [86] and Belytschko and Hsieh [7]. The idea of a CR frame attached to individual elements was introduced by Horrigmoe and Bergan [39,40]. This activity continued briskly under Bergan at NTH-Trondheim with contributions by Kr keland [46], a Nyg rd [15,56,57], Mathisen [48,49], Levold [47] and Bjrum [18]. It was summarized in a 1989 review a article [56]. Throughout this work the CR conguration is labeled as either shadow element or ghostreference. As previously noted the device is not mathematically necessary but provides a convenient visualization tool to explain CR. The shadow element functions as intermediary that separates rigid and deformational motions, the latter being used to determine the element energy and internal force. However the variation of the forces in a rotating frame was not directly used in the formation of the tangent stiffness, leading to a loss of consistency. Criseld [2224] developed the concept of consistent CR formulation where the stiffness matrix appears as the true variation of the internal force. An approach blending the TL, UL and CR descriptions was investigated in the mid-1980s at Chalmers [5052]. In 1986 Rankin and Brogan at Lockheed introduced [63] the concept of element independent CR formulation or EICR, which is further discussed below. The formulation relies heavily on the use of projection operators, without any explicit use of shadow congurations. It was further rened by Rankin, Nour-Omid and coworkers [54,6467], and became essential part of the nonlinear shell analysis program STAGS [68]. The thesis of Haugen on nonlinear thin shell analysis [37] resulted in the development of the formulation discussed in this article. This framework is able to generate a set of hierarchical CR formulations. The work combines tools from the EICR (projectors and spins) with the shadow element concept and assumed strain element formulations. Spins (instead of rotations) are used as incremental nodal freedoms. This simplies the EICR front end and facilitates attaining consistency. Battini and Pacoste at KTH-Stockholm [2,3,58] have recently used the CR approach, focusing on stability applications. The work by Teigen [79] should be cited for the careful use of offset nodes linked to element nodes by eccentricity vectors in the CR modeling of prestressed reinforced-concrete members. 2

Corotated (a.k.a. dynamic reference) frame

PR

CR P Corotated (shadow) configuration C R C Deformed configuration C D (Shown separate from C R for visualization convenience)

P0
Base (a.k.a. inertial) frame

C0

Base (initial) configuration C 0

Figure 2.

The concept of separation of base (a.k.a. inertial) and CR (a.k.a. dynamic) congurations in aircraft dynamics. Deformed conguration (with deformations grossly exaggerated) shown separate from CR conguration for visibility. In reality points C and C R coincide.

2.3. Shadows of the Past The CR approach has also roots on an old idea that preceeds FEM by over a century: the separation of rigid body and purely deformational motions in continuum mechanics. The topic arose in theories of small strains superposed on large rigid motions. Truesdell [80, Sec. 55] traces the subject back to Cauchy in 1827. In the late 1930s Biot advocated the use of incremental deformations on an initially stressed body by using a truncated polar decomposition. However this work, collected in a 1965 monograph [16], was largely ignored as it was written in an episodic manner, using full notation by then out of fashion. A rigurous outline of the subject is given in [83, Sec. 68] but without application examples. Technological applications of this idea surged after WWII from a totally different quarter: the aerospace industry. The rigid-plus-deformational decomposition idea for an entire structure was originally used by aerospace designers in the 1950s and 1960s in the context of dynamics and control of orbiting spacecraft as well as aircraft structures. The primary motivation was to trace the mean motion. The approach was systematized by Fraeijs de Veubeke [25], in a paper that essentially closed the subject as regards handling of a complete structure. The motivation was clearly stated in the Introduction of that article, which appeared shortly before the authors untimely death:
The formulation of the motion of a exible body as a continuum through inertial space is unsatisfactory from several viewpoints. One is usually not interested in the details of this motion but in its main characteristics such as the motion of the center of mass and, under the assumptions that the deformations remain small, the history of the average orientation of the body. The last information is of course essential to pilots, real and articial, in order to implement guidance corrections. We therefore try to dene a set of Cartesian mean axes accompanying the body, or dynamic reference frame, with respect to which the relative displacements, velocities or accelerations of material points due to the deformations are minimum in some global sense. If the body does not deform, any set of axes xed into the body is of course a natural dynamic reference frame.

Clearly the focus of this article was on a whole structure, as illustrated in Figure 2 for an airplane. This will be called the shadowing problem. A body moves to another position in space: nd its mean rigid body motion and use this information to locate and orient a corotated Cartesian frame. Posing the shadowing problem in three dimensions requires fairly advanced mathematics. Using two best t criteria Fraeijs de Veubeke showed that the origin of the dynamic frame must remain at the center of mass of the displaced structure: C R in Figure 2. However, the orientation of this frame leads to 3

Deformed (bar)

Deformed (beam)

Deformed (membrane)

Deformed (shell)

Base (a) (b)

;; ; ;; ;;
Base

Figure 3.

Geometric tracking of CR frame: (a) Bar or beam element in 2D; (b) Membrane or shell element in 3D.

an eigenvalue problem that may exhibit multiple solutions due to symmetries, leading to non uniqueness. (This is obvious by thinking of the polar and singular-value decompositions, which were not used in that article.) That this is not a rare occurrence is demonstrated by considering rockets, satellites or antennas, which often have axisymmetric shape.
Remark 1. Only C (shown in darker shade in Figure 2) is an actual conguration taken by the pictured aircraft 0 R structure. Both reference congurations C and C are virtual in the sense that they are not generally occupied by the body at any instance. This is in contrast to the FEM version of this idea.
D

2.4. Linking FEM and CR The practical extension of Fraeijs de Veubekes idea to geometrically nonlinear structural analysis by FEM relies on two modications: 1. 2. Multiple Frames. Instead of one CR frame for the whole structure, there is one per element. This is renamed the CR element frame. Geometric-Based RBM Separation. The rigid body motion is separated directly from the total element motion using elementary geometric methods. For example in a 2-node bar or beam one axis is dened by the displaced nodes, while for a 3-node triangle two axes are dened by the plane passing through the points. See Figure 3.

The rst modication is essential to success. It helps to fulll assumption (1): the element deformational displacements and rotations remain small with respect to the CR frame. If this assumption is violated for a coarse discretization, break it into more elements. Small deformations are the key to element reuse in the EICR discussed below. If intrinsically large strains occur, however, the breakdown prescription fails. In that case CR offers no advantages over TL or UL. The second modication is inessential. Its purpose is to speed up the implementation of geometrically simple elements. The CR frame determination may be rened later, using more advanced tools such as polar decomposition and best-t criteria, if warranted.
Remark 2. CR is ocassionally confused with the convected-coordinate description of motion, which is used in branches of uid mechanics and rheology. Both may be subsumed within the class of moving coordinate kinematic descriptions. The CR description, however, maintains orthogonality of the moving frame(s) thus achieving an exact decomposition of rigid-body and deformational motions. This property enhances computational efciency as transformation inverses become transposes. On the other hand, convected coordinates form a curvilinear system that ts the change of metric as the body deforms. The difference tends to disappear as the discretization becomes progressively ner, but the fact remains that the convected metric must encompass deformations. Such deformations are more important in solid than in uid mechanics (because classical uid models forget displacements). The idea nds more use in UL descriptions, in which the individual element metric is updated as the motion progresses.

Assembler

Global element equations

Incorporate rigid body motions

Deformational element equations

Form element mass, stiffness & forces

System equations of motion


Extract deformational motions Evaluate element stresses

Solver Total displacements

Deformational displacements

CR "Filters"

Finite Element Library

Figure 4.

The EICR as a modular interface to a linear FEM library. The owchart is mainly conceptual. For computational efciency the interface logic may be embedded with each element through inlining techniques.

2.5. Element Independent CR As previously noted, one of the sources of the present work is the element-independent corotational (EICR) description developed by Rankin and coworkers [54,6367]. Here is a summary description taken from the Introduction to [54]:
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 element type (except for slight distinctions between beams, triangular and quadrilateral elements).

Why is the EICR worth study? The question ts in a wider topic: why CR? That is, what can CR do that TL or UL cannot? The topic is elaborated in the Conclusions section, but we advance a practical reason: reuse of smallstrain elements, including possibly materially nonlinear elements. The qualier element independent does not imply that the CR equations are independent of the FEM discretization. Rather it emphasizes that the key operations of adding and removing rigid body motions can be visualized as a front end lter that lies between the assembler/solver and the element library, as sketched in Figure 4. The lter is purely geometric. For example, suppose that a program has four different triangular shell elements with the same node and degree-of-freedom conguration. Then the front end operations are identical for all four. Adding a fth small-strain element of this type incurs relatively little extra work to make it geometrically nonlinear. This modular organization is of interest because it implies that the element library of an existing FEM program being converted to the CR description need not be drastically modied, as long as the analysis is conned to small deformations. Since that library is typically the most voluminous and expensive part of a production FEM code, element reuse is a key advantage because it protects a signicant investment. For a large-scale commercial code, the investment may be thousands of man-years. Of course modularity and computational efciency can be conicting attributes. Thus in practice the front end logic may be embedded with each element through techniques such as code inlining. If so the owchart of Figure 4 should be interpreted as conceptual. 3. Corotational Kinematics This section outlines CR kinematics of nite elements, collecting the most important relations. Mathematical derivations pertaining to nite rotations are consigned to Appendix A. The presentation assumes static analysis, with deviations for dynamics briey noted where appropriate. 5

Table 1

Congurations in Nonlinear Static Analysis by Incremental-Iterative Methods


Equilibrium Required? No No No Identication

Name Generic Perturbed Deformed

Alias Admissible

Explanation A kinematically admissible conguration Kinematically admissible variation of a generic conguration.

C C + C CD

Current Spatial

Actual conguration taken during the analysis process. Contains others as special cases.

Base

Initial The conguration dened as the Undeformed origin of displacements. Material Conguration to which computations are referred Conguration taken at the k th iteration of the n th increment step Equilibrium conguration accepted at the n th increment step Body or element-attached conguration 0 obtained from C through a rigid body motion (CR description only) Corotated conguration forced to align with the global axes. Used as connector in explaining the CR description.

Yes

C0

Reference Iterated Target Corotated Shadow Ghost Globallyaligned

TL,UL: Yes. TL: C , UL: C , R 0 R 0 CR: C no, C yes CR: C and C


0 n1

No Yes No

Cn k Cn CR

Connector

No

CG

C0 is often the same as the natural state in which body (or element) is undeformed and stress-free.

Used only in Part II [38] in the description of solution procedures. 0 R In dynamic analysis C and C are called the inertial and dynamic-reference congurations, respectively, when they apply to the entire structure.

3.1. Congurations To describe Lagrangian kinematics it is convenient to introduce a rich nomenclature for congurations. For the readers convenience those used in geometrically nonlinear static analysis using the TL, UL or CR descriptions are collected in Table 1. Three: base, corotated and deformed, have already been introduced. Two more: iterated and target, are connected to the incremental-iterative solution process covered in Part II [38]. The generic conguration is used as placeholder for any kinematically admissible one. The perturbed conguration is used in variational derivations of FEM equations. Two remain: reference and globally-aligned. The reference conguration is that to which element computations are referred. This depends on the description chosen. For Total Lagrangian (TL) the reference is base conguration. For Updated Lagrangian (UL) it is the converged or accepted solution of the previous increment. For corotational (CR) the reference splits into CR and base congurations. The globally-aligned conguration is a special corotated conguration: a rigid motion of the base that makes the body or element align with the global axes introduced below. This is used as a connector 6

C is element centroid in statics, Deformed (current) Base udP but center of mass in dynamics. (initial, R P P R xPC undeformed) uP CR C R 0 _ uP P0 xPC xP x2 c R C0 xP X2,x2 0 xP b Corotated a ~ ~ O x1 ~ //x 1 //x 2 X ,x ~ _ x2 global frame 1 1 x 1 element element (with material & rigid body base frame CR frame spatial coalesced) rotation

Figure 5.

CR element kinematics, focusing on the motion of generic point P. Two-dimensional kinematics pictured for visualization convenience.

R (depends on P) P0 C0
~ x
1

deformational rotation Rd (a "drilling rotation" in 2D)

c
T T0

CR _ C

x2

X2,x2

a
~ x
2

T TR

element base frame

O global X1,x1
frame

x1

element _ CR frame

R0

Figure 6.

CR element kinematics, focusing on rotational transformation between frames.

device to teach the CR description, and does not imply the body ever occupies that conguration. The separation of rigid and deformational components of motion is done at the element level. As noted previously, techniques for doing this have varied according to the taste and background of the investigators that developed those formulations. The approach covered here uses shadowing and projectors. 3.2. Coordinate Systems A typical nite element, undergoing 2D motion to help visualization, is shown in Figure 5. This diagram as well as that of Figure 6 introduces kinematic quantities. For the most part the notation follows that used by Haugen [37], with subscripting changes. Congurations taken by the element during the response analysis are linked by a Cartesian global frame, to which all computations are ultimately referred. There are actually two such frames: the material global frame with axes {X i } and position vector X, and the spatial global frame with axes {xi } and position vector x. The material frame tracks the base conguration whereas the spatial frame tracks the CR and deformed (current) congurations. The distinction agrees with the usual conventions of dual-tensor continuum mechanics [81, Sec. 13]. Here both frames are taken to be identical, since for small strains nothing is gained by separating them (as is the case, for example, in the TL description). Thus only one set of global axes, with dual labels, is drawn in Figures 5 and 6. Lower case coordinate symbols such as x are used throughout most of the paper. Occasionally it is convenient for clarity to use upper case coordinates for the base conguration, as in Appendix C. The global frame is the same for all elements. By contrast, each element e is assigned two local Cartesian frames, one xed and one moving: {xi } The element base frame (blue in Figure 5). It is oriented by three unit base vectors ii0 , which are 7

(a)
~ x2

~ x1

(b) C
R

(c) C
0

C0

(moving)

0 c a

(fixed)

_ x2 _ x1

CD C0 R CG

X2, x2

CR

CG

(fixed)
Figure 7.

X1, x1

CG

Further distillation to essentials of Figure 5. A bar moving in 2D is shown: (a) Rigid motion from globally-aligned to base and corotated congurations; (b) key geometric quantities that dene rigid motions in 2D; (c) as in (a) but followed by a stretch from corotated to deformed. The globally-aligned conguration is ctitious: only a convenient link up device.

T rows of a 3 3 orthogonal rotation matrix (rotator) T0 , or equivalently columns of T0 .

