Sei sulla pagina 1di 38

COMPUTER

METHODS

IN APPLIED

MECHANICS

AND ENGINEERING

72 (1989) 267-304

NORTH-HOLLAND

ON A STRESS RESULTANT GEOMETRICALLY EXACT SHELL MODEL. PART I: FORMULATION AND OPTIMAL PARAMETRIZATION J.C. SIMO and D.D. FOX
Division of Applied Mechanics, Department of Mechanical Engineering, CA, U.S.A. Stanford University, Stanford,

Received 2 October 1987 Revised manuscript received 7 June 1988

1. Introduction and overview


1.1. Overview

Over the past two decades computational shell analysis has been, to a large extent, dominated by the so-called degenerated solid approach, which finds its point of departure in the paper of Ahmad, Irons and Ziekiewicz [3]. The works of Ramm [40], Parish [39], Hughes and Liu [28,29], Hughes and Carnoy 1301, Bathe and Dvorkin [13], Hallquist, Benson and Goudreau [27], Parks and Stanley [38], and Liu, Law, Lam and Belytschko [33], among many others, constitute representative examples of this methodology carried over in its full generality to the nonlinear regime. The thesis of Stanley [48], and the books of Bathe [12], Hughes 1211, and Crisfield {20], offer comprehensive overviews of the degenerated solid approach and related methodologies which involve some type of reduction to a resultant formulation. An alternative approach to the development of shell elements is found in the pioneering work of Argyris et al. [&lo], which makes use of the classical matrix displacement method with high order interpolations (5th- and Ith-order polynominals) within the context of the authors natural approach. By contrast, the present work, the first part of a series of papers, constitutes a departure from the aforementioned methodology. In a sense, the proposed approach represents a return to the origins of classical nonlinear shell theory, which has its modern point of departure in the pioneering work of the Cosserats 1191, subsequently rediscovered by Ericksen and Truesdell [21], and further elaborated upon by a number of authors; notably Green and Laws 1231, Green and Zerna [25], or Cohen and DaSilva [18]. We refer to [35] for a compehensive review including many references to the classical literature and historical overviews, to [4,5] for a careful analysis the mathematical foundations of classical shell theory, and to [44] for a discussion of the underlying Hamiltonian structure. Although the hypothesis underlying the degenerated solid approach and classical shell theory are essentially the same, the reduction to resultant form is typically carried out nume~~ally in the former, and analytically in the latter. conceptually, this appears to be the only essential difference between the two approaches. A point frequently made concerning the degenerated approach is that it avoids the mathematical complexities associated with classical
~45-7825/89/$3.50 @ 1989, Elsevier Science Publishers

B.V. (North-HoIland)

268

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

shell theory, and hence is better conditioned for numerical implementation. A main thrust of the present work is to demonstrate that classical shell theory, phrased as one-director Cosserat surface, leads itself to an efficient numerical implementation which is free from mathematical complexities and suitable for large scale computation. As an illustration, Figs. 1, 2, and 3 contain simulations involving extremely large displacements and rotations obtained with the formulation described in this paper. In Part II of this paper, it will be shown that the present approach is able to reproduce the exact solutions of standard benchmark linear problems often used to assess the performance of numerical formulations based on the degenerated solid approach.

Fig. 1. Large deformation

of a pinched hemisphere,

a standard benchmark

test Problem for linear shell elements.

Mesh Fig.

Clamped

Deformation

2. Large deflection of one quarter obtained in one single step loading.

of a cylinder shell. The deformation

shown (with no magnification)

is

J.C.

Simo, D.D.

Fox, Stress resultant geometrically exact shell model. Part I

269

Fig. 3. Numerical

simulation

involving

180 degree torsion of a thin, initially flat, shell.

To conclude this overview, we observe that most of the existing numerical formulations employing the degenerated approach are inspired by numerical analysis of classical stress resuZtant models; such as the Kirchhoff-Love, Euler, and Timoshenko beam models, and the Poisson-Kirchhoff and the Mindlin-Reissner plate models. Techniques for dealing with phenomena such as shear and membrane locking are typically borrowed from the numerical analysis of these models (or related stress-resultant models i.e., Marguere shallow beam and shell equations). As with these classical linear stress-resultant models, we believe that a deeper understanding of the numerical analysis issues involved in finite element shell analysis can be gained only from a careful formulation of the weak form of the momentum equations in resultant form. A case in point is the analysis of the Koiter-Sanders shell model by Bernadou and Boisserie [15].
1.2. Scope of the paper