{xi } The element corotated or CR frame (red in Figure 5). It is oriented by three unit base vectors iiR , T which are rows of a 3 3 orthogonal rotation matrix (rotator) T R , or equivalently columns of T0 . Note that the element index e has been suppressed to reduced clutter. That convention will be followed throughout unless identication with elements is important. In that case e is placed as supercript. The base frame {xi } is chosen according to usual FEM practices. For example, in a 2-node spatial beam element, x1 is dened by the two end nodes whereas x2 and x3 lie along principal inertia directions. An important convention, however, is that the origin is always placed at the element centroid C 0 . For each deformed (current) element conguration, a tting of the base element denes its CR conguration, also known as the element shadow. Centroids C R and C C D coincide. The CR frame {xi } originates at C R . Its orientation results from matching a rigid motion of the base frame, as discussed later. When the current element conguration reduces to the base at the start of the analysis, the base and CR frames coalesce: {xi xi }. At that moment there are only two different frames: global and local, which agrees with linear FEM analysis. Notational conventions: use of G, 0, R and D as superscripts or subscripts indicate pertinence to the globally-aligned, base, corotated and deformed congurations, respectively. Symbols with a overtilde or overbar are measured to the base frame {xi } or the CR frame {xi }, respectively. Vectors without a superposed symbol are referred to global coordinates {xi xi }. Examples: x R denote global coordinates of a point in C R whereas xG denote base coordinates of a point in CG . Symbols a, b and c = b a are abbreviations for the centroidal translations depicted in Figure 5, and more clearly in Figure 7(b). A generic, coordinate-free vector is denoted by a superposed arrow, for example u, but such entities rarely appear in this work. The rotators T0 and T R are the well known local-to-global displacement transformations of FEM analysis. Given a global displacement u, u = T0 u and u = T R u. 3.3. Coordinate Transformations Figures 5 and 6, although purposedly restricted to 2D, are still too busy. Figure 7, which pictures the 2D motion of a bar in 3 frames, displays essentials better. The (ctitious) globally-aligned conguration CG is explicitly shown. This helps to follow the ensuing sequence of geometric relations. Begin with a generic point xG in CG . This point is mapped to global coordinates x0 and x R in the base and corotated congurations C0 and C R , respectively, through
T x0 = T0 xG + a,

x R = TT xG + b, R 8

(2)

in which rotators T0 and T R were introduced in the previous subsection. To facilitate code checking, for the 2D motion pictured in Figure 7(b) the global rotators are T0 = c0 s0 0 s0 c0 0 0 0 , 1 TR = cR s R 0 sR cR 0 0 0 , 1 c0 = cos 0 , s0 = sin 0 , etc. (3)

When (2) are transformed to the base and corotated frames, the position vector xG must repeat: x0 = xG and x R = xG , because the motion pictured in Figure 7(a) is rigid. This condition requires x0 = T0 (x0 a), x R = T R (x R b). (4)

These may be checked by inserting x0 and x R from (2) and noting that xG repeats. 3.4. Rigid Displacements The rigid displacement is a vector joining corresponding points in C0 and C R . This may be referred to the global, base or corotated frames. For convenience call the C0 C R rotator R0 = TT T0 . Also introduce R T c = T0 c and c = TT c. Some useful expressions are R
T T T ur = x R x0 = (TT T0 )xG + c = (R0 I)T0 xG + c = (R0 I)T0 x0 + c R T T = (R0 I)T0 x R + c = (R0 I)(x0 a) + c = (I R0 )(x R b) + c, T x ur = T0 ur = T0 (R0 I)T0 x0 + c = (R0 I) 0 + c, T T x ur = T R ur = T R (I R0 )TT x R + c = (I R0 ) R + c. R T Here I is the 3 3 identity matrix, whereas R0 = T0 R0 T0 and R0 = T R R0 TT denote the C0 C R R rotator referred to the base and corotated frames, respectively.

(5)

3.5. Rotator Formulas Traversing the links pictured in Figure 8 shows that any rotator can be expressed in terms of the other two:
T T T T0 = T R R0 , T R = T0 R0 , R0 = TT T0 , R0 = T0 T R . R

T R0

(6)

In the CR frame: R0 = T R R0 TT , whence R R 0 = T0 T T , R


T T R0 = T R T 0 .

element base frame ~ X

R0

element CR frame _ x

(7)

T0

TT 0

T R
global frame X, x

TT R

Notice that T0 is xed since CG and C0 are xed throughout the analysis, whereas T R and R0 change. Their variations of these rotators are subjected to the following constraints:
T T0 = T0 = 0, T T R0 = T0 T R , T T R = T0 R0 , T TT = R0 T0 , R

Figure 8.

Rotator frame links.

R0 = TT T0 , R

TT T R + TT T R = 0, R R

T T R0 R0 + R0 R0 = 0.

(8)

T The last two come com the orthogonality conditions TT T R = I and R0 R0 = I, respectively, and provide R T T T T R0 = R0 R0 R0 , R0 = R0 R0 R0 , etc. We denote by and the axial vectors of R0 and R0 , respectively, using the exponential map form of the rotator described in Section A.10. The variations and are used to form the skew-symmetric

Table 2. Degree of Freedom and Conjugate Force Notation Notation v = [ v1 . . . v N ]T ua with va = Ra v = [ v1 . . . v N ]T ua with va = a ve = [ ve . . . ve e ]T 1 N e ua e with va = a e d ve = [ ve . . . ve N e ]T d d1 da ue e with vda = e
da

Frame Global

Level Structure

Description Total displacements and rotations at structure nodes. Translations: ua , rotations: Ra , for a = 1, . . . N . Incremental displacements and spins at structure nodes used in incremental-iterative solution procedure. Translations: ua , spins: a ; conjugate forces: na and ma , respectively, for a = 1, . . . N . Localization of above to element e in CR frame. Transe e e e lations: ua , spins: a ; conjugate forces: na and ma , e respectively, for a = 1, . . . N . Deformational displacements and rotations at element e da nodes. Translations: ue , rotations: da ; conjugate forces: na and ma , respectively, for a = 1, . . . N e .

Global

Structure

Local CR

Element

Local CR

Element

N = number of nodes in structure; N e = number of nodes in element e; a, b : node indices.


T spin matrices Spin() = R0 R0 = Spin()T and Spin( ) = R0 R0 = Spin( )T . These matrices are connected by congruential transformations: T T Spin() = T0 Spin( ) T0 , T Spin( ) = T0 Spin() T0 .

(9)

Using these relations the following catalog of rotator variation formulas can be assembled:
T T T T R = T0 R0 = T R R0 R0 = T R Spin() = R0 Spin( ) T0 , T T T TT = R0 T0 = R0 R0 TT = Spin() TT = T0 Spin( ) R0 . R R R T T T R0 = T R T0 = R0 R0 R0 = Spin() R0 = T0 Spin( ) R0 T0 , T T T T T T T R0 = T0 T R = R0 R0 R0 = R0 Spin() = T0 R0 Spin( ) T0 , T T T T R0 = T0 R0 T0 = T0 R0 R0 T R = T0 Spin() T R = Spin( ) R0 , T T T T T T T R0 = T0 R0 T0 = T R R0 R0 T0 = T R Spin() T0 = R0 Spin( ).

(10)

3.6. Degrees of Freedom For simplicity it will be assumed that an N e -node CR element has six degrees of freedom (DOF) per node: three translations and three rotations. This assumption covers the shell and beam elements evaluated in 0 Part II [38]. The geometry of the element is dened by the N e coordinates xa , a = 1, . . . N e in the base (initial) conguration, where a is a node index. The notation used for DOFs at the structure and element level is collected in Table 2. If the structure has N nodes, the set {ua , Ra } for a = 1, . . . N collectively denes the structure node displacement vector v. Note, however, that v is not a vector in the usual sense because the rotators Ra do not transform as vectors when nite rotations are considered. The interpretation as an array of numbers that denes the deformed conguration of elements is more appropriate. The element total node displacements ve are taken from v in the usual manner. Given ve , the key CR operation is to extract the deformational components of the translations and rotations for each node. 10

Table 3. Forming the Deformational Displacement Vector. Step 1. Operation for each element e and node a = 1, . . . a
e From the initial global nodal coordinates xa compute centroid position ae = Ne e e e xC0 = (1/N e ) a=1 xa . Form rotator T0 as per element type convention. e e Compute node coordinates in the element base frame: xa = Te (xa ae ). 0 e e e Compute node coordinates in deformed (current) conguration: xa = xa + ua Ne e e and the centroid position vector be = xC = (1/N e ) a=1 xa . Establish the e e deformed local CR system T by a best-t procedure, and R0 = Te (Te )T . Form 0 e Ra local-CR node coordinates of CR conguration: xe = Te (xa be ). T e Ra Compute the deformational translations uda = xa xe . Rd = Tn Ra T0 Come e pute the deformational rotator Rda = Te Ra (Te )T . Extract the deformational 0 e e . angles da from the axial vector of Rda

2.

3.

That sequence of operations is collected in Table 3. Note that the computation of the centroid is done by simply averaging the coordinates of the element nodes. For 2-node beams and 3-node triangles this is appropriate. For 4-node quadrilaterals this average does not generally coincides with the centroid, but this has made little difference in actual computations. 3.7. EICR Matrices Before studying element deformations, it is convenient to introduce several auxiliary matrices: P = Pu P , S, G, H and L that appear in expressions of the EICR front-end. As noted, elements treated here possess N e nodes and six degrees of freedom (DOF) per node. The notation and arrangement used for DOFs at different levels is dened in Table 2. Subscripts a and b denote node indices that run from 1 to N e . All EICR matrices are built node-by-node from node-level blocks. Figure 9(a) illustrates the concept of perturbed conguration C D + C, whereas Figure 9(b) is used for examples. The CR and deformed conguration are frozen; the latter being varied in the sense of variational calculus. The translational projector matrix Pu or simply T-projector is dimensioned 6N e 6N e . It is built from 3 3 numerical submatrices Uab = (ab 1/N e ) I, in which I is the 3 3 identity matrix and ab the Kronecker delta. Collecting blocks for all N e nodes and completing with 3 3 zero and identity blocks as placeholders for the spins and rotations gives a 6N e 6N e matrix Pu . Its conguration is illustrated below for N e = 2 (e.g., bar, beam, spar and shaft elements) and N e = 3 (e.g., triangular shell elements): 2 I 0 1I 0 1I 0 3 3 3 1 0 I 0 1I 0 I 0 0 0 0 2 2 1I 0 2 1 0 I 0 0 I 0 3I 0 3 3 , Ne = 3: Pu = Ne = 2: Pu = 1 . (11) 1 I 0 0 0 0 I 0 0 I 0 2 2 1 2 I 0 1I 0 I 0 0 0 0 I 3 3 3 0 0 0 0 0 I For any N e 1 it is easy to verify that P2 = Pu , with 5N e unit eigenvalues and N e zero eigenvalues. Thus u Pu is an orthogonal projector. Physically, it extracts the deformational part from the total translational displacements. Matrix S is called the spin-lever or moment-arm or matrix. It is dimensioned 3N e 3 and has the conguration (written in transposed form to save space):
T S = [ S1

T S2

I . . . ST e N 11

I ]T ,

(12)

(a)

Instantaneous rotation axis CR C


xD a

Corotated CR Deformed CD C + C

_ x3 _ x2 _ x1
a
v a (includes displacements and spins)

(b) Perturbed C D + C
C + C

_ u2j
jR

c perturbed a

x D xa

_ 3
iR

CR

D xa D Perturbed C + C

_ u2i iD _ Deformed CD _ u1i L u3i (+up)

_ x1 C _ x3 (+up) Corotated CR

_ x2

jD _ u1j u3j(+up)

Figure 9.

Concept of perturbed conguration to illustrate derivation of EICR matrices: (a) facet triangular shell element moving in 3D space; (b) 2-node bar element also in 3D but depicted in the {x1 , x2 } plane of its CR frame. Deformations grossly exaggerated for visualization convenience; strains and local rotations are in fact innitesimal.

in which I is the 3 3 identity matrix and Sa are node spin-lever 3 3 submatrices. Let xa = [ x1a x2a x3a ]T generically denote the 3-vector of coordinates of node a referred to the element centroid. Then Sa = Spin(xa ). The coordinates, however, may be those of three different congurations: C0 , C R and C D , referred to two frame types: global or local. Accordingly superscripts and overbars (or tildes) are used to identify one of six combinations. For example 0 0 0 x3a a3 x2a a2 0 x3a x2a R R 0 x3a x2a D D 0 0 0 R D Sa = x3a a3 0 x1a a1 , Sa = x3a R 0 x1a , Sa = x3a R D 0 x1a , D 0 0 x2a a2 x1a a1 0 x2a x1a R R 0 x2a x1a D D 0 (13) are node spin-lever matrices for base-in-global-frame, CR-in-local-frame and deformed-in-local-frame, R D respectively. The element matrix (12) inherits the notation; in this case S0 , S and S , respectively. For instance, S matrices for the 2-node space i j bar element pictured in Figure 9(b) are 12 3. If the length of the bar in C D is L, the deformed bar spin-lever matrix referred to the local CR frame is S = 1L 2
D

0 0 0 0 0 1

0 0 0 0 0 0 0 1 0 0 0 0 0 1 0 0 0 0 0 1 0

0 0 0 0 0 0 0 0 0

(14)

The rst row is identically zero because the torque about the bar axis x1 vanishes in straight bar models. Matrix G, introduced by Haugen [37], is dimensioned 36N e , and will be called the spin-tter matrix. It links variations in the element spin (instantaneous rotations) at the centroid of the deformed conguration in response to variations in the nodal DOFs. See Figure 9(a). G comes in two avors, global and local: = G ve =
def e G va , a a def = G ve =

e G va , a a

with

Ne a=1

(15)

Here the spin axial vector variation e denotes the instantaneous rotation at the centroid, measured in the global frame, when the deformed conguration is varied by the 6N e components of ve . When referred to the local CR frame, these become e and ve , respectively. For construction, both G and G a shown above. As may be split into node-by-node contributions using the 3 6 submatrices Ga and G an example, G matrices for the space bar element shown in Figure 9(b) is 3 12. The spin-lever matrix 12

in C D referred to the local CR frame is 1 D G = L 0 0 0 0 0 1 0 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 1 0 0 0 0 0 0 0 . 0 0 0 (16)

The rst row is conventionally set to zero as the spin about the bar axis x1 is not dened by the nodal freedoms. This torsion spin is dened, however, in 3D beam models by the end torsional rotations. Unlike S, the entries of G depend not only on the element geometry, but on a developers decision: how the CR conguration C R is tted to C D . For the triangular shell element this matrix is given in Appendix B. For quadrilateral shells and space beam elements it is given in Part II [38]. D D Matrices S and G satisfy the biorthogonality property G S = D. (17)

where D is a 3 3 diagonal matrix of zeros and ones. A diagonal entry of D is zero if a spin component is DD undened by the element freedoms. For instance in the case of the space bar, the product G S of (14) and (16) is diag(0, 1, 1). Aside from these special elements (e.g., bar, spars, shaft elements), D = I. This property results from the fact that the three columns of S are simply the displacement vectors associated with the rigid body rotations i = 1. When premultiplied by G one merely recovers the amplitudes of those three modes. The rotational projector or simply R-projector is generically dened as P = S G. Unlike the T-projector Pu such as those in (11), the R-projector depends on conguration and frame of reference. Those are R R R identied in the usual manner; e.g., P = S G . This 6N e 6N e matrix is an orthogonal projector of rank equal to that of D = G S. If G S = I, Pr has rank 3. The complete projector matrix of the element is dened as (18) P = Pu P . This is shown to be a projector, that is P2 = P, in Section 4.2. Two additional 6N e 6N e matrices, denoted by H and L, appear in the EICR. H is a block diagonal matrix built of 2N e 3 3 blocks: H = diag [ I H1 I H2 . . . I H N e ] , Ha = H(a ), H() = /. (19)

Here Ha denotes the Jacobian derivative of the rotational axial vector with respect to the spin axial vector evaluated at node a. An explicit expression of H() is given in (101) of Appendix A. The local version in the CR frame is H = diag [ I Hd1 I Hd2 . . . I Hd N e ] , Hda = H(da ), H(d ) = d / d . (20)

L is a block diagonal matrix built of 2N e 3 3 blocks: L = diag [ 0 L1 0 L2 . . . 0 L N e ] , La = L(a , ma ). (21)

where ma is the 3-vector of moments (conjugate to a ) at node a. The expression of L(, m) is provided in (102) of Appendix A. The local form L has the same block organization with La replaced by La = L(da , ma ). 13

3.8. Deformational Translations Consider a generic point P 0 of the base element of Figure 5, with global position vector x0 . P 0 rigidly P moves to P R in C R with position vector x R = x0 + u R = x0 + c + x R . Next the element deforms to P P P P PC occupy C D . P R displaces to P, with global position vector x P = x0 + u P = x0 + c + x R + ud P . P P PC P The global vector from C 0 to P 0 is x0 a, which in the base frame becomes x0 = T0 (x0 a). The P P R R P global vector from C C to PR is x P b, which in the element CR frame becomes x R = T R (x R b). PC P P But x0 = x R since the C0 C R motion is rigid. The global vector from PR to P is ud P = x P x R , P which represents a deformational displacement. In the CR frame this becomes ud P = T R (x P x R ). P The total displacement vector is the sum of rigid and deformational parts: u P = ur P + ud P . The rigid displacement is given by expressions collected in (5), of which ur P = (R0 I) (x0 a) + c is the most P useful. The deformational part is extracted as ud P = u P ur P = u P c + (I R0 ) (x0 a). Dropping P P to reduce clutter this becomes ud = u c + (I R0 ) (x0 a). The element centroid position is calculated by averaging its node coordinates. Consequently c = (1/N e ) u , b b ua c = U u , b ab b with
b

(22)

Ne b=1

(23)

in which Uab = (ab 1/N e ) I is a building block of the T-projector introduced in the foregoing subsection. Evaluate (22) at node a, insert (23), take variations using (10) to handle R0 , use (2) to map R0 (x0 a) = x R b, and employ the cross-product skew-symmetric property (56) to extract :
0 uda = (ua c) R0 (xa a) = b 0 Uab ub Spin() R0 (xa a) b R Uab ub + Spin(xa b)

= =

b b

R Uab ub Spin() (xa b) =

(24)

Uab ub +

R Sa Gb vb .

Here matrices S and G have been introduced in (12)(15). The deformational displacement in the element T x CR frame is ud = T R ud . From the last of (5) we get ud = u c (I R0 ) R , where R0 = T R R0 TT . R Proceeding as above one gets uda =
b

Uab ub +
D

D Sa Gb vb .

(25)

R The node lever matrix Sa of (24) changes in (25) to Sa , which uses the node coordinates of the deformed element conguration.

3.9. Deformational Rotations Denote by R P the rotator associated with the motion of the material particle originally at P 0 ; see Figure 6. Proceeding as in the translational analysis this is decomposed into the rigid rotation R0 and a deformational rotation: R P = Rd P R0 . The sequence matters because Rd P R0 = R0 Rd P . The order Rd P R0 : rigid rotation follows by deformation, is consistent with those used by Bergan, Rankin and coworkers; e.g. [54,56]. (From the standpoint of continuum mechanics based on the polar decomposition T theorem [81, Sec. 37] the left stretch measure is used.) Thus Rd P = R P R0 , which can be mapped to the local CR system as Rd = T R Rd TT . Dropping the label P for brevity we get R
T T Rd = R R0 = R T0 T R , T R d = T R Rd T T = T R R T0 . R

(26)

14

The deformational rotation (26) is taken to be small but nite. Thus a procedure to extract a rotation axial vector d from a given rotator is needed. Formally this is d = axial Loge (Rd )], but this can be prone to numerical instabilities. A robust procedure is presented in Section A.11. The axial vector is evaluated at the nodes and identied with the rotational DOF. Evaluating (26) at a node a, taking variations and going through an analysis similar to that carried out in the foregoing section yields da = da da da da = da da b = Ha b da b = Ha b ab [ 0 ab [ 0 I ] Gb vb . I ] G b vb . (27)

where Gb is dened in (15) and Ha in (19). 4. Internal Forces e The element internal force vector pe and tangent stiffness matrix K are computed in the CR conguration based on small deformational displacements and rotations. Variations of the element DOF, collected in ve as indicated in Table 2, must be linked to variations in the global frame to esh out the EICR interface d of Figure 4. This section develops the necessary relations. 4.1. Force Transformations Consider an individual element e with N e nodes with six DOF (three translations and three rotations) at each. Assume the element to be linearly elastic, undergoing only small deformations. Its internal energy d is assumed to be a function of the deformational displacements: U e = U e ( e ), with array ve organized vd e as shown in Table 2. U is a frame independent scalar. The element internal force vector pe in the CR d frame is given by pe = U e / ve . For each node a = 1, . . . N e : e pa = U , da ve
e

or

ua pe a pe

U e uda = U e da

(28)

where the second form separates the translational and rotational (moment) forces. To refer these to the global frame we need to relate local-to-global kinematic variations: ue Jab
b=1 e ua e , a db e ua e db e ua

da ue e da

Jab =

db ue e a e . db e a

(29)

e d u From virtual work invariance, (pe )T ue + (pe )T d = (pe )T ue + (pe )T e , whence u pe ua pe a


Ne

=
b=1

T Jab

ua pe , a pe

a = 1, . . . N e .
T T

(30)

T T It is convenient to split the Jacobian in (29) as Jab = Hb Pab Ta and Jab = Ta Pab Hb . These matrices are

15

provided from three transformation stages, owcharted in Figure 10: db ue e db b ue e b e ua a e = I 0 db 0 H e ua , a e


e ua e a

b ue , e b

= Pab = Ta

e a ab = ue with P db e ua e TR 0 ua = e , 0 TR a

with Hdb = ue
db

db e db db ue a e e db a e
e

, , (31)

The 3 3 matrix L is the Jacobian derivative already encountered in (19). An explicit expression in terms of is given in (101) of Appendix A. To express compactly the transformations for the entire element it is convenient to assemble the 6N e 6N e matrices P11 P P = 21 ... PN e1 P12 P22 ... PN e2 ... ... ... ... P1N e P2N e , ... PN e N e
_ _ HPT

(32)

CR deformational displac_ rotations & _ ud , d

global total displac & spins u, global-to-CR frame rotator

T = diag [ T R T R . . . T R ] . and H is dened in (20). Then the element transformations can be written ve = H P T ve , d T T pe = TT P H pe . (33)

_ rotation-toH spin Jacobian


CR deformational displac & _ _ spins ud , d

_ _ _ P = Pu + P

Projector

CR total displac &_ _ spins u,

Figure 10.

Staged transformation sequence from deformed to global DOFs.

The 6 6 matrix Pab in (31) extracts the deformational part of the displacement at node b in terms of the d total displacement at node a, both referred to the CR frame. At the element level, ve = P ve extracts is called a projector. the deformational part by projecting out the rigid body modes. For this reason P As noted in Section 3.5, P may be decomposed into a translational projector or T-projector Pu and a rotational projector or R-projector P , so that P = Pu + P . Each has a rank of 3. The T-projector is a purely numeric matrix exemplied by (11). The R-projector can be expressed as P = SG, where S is dened in (12) and G in (15). Additional properties are studied below.
Remark 3. Rankin and coworkers [54,6367] use an internal force transformation in which the incremental nodal 1 rotations are used instead of the spins. This results in an extra matrix, H appearing in the sequence (33). The projector derived in those papers differs from the one constructed here in two ways: (1) only the R-projector is considered, and (2) the origin of the CR frame is not placed at the element centroid but at an element node dened by local node numbering. Omitting the T-projection is inconsequential if the element is clean with respect to translational rigid body motions [30, Sec. 5].

4.2. Projector Properties In this section the bar over P, etc is omitted for brevity, since the properties described below are frame independent. In Section 3.5 it was stated without proof that (18) veries the orthogonal projector property P2 = P. Since P2 = (Pu P )2 = P2 2Pu P + P2 , satisfaction requires P2 = Pu , P2 = P , and u u Pu P = 0. Verication of P2 = Pu is trivial. Recalling that P = SG we get u P2 = S (G S) G = S I G = S G = P . 16 (34)

This assumes D = I in (17); verication for non-identity D is immediate upon removal of zero rows and columns. The orthogonality property Pu P = Pu S G = 0 follows by observing that Spin(xC ) = 0, where xC are the coordinates of the element centroid in any frame with origin at C. In the derivation of the consistent tangent stiffness, the variation of PT contracted with a force vector f, where f is not varied, is required. The variation of the projector can be expressed as P = Pu P = P = S G S G. (35)

For the tangent stiffness one needs PT f. This vector can be decomposed into a balanced (selfequilibrated) force fb = Pf and an unbalanced (out of equilibrium) force fu = (I P)f. Then PT f = (GT ST + GT ST ) (fb + fu ) = GT ST PT f (GT ST + GT ST ) fu = GT ST PT f + PT fu . (36)

where ST fb = 0 was used. This comes from the fact that the columns of S are the three rotational rigid body motions, which do not produce work on an self-equilibrated force vector. The term PT fu will be small if element congurations C R and C D are close because in this case fu will approach zero. If G has the factorizable form shown below, however, we can show that PT fu = 0 identically, regardless of how close C R and C D are, as long as f is in translational equilibrium. Assume that G can be factored as G = , with G = . (37) where is a coordinate dependent invertible 3 3 matrix, and is a constant 3 6N e matrix. Since GS = I as per (17), 1 = S, and G S + G S = S + G S = 0, whence = G S . Then PT fu = (GT ST + T T ST ) fu = (GT ST T T ST , GT ST ) fu
T T T = (GT ST GT ST P ) fu = (GT ST (I P ) fu = GT ST (I P ) fu

(38)

= G S (I
T T

T P )(I

P ) f = 0,
T

T if f is in translational equilibrium: f = P f. This is always satised for any element that represents rigid body translations correctly [30, Sec. 5].

5. Tangent Stiffness We consider here only the stiffness derived from the internal energy. The load stiffness due to nonconservative forces, such as aerodynamic pressures, has to be treated separately. 5.1. Denition The consistent tangent stiffness matrix Ke of element e is dened as the variation of the internal forces with respect to element global freedoms: pe = Ke ve ,
def

whence Ke =

pe . ve

(39)

Taking the variation of pe in (33) gives rise to four terms: T T T T T T T pe = TT P H pe + TT P H pe + TT P HT pe + TT P H pe = (Ke R + Ke P + Ke M + Ke ) ve . G G G M (40)

The four terms identied in (40) receive the following names. K M is the material stiffness, KG M the moment-correction geometric stiffness, KG P the equilibrium projection geometric stiffness, and KG R the rotational geometric stiffness. If nodal eccentricities treated by rigid links are considered, one more term appears, called the eccentricity geometric stiffness. This term is studied in great detail in [37]. 17

5.2. Material Stiffness The material stiffness is generated by the variation of the element internal forces pe : T T Ke ve = TT P H pe . M R (41)

d The linear stiffness matrix in terms of the deformational freedoms in ve is dened as the Hessian of the internal energy: 2U e pe e (42) K = e e = e vd vd vd Using the transformation of ve in (33) gives T T e Ke = TT P H K H P T. M
e

(43)

Thus the material stiffness is given by a congruential transformation of the local stiffness K to the global frame. This is formally the same as in linear analysis but here the transformation terms depend on the d state. The expression (43) is valid only if Ke is independent of the deformational ve freedoms. 5.3. Geometric Stiffness To express compactly the geometric stiffness components it is convenient to introduce the arrays ne 1 1 me T T P pe = P H pe = . , . . e nN e N me e 1 Spin(ne ) 0 . , . Fn = . Spin(ne e ) N 0 Spin(ne ) 1 1 Spin(me ) . . . = . e Spin(n N e ) N Spin(me e )

Fmn

(44)

P These are lled with the projection node forces pe . Only the nal form of the geometric stiffness components is given below, omitting the detail derivations of [37]. The rotational geometric stiffness is T T generated by the variation of T: Ke ve = TT P H pe and can be expressed as RG KG R = TT Fnm G T, (45)

Ke is the gradient of the internal force vector with respect to the rigid rotation of the element. This RG interpretation is physically intuitive because a rigid rotation of a stressed element necessarily reorients the stress vectors by that amount. Consequently the internal element forces must rigidly rotate to preserve equilibrium. The moment-correction geometric stiffness is generated by the variation of the jacobian H: Ke ve = RG T T T e T R P H p . It is given by T Ke M = TT P L PT. (46) G where L is dened in Section 3.5. The equilibrium projection geometric stiffness arises from the variation of the projector P with respect e T T T e e to the deformed element geometry: KG P v = T R P H p . As in Section 4.2, decompose pe into a e T pe and an unbalanced force pe = pe pe . If PT pe is either u u b balanced (self-equilibrated) force pb = P identically zero or may be neglected as discussed in Section 4.2, Ke P is given by G T KG P = TT G Fn P T, 18 (47)

b in which the balanced force pe is used in (44) to get Fn . u If TT P pe cannot be neglected, as may happen in highly warped shell elements in a coarse mesh, the following correction term may be added to Ke : RG
T

KG P = TT

T G S pe u G Fnu P + v

T,

(48)

u where Fnu is Fn of (44) when pe is inserted instead of pe . In the computations reported in Part II [38] this term was not included. Ke expresses the variation of the projection of the internal force vector pe as the element geometry RG changes. This can be interpreted mathematically as follows: In the vector space of element force vectors the subspace of self-equilibrium force vectors changes as the element geometry changes. The projected force vector thus has a gradient with respect to the changing self-equilibrium subspace, even though the element force e does not change. f The complete form of the element tangent stiffness, excluding correction terms (48) for highly warped elements, is T T e T T T e Ke = TT P H K H P + P L P Fnm G G Fn P T = TT K P T, (49)

e in which K R , which is the local tangent stiffness matrix (the tangent stiffness matrix in the local CR frame of the element) is given by the parenthesized expression. 5.4. Consistency Verication e The local tangent stiffness matrix K R given in (49) has some properties that may be exploited to verify the computer implementation [54,37]: e K R S = Fnm , KG P S = 0,
T T e S K R = Fn ,

K M S = 0,

KG R S = Fnm ,

T S K M = 0,

T T S (KG R + KG P ) = Fn .

(50)

e In addition, rigid-body-mode tests on the linear stiffness matrix K using linearized projectors are discussed in [30]. The set (50) tests the programming of the nonlinear projector P since it checks the null space of P. It also indicates whether the projector matrix is used correctly in the stiffness formulation. However, satisfaction does not fully guarantee consistency between the internal force and the tangent stiffness because H and L are left unchecked. Full verication of consistency can be numerically done through nite difference techniques. 6. Three Consistent CR Formulations From the foregoing unied forms of the internal force and tangent stiffness, three CR consistent formulations can be obtained by making simplifying assumptions at the internal force level. These satisfy self-equilibrium and symmetry to varying degree. The following subsections describe the three versions in order of increasing complexity. For all formulations one can take in account DOFs at eccentric nodes as described in [37]. 6.1. Consistent CR formulation (C) This variant is that developed by Bergan and coworkers in the 1980s at Trondheim and summarized in the review article [56]. The internal force (33) is simplied by taking H = I and P = I, while retaining 19

d ve = H P ve for recovery of deformational DOFs. Since P = H = 0, the expression for the tangent stiffness of (40) simplies to the material and rotational geometric stiffness terms: e d pe = TT K ve , e Ke = TT (K H P Fnm G)T. (51)

e Here Fnm is computed according to (44) with pe = K vd . The internal force is in equilibrium with respect to the CR conguration C R . For a shell structure, the material stiffness approaches symmetry as the element mesh is rened if the membrane strains are small. As the mesh is rened, the deformational rotation axial vectors da become smaller and approach vector properties that in turn make H(da e approach the identity matrix. With small membrane strains K is indifferent with respect to post e e multiplication with P because the C R and C congurations will be close and K P K . The consistent geometric stiffness is always unsymmetric, even at equilibrium. Because of this fact one cannot expect quadratic convergence for this formulation if a symmetric solver is used. This formulation may be unsatisfactory for warped quadrilateral shell elements since the C R and C D reference congurations may be far apart. Only in the limit of a highly rened element mesh will the C R and C D references in general be close, and satisfactory equilibrium ensured. 6.2. Consistent Equilibrated CR Formulation (CE) The internal force (33) is simplied by taking H = I so H = 0, but the projector P is retained. This gives T e T e d T T (52) pe = TT P K ve , Ke = TT (P K H P Fmn G G Fn P)T. T e d where Fnm and Fn are computed according to (44) with pe = P K ve . Due to the presence of P on both sides, the material stiffness of the CE formulation approaches symmetry as the mesh is rened regardless of strain magnitude. The geometric stiffness at the element level is nonsymmetric, but the assembled global geometric stiffness will become symmetric as global equilibrium is approached, provided that there are no applied nodal moments and that displacement boundary conditions are conserving. A symmetrized global tangent stiffness maintains quadratic convergence for rened element meshes with this formulation. 6.3. Consistent Symmetrizable Equilibrated CR Formulation (CSE) All terms in (33) are retained, giving T T e d pe = TT P H K ve , T T e T T T Ke = TT (P H K H P + P L P Fnm G G Fn P)T. (53)

T e f d where Fnm and Fn are computed according to (44) with = P H K ve . The assembled global geometric stiffness for this formulation becomes symmetric as global equilibrium is approached, as in the CE case, as long as there are no applied nodal moments and the loads as well as boundary conditions are conserving. Since the material stiffness is always symmetric, quadratic convergence with a symmetrized tangent stiffness can be exppected without the rened-mesh-limit assumption of the CE formulation.