In this paper, we adopt a classical point of view and regard the shell as an inextensible one-director Cosserat surface. The objective is a precise formulation of the local balance laws, the local constitutive equations, and the weak form of the momentum equations in a form suitable for numerical analysis and finite element implementation. Although many expositions of classical theory are available, notably the aforementioned reviews of Naghdi [35,36] and Antman [50], the present work places special emphasis on the geometric structure underlying stress resultant (Cosserat) shell models. We believe that this geometric point of view is essential in a numerical implementation for the following reasons: (i) A singularity-free parametrization of the rotation field is developed by identifying the geometric structure of the configuration space; essentially a differentiable manifold (and not a linear space!) modeled on the unit sphere (S*) and [w3.This parametrization is optimal in the sense that the number of parameters describing the rotation field cannot be further reduced without introducing points of singularity. The situation parallels that found in nonlinear (geometrically exact) rod models; see [39,41-471. An important difference, however, concerns the role of drilZ rotations (rotations about the director) which, in sharp contrast with rods, are irrelevant in shells. (ii) Update procedures for the rotation field, which are exact and preserve objectivity for any magnitude of the director rotation increment, can be constructed merely by employing the discrete version of the exponential map in the unit sphere. We address these procedures in detail in Part II of this work.

270

J. C. Simo, D. D. Fw,

Stress resultant geometrically exact shell model. Part I

(iii) Constraints such as inextensibility of the director field are enforced exactly without resorting to penalty methods by exploiting the connection between the unit sphere and a certain subset of the rotation group. In particular, given two vectors in the unit sphere there is a remarkably simple expression for the orthogonal transformation that rotates one vector into the other without drill. (iv) The underlying geometric structure associated with an inextensible one-director Cosserat surface makes the fact that there is no couple stress about the director transparent. This is at variance with some theories based on the use of the rotation vector, as in [32,41,42]; see also the discussion in [24]. Additional noteworthy features emphasized in the present work, which are motivated by computational considerations, are the following: (v) In classical shell theory, it is customary to resolve the surface displacement and surface director in components relative to a Gaussian (intrinsic) frame. In a computational context, however, it is far more convenient to retain the components of the surface displacement relative to a fixed inertial frame. As for the components of the rotation fields, the most convenient resolution is in terms of a local Cartesian frame, denoted by {tl, t,, t3}, and obtained by exploiting the correspondence between the unit sphere and the rotation group. (vi) The weak form of the momentum equations is parametrized in a way that avoids explicit appearance of the Reimannian connection of the mid-surface. In particular, objects such as Christoffel symbols or the second fundamental form, which are not readily accessible in a computational context, do root explicitly appear in the fo~ulation. (vii) The restriction that balance of angular momentum places on the admissible form of the constitutive equations leads to a symmetric form of the consistent tangent operator even away from equilibrium. (viii) Finally, unnecessary assumptions frequently made in the context of the degenerated approach concerning the form and symmetry of the stress resultants are not made. In particular, only a certain symmetric combination of the stress resultants (known as the effective stress resultants) appears in the weak form of the momentum balance equations, An outline of the remainder of this paper is as follows. First, we develop in detail some geometric preliminaries that play a central role in the formulation of the theory and in its numerical implementation, Next, we focus our attention on the basic kinematics of the model, including the precise (classical) definitions of the resultant linear, angular, and director momentum. We then introduce the local momentum balance equations, formulated in terms of rest&ants, and make physical definitions of these resultants within the context of the three-dimensional theory, Properly invariant constitutive equations in terms of stress resultants and conjugate strain measures are developed by exploiting the exact expression for the stress power. We conclude our presentation with the formulation of the weak form of the momentum balance equations and the exact expression for the linearized weak form about an arbitrary (not necessarily equilibrated) configuration.

2. Geometric preliminari~.

Parametri~tion

In this section, we summarize some basic properties of the rotation group, 50(3), and the For further details we refer to [l, unit sphere, S, needed for subsequent developments. Section 4.1; 2; 17, pp. 181-1941.

J.C.

Simo, D. D. Fox, Stress redtant

geometrically exact shell model. Part I

271

2.1.

The rotation group and its Lie algebra

Following standard notation, i.e.

we denote by SO(3) the group of orthogonal

transformations,

SO(3) := {A: R3 + R3 1 A = A-

and det A = t-1) .