Remark 4. The relative importance of including the H matrix, which is neglected by most authors, and the physical signicance of this Jacobian term are discussed in Part II [38].

6.4. Formulation Requirements It is convenient to set forward a set of requirements for geometrically nonlinear analysis with respect to which different CR formulations can be evaluated. They are listed below in order of decreasing importance. 20

Table 4. Attributes of Corotated Formulations C, CE and CSE. Formulation Self-equil.(1) Consistent(2) Invariant(3) Symmetriz.(4) Elem.Indep.(5) C CE CSE checked if element is in self-equilibrium in deformed conguration C . checked if tangent stiffness is the v gradient of the element internal force. checked if the formulation is insensitive to choice of node numbering. checked if formulation maintains quadratic convergence of a true Newton solution algorithm with a symmetrized tangent stiffness matrix. (5) checked if the matrix and vector operations that account for geometrically nonlinear effects are the same for all elements with the same node and DOF conguration. (1) (2) (3) (4)
D

Equilibrium. By this requirement is meant: to what extent the nite element internal force vector p is in self-equilibrium with respect to the deformed conguration C D ? This is a fundamental requirement for tracing the correct equilibrium path in an incremental-iterative solution procedure. Consistency. A formulation is called consistent if the tangent stiffness is the gradient of the internal forces with respect to the global DOF. This requirement determines the convergence rate of an incrementaliterative solution algorithm. An inconsistent tangent stiffness may give poor convergence, but does not alter the equilibrium path since this is entirely prescribed by the foregoing equilibrium requirement. However, lack of consistency may affect the location of bifurcation (buckling) points and the branch switching mechanism for post-buckling analysis. In other words, an inconsistent tangent stiffness matrix may detect (see) a bifurcation where equilibrium is not satised from the residual equation. Subsequent traversal by branch-switching will then be difcult because the corrector iterations need to jump to the secondary path as seen by the residual equation. Invariance. This requirement refers to whether the solution is insensitive to internal choices that may depend on node numbering. For example, does a local element-node reordering give an altered equilibrium path or change the convergence characteristics for the analysis for an otherwise identical mesh? The main contributor to lack of invariance is the way the deformational displacement vector is extracted from the total displacements, if the extraction is affected by the choice of the local CR frame. If lack of invariance is observed, it may be usually traced to the matrix G, which links the variation of the rigid body rotation to that of the nodal DOF degrees of freedom. Symmetrizability. This means that a symmetrized K can be used without loss of quadratic convergence rate in a true Newton solver even when the consistent tangent stiffness away from equilibrium is not symmetric. In the examples studied in Part II [38] this requirement was met when the material stiffness of the formulation was rendered symmetric. Element Independence. This is used in the sense of the EICR discussed in Section 2.5. It means that the matrix and vector operations that account for geometrically nonlinear effects are the same for all elements that possess the same node and DOF conguration. Attributes of the C, CE and CSE formulations in light of the foregoing requirements are summarized in Table 4. 6.5. Limitations of the EICR Formulation The present CR framework, whether used in the C, CE or CSE formulation variants, is element independent in the EICR sense discussed in Section 2.5 since it does not contain gradients of intrinsically element dependent quantities such as the strain-displacement relationship. This treatment is appropriate 21

for elements where the restriction to small strains automatically implies that the CR and deformed element congurations are close. This holds automatically for low order models such as two-node straight bars and beams, and three-node facet shell elements. The main reason for limiting element independence to low-order elements is the softening effect of the nonlinear projector P. The use of P to restore the correct rigid body motions, and hence equilibrium with respect to the deformed element geometry, effectively reduces the eigenvalues of the material stiffness relative to the CR material stiffness Ke before projection. This softening effect becomes signicant if R D the C and C geometries are far apart. Such softening effects are noticeable in four-node initially-warped shell elements. Assume that the element is initially warped with positive warping, and consider only the effect of P. The element T material stiffness of this initial positive warping is then K+ = P+ Ke P+ = Ke . Apply displacements that switch this warping to the opposite of the initial one; that is, a negative warping. The new T element material stiffness then becomes K = P Ke P = Ke One will intuitively want the two element congurations to have the same rigidity in the sense of the dominant nonzero eigenvalues of the tangent stiffness matrix. But it can be shown that the eigenvalues of the projected material stiffness matrix K can be signicantly lower than those of the initial stiffness matrix K+ . If the element stiffness Ke is referred to the at element projection, one can restore symmetry of K+ and K with respect to dominant nonzero eigenvalues, but it is not possible to remove the softening effect. This argument also carries over to higher order bar, arch and shell elements that are curved in the initial reference conguration. It follows that the EICR is primarily useful for low-order elements of simple geometry. 7. Conclusions This article presents a unied formulation for geometrically nonlinear analysis using the CR kinematic description, assuming small deformations. Although linear elastic material behavior has been assumed for brevity, extension to materially nonlinear behavior such as elastoplasticity and fracture within the connes of small deformations, is feasible as further discussed below. All terms in the internal force and tangent stiffness expressions are accounted for. It is shown how dropping selected terms in the former produces simpler CR versions used by previous investigators. These versions have been tested on thin shell and exible-mechanism structures, as reported in Part II [38]. Shells are modeled by triangle and quadrilateral elements. The linear stiffness of these elements is obtained with the ANDES (Assumed Natural DEviatoric Strains) formulation of high performance elements [2732,53]. Test problems include benchmarks in buckling, nonlinear bifurcation and collapse. Does the unied formulation close the book on CR? Hardly. Several topics either deserve further development or have been barely addressed: (A) Relaxing the small-strain assumption to allow moderate deformations. (B) Robust handling of extremely large rotations involving multiple revolutions. (C) Integrating CR elements with rigid links and joint elements for exible multibody dynamics. (D) Using substructuring concepts for CR modeling of structural members with continuum elements. (E) Achieving a unied form for CR dynamics, including nonconservative effects and multiphysics. Topic (A) means the use of CR for problems where strains may locally reach moderate levels, say 110%, as in elastoplasticity and fracture, using appropriate strain and stress measures in the local frame. The challenge is that change of metric of the CR conguration should be accounted for, even if it means dropping the EICR property. Can CR compete against the more established TL and UL descriptions? It seems unreasonable to expect that CR can be of use in overall large strain problems such as metal 22

forming, in which UL reigns supreme. But it may be competitive in localized failure problems, where most of the structure remain elastic although undergoing nite rotations. Some data points are available: previous large-deformation work presented in [46,47,50,51,57]. More recently Skallerud et al. reported [75] that a submerged-pipeline failure shell code using the ANDES CR quadrilateral of [37] plus elastoplasticity [74] and fracture mechanics [21] was able to beat a well known commercial TL-based code by a factor of 600 in CPU time. This speedup is of obvious interest in inuencing both design cycle and deployment planning. Topic (B) is important in applications where a oating (free-free) structure undergoes several revolutions, as in combat airplane maneuvers, payload separation or orbital structure deployment. The technical difculty is that expressions presented in the Appendix cannot handle nite rotations beyond 2 , and thus require occasional resetting of the base conguration. While this can be handled via restarts for structures such as full airplanes, it can be more difcult when the relative rotation between components exceeds 2 , as in separation, fragmentation or deployment problems. Topics (C) and (D) have been addressed in the FEDEM program developed by SINTEF at Trondheim, Norway. This program combines CR shell and beam elements of [37], grouped into substructures, with kinematic objects typical of rigid-body dynamics: eccentric links and joints. Basic tools used in FEDEM for combining joint models with exible continuum elements are covered in a recent book [73]. Finally, topic (E) is fertile ground for research. The handling of model components such as mass, damping and nonconservative effects in uid-structure interaction and aeroelasticity is an active ongoing research topic. For example a recent paper [26] describes ight maneuver simulations of a complete F-16 ghter using CR elements to model the aircraft. As in statics, a key motivation for CR in dynamics is reuse of linear FEM force-stiffness libraries. Can that reuse extend to mass and damping libraries? And how do standard time integration methods perform when confronted with unsymmetric matrices? These topics have barely been addressed.

23

Appendix A. THE MATHEMATICS OF FINITE ROTATIONS This Appendix collects formulas and results for the mathematical treatment of nite rotations in 3D space. Emphasis is placed on representations useful in the CR description of FEM. Several of the results are either new or simplications of published results. Some statements in the literature concerning the exponential and Cayley maps are shown to be erroneous. One difculty for investigators and students learning the CR description is that different normalizations have been used by different authors. This topic is reviewed in some detail, and summary help provided in the form of Table 5. A.1 Spatial Rotations

Plane rotations are easy. A rotation in, say, the {x1 , x2 } plane, is dened by just a scalar: the rotation angle about x3 . Plane rotations commute: 1 + 2 = 2 + 1 , because the s are numbers. The study of spatial (3D) rotations is more difcult. The subject is dominated by a fundamental theorem of Euler: The general displacement of a rigid body with one point xed is a rotation about some axis which passes through that point. (54)

Thus spatial rotations have both magnitude: the angle of rotation, and direction: the axis of rotation. These are nominally the same two attributes that categorize vectors. Not coincidentally, rotations are sometimes pictured as vectors but with a double arrow in Mechanics books. Finite 3D rotations, however, do not obey the laws of vector calculus, although innitesimal rotations do. Most striking is commutation failure: switching two successive rotations does not yield the same answer unless the rotation axis is kept xed. Within the framework of matrix algebra, nite rotations can be represented in two ways: (a) As 3 3 real orthogonal matrices R called rotators. (An abbreviation for rotation tensor). (b) As 3 3 real skew-symmetric matrices called spinors. (An abbreviation for spin tensor). The spinor representation is important in theory and modeling because the matrix entries are closely related to the ingredients of Eulers theorem (54).
Not unique: adjustable by a scale factor Spin() Spinor axial() Unique Rot() Skew(R)

The rotator representation is convenient for numer- Axial vector or ical work, as well as being naturally related to the pseudovector polar factorization of a transformation matrix. The two representations are connected as illustrated in Figure 11. Figure 11, which also includes the axial vector introduced below. Of the R links, the RodriguesCayley version is historically the rst, as discussed in [20], although not the most important one.

Rotator R

Representations of nite spatial rotations. Note: Some authors write e and Loge (R) for Rot() and Skew(R), respectively. This is correct for a particular scale factor ; cf. Table 5.

A 3 3 skew-symmetric matrix such as is dened by three scalar parameters. These three numbers can be arranged as components of an axial vector . Although looks like a 3-vector, it violates certain properties of classical vectors such as the composition rule. Therefore the term pseudovector is sometimes used for . In the EICR formulation presented here, axial vectors of rotations and spins function as incremental quantities directly connected to variations. Rotator matrices are used to record the global structure motions. See Table 2. This Appendix intends to convey that nite 3D rotations can appear in alternative mathematical representations, as diagramed in Figure 11. The following exposition expands on this topic, and studies the links shown there. A.2 Spinors

Figure 12(a) depicts a 3D rotation in space {x1 , x2 , x3 } by an angle about an axis of rotation . For convenience the origin of coordinates O is placed on . The rotation axis is dened by three directors: 1 , 2 , 3 , at least one of which must be nonzero. These components may be scaled by an nonzero factor through which the vector may be normalized in various ways discussed later. The rotation takes an arbitrary point P(x), located by its position vector x, into Q(x, ), located by its position vector x . The center of rotation C is dened by projecting P on the rotation axis. The plane of rotation CPQ is normal to that axis at C.

24

The radius of rotation is vector r of magnitude r from C to P. As illustrated in Figure 12(b) the distance between P and Q is 2r sin 1 . 2 The positive sense of obeys the RHS screw rule: positive counterclockwise if observed from the tip of the rotation axis. The angle shown in Figure 12 is positive. A.3 Spin Tensor and Axial Vector
x3
O

n, C
r Rotation axis defined by axial vector Can be normalized in various ways

(a)

Given the three directors 1 , 2 , and 3 of the axis , we can associate with it a 3 3 skew-symmetric matrix , called a spin tensor, spin matrix or briey spinor, by the rule = Spin() = 0 3 2 3 0 3 2 1 0 = T . (55)

P(x) x

+ x Q(x,)

(b)
2r cos /2 /2 /2

Q(x,)
2r sin /2

C r

x1

x2

P(x)

Figure 12. The rotation angle is positive as shown, obeying the right-hand screw rule about the rotation axis.

The space of all s form the Lie algebra so(3). Premultiplication of an arbitrary 3-vector v by is equivalent to the cross product of and v: v = v = Spin() v = Spin(v) = v . (56)

In particular = 0, and vT v = 0, as may be directly veried. The operation converse to (55) extracts the 3-vector , called axial vector or pseudovector, from a given spin tensor: = axial() = The length of this vector is denoted by :
2 2 2 = || = + 1 + 2 + 3 .

1 2 3

(57)

(58)

As general notational rule, we will use corresponding upper and lower case symbols for the spinor and its axial vector, respectively, if possible. For example, N and n, and , V and v. Exceptions are made in case of notational conicts; for example the spinor built from the quaternion vector pi is not denoted as P because that symbol is reserved for projectors. A.3 Spinor Normalizations

As noted, and can be multiplied by a nonzero scalar factor to obtain various normalizations. In general has the form g()/, where g(.) is a function of the rotation angle . The goal of normalization is to simplify the connections Rot and Skew to the rotator, to avoid singularities for special angles, and to connect the components 1 , 2 and 3 closely to the rotation amplitude. This section overviews some normalizations that have practical or historical importance. Choosing = 1/ we obtain the unit axial-vector and unit spinor, which are denoted by n and N, respectively: n= n1 n2 n3 = 1 / 2 / 3 / = , N = Spin(n) = = 0 n3 n 2 n 3 0 n3 n2 n 1 0 . (59)

Taking = tan 1 / is equivalent to multiplying the n i by tan 1 . We thus obtain the parameters bi = n i tan 1 , 2 2 2 i = 1, 2, 3 attributed to Rodrigues [69] by Cheng and Gupta [20]. These are collected in the Rodrigues axial-vector b with associated spinor : b = tan 1 n = (tan 1 /) , 2 2 = Spin(b) = tan 1 N = (tan 1 /) . 2 2 (60)

25

This representation permits an elegant formulation of the rotator via the Cayley transform studied later. However, it collapses as nears 180 since tan 1 . One way to circumvent the singularity is through the use of the 2 four Euler-Rodrigues parameters, also called quaternion coefcients: p0 = cos 1 , 2 pi = n i sin 1 = i / sin 1 , 2 2 i = 1, 2, 3. (61)

2 2 2 2 under the constraint p0 + p1 + p2 + p3 = 1. This set is often used in multibody dynamics, robotics and control. It comes at the cost of carrying along an extra parameter and an additional constraint.