(2-l)

Any A E SO(3) possesses an eigenvector + E R3 such that A$ = a,k. Geometrically, represents a rotation about +. The tangent space to SO(3) at the identity, denoted T,S0(3), is SO(~), the set of skew-symmetric tensors, and is defined as
so(3) = { 6: R3 --&~~+8=0}.

by

(2.2)

Any & E SO(~) processes an eigenvector ism A: SO(~)+ R3 by the relation


&h = 0 x h

0 E R3 such that 60

= 0. This defines an isomorph-

for any h E R3 ,

cm

In what follows, we denote by {E,},=,,,,, basis, i.e.

an inertial (fixed) basis, chosen to be the standard

E1=[Bi, E;LIB}, E;={Hi.


Thus, E, = E. For any & E SO(~) we have the matrix representation [&,:]=
[

(2.4) (relative to {E,})

o3 -02

-p 01

->l],

(O>

f].

(2.5)

Any 6 E SO(~) represents an i~~~ites~rna~ rotation about 0 E R. The tangent A E SO(3) is defined (by either left or right translations of SO(~)) as T,S0(3) := {A&l = h4 1 6 E SO(~) or 6 = A&A E SO(~)} .

space at any

(2.6)

Note that A& ET,S0(3) can be Athought of as a finife rotation superposed onto an infinitesimal rotation. Alternatively, f?A E T,S0(3), where 8 = A@A E SO(~) may be thought of as an infinitesimal rotation 0 superposed onto a finite rotation, Finally, the exponential mapping exp: T,S0(3)+ SO(3) maps infinitesimal rotations into finite rotations according to
4 E so(3) -

exp[&] : = kgo $

E SOf3)

(2.7)

One speaks of iA E T,S0(3) SO(3) at A.

and A& E T,S0(3)

as the right and left representation

of the tangent space of

272

J.C.

Simo, D.D. Fox, Stress resultant geometrically exact shell model. Part I

(See Fig. 4 for a geometric illustration of the exponential map.) Remarkably, following closed-form explicit characterization for the exponential map:

one has the

(2.8a) Alternatively, since d2h = 0 X (0 X h) = (0 . h)O - I/ 0 112h, making use of the half-angle formulae we have the following equivalent expression for the exponential map in SO(3):

where e:= O/l]@]] an d b is the skew-symmetric tensor defined by the isomorphism (2.3). Formulae (Ua), (28b) are often accredited to Rodrigues (see [22, p. 165; 61 and play a central role in our developments. For a historical review, see [14]. 2.2. irhe unit
sphere

S. Relation to SO(3)

In the development of shell theory as an inextensible one-director sphere, denoted by S, plays a central role. We set

Cosserat surface, the unit

s2:= {tE R31lItI = l}

(2.9)

The tangent space at t E S2 is the linear space of vectors tangent to curves E H t, E S2, the telgzO = t, and is obtained by noting that (2.10)

Fig.

4. The

T,S0(3)

special orthogonal group SO(3) and its tangent spaces. --+ SO(3) maps skew-symmetric tensors into orthogonal tensors.

The

exponential

mapping

exp:

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

273

consequently,

T,S2 is defined as (2.11)

T,S2:={&R3]W~t=0}.

The unit sphere S2 is in a one-to-one correspondence with a subset Si C SO(3) defined as follows. Let E E R3 be a fixed but otherwise arbitrary vector. Set S;:= (AESO(3)
1w h ere

J/ E lR3, such that A+ = 9, satisfies r,k*E = 0) ;

(2.12)

that is, SF is the set of (finite) rotations whose rotation axis is perpendicular space T1 S,at the identity 1 E Si is then given by T&:= {&SO(~) ] @*E=O}.

to E. The tangent

(2.13)

The tangent space at any arbitrary element A E 5; is defined by either left or right translations of the tangent space at the identity, T,Si, by the expression (2.14) Next, we prove a fundamental PROPOSITION such that t=AE, where
A:=(t.E)l+[E=]+ 1+:-E (Ext)@(Exf)*

result concerning

the relation between S2 and 5:.

2.1. Given any two vectors E, t E S2, with t # -E there exists a unique A E Si

(2.15a) (2.15b)

PROOF. Let & E T,Si, so that 0. E = 0. We show that & E T,Si is uniquely determined from t E: S2 (and E ES) and that (2.15b) follows from (2.8b). Let t = AE. Since e *E = 0, where e:= O/O and O:= I]@(], from (2.8b) we let A = exp[&]. Thus, we have cos@=t*E =$ @=cos-(t*E). we have (2.17) (2.16)

It also follows from (2.8b) that t = AE = cos OE + sin Oe x E. Consequently, Ext=sinOEx(exE)=sinOfe-(E+e)E]=sinOe,

From e and 0, 0, or equivalently 6, is determined and the result follows by substituting (2.16) and (2.17) into (2.8b). Uniqueness of A follows at once from the construction. cl The following correspondence between tangent fields in the unit sphere, S, and tangent fields in $. also plays an important role in subsequent developments of the shell theory.

274

J.C. Simo, D.D.

Fox, Stress resultant geometrically exact shell model. Part I

COROLLARY 2.2. T,,Si and T,S* are in one-to-one correspondence &A ET,& t-+ GETS*
with W = w x t .

through the isomorphism (2.18)

PROOF. (i) Let &A ET,Si. Hence w - AE = 0. Define W E R3 through the relation O:=w X t, where t=AE. Then t-6=0 and GETS*. (ii) Conversely, let 6 E T,S*, so that W t = 0. Define w E R3 through the relation o := t x 0. Since
l

m.AE=w.t=(tX&)nt=o

cjllCJ&,

(2.19)

the result follows.

El we obtain the one-to-one corre-

REPARK 2.3. By particularizing (2.18) to the identity, spondence between T,Si and T,S* jl! ET& where n*E=a*E=O. A geometric illustration r--) &T,S* q with fi = a X E ,

(2.20)

of the preceding concepts is contained

in Fig. 5.

2.3. The exponential map in Sz Making use of the correspondence established in Proposition 2.1 between the unit sphere S* we derive below the closed-form expression for the and the subset S: (or Si)CSO(3), exponential map in S. Let t E S2 be any fixed (but otherwise arbitrary) vector in S*. We have the following. PROPOSITION 2.4. The exp,: T,S2 -+ S* holds:
following closed-form expression for the exponential map

Fig. 5. The S2-sphere and its tangent spaces. A E S, is the orthogonal

transformation

mapping E ++ t = AE E S.

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

275

sin]]811 e exp,[6] = cos]]ti]]t + ~

ll~ll

(2.21)

where

6 E T,S2.

PROOF. By Corollary 2.2, 8 := t X 6 is such that e E T,S:. Consequently, A := exp[e] is in S:. By Proposition 2.1, this defines a unique t E S2 through the relation t = At. Thus set
t =

exp[i]t =: exp,[ti] . formula (2.8b), since 0 *t = 6 *t = 0, we have

(2.22)

By Rodrigues

exp,[ti] = cos]]8]] t + ~

si4Pll ()x t ll@ll *


x t. Cl

(2.23)

The result then follows by noting that ]I011= ]]&]I along with ti = 0

REMARK 2.5. In particular, Proposition 2.4 may be used to parameterize the unit sphere as follows. Choose {E,}I=,,2,, such that E, = E defines the north Dole. We then have the following characterization of exp,: T,S2+ S2, for E ES2, 1

0 ET,S2
where 0 = E
X 6 =

exp,[ti]

:= COSII~IIE

sinll@II (gE +lloll


n -

s2

(2.24)

OE, + OE, with & ET,,,.c2

The geometric notions in this section are key to the development of a well-conditioned singularity-free parametrization of the director field t E S2 and its associated orthogonal transformation A. We shall address these issues in Part II of this work where computational aspects are examined in detail.

3. Kinematic description of the shell In this section we consider the basic kinematic 3.1. The basic kinematic assumption. results underlying the shell model.

Configurations assumption, we define the set % (a differentiable

To state precisely the basic kinematic manifold) as


% :=

{(q,

t):

c lR2d lR3x S} .
boundary aa, compact

(34
closure L$, and points

Here,

&

C R2

is an open set with smooth

276

J.C.

Sirno, D. D. Fox, Stress resultant geometrically exact shell model. Part I

denoted

by 5 E SL We set
g =

tlEl + t*E, , ([l, ()

R* .

(3.2)

The basic kinematic assumption is that of an inextensible one-director Cosserut surface. Accordingly, any configuration of the shell is described by a pair (9, t) E %, where: (i) The map 9: ~4--, [w3defines the position of the mid-surface of the shell. (ii) The map t: &+ S* defines a unit vector field at each point of the surface, referred to as the director (or fiber) field. One is then led to the following kinematic hypothesis.
BASIC ASSUMPTION.

Any configuration

of the shell Y

C R3

is assumed to be defined as (3.3)

Y:={xE[W3]X=~+& In particular,

where (q,t)E%

and ~E[!z-,h]}.

the reference configuration is exactly described as*


P-4)

9 := {x0 E Iw31x,, = q,, + &, with (q,,, to) E % and 5 E [h-, h]} . Here, [h-, h+] C R, with h > h- and h = h - h- is the thickness of the shell. We shall often use the notation Cl

x=

@(,l, 5*,t> := 10(5, t) + ml, 5).