A related singularity-free normalization introduced by Fraeijs de Veubeke [25] takes = sin 1 / and is equivalent 2 to using only the last three parameters of (61): p = sin 1 n = (sin 1 /) , 2 2 p = Spin(p) = sin 1 N = (sin 1 /) . 2 2 (62)

Fraeijs de Veubeke calls this representation Rodrigues-Hamilton without explanatory references. Finally, an important normalization that preserves three parameters while avoiding singularities is that associated with the exponential map. Introduce a rotation vector dened as = n = (/) , = Spin() = N = (/) . (63)

For this normalization the angle is the length of the rotation vector: = || = sign of is a matter of convention. A.4 Spectral Properties

2 2 2 1 + 2 + 3 . The selection of the

The study of the spinor eigensystem zi = i zi is of interest for various developments. Begin by forming the characteristic equation (64) det( I) = 3 2 = 0 where I denotes the identity matrix of order 3. It follows that the eigenvalues of are 1 = 0, 2,3 = i. Consequently is singular with rank 2 if = 0, whereas if = 0, is null. The eigenvalues are collected in the diagonal matrix = diag(0, i, i) and the corresponding right eigenvectors zi in columns of Z = [ z1 z2 z3 ], so that Z = Z. A cyclic-symmetric expression of Z, obtained through Mathematica, is 1 1 s 2 + i(2 3 ) 1 s 2 i(2 3 ) (65) Z = 2 2 s 2 + i(3 1 ) 2 s 2 i(3 1 ) 3 3 s 2 + i(1 2 ) 3 s 2 i(1 2 ) where s = 1 + 2 + 3 . Its inverse is
2 2 2 2 2 + 3 2 + 3 1 2 2 s 2 + i( ) 1 s i(2 3 ) 1 2 3 2 2 2 2 1 3 + 1 3 + 1 1 1 1 (66) Z = 2 2 2 2 2 2 2 s i(3 1 ) 2 s + i(3 1 ) 2 2 2 2 1 + 2 1 + 2 1 3 1 2 s 2 i( ) 2 s 2 + i( ) 3 1 2 3 1 2 The real and imaginary part of the eigenvectors z2 and z3 are orthogonal. This is a general property of skewsymmetric matrices; cf. Bellman [6, p. 64]. Because the eigenvalues of are distinct if = 0, an arbitrary matrix function F() can be explicitly obtained as

1 2

F() = Z

f (0) 0 0

0 f (i) 0

0 0 f (i)

Z1 ,

(67)

where f (.) denotes the scalar version of F(.). One important application of (67) is the matrix exponential, for which f (.) e(.) . The square of , computed through direct multiplication, is =
2 2 2 2 + 3 1 2 1 3

1 2 2 2 3 + 1 2 3

1 3 2 3 2 2 1 + 2

= T 2 I = 2 (nnT I).

(68)

26

This is a symmetric matrix of trace 22 whose eigenvalues are 0, 2 and 2 . By the Cayley-Hamilton theorem, satises its own characteristic equation (64): 3 = 2 , 4 = 2 2 , . . . and generally n = 2 n2 , n 3. (69)

Hence if n = 3, 5, . . . the odd powers n are skew-symmetric with distinct purely imaginary eigenvalues, whereas if n = 4, 6 . . ., the even powers n are symmetric with repeated real eigenvalues. The eigenvalues of I + and I , are (1, 1 i) and (1, 1 i), respectively. Hence those two matrices are guaranteed to be nonsingular. This has implications in the Cayley transform (82). Example. Consider the pseudo-vector = [ 6 2 3 ]T , for which = 62 + 22 + 32 = 7. The associated spin matrix and its square are = 0 3 2 3 0 6 2 6 0 , 2 = 13 12 18 12 45 6 18 6 40 , 3 = 49 , . . . (70)

The eigenvalues of are (0, 7i, 7i) while those of 2 are (0, 7, 7). A.5 From Spinors To Rotators

Referring to Figure 12, a rotator is a linear operator that maps a generic point P(x) to Q(x ) given the rotation axis and the angle . We consider only rotator representations in the form of 3 3 rotation matrices R, dened by x = R x, x = RT x . (71)

The rotation matrix is proper orthogonal, that is, RT R = I and det(R) = +1. It must reduce to I if the rotation vanishes. The space of all Rs form the rotation group SO(3). Remark 5. The denition (71) is taken to agree with the convention for positive rotation angle illustrated in Figure 12 and the denition of Spin() given in (55). Several books, e.g. Goldstein [35], dene as spin matrix the transpose of ours. The denition used here agrees with that of the historical review article by Cheng and Gupta [20]. Readers consulting the literature or implementing nite rotation analysis are advised to check sign conventions carefully. A recommended verication is to work out the 2D case by hand, since the specialization to plane rotation should produce well known coordinate transformations. A.6 Rotator Parametrizations

A key attribute of R is the trace property trace(R) = 1 + 2 cos , (72)

proofs of which may be found for example in Goldstein [35, p. 123] or Hammermesh [36, p. 326]. (It follows from the fact that the eigenvalues of R are 1, ei and ei and their sum is (72).) The problem considered here is the construction of R from rotation data. The inverse problem: given R, extract spin and/or rotation angles, is treated later. Now if R is assumed to be analytic in it must have the Taylor expansion R = I + c1 + c2 2 + c3 3 + . . ., where all ci must vanish if = 0. But because of the Cayley-Hamilton theorem (69), all powers of order 3 or higher may be eliminated. Thus R must be a linear function of I, and 2 . For convenience this will be written R = I + ( ) + ( )2 . (73)

Here is the spinor normalization factor whereas and are scalar functions of and of invariants of or . Since the only invariant of the latter is we may anticipate that = (, ) and = (, ), both vanishing if = 0. Two techniques to determine those coefcients for = 1 are discussed in the next subsections. Table 5 collects several representations of a rotator in terms of the scaled , used by different authors.

27

Table 5. Rotator Forms for Several Spinor Normalizations Parametrization None (unscaled) Unit axial-vector 1 sin sin 2 cos2 1 2 2 cos 1 2 sin 2 sin2 1 2 2 2 sin2 1 2 2 cos2 1 2 2 2 sin2 1 2 2 Spinor N = = Rotator R 2 sin2 1 2 2 I + sin + 2 I + sin N + 2 sin2 1 N2 , 2 I + 2 cos2 1 ( + 2 ) = (I+)(I)1 2 2 sin2 1 2 2 = e = e N I + sin + 2

1 tan 1 2 Rodrigues-Cayley 1 sin 2 Fraeijs de Veubeke Exponential map

p = I + 2 cos 1 p + 22 p 2 =

A.6.1

Rotator From Algebra

It is possible to nd and for = 1 (the unscaled spinor) directly from algebraic conditions. Taking the trace of (73) for = 1 and applying the property (72) requires 3 2 = 1 + 2 cos ,
2

2 sin2 1 1 cos 2 whence = = . 2 2

(74)

The orthogonality condition I = RT R = (I + 2 )(I + + 2 ) = I + (2 2 )2 + 2 4 = I + (2 2 2 2 )2 leads to 2 2 2 2 = 0, Therefore whence = sin . (75)

2 sin2 1 2 sin 1 cos 2 sin 2 =I+ . (76) + + 2 2 From a numerical standpoint the sine-squared form should be preferred to avoid the cancellation in computing 1 cos for small . Replacing the components of and 2 gives the explicit rotator form R=I+

R=

1 2 2 2 21 2 sin2 1 + 3 sin 2 + (3 + 1 ) cos 22 3 sin2 1 1 sin . 2 2 2 2 2 2 21 3 sin2 1 2 sin 22 3 sin2 1 + 1 sin 3 + (1 + 2 ) cos 2 2

2 2 2 1 + (2 + 3 ) cos

21 2 sin2 1 3 sin 21 3 sin2 1 + 2 sin 2 2

(77)

If = 1 but nonzero, the answers are = sin /( ) and = (1 cos )/( 2 2 ). It follows that (76) and (77) are independent of , as was to be expected. A.6.2 Rotator from Geometry

The vector representation of the rigid motion depicted in Figure 12 is x = x cos + (n x) sin + n(n x)(1 cos ) = x + (n x) sin + n (n x) (1 cos ). (78)

where n is normalized to unit length as per (63). This can be recast in matrix form by substituting n x Nx = x/ and x = Rx. On cancelling x we get back (76). A.7 Rotators for All Seasons

If is unit-length-normalized to n as per (59), = 1/ and R = I + sin N + (1 cos ) N2 . This is the matrix form of (78). Because N2 = nnT I, an ocassionally useful variant is R = R + (1 cos ) n nT , R = cos I + sin N. (79)

28

In terms of the three Rodrigues-Cayley parameters bi introduced in (60), = = 2 cos2 1 and R = I + 2 2 cos2 1 ( + 2 ). This can be explicitly worked out to be 2 1 R= 2 2 2 1 + b 1 + b2 + b 3
2 2 2 1 + b1 b2 b3 2(b1 b2 b3 ) 2(b1 b3 + b2 ) 2 2 2 2(b1 b2 + b3 ) 1 b1 + b2 b3 2(b2 b3 b1 ) 2 2 2 2(b1 b3 b2 ) 2(b2 b3 + b1 ) 1 b1 b2 + b3

(80)

This form was rst derived by Rodrigues [69] and used by Cayley [19, p. 332336] to study rigid body motions. It has the advantage of being obtainable through an algebraic matrix expression: the Cayley transform, presented below. It becomes indeterminate, however, as 180 , since all terms approach 0/0. This indeterminacy is avoided by using the four Euler-Rodrigues parameters, which are also the quaternion coefcients, dened in (61). In terms of these we get 2 2 p0 + p1 1 p1 p2 p0 p3 p1 p3 + p0 p2 2 2 2 R = 2 p1 p2 + p0 p3 p0 + p2 1 p2 p3 p0 p1 (81) 2 2 2 p1 p3 p0 p2 p2 p3 + p0 p1 p0 + p3 1 2 This expression cannot become singular. This is paid, however, at the cost of carrying along an extra parameter in 2 2 2 2 addition to the constraint p0 + p1 + p2 + p3 = 1. The normalization of Fraeijs de Veubeke [25] introduced in (62): pi = (i /) sin 1 , leads to = 2 cos 1 and 2 2 = 2. Hence R = I + 2 cos 1 p + 22 , with p = (sin 1 /). p 2 2 A.8 The Cayley Transform

Given any skew-symmetric real matrix = T , we can apply the transformation Q = (I + )(I )1 . (82)

[Note: this mapping can be written in many different ways; some authors use (I )(I + )1 , some switch I and , while others prefer the involutory form (I + )1 (I ).] Q is a proper orthogonal matrix, that is QT Q = I and det Q = +1. This is stated in several texts, e.g., Gantmacher [33, p. 288] and Turnbull [84, p. 156] but none gives a proof for general order. Here is the orthogonality proof: QT Q = (I + )1 (I )(I + )(I )1 = (I + )1 (I + )(I )(I )1 = I2 = I because I + and I commute. The property det Q = +1 can be shown to hold from the spectral properties of skew-symmetric matrices. The inverse transformation = (Q + I)1 (Q I), (83)

produces skew-symmetric matrices from a source proper-orthogonal matrix Q. Equations (82) and (83) are called the Cayley transforms after Cayley [20]. These formulas are sometimes useful in the construction of approximations for moderate rotations. An interesting question is: given and , can (82) be used to produce the exact R? The answer is: yes, if is scaled by a specic = (, ). We thus investigate whether R = (I + )(I )1 exactly for some . Premultiplying both sides by I and representing R by (73) we require (I )(I + + 2 2 ) = I + ( + 2 ) + 2 ( )2 = I + (84)

Identifying coefcients and assuming = 0 we get the conditions = and 1 + 2 2 = 1, from 1 which = = 2/(1 + 2 2 ). Equating to = 2 sin2 2 /( 2 2 ) and solving for gives as only solutions 1 = tan 2 /. Adopting the + sign, this becomes the normalization (60). Consequently R = (I + )(I )1 , The explicit calculation of R in terms of the bi leads to (80). = tan 1 2 . (85)

29

A.9

Exponential Map

This is a nal representation of R that has theoretical and practical importance. Given a skew-symmetric real matrix W, the matrix exponential Q = eW = Exp(W) (86) is proper orthogonal. Here is the simple proof of Gantmacher [33, p 287]. First, QT = Exp(WT ) = Exp(WT ) = = trace(W) = 0. The eigenvalues of Q are i = exp(i ); Q1 . Next, if the eigenvalues of W are i , i i thus det(Q) = i i = exp( i i ) = exp(0) = +1. The transformation (86) is called an exponential map. The converse is of course W = Loge (Q) with the principal value taken. As in the case of the Cayley transform, one may pose the question of whether we can get the rotator R = Exp( ) exactly for some factor = (, ). To study this question we need an explicit form of the exponential. This can be obtained from the spectral form (67) in which the spin function is Exp so that the diagonal matrix entries are 1 and exp( i) = cos i sin . The following approach is more instructive and leads directly to the nal result. Start from the denition of the matrix exponential Exp( ) = I + + 2 2 3 3 + + ... 2! 3! (87)

and use the Cayley-Hamilton theorem (69) to eliminate all powers of order 3 or higher in . Identify the coefcient series of and 2 with those of the sine and cosine, to obtain Exp( ) = I + sin( ) 1 cos( ) 2 . + 2 2 (88)

Comparing this to (76) requires = , or = /. Introducing i = i / and = Spin() = N = (/) as in (63), we nd R = Exp() = I + 2 sin2 1 2 sin sin 1 cos 2 2 =I+ . + + 2 2 (89)

On substituting = N this recovers R = I + sin N + (1 cos ) N2 , as it should. This representation has several advantages: it is singularity free, the parameters i are exactly proportional to the angle, and the differentiation of R is simplied. Because of these favorable attributes the exponential map has become a favorite of implementations where very large rotations may occur, as in orbiting structures and robotics. Remark 6. Some authors state that the exponential map is exact whereas the Cayley transform is a (1,1) Pad e approximation to it. The foregoing treatment shows that the statement is incorrect. Both are exact for specic but different spinor normalizations, and inexact otherwise. A.10 Skew-Symmetric Matrix Relations The following relations involving spinors are useful in some derivations. If v and w are two 3-vectors, and V = Spin(v) and W = Spin(w) the associated spinors, VW WV is skew-symmetric and axial(VW WV) = Vw = Wv. (90)

Let Q be an arbitrary nonsingular 3 3 matrix whereas W = Spin(w) is skew-symmetric. It can be easily veried that QT WQ is skew-symmetric. Then det(Q) Q1 w = axial(QT WQ), det(Q) QT w = axial(QWQT ). (91)

If Q is proper orthogonal, Q1 = QT and det(Q) = 1, in which case QT w = axial(QT WQ), Qw = axial(QWQT ). (92)

The inverse of Q = I + W + W2 , in which W = Spin(w) is skew-symmetric, is Q1 = I + 2 + (w2 1) W+ 2 2 W2 , 2 w 2 + (w 2 1)2 w + (w 2 1)2


2 2 2 with w2 = ||w||2 = w1 + w2 + w3 .