R. A deformation of the shell, then, is a mapping

W)

It follows that @: L&X [h-, h]+

An illustration

of the basic kinematics

of the shell model is contained


Deformation gradient

in Fig. 6.

3.2. The tangent map at a configuration.

Given a configuration @: & X [h-, h+]* R3 the tangent map is the Frechet denoted by V@. Relative to {E,},=,,,,, one has (with t3 = 6)

derivative,

V@:=$@E=g,@EE.

(3.7)

One refers to g, = ~@/a( as the convected basis. By the chain rule, the deformation gradient associated with a deformation x: 93+ Y is F = T,y := V@o(V@J
.

(3.8) in the following.

An explicit expression for V@ is contained


*This parameterization

is often termed normal coordinate chart; i.e. see [3O].

PROPOSITION

3.1. Th.e tangent mq

V@ associated with Qs: A? x [h-, II+]-+ Iw3is given by

(3.9) (3.10)

PROP,

From (3.5) and (3.7) onr: has immediately V~=fvJ,,+rt,,)~E~t~PE. (3.11) of (3.10) into (3.31)

Since by Proposition 2-5. we have 1= MC, where 6, = E3 =E, substitution yields the result. q

R~~~AR~~ 3.ii?. (1) In the assumpt~u~ of smaff strains anst large rotations one has VQ, = A.,
i.e. an ~rth~g~na~ tra~sf~rmat~u~_ Such an assumption wit1 not be made here. (2) In addition to the convected basis {gI}l=1_2,3 (where g3 = t) one defines the recipracorl by the standard relation gr - gJ = tii- Thus basis {gL,,~J

278

J.C. Simo, D.D. FQX, Stress resultant geometrically exact shell model. Part I

(3) We shall use the following notation: j : = de@@] ,


j. : =

de@@,,] ,

J = j/j0 ,

(3.13)

where J := det[F] is the Jacobian of the deformation (4) Observe that while g, = t, g3 is given by
g3 = ; g,

gradient which is given by (3.8).

x g, .
but

(3.14)

Thus, llttl = Ikll = 1, andj=[glxg2]*g3,


3.3. Reference frames on the mid-surface

llgl1+1- q

In addition to the fixed inertial frame {E,},,,,,,,, we define two reference frames on the mid-surface that play an important role in the subsequent developments. (i) Surface convected frame. Denote by {a,}I=1,2,3 the surface convected frame defined as a, =g,(5_0. Note that {a,},= 1,2 span the tangent space to the mid-surface. It follows from (3.5) and (3.7) that

a, = pa and

a3 = g, =: t . form) on the reference surface is then

(3.15)

The metric tensor (first fundamental

denotes the dual surface convected basis defined by the standard relation where a, a = Si. (ii) Director orthogonal frame. Denote by (t,},=l,,,, the director orthogonal frame defined through the orthogonal transformation A: & C Iw2--, Si as

WL1,2,3

t, =

AE, ,

t,= AE,=t.

(3.17) to

Note that (t,},=, 2 3 is the o~tho~~rrna~ basis which, by virtue of (3.9), is the closest In addition, {tr, 2,) span T,S2, the tangent space normal to t, since t *t, = 0. ~%L1,2,3
REMARK 3.3. Our definition

by expression computational

of the orthogonal basis {tl}I_,,2,3 is intrinsic, and is motivated (3.9) for V@. This definition bypasses ad-hoc constructions often made in the literature; see [31, Ch. 61. Cl

These frames are illustrated in Fig. 7. Finally, we recall that the element of mid-surface area is given by the differential (two-form) d& :=
a, x a2

dtl dt2.

(3.18a)

J.C.

Sirno, 13.D. Fox, Stress resultant geometrically exact shell model. Part I

279

Fig. 7. Reference

frames on the mid-surface.

We use the notation

Ai:= I/% x %*ll


to designate respectively. the mid-surface

i:=

Ila, x a,11 and

j:=

i/,&,

(3.18b) Jacobian,

Jacobians

in 3 and Y, and the relative mid-surface

3.4. Resultant linear, angular, and director momentum We conclude this section with a derivation of the expressions for resultant linear, angular, and director momentum. A motion is a curve of configurations; that is, a mapping alp,: & x [h-, h+] x [w, + [w3. Associated with a motion, we have the mapping t t-+ ((9, t,) E %, which defines the motion of the mid-surface and the director. 3.4.1. Angular velocity of the director field By Proposition 2.1, there is a unique A,: & X R, -+ Si such that t, = _4,E. Time differentiation yields 1 . f t, = A,E = AMWAY = DIt f (3.19) where I@t= A:& and I+, := &I: (3.20) are skew-symmetric tensors. Observe that G;,A, E T,,Si and I$( E T,S& whereas i E T,S* since t, - i, = 0. Consequently i, = w, x t, = A,[W, x E], with w; t, =0 and W;E =O. (3.21b) One refers to w, and Vv, as the spatial and rotated velocity of the director field, respectively. (3.21a)