(93)

30

A.11

From Rotators to Spinors

If R(n, ) is given, the extraction of the rotation angle and the unit pseudo-vector n = / is often required. The former is easy using the trace property (72): cos =
1 2

trace(R) 1

(94)

Recovery of n is straightforward from the unit-axial-vector form R = I + sin N + (1 cos )N2 since R RT = 2 sin N, whence n1 R RT (95) N= , n = n 2 = axial(N) 2 sin n3 One issue is the sign of since (94) is satised by . If the sign is reversed, so is n. Thus it is possible to select 0 if no constraints are placed on the direction of the rotation axis. The foregoing formulas are prone to numerical instability for angles near 0 , 180 , etc., because sin vanishes. A robust algorithm is that given by Spurrier [76] in the language of quaternions. Choose the algebraically largest of trace(R) and Rii , i = 1, 2, 3. If trace(R) is the largest, compute p0 = cos 1 = 2
1 2

1 + trace(R),

pi = n i sin 1 = 1 (Rk j R jk )/ p0 , 2 4

i = 1, 2, 3,

(96)

where j and k are the cyclic permutations of i. Otherwise let Rii be the algebraically largest diagonal entry, and again denote by i, j, k its cyclic permutation. Then use pi = n i sin 1 = 2
1 2

Rii + 1 (1 trace(R)), 4 m = j, k.

p0 = cos 1 = 1 (Rk j R jk )/ pi , 2 4

pm = 1 (Rm,i + Ri,m )/ pi , 4

(97)

From p0 , p1 , p2 , p3 it is easy to pass to , n 1 , n 2 , n 3 once the sign of is chosen as discussed above. The theoretical formula for the logarithm of a 3 3 orthogonal matrix is = Loge (R) = arcsin (R RT ), 2 = , N= , (98)

where = 1 ||axial(RRT )||2 . The formula fails, however, outside the range [/2 /2] and is numerically 2 unstable near = 0. A.12 Spinor and Rotator Transformations

Suppose is a spinor and R = Rot() the associated rotator, referred to a Cartesian frame x = {xi }. It is required to transform R to another Cartesian frame x = {xi } related by Ti j = xi / x j , where Ti j are entries of a 3 3 orthogonal matrix T = x/x. Application of (71) yields R = TRTT , T R = TRT TT , R = TT RT, T R = TT R T. (99)

More details may be found in Chapter 4 of [35]. Pre and post-multiplying (95) by T and TT , respectively, yields the T transformed spinor N = T N TT , which is also skew-symmetric because N = T NT TT = N. Likewise for the other spinors listed in Table 5. Relations (92) with Q T show how axial vectors transform. A.13 Axial Vector Jacobian

The Jacobian matrix H() = / of the rotational axial vector with respect to the spin axial vector , and its inverse H()1 = /, appear in the EICR. The latter was rst derived by Simo [71] and Szwabowicz [77], and rederived by Nour-Omid and Rankin [54, p. 377]: H()1 = sin T 1 cos sin 2 sin 1 cos + = I + + . = I+ 2 3 2 3 (100)

31

The last expression in (100), not given by the cited authors, is obtained on replacing T = 2 I + 2 . Use of the inversion formula (93) gives H() = = I 1 + 2 2 with = 1 1 cot( 1 ) 2 2 = 2
1 12

1 2 720

1 4 30240

1 6 1209600

+ ...

(101)

The given in (101) results by simplifying the value = [sin (1 + cos )]/[ 2 sin ] given by previous investigators. Care must be taken on evaluating for small angle because it approaches 0/0; if | | < 1/20, say, the series given above may be used, with error < 1016 when 4 terms are retained. If is a multiple of 2, blows up since cot( 1 ) , and a modulo-2 reduction is required. 2 In the formulation of the tangent stiffness matrix the spin derivative of H()T contracted with a nodal moment vector m is required: L(, m) = in which = d/d 2 + 4 cos + sin 4 = = 4 4 sin2 ( 1 ) 2
1 360

H()T :m = [H()T m] H() = [(T m) I + mT 2m T ] + 2 m T 1 Spin(m) H(). 2 (102) +


1 2 7560

1 4 201600

1 6 5987520

+ ...

(103)

This expression of was obtained by simplifying results given in [54, p. 378]. A.14 Spinor and Rotator Differentiation

Derivatives, differentials and variations of axial vectors, spinors and rotators with respect to various choices of independent variables appear in applications of nite rotations to mechanics. In this section we present only expressions that are useful in the CR description. They are initially derived for dynamics and then specialized to variations. Several of the formulas are new. A.14.1 Angular Velocities We assume that the rotation angle (t) = (t)n(t) is a given function of time t, which is taken as the independent variable. The time derivative of (t) is = axial(). To express rotator differentials in a symmetric manner we and a spinor = Spin(), related to congruentially through the rotator: introduce an axial vector = Spin() = 0 3 2 3 2 0 1 1 0 = RRT , = Spin() = 0 3 2 3 0 1 2 1 0 = RT R. (104)

From (92) it follows that the axial vectors are linked by = RT , = R. (105)

Corresponding relation between variations or differentials, such as = R RT or d = R d RT are incorrect as shown in below. In CR dynamics, is the vector of inertial angular velocities whereas is the vector of dynamic angular velocities. Repeated temporal differentiation of R = Exp() gives the rotator time derivatives R = R, = R , R 2 R = ( + )R, 2 R = R( + ),
T

... ... 3 R = ( + 2 + + )R, . . . ... ... 3


R = R( + 2 + + ), . . .
T

(106) (107)

The following four groupings appear often: R R , R RT , RT R and R R. From the identities RRT = I and RT R = I it can be shown that they are skew-symmetric, and may be associated to axial vectors with physical meaning. For T T example, taking time derivatives of RRT = I yields R R + R RT = 0, whence R R = R RT = (R R)T . Pre T and postmultiplication of R in (106) and (107) by R and R furnishes T R RT = R R = , T RT R = R R = . (108)

Note that the general integral of R = R aside from a constant, is R = Exp(), from which (t) can be extracted. On the other hand there is no integral relation dening (t); only the differential equation R = R .

32

A.14.2

Angular Accelerations

Postmultiplying the second of (106) by RT yields T 2 RRT = + = + ( )2 I = 0 3 3 0 2 1 2 1 0 + 2 2 1 2 2 3 1 3 2 2 1 2 3 1 2 3 2 2 1 3 2 3 1 2 , (109)

T 2 2 2 in which ( )2 = 1 + 2 + 3 . When applied to a vector r, (+ )r = r+ r( )2 r. This operator appears in the expression of particle accelerations in a moving frame. The second and third term give rise to the Coriolis and centrifugal forces, respectively. Premultiplying the second derivative in (107) by R yields RR = + , and so on. A.14.3 Variations

Some of the foregoing expressions can be directly transformed to variational and differential forms while others cannot. For example, varying R = Exp() gives R = R, RT = RT . (110)

This matches R = R and RT = RT from (106) on replacing () by . On the other hand, the counterparts = R and RT = R are not R = R and RT = RT , a point that has tripped authors of (107): R unfamiliar with moving frame dynamics. Correct handling requires the introduction of a third axial vector : R = R , RT = R, in which = Spin() = 0 3 2 3 0 1 2 1 0 (111)

Axial vectors and can be linked as follows. Start from R = R and R = R . Time differentiate the former: R = R + R , and vary the latter: R = R + R , equate the two right-hand sides, premultiply by RT , replace RT R and RT R by and , respectively, and rearrange to obtain = + , or = + = . (112)

The last transformation is obtained by taking the axial vectors of both sides, and using (90). It is seen that and match if and only if and commute. Higher variations and differentials can be obtained through similar techniques. The general rule is: if there is a rotator integral such as R = Exp(), time derivatives can be directly converted to variations. If no integral exists, utmost care must be exerted. Physically quantities such as and are related to a moving frame, which makes differential relations rheonomic.

33

Appendix B. CR MATRICES FOR TRIANGULAR SHELL ELEMENT The element types tested in Part II [38] are shown in Figure 13. Each element has 6 degrees of freedom (DOF) per node: three translations and three rotations. Shells include the drilling DOF. The linear internal force and stiffness matrix of the shell elements are constructed with the Assumed Natural Deviatoric Strain (ANDES) formulation [27 32,53] in terms of the deformational displacements and rotations. ANDES is a direct descendent of the Assumed Natural Strain (ANS) formulation of Park and Stanley [59] and the Free Formulation of Bergan and coworkers [1114]. The derivation of the linear models is outlined in Part II [38]. The matrices required to implement the EICR are T, P, S, G, H and L. None of these depend on how the internal e element force pe and stiffness matrix K of the small-strain linear element are formed. In terms of implementation the EICR matrices can be classied into two groups: (i) T, H and L, as well as the T-projector component Pu of P, are block diagonal matrices built up with 3 3 node blocks. These blocks can be formed by standard modules which are independent of the element type, as long as the element has the standard 6 DOFs per node. The only difference is the number of nodes. The R-projector component of P, which is P = SG, does depend on element type, geometry and choice of CR frame through matrix G. These must be recoded for evry change in those attributes.

(ii)

In this Appendix we give the G and S matrices that appear in the front end of the EICR for the triangular shell element, as that illustrates the effect of CR frame selection. Matrices for the beam and quadrilateral element are given in Part II. In the following sections, element axes labels are changed from {x1 , x2 , x3 } to {x, y , z } to unclutter nodal subscripting. Likewise the displacement components {u 1 , u 2 , u 3 } are relabeled {u x , u y , u z }.

2 1

3 4 3

1 2 2

Figure 13. Elements tested in Part II.

B.1

Matrix S

This 18 3 lever matrix is given by (written in transposed form to save space) S = [ Spin( 1 ) x I Spin( 2 ) x I Spin( 3 ) x I ]T (113)

where I is the 3 3 identity matrix, and xa = [ xa , ya , z a ]T , the position vector of node a in the deformed (current) conguration, measured in the element CR frame. B.2 Matrix G

This 3 18 matrix connects the variation in rigid element spin to the incremental translations and spins at the nodes, both with respect to the CR frame. G decomposes into three 3 6 submatrices, one for each node: r = G v, G = [ G1 G2 G3 ] , v = v1 v2 v3 , va = ua , a a = 1, 2, 3. (114)

Submatrices Ga depend on how the element CR frame is chosen. The origin of the frame is always placed at the element centroid. But various methods have been used to direct the axes {xi }. Three methods used in the present research are described.

34

B.2.1

G by Side Alignment

This procedure is similar to that used by Rankin and coworkers [54,6366]. They select side 13 for x2 and node 1 as frame origin. The approach used here aligns x1 with side 12 and picks the centroid as origin. Then 1 G1 = 2A 0 0 x32 0 0 0 0 0 y32 0 0 0 0 h 3 0 0 0 0 , 1 G2 = 2A 0 0 0 0 x13 0 0 0 0 y13 0 0 0 h3 0 0 0 0 , 1 G3 = 2A 0 0 x32 0 0 0 0 0 y32 0 0 0 0 0 0 0 0 0 ,

(115) where h 3 = 2A/L 3 is the distance of corner 3 to the opposite side, A the triangle area, and L 3 the length of side 12. This choice satises the decomposition property (37) that guards against unbalanced force effects while iterating for equilibrium. On the other hand it violates invariance: the choice of CR frame depends on node numbering, and different results may be obtained if the mesh is renumbered. B.2.2 G by Least Square Angular Fit
R

Nyg rd [57] and Bjrum [18] place C in the plane of the dea formed element with origin at node 1. The inplane orientation of the CR element is determined by a least square t of the side angular errors. Referring to Figure 14 the squared error is 2 2 d 2 = 1 + 2 + 2 . Rotating by an additional angle this be2 comes d ( ) = (1 + )2 +(2 + )2 +( +)2 . Miniminization respect to : d 2 / = 0 yields = (1 + 2 + 3 )/3. Consequently the optimal in-plane position according to this criterion is given by the mean of the side angular errors. This condition yields for the nodal submatrices 0 1 0 Gi = 2A 2A ( s j y + sky ) lj Lk 3

1+

2 1

2 2+ 1

2 3+

Figure 14. Side angular error measure for triangular shell element.

2A ( s j x skx ) Lk 3 lj

0 0

xk j yk j 0

0 0 0 0 0 0. 0 0 0

(116)

where L i is the side length opposite corner i and {skx , sky } the projections of that side on the {x, y} axes. An advantage of this tting method is that it satises invariance with respect to node numbering. On the other hand, the rotator gradient matrix cannot be decomposed as in (37) leading to an approximate projector away from equilibrium. This disadvantage is not serious for a at triangular element, however, since the CR and deformed congurations remain close on the assumption of small membrane strains. A more serious shortcoming is that the procedure d d reintroduces the problem of spurious normal-to-theplane rotations when an element with drilling freer doms is subjected to pure stretch. The difculty is d ~ illustrated in Figure 15, where a two triangle patch Y _ is subject to uniform stretch in the y direction. Un y r d der this state all rotations should vanish. The ele~ d ments do rotate, however, because of the in-plane d X _ skewing of the diagonal. A deformational rotation x Undeformed (base) Deformed and configuration CR configurations is picked up since a predictor step gives no drilling rotations at the nodes, whereas the deformational Figure 15. Finite stretch patch test for two triangle drilling rotation is the total minus the rigid body roelements equipped with corner drilling freedoms. tation: d = r . The problem is analogous to that discussed by Irons and Ahmad [45, p. 289] when dening node drilling freedoms as the mean of rotations of element sides meeting at that node; such elements grossly violate the patch test. This difculty was overcome by Bergan and Felippa [14] by dening the node drilling freedom as the continuum mechanics rotation z = 1 ( u y / x u x / y ) at the node. It is seen that the problem of spurious drilling rotations has been 2 reintroduced for the nonlinear case by the choice of CR frame positioning. In fact this problem becomes even more serious with the side alignment procedure described in the foregoing subsection.

35

B.2.3

Fit According to CST Rotation

The innitesimal drilling rotation of a plane stress CST element (Turner triangle, linear triangle), is given by [27,32] lin C ST =
1 2

uy ux x y

1 x23 u x1 + x31 u x2 + x12 u x3 + y23 u y1 + y31 u y2 + y12 u y3 . 4A

(117)

in which xi j = xi x j , yi j = xi y j , etc. The extension of this result to nite rotations can be achieved through a mean nite strain rotation introduced by Novozhilov [55, p. 31]; cf. also [81, Sec. 36]: tan C ST = lin C ST (1 + x x )(1 + yy ) 1 x2y 4 (118)

in which x x = u x / x, yy = u y / y and x y = u y / x + u x / y are the innitesimal strains (constant over the triangle) computed from the CST displacements. Novozhilov proves that this rotation measure is invariant with respect to the choice of CR axes {x, y } since it is obtained as a rotational mean taken over a 2 sweep about z . If this result is applied to the nite stretch path test of Figure 15 it is found that the CR frames of both elements do not rotate, and the test is passed. The rotation gradient submatrices are 1 Gi = 2A 0 0 1 xk j 2 0 0 1 yk j 2 xk j yk j 0 0 0 0 0 0 0 0 0 0 (119)

where j, k denote cyclic permutations of i = 1, 2, 3. The resulting G matrix satises the geometric separabilty condition. B.2.4 Best Fit by Minimum LS Deformation

The best t solution by least-squares minimization of relative displacements given in Appendix C as equation (127) was obtained in 2000 [29]. Unlike the previous ones it has not been tested as part of a CR shell program. The measure is invariant, but it is not presently known whether it passes the stretch patch test, or if the associated G satises the geometric separability condition (37).

Appendix C. BEST FIT CR FRAME To dene the CR frame of an individual element e we must nd c and R0 from the following information. (In what follows the element index e is suppressed for brevity.) (i) (ii) The geometry of the base element. This is dened through x, which is kept xed. The motion, as dened by the total displacement eld v(x) recorded in the global frame as stated in Table 2.
0

In dynamic analysis there is a third source of data: (iii) Mass distribution. Associated with each x in C there is a volume element d V and a mass density giving a mass element dm = d V . This mass may include nonstructural components. The case of time-varying mass because of fuel consumption or store drop is not excluded. Finding c and R0 as functions of x and v, plus the mass distribution in case of dynamics, is the shadowing problem introduced previously. What criterion should be used to determine c and R0 ? Clearly some nonnegative functional of the motion (but not the motion history) should be minimized. Desirable properties of the functional are: Versatility. It must work for statics, dynamics, rigid bodies and nonstructural bodies (e.g., a fuel tank). These requirements immediately rule out any criterion that involves the strain energy, because it cannot be used for a nonstructural or rigid body. Invariance. Should be insensitive to element node numbering. It would be disconcerting for a user to get different answers depending on how mesh nodes are numbered. Rigid Bodies. For a rigid body, it should give the same results as a body-attached frame. This simplies the coupling of FEM-CR and multibody dynamics codes.

36

Finite Stretch Patch Test Satisfaction. If a group of elements is subjected to a uniform stretch, no spurious deformational rotations should appear. This test is particularly useful when considering shell elements with drilling freedoms, as discussed in Section B.2 for the triangle. Fraeijs de Veubeke [25] proposed two criteria applicable to exible and rigid bodies: minimum kinetic energy of relative motion, and minimum Euclidean norm of the deformational displacement eld. As both lead to similar results for dynamics and the rst one is not applicable to statics, the second criterion is used here. An open research question is whether a universal best t criterion based on the polar decomposition of the element motion could be developed. C.1 Minimization Conditions

In statics the density may be set to unity and mass integrals become volume integrals. These are taken over the base conguration, whence . . . d V means 0 . . . d V . Furthermore the element global frame is made to coincide C with the base frame to simplify notation. Thus x x, u u and a = 0. Coordinates x0 and x R of the base are relabeled X and x = X + u, respectively, to avoid supercripts clashing with transpose signs. Because C 0 is by denition at the base element centroid X d V = 0. The centroid C of C is located by the condition x d V = (X + u) d V = u d V = c V , where V = the element volume. From (22) with a = 0 and x0 X, the best-t functional is
D

(120) d V is

J [c, R0 ] =

1 2

T ud ud d V =

1 2

u c + (I R0 ) X

u c + (I R0 ) X d V

(121)

T Minimizing ud ud leads to the same result, since ud and ud only differ by a rotator. C.2 Best Origin c the distance between C R and C. Taking the variation of J with respect to c yields (c)T ud d V = cT c d V = cV = u c + (I R0 ) X d V = cT c) V.
R

Denote by

J c = c

(u c) d V = 0, (122)

u d V = (c +

where (120) has been used. Consequently C.3 Best Rotator

c = 0. That is, the centroids of C and C must coincide: C R C.

The variation with respect to R0 gives the condition J R0 = R0 XT R0 u c + (I R0 ) X d V = 0. XT R0 c d V = 0 yields (124) (123)

Replacing R0 = R0 and noting that XT X = 0 and XT R0 X + u d V =

XT R(X + u) d V = 0.

Applying (124) to each angular variation: 1 , 2 and 3 of in turn, provides three nonlinear scalar equations to determine the three parameters that dene R0 . For the two-dimensional case, in which only 3 is varied and R0 depends only on = 3 , the following equation is obtained: (X 1 + u 1 )(x2 cos 3 X 1 sin 3 ) (X 2 + u 2 )(x1 cos 3 + X 2 sin 3 ) d V = 0, (125)

37

whence tan 3 =

(X 2 u 1 X 1 u 2 ) d V X 1 (X 1 + u 1 ) + X 2 (X 2 + u 2 ) d V

(X 2 u 1 X 1 u 2 ) d V (X 1 x1 + Y2 x2 ) d V

(126)

For the three-dimensional case a closed form solution is not available. This is an open research topic. Fraeijs de Veubeke [25] reduces (124) to an eigenvalue problem in the axial vector of R0 , with a subsidiary condition handled by a Lagrange multiplier. Rankin [67] derives a system equivalent to (124) for an individual nite element by lumping the mass equally at the nodes whereupon the integral over the body is reduced to a sum over nodes. The nonlinear system is solved by Newton-Raphson iteration. C.4 Linear Triangle Best Fit

Consider a constant thickness three-node linear plane stress triangle (also known as Turner triangle and CST), displacing in two dimensions as illustrated in Figure 16. To facilitate node subscripting, rename X 1 X , X 2 Y , x1 x, x2 y, u 1 u x , u 2 u y . Using triangular coordinates {1 , 2 , 3 } position coordinates and displacements are interpolated linearly: X = X 1 1 + X 2 2 + X 3 3 , Y = Y1 1 + Y2 2 + Y3 3 , x = x1 1 + x2 2 + x3 3 , y = y1 1 + y2 2 + y3 3 , u x = u y1 1 + u x2 2 + u x3 3 and u y = u y1 1 + u y2 2 + u y3 3 , in which numerical subscripts now denote node numbers. Insert into (126) and introduce side conditions X 1 + X 2 + X 3 = 0, Y1 + Y2 + Y3 = 0, u x1 + u x2 + u x3 = x1 + x2 + x3 and u y1 + u y2 + u y3 = y1 + y2 + y3 for centroid positioning. Mathematica found for the best t angle x1 Y1 + x2 Y2 + x3 Y3 y1 X 1 y2 X 2 y3 X 3 tan 3 = . (127) x1 X 1 + x2 X 2 + x3 X 3 + y1 Y1 + y2 Y2 + y3 Y3 where xa = xa + u xa , ya = ya + u ya , a = 1, 2, 3. The same result can be obtained with any 3-point integration rule where the Gauss points are placed on the medians. In particular, Rankins idea of using the three corners as integration points [67] gives the same answer.
Y, y, ~ y 2 (X 2 ,Y2 ) 2 (x 2 ,y2)

u2

global frame = element base frame for convenience

_
y

x element CR frame 3

3 (X3 ,Y3 )

C0

~ X, x, x

u3 u1
1 (X 1 ,Y1 )

C
3 (x 3 ,y3)

1 (x 1 ,y1)

Figure 16. Best CR frame t to 3-node linear triangle in 2D. Axes relabeled as explained in the text.

This solution is interesting in that it can be used as an approximation for any 3-node triangle undergoing three dimensional motion (for example, a thin shell facet element) as long as axis x3 is preset normal to the plane passing through the three deformed corners. The assumption is that the best t is done only for the in-plane displacements. C.5 Rigid Body Motion Verication

Suppose the element motion u = ur is rigid (ud = 0) with rotator Rr . If so ur = c (I Rr )X. Inserting into (121) yields J=
1 2

XT (Rr R0 )T (Rr R0 )X dV.

(128)

in which X = xG a. This is clearly minimized by R0 = Rr , so the body attached frame is a best t CR frame. (There may be additional solutions if (Rr R0 )X 0 for some R0 = Rr , for example by relabeling axes.) This check veries that the functional (121) also works as expected for rigid body components of a structural model.

38

Appendix D. NOMENCLATURE The notation used by different investigators working in CR formulations has not coalesced, since the topic is in ux. This Appendix identies the symbols used in the present article. General scheme of notation is Non-bold letters: scalars. Uppercase bold letters (roman and greek): matrices and tensors. Lowercase bold letters (roman and greek): column vectors, but there are occasional exceptions such as X and Y. Symbol a b bi ci d e ei j fi hi i j k l m n ni pi r skx , sky t u ui , ui , ui u di , u di , u di u ri , u ri , u ri x xi xi xi y z A C C 0 ,C R ,C L Lk N Ne O T U W X Xi Y Z a b Meaning As subscript: generic nodal index As subscript: generic nodal index Components of Rodrigues-Cayley axial vector. Generic coefcients. Components of vector c. As subscript: deformational Base of natural logarithms. As superscript: element index. Strains. Components of force vector f Triangle heights. As subscript: generic index. Imaginary unit in complex numbers. As subscript: generic index As superscript: iteration count As subscript: generic index As subscript: generic index As subscript: incremental step number Components of vector n. Quaternion components. As subscript: rigid Projections of triangle side L k on {x, y} axes. Time. Displacement magnitude Displacement components in xi , xi , xi axes, respectively. Deformational displacement components in xi , xi , xi axes, respectively. Rigid displacement components in xi , xi , xi axes, respectively. x1 axis when using {x, y, z} notation Components of global-frame position vector x Components of element base-frame position vector x. Components of element CR-frame position vector x. x2 axis when using {x, y, z} notation x3 axis when using {x, y, z} notation Area. Generic centroid of body or element in statics. Center of mass in dynamics. Body or element centroid in base, CR and deformed congurations, respectively. Length of 2-node element. Generic length. Length of triangle side opposite node k. Number of structure nodes. Number of nodes of element e. Origin of global frame. Kinetic energy Internal energy, strain energy. External work. x1 axis when using {X, Y, Z } notation Material global Cartesian axes x2 axis when using {X, Y, Z } notation x3 axis when using {X, Y, Z } notation Position vector of C 0 from origin of global frame. Generic vector. Position vector of C R C from origin of global frame. Generic vector.

39

c d d e f f fb fu i m ma n na 0 p pe pe q r u v

w v x x x x y y zi A Bd C D E F Fn , Fnm G Ga H Ha H() I J Jab K Ke e K e KR KG M KG P KG R

Displacement of base element centroid C 0 to C R C. Deformational displacement. Array of nodal DOF, collecting translations and rotations. Array of nodal deformational DOF, collecting translations and rotations. Strains arranged as vector. Generic force vector. External force vector. External force vector in element CR frame. Balance (self-equilibrated) force vector. Unbalanced (out of equilibrium) force vector. Base unit vector Generic nodal moment vector. Nodal moment components at node a. Unit normal. Direction of rotation vector in 3D. Translational force components at node a. Null matrix or vector. Internal force vector of structure. Element internal force vector in global frame. Element internal force vector in CR frame. Not used Force residual vector. Displacement vector. Also generic vector: meaning from context. Denotes vector collecting nodal displacements and rotations, displacements and spins, or displacements and rotators. See Table 2. Also generic vector: meaning from context. Generic vector: meaning from context. Global DOF vector for complete structure. Global-frame spatial position vector. Element base-frame position vector. Element CR-frame position vector. Frame independent position vector. Alternative notation for position vector. Generic vector. Frame independent position vector. Spinor eigenvector. Generic matrix. Matrix relating strains to deformational displacements. Generic matrix. Strain-stress (compliance) matrix Diagonal matrix of zeros and ones. Generic diagonal matrix, size from context. Stress-strain (elasticity) matrix Generic matrix function Auxiliary matrices in tangent stiffness derivations. e Element spin-tter matrix linking r to ve . Component of G associated with node a. Block diagonal matrix built with Ha and I blocks. Evaluation of H() at node a. Jacobian of with respect to . Identity matrix, size from context. Jacobian matrix in general. Jacobian matrix relating quantities at nodes a and b. Tangent stiffness matrix of structure. Element tangent stiffness matrix in global frame. Linear element stiffness matrix in local CR frame. Element tangent stiffness matrix in local CR frame. Moment correction geometric stiffness matrix. Equilibrium projection geometric stiffness matrix. Rotational geometric stiffness matrix.

40

KM L La L(, m) M N P Pu P Q R R0 R0 , R0 S Sa T T0 TR Uab V W X X X X Y Z ab i i j ,,,


a, b

Material stiffness matrix. Block diagonal built with La and 0 blocks. Evaluation of L(, m) at node a. Contraction of H()T /) with vector m Mass matrix. Spinor for vector n. Projector matrix. Translational projector matrix, a.k.a. T-projector. Rotational projector matrix, a.k.a. R-projector. Generic orthogonal matrix. Orthogonal rotation matrix: the matrix representation of a rotator. Transformation rotator between element base frame and CR frame. Rotator R0 referred to element base and CR frames, respectively. Spin-lever (a.k.a. moment-arm) matrix built up of Sa blocks. Generic skew-symmetric matrix. Spin-lever matrix for node a. Element CR-to-global transformation matrix built from T R blocks. Transformation rotator from element base frame to global frame. Transformation rotator from element CR frame to global frame. Building block of translational projector Pu . Generic skew-symmetric matrix. Generic skew-symmetric matrix. Global position vector with X i as components. Coordinate free material position vector. Element base position vector withxi as components. Element CR position vector with xi as components. Position vector collecting yi as components. Matrix of right eigenvectors of spinor. Coefcient in parametrized rotator representation. Coefcient in parametrized rotator representation. Spinor normalization factor. Variation symbol. Kronecker delta. Small scalar Triangular coordinates Coefcient in H(). Rotation angle in general. Magnitude of rotation vector . Curvature Lagrange multiplier Coefcient in L(, m). Poissons ratio Mass density. Stresses. Coefcient in Loge (R) formula. Generic angle symbols. Generic angle. Magnitude of spin vector . Summation symbol. Ne Ne Abbreviations for a=1 and b=1 . Rotation axial vector (a.k.a. rotation pseudovector). Axial vector built from i angles. Axial vector built from i angles. Stresses arranged as vector. Initial stresses (in the base conguration) arranged as vector. Spin axial vector (a.k.a. spin pseudovector).

41

CR DOF EICR Rotator Spinor TL UL (.) (.)0 or (.)0 (.) D or (.) D (.)G or (.)G (.) R or (.) R (.)d (.)e (.)r ()T ()1 axial(.) e(.) Exp(.) Loge (.) Rot(.) Skew(.) Spin(.) trace(.)