280

J.C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

3.4.2. Density and inertia Let pO(el, t*, 5) and p( [l, S, 5, t) be the mass density in the reference and current configurations, % and Sp, respectively. Recall that conse~ation of mass requires p0 = Jp, where J = det F, = j/j0 is the Jacobian of the deformation gradient F, = [V@t][V@O]- associated with the motion xt = @t0 @i: ?$ --, 9. We select the mid-surface v,,: sllz + R3 (which does not necessarily correspond to the geometric center of the cross-section) such that (3.22) Define the surface densities &: &+ Poj,,dt In addition, and Iw and p: SQ * Iw by the expressions 1 @:= T I R and fP := &--, Iw as (3.24) The significance of these definitions will become apparent in what follows. By (3.23), (3.24) and the conservation of mass relation p0 = Jp, we have Is, = .&S and which constitutes &, = J, , of conservation of mass for shells. (3.25) (3.23)

define the surface inertias joP: &+

the statement

3.4.3. ~esuitant linear, a~gu~ar~ and director ~o~ent~~ For a motion et: & x [h-, k+] X II%+ + R3, time di~erentiation

of expression

(3.5) yields (3.26)

Define the resultant linear ~o~en~~ expressions

pt, and the resultant angular momentum m, by the

(3.27) Making use of (3.22), (3.26), the kinematic hy~thesis and (3.24) we have (3.5) and introducing definitions (3.23)

(3.28) where the relation w, tt = 0 was employed.


l

J.C.

Simo, D.D.

Fox, Stress resultant geometrically exact shell model. Part I t,;

281

Since mZ- t, = 0, by Corollary 2.2, there is a ~~~~~e +, E T,S* such that tit = rr* x - . Gt := Tr, x t, = zpw* x t, = zptt . #f is referred to as the resultant director momentum.

in fact (3.29)

4. Stress resultants and stress couples. Local balance laws In this section we define stress resultants and stress couple resultants from the threedimensional theory and develop the balance laws in terms of these resultants. Given a rnot~o~ xl: $32x IR, -+ Y, where x, = @!0 @P,, we denote by ft the symmetric Cauchy stress tensor in Y, and by P the nonsymmetric (first) Piola-Kirchhoff stress tensor relative to 5%and Y. 4.1, Definitions for the three-dimensional
theory

Consider sections in the current configuration Y := {Xf R3 1x = @(5=const,} , The (one-form)


Q =

Y C iR3 defined as 1,2. (4.1)

field normal to Y is given by


(4.2)

dY* := j[V@]-E 65 d,$ = jg dt2 dr , and the analogous expression holds for dY2. Consequently, coordinate length 5 is R := Similarly, the torque acting on Y1 per unit of coordinate
y-l:=

the force acting on .Yi per unit of

length 5 is dt .

(x - 40)x u

$ =lh: (x -

4p)x ugj

(4.4)

We define the stress resultant n, and the stress couple ma, by normalizing R* and T with the surface Jacobian i= [/al x a2 11.Accordingly, we set 1
h+

n *

:=

= j

I h-

4j

de

(4Sa)

1 ma := - ::(X-yl)xMjdL I f

(4Sb)

282

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

where x = @( 5 , r*, 6). Finally, one defines the across-the-thickness stress resultunt, denoted by 1, by the expression

(4.6) where y = (g, x g2) / 11 g, x g, 11 = ( jg3) / IIg, x g, II is the (one-form) surface. REMARKS normal to the laminae

4.1. (1) For the stress resultant n we have the equivalent 1


h+
h.I-

expressions

na=

y
I

ug"j

dt = f
I

,,y+Pgtj, dt

(43
where F =V@[WO]- is the

the second of which follows by recalling the relation P = JaF-, deformation gradient and J : = det F = j/jO, and noting that juVW_ = jOPV@ .