Auxiliary matrix used in the decomposition of G Spinor built from rotation axial vector. Diagonal matrix of eigenvalues of . Auxiliary matrix used in the decomposition of G. Spinor built from axial vector . Spinor built from axial vector . Spinor built from the Rodrigues-Cayley parameters bi . Spinor built from spin axial vector . Spinor build from quaternion parameters pi . Conguration: see Table 1 for further identication by superscripts. Abbreviation for corotational (a.k.a. corotated) kinematic description. Abbreviation for degree of freedom. Abbreviation for Element Independent Corotational formulation. Abbreviation for rotation tensor (or 3 3 rotation matrix). Abbreviation for spin tensor (or 3 3 spin matrix). Abbreviation for Total Lagrangian kinematic description. Abbreviation for Updated Lagrangian kinematic description. Abbreviation for d(.)/dt 0 (.) pertains to base (initial) conguration C . D (.) pertains to deformed conguration C . G (.) pertains to globally aligned conguration C . R (.) pertains to corotated conguration C . (.) is deformational. (.) pertains to element e. (.) is rigid. Matrix or vector transposition. Matrix inverse. Extraction of axial vector from skew-symmetric matrix argument Matrix exponential if (.) is a square matrix Alternative form for matrix exponential Matrix natural logarithm Construction of rotator from spinor argument Extraction of spinor from rotator argument Construction of spinor from axial vector argument Sum of diagonal entries of matrix argument

42

Acknowledgements Preparation of this paper was supported by the National Science Foundation under Grant CMS0219422, by the Finite Elements for Salinas contract with Sandia National Laboratories, and by the Air Force Ofce of Scientic Research under grant F49620-99-1-0007. References [1] [2] [3] [4] [5] [6] [7] [8] [9] J. H. Argyris, An excursion into large rotations, Comp. Meths. Appl. Mech. Engrg., 32, 85155, 1982. J.-M. Battini, C. Pacoste, Co-rotational beam elements with warping effects in instability problems, Comp. Meths. Appl. Mech. Engrg., 191, 17551789, 2002. J.-M. Battini, C. Pacoste, Plastic instability of beam structures using co-rotational elements, Comp. Meths. Appl. Mech. Engrg., 191, 58115831, 2002. K.-J. Bathe, Finite Element Procedures in Engineering Analysis, Prentice-Hall, 1982. K.-J. Bathe, E. Ramm and E. L. Wilson, Finite element formulations for large deformation dynamic analysis, Int. J. Numer. Meth. Engrg., 7, 255271, 1973. R. Bellman, Introduction to Matrix Analysis, McGraw-Hill, New York, 1960. T. Belytschko and B. J. Hsieh, Nonlinear transient nite element analysis with convected coordinates, Int. J. Numer. Meth. Engrg., 7, 255-271, 1973. T. Belytschko and B. J. Hsieh, Application of higher order corotational stretch theories to nonlinear nite element analysis, Computers & Structures, 11, 175182, 1979. T. Belytschko and B. J. Hsieh, An overview of semidiscretization and time integration operators, Chapter 1 in Computational Methods for Transient Analysis, T. Belytschko and T. J. R. Hughes, (eds.), North-Holland, Amsterdam, 166, 1983. P. G. Bergan and G. Horrigmoe, Incremental variational principles and nite element models for nonlinear problems, Comp. Meths. Appl. Mech. Engrg., 7, 201217, 1976. P. G. Bergan and L. Hanssen, A new approach for deriving good nite elements, MAFELAP II Conference, Brunel University, 1975, in The Mathematics of Finite Elements and Applications Volume II, ed. by J. R. Whiteman, Academic Press, London, 483497, 1976. P. G. Bergan, Finite elements based on energy orthogonal functions, Int. J. Numer. Meth. Engrg., 15, 1141 1555, 1980. P. G. Bergan and M. K. Nyg rd, Finite elements with increased freedom in choosing shape functions, Int. J. a Numer. Meth. Engrg., 20, 643664, 1984. P. G. Bergan and C. A. Felippa, A triangular membrane element with rotational degrees of freedom, Comp. Meths. Appl. Mech. Engrg., 50, 2569, 1985. P. G. Bergan and M. K. Nyg rd, Nonlinear shell analysis using Free Formulation nite elements, in Finite a Element Methods for Nonlinear Problems, Springer Verlag, Berlin, 317-338, 1989. M. A. Biot, The Mechanics of Incremental Deformations, McGraw-Hill, New York, 1965. R. B. Bird, R. C. Amstrong and O. Hassager, Dynamics of Polymeric Liquids. Vol 1: Fluid Dynamics, Wiley, New York, 1977. R. O. Bjrum, Finite element formulations and solution algorithms for buckling and collapse analysis of thin shells. Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1992. A. Cayley, Collected Works, Cambridge Univ. Press, 1898. H. Cheng and K. C. Gupta, An historical note on nite rotations, J. Appl. Mech., 56, 139145, 1989. M. Chiesa, B. Skallerud and D. Gross, Closed form line spring yield surfaces for deep and shallow cracks: formulation and numerical performance, Computers & Structures, 80, 533545, 2002. M. A. Criseld, A consistent corotational formulation for nonlinear three-dimensional beam element, Comp. Meths. Appl. Mech. Engrg., 81, 131150, 1990. M. A. Criseld and G. F. Moita, A unied co-rotational for solids, shells and beams, Int. J. Solids Struc., 33, 29692992, 1996.

[10] [11]

[12] [13] [14] [15] [16] [17] [18]

[19] [20] [21] [22] [23]

43

[24] M. A. Criseld, Nonlinear Finite Element Analysis of Solids and Structures. Vol 2: Advanced Topics, Wiley, Chichester, 1997. [25] B. M. Fraeijs de Veubeke, The dynamics of exible bodies, Int. J. Engrg. Sci., 14, 895-913, 1976. [26] C. Farhat, P. Geuzaine and G. Brown, Application of a three-eld nonlinear uid-structure formulation to the prediction of the aeroelastic parameters of an F-16 ghter, Computers and Fluids, 32, 329, 2003. [27] C. A. Felippa and C. Militello, Membrane triangles with corner drilling freedoms: II. The ANDES element, Finite Elem. Anal. Des., 12, 189201, 1992. [28] C. A. Felippa and S. Alexander, Membrane triangles with corner drilling freedoms: III. Implementation and performance evaluation, Finite Elem. Anal. Des., 12, 203239, 1992. [29] C. A. Felippa, A systematic approach to the Element-Independent Corotational dynamics of nite elements. Report CU-CAS-00-03, Center for Aerospace Structures, College of Engineering and Applied Sciences, University of Colorado at Boulder, January 2000. [30] C. A. Felippa and K. C. Park, The construction of free-free exibility matrices for multilevel structural analysis, Comp. Meths. Appl. Mech. Engrg., 191, 21112140, 2002. [31] C. A. Felippa, Recent advances in nite element templates, Chapter 4 in Computational Mechanics for the Twenty-First Century, ed. by B.H.V. Topping, Saxe-Coburn Publications, Edinburgh, 7198, 2000. [32] C. A. Felippa, A study of optimal membrane triangles with drilling freedoms, Comp. Meths. Appl. Mech. Engrg., 192, 21252168, 2003. [33] F. R. Gantmacher, The Theory of Matrices, Vol. I, Chelsea, New York, 1960. [34] M. Geradin, Finite element approach to kinematic and dynamic analysis of mechanisms using Euler parameters, in Numerical Methods for Nonlinear Problems II, ed. by C. Taylor, E. Hinton and D. R. J. Owen, Pineridge Press, Swansea, 1984. [35] H. Goldstein, Classical Mechanics, Addison-Wesley, Reading, Massachussets, 1950. [36] M. Hammermesh, Group Theory and Its Application to Physical Problems, Dover, New York, 1989. [37] B. Haugen, Buckling and stability problems for thin shell structures using high-performance nite elements, Ph. D. Dissertation, Dept. of Aerospace Engineering Sciences, University of Colorado, Boulder, CO, 1994. [38] B. Haugen and C. A. Felippa, A unied formulation of small-strain corotational nite elements: II. Applications to shells and mechanisms, in preparation. [39] G. Horrigmoe, Finite element instability analysis of free-form shells, Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1977. [40] G. Horrigmoe and P. G. Bergan, Instability analysis of free-form shells by at nite elements, Comp. Meths. Appl. Mech. Engrg., 16, 1135, 1978. [41] A. S. Householder, The Theory of Matrices in Numerical Analysis, Blaisdell, New York, 1964. [42] T. J. R. Hughes and W.-K. Liu, Nonlinear nite element analysis of shells: Part I. Three-dimensional shells, Comp. Meths. Appl. Mech. Engrg., 26, 331362, 1981. [43] T. J. R. Hughes and W.-K. Liu, Nonlinear nite element analysis of shells: Part II. Two-dimensional shells, Comp. Meths. Appl. Mech. Engrg., 27, 167182, 1981. [44] T. J. R. Hughes and J. Winget, Finite rotation effects in numerical integration of rate constitutive equations arising in large deformation analysis, Int. J. Numer. Meth. Engrg., 15, 18621867, 1980. [45] B. M. Irons and S. Ahmad, Techniques of Finite Elements, Ellis Horwood Ltd, 1980. [46] B. Kr keland, Large displacement analysis of shells considering elastoplastic and elastoviscoplastic materials, a Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1977. [47] E. Levold, Solid mechanics and material models including large deformations, Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1990. [48] K. M. Mathisen, Large displacement analysis of exible and rigid systems considering displacement-dependent loads and nonlinear constraints. Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1990. [49] K. M. Mathisen, T. Kvamsdal and K. M. Okstad, Adaptive strategies for nonlinear nite element analysis of shell structures, In: Numerical Methods in Engineering 92, C. Hirchs et al. (Eds.), Elsevier Science Publishers B. V., 1992.

44

[50] K. Mattiasson, On the corotational nite element formulation for large deformation problems, Dr. Ing. Thesis, Dept. of Structural Mechanics, Chalmers University of Technology, G teborg, 1983. o [51] K. Mattiasson and A. Samuelsson, Total and updated Lagrangian forms of the co-rotational nite element formulation in geometrically and materially nonlinear analysis, in Numerical Methods for Nonlinear Problems II, ed. by C. Taylor, E. Hinton and D. R. J. Owen, Pineridge Press, Swansea, 134151, 1984. [52] K. Mattiasson, A. Bengston and A. Samuelsson, On the accuracy and efciency of numerical algorithms for geometrically nonlinear structural analysis, in Finite Element Methods for Nonlinear Problems, ed. by P. G. Bergan, K. J. Bathe and W. Wunderlich, Springer-Verlag, 323, 1986. [53] C. Militello and C. A. Felippa, The rst ANDES elements: 9-DOF plate bending triangles, Comp. Meths. Appl. Mech. Engrg., 93, 217246, 1991. [54] B. Nour-Omid and C. C. Rankin, Finite rotation analysis and consistent linearization using projectors, Comp. Meths. Appl. Mech. Engrg., 93, 353384, 1991. [55] V. V. Novozhilov, Foundations of the Nonlinear Theory of Elasticity, Graylock Press, Rochester, 1953. [56] M. K. Nyg rd and P. G. Bergan, Advances in treating large rotations for nonlinear problems, Chapter 10 in a State-of the Art Surveys on Computational Mechanics, ed. by A. K. Noor and J. T. Oden, ASME, New York, 305332, 1989. [57] M. K. Nyg rd, Large displacement analysis of exible and rigid systems considering displacement-dependent a loads and nonlinear constraints. Dr. Ing. Thesis, Div. of Structural Mechanics, Norwegian Institute of Technology, Trondheim, Norway, 1990. [58] C. Pacoste, Co-rotational at facet triangular elements for shell instability analysis, Comp. Meths. Appl. Mech. Engrg., 156, 75-110, 1998. [59] K. C. Park and G. M. Stanley, A curved C 0 shell element based on assumed natural-coordinate strains, J. Appl. Mech., 53, 278290, 1986. [60] W. Pietraszkiewicz, Finite rotations in shells, in Theory of Shells, Proc. 3rd IUTAM Symp on Shell Theory, Tsibili 1978, 445471, North Holland, Amsterdam, 1980. [61] W. Pietraszkiewicz (ed.), Finite Rotations in Structural Mechanics, Lecture Notes in Engineering 19, SpringerVerlag, Berlin, 1985. [62] E. Ramm, A plate/shell element for large deections and rotations, Proc. US-Germany Symposium on Formulations and Algorithms in Finite Element Analysis, MIT-Cambridge, Mass., 264293, 1976. [63] C. C. Rankin and F.A. Brogan, An element-independent corotational procedure for the treatment of large rotations, ASME J. Pressure Vessel Technology, 108, 165174, 1986. [64] C. C. Rankin, Consistent linearization of the element-independent corotational formulation for the structural analysis of general shells, NASA Contractor Report 278428, Lockheed Palo Alto Res. Lab., CA, 1988. [65] C. C. Rankin and B. Nour-Omid, The use of projectors to improve nite element performance, Computers & Structures, 30, 257267, 1988. [66] C. C. Rankin, Consistent linearization of the element-independent corotational formulation for the structural analysis of general shells, NASA Contractor Report 278428, Lockheed Palo Alto Research Laboratory, Palo Alto, CA, 1988. [67] C. C. Rankin, On choice of best possible corotational element frame, in Modeling and Simulation Based Engineering, ed by S. N. Atluri and P. E. ODonoghue, Tech Science Press, Palmdale, CA, 1998. [68] C. C. Rankin, F. A. Brogan, W. A. Loden and H. Cabiness, STAGS User Manual, LMMS P032594, Version 3.0, January 1998. [69] O. Rodrigues, Des lois g ometriques qui regissent les d placements dun syst me solide dans lespace, et de e e e la variation des coordonn es provenant de ces d placement consider es ind pendent des causes qui peuvent e e e e les produire, J. de Math matiques Pures et Appliqu es, 5, 380400, 1840. e e [70] J. G. Simmonds and D. A. Danielson, Nonlinear shell theory with nite rotation and stress function vectors, J. Appl. Mech., 39, 10851090, 1972. [71] J. C. Simo, A nite strain beam formulation. Part I: The three-dimensional dynamic problem, Comp. Meths. Appl. Mech. Engrg., 49, 5570, 1985. [72] J. C. Simo and L. Vu-Quoc, A nite strain beam formulation. Part II: Computational aspects, Comp. Meths. Appl. Mech. Engrg., 58, 79116, 1986.

45

[73] O. I. Sivertsen, Virtual Testing of Mechanical Systems: Theories and Techniques, Swets & Zeitlinger Pubs., Heereweg, Netherlands, 2001. [74] B. Skallerud and B. Haugen, Collapse of thin shell structures: Stress resultant plasticity modeling within a co-rotated ANDES nite element formulation, Int. J. Numer. Meth. Engrg., 46, 19611986, 1999. [75] B. Skallerud, K. Holthe and B. Haugen, Combining haigh-performance thin shell and surface crack nite elements for simulation of combined failure modes, Proc. 7th US Nat. Congress in Computational Mechanics, Albuquerque, NM, July 2003. [76] R. A. Spurrier, A comment on singularity-free extraction of a quaternion from a direction cosine matrix, J. Spacecrafts & Rockets, 15, p. 255, 1978. [77] M. L. Szwabowicz, Variational Formulation in the Geometrically Nonlinear Thin Elastic Shell Theory, Int. J. Solids Struc., 22, 11611175, 1986. [78] R. Tanner, Engineering Rheology, Oxford University Press, 1985. [79] J. G. Teigen, Nonlinear analysis of concrete structures based on a 3D shear-beam element formulation. Ph.D Dissertation, Department of Mathematics, Mechanics Division, University of Oslo, Oslo, Norway, 1994. [80] C. Truesdell, Continuum Mechanics I: The Mechanical Foundations of Elasticity and Fluid Dynamics, Gordon & Breach, New York, 1966; corrected reprint of ibid., J. Rational Mech. Anal., 1, 125300, 1952. [81] C. Truesdell and R. A. Toupin. The classical eld theories, in S. Flgge, editor, Handbuch der Physik, vol. III/1, 226790, Springer Verlag, Berlin, 1960. [82] C. Truesdell, Hypoelasticity, J. Rational Mech., 4, 83133, 10191020, 1955. Reprinted in Foundations of Elasticity Theory, Gordon & Breach, New York, 1965. [83] C. Truesdell and W. Noll, The Nonlinear Field theories of Mechanics, in in S. Flgge, editor, Handbuch der Physik, Vol. III/3, Springer-Verlag, 1965. [84] H. W. Turnbull, The Theory of Determinants, Matrices and Invariants, Blackie and Son, London, 1929. [85] K. C. Valanis, A theory of viscoplasticity without a yield surface, Part II: Application to mechanical behavior of metals, Arch. Mech., 23, 535551, 1971. [86] G. A. Wempner, Finite elements, nite rotations and small strains of exible shells, Int. J. Solids Struc., Vol 5, 117153, 1969

46

Вам также может понравиться