(4.8)

(2) Similarly, noting that (X - 9) = [t, for the stress couple tnn we have the alternative expressions in terms of the Piola-Kirchhoff stress 1
f

m=tx

7
I

h:

&rgci

d& =

t x

r, j

l+ h-

@gGjo

dS

@*9)
resultant can be written

(3) In terms of the Piola-Kirchhoff


1

stress, the across-the-thickness

= y
j

h+ I hag3j

h+

d[ = )

f h-

&j,

dS

(4.10)

(4) Observe that ma *t = 0. Hence, there is no component of the stress couple along the director t. This is at variance with some formulations of shell theory employing the rotation vector, as in [32]. AIternatively, one may define the director stress couple, &, according to the expression

It should be noted that ~6 *t # 0. However, because of the constraint t E S, the component of &* along t does not enter explicitly in the subsequent developments. In fact, this component could be eliminated completely by defining ti := hi - (6 t)t, so that iii *t = 0, and (4.11) would be of the form ma = t X ii. Cl
l

J.C. Simo, D.D. Fox, Stress resultant geometrically exact shell model. Part I

283

Starting with the momentum balance equations of the three-dimensional shown (see Appendix A) that the resultant local form of the momentum take the following form:

theory, it can be balance equations

(4.12a) (4.12b)

where $ and niare the applied resultant force and applied resultant coupler per unit length as defined in Appendix A. We note that m.t=O and m*f=O. (4.13)

Equations (4.12) are in the form considered by several authors; see Green and Zerna [25, pp+ 379-3801 or Libai and Simmonds 1321. These authors, however, immediately proceed to derive component equations relative to the Gauss frame {a,, a2, t} leading, inevitably, to the explicit appearance of the Christoffel symbols associated with the Riemannian connection of the mid-surface. In this regard, see also ]35], We show below that the direct use of the vector equations (4.12) is ali that is needed in the weak formulation of the equations. The Riemannian connection of the mid-surface does not enter ~~~Zicitly in the formulation.

The balance of angular momentum of the three-dimensional theory is expressed by the symmetry of the Cauchy stress tensor, CT= a: In what follows, this balance law is interpreted as a restriction balance of angular momentum places on the admissible form of the constitutive equations for the resultants 1~&, ma (or equivalently m), and 1. Expressing q in components relative to the convected basis g, we have o = crzJg, $r g,. The symmetry condition CI = 0 implies g, X gJarJ= g, X rrg = 0. Integration of the latter relation over [h-, h] C R and use of the expressions g, = 8, + St,, and g, = t yields
k*

crgjd[ Introducing definitions

+ E,, x

@gjd&

+ tx
f

h_ crgjdt==O.

(4.14)

(4.5a), (4.6) and (4.11), (4.14) reduces to (4.15) places on the admissible

q&3~Xnn+t,*X~iia+tX1=0.

Equation (4.15) is the restriction that balance of angular momentum form of the constitutive equations.

284

J.C. Simo, D.D.

Fox, Stress resultant geometrically exact shell model. Fart I

4.3.2. Alternative form of the momentum equations With the help of (4.15), the angular momentum following form. From (4.11) we have

equation

(4.12b) can be recast in the

1
7

(P ),a= t,,
x i,

Tu

x rn+tx

;ci;;l,,, *
of (4.15) and (4.16) into (4.12b) yields

(4.16)

Recalling that W= t
t

substitution

[i

1(jIGi),, -1+&Q
* ;ni=mxt.

=o,

(4.17)

where g is the applied director couple per unit length which satisfies &t=o Consequently, tions (4.18) system of resultant local momentum balance equa-

we obtain the equivalent

(4.19) where I= 1 -t ht, and A: ,YY+ IR is an undetermined director pressure, whose significance is analogous to the hydrostatic pressure in incompressible elasticity. Further elaboration on the significance of tis given below. 4.4. Further reduction. The across-the-thickness stress resultant By making use of the constraint condition [Itll = 1, we derive an explicit expression for the across-the-thickness stress resultant and obtain a further reduction of the constitutive restriction (4.15). These expressions play a fundamental role in the variational formulation of the momentum equations considered in Section 6. 4.4.1. Further reduction of the co~titutive restriction We now consider component expressions relative to the surface convected the development by setting
t ,4 = A~$o, + /iit.

basis. We start

(4.20)
/It11 =

Using the constraint


t * t,, = 0 as

condition

3 is determined 1, A,

in terms of A: from the condition

A; = -A&,

. t= -hEy,

(4.21) along (p,,, t} as

Next, we resolve the resultants

nff and G into components

f. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

285

na= n@+$+ qt
Restriction

rii =

#=+$ + 6i3*t.

(4.22)

(4.15) then becomes (n@ - hP,yii)$Q x p,, + t X [Z- (q + A;??? - h:6?)$2J


= 0.

(4.23)

By taking the dot product of (4.23) with t we obtain j;rag(nPa - A,PSa) = 0, where e,@ is the surface alternator tensor. Consequently, (4.24) we define the ~y~~e~~~c resultant (4.25) for 2 (4.26)

$ := lZPa_ APea E n@ @ In addition, from (4.23) and (4.25) one obtains the explicit expression

I = At + (q= + A;& - h:Yii)+~+ . Introducing the definition

expression

(4.26) now becomes


1= At + TV,,

+ A;rii3'+,

(4.28)

We shall refer to iP and T as the effective membrane and effective shear stress resultants. The significance of these definitions will become apparent in the following section.

5. Local (elastic) constitutive equations In this section, we derive properly invariant elastic constitutive equations for the effective stress resultants n and 4 and for the resultant stress couple iii. To this end, we first obtain appropriate conjugate strain measures by reducing the general expression for the stress power of the three-dimensional theory by means of the basic kinematic assumption (3.5).
5.1. Reduced stress power. Conjugate strain measures

The main result to be exploited in the formulation the following.

of constitutive

equations is contained

in

PROP~S~~~~~ 5.1. By making use of the basic kinematic ~sumption of the three-dimensional theory is expressed in the form

(3.5)) the stress power

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

where dp = id,$l dt2 is the current surface area measure, P is the first Piola-Kirchhoff tensor, and F is the deformation gradient given by (3.8). PROOF.

stress

Using (3.8), along with Proposition ti = [V @]V@, )

3.1, time differentiation

yields

(5.2)
We can now write P : iTi= PV@,E*
l

(+,, + &t,,) f PV@,E3

t .

(5.3)

Thus the stress power relation is expressed rh+ i-h+


1

The result (5.1) then follows by recalling definitions (4.7), (4.10), and (4.11) of the stress and III stress couple resultants in terms of the Piola-Kirchhoff stress tensor. An alternative form of the stress-power relation (5.1) is obtained by introducing the effective stress resultants (4.25) and (4.27). This result is summarized in Corollary 5.2. First, we make the following definitions. Define the following spatial tensors:

cr:= qnaa
)

(5-5)

_ := rFiPap @a, m

Define kinematic variables as follows:


a0np =

4P0,a 4c,,,
l

(5.6a) (5.6b)

YOa

Y)o,a

to

KO,,

vo,a

to,p

(5.6c)

The corresponding as

relative strain measures are defined relative to the dual spatial surface basis

J.C.
& :=

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

287

$(a,, - @,,p)ua @taP ,

With this notation


COROLLARY 5.2. equivalent form

at hand, we present the following corollary to Proposition


The stress power of the three-dimensional

5.1.

theory may be expressed in the

(5.8a)
= 3[n: L++L,S+fi:

L,,p]d/+

(Mb)

where L, represents the convected time derivative. PROOF.

First, we use the component

expressions

(4.22) to rewrite (5.1) as

From (5.6~) and the constraint

]]t]] = 1,
t,,
7

= d,p*t,a+ Q,p %a
tot,, =o Recalling substitute *

(5.10)
.

t*i,, = -i*t,,

the expression (4.20) for t,a in components and the expression (4.28) for 1, we (5.10) into (5.9) to find that terms involving yii3n cancel, to Ieave

Result (5.8a) follows immediately from the symmetry of the effective membrane stress resultant (4.25), the de~nition of the effective shear stress resultant (4.27), and time differentiation of y, . Result (5.8b) follows from the definition of the convected time derivative

(5.12)

288

.I. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

REMARK derivative?

5.3. As defined, IJ

the convected

time derivative

is a particular

form of the Lie

5.2. General elastic constitutive equations Within the context of the purely mechanical theory, hyperelastic constitutive equations may be formulated by postulating the existence of a stored energy function of the form @( 5, a, Y, 7 t,, ; *o, yo, Pt 0,a ) or equivalently q( 5, E, S, p; a,, yO,, to,,). We now reduce the form of !P by exploiting standard invariance requirements under superposed rigid-body motions. Given any superposed rigid-body rotation t I.+ Q(t) E SO(3), E, 8, p, $, n, 6, &, and 6 must transform according to

&+= Q(+Q"(t) 6 = Q(t)6 ,

ii = Q(t)ZQ(t) 4+ = Q(t)@,

(5.13a) (513b)

P+= Q(t)pQ'(t)

2 = Q(+Q'(t)

(513c) (513d)

Since the base vectors a, and the dual vectors an must transform

according to (5.14)

4 = Q(t)aa

aa+ = Q(t)u

it follows from (5.13a) and (5.14) that

EC= &ipaCl+ @a P+
= Q(t)(c&an @ aP)Q(t)
= Q(t)(&~~u~ C3~~)Qt(t) .

(5.15) for S and p we conclude that (5.16)

Result (5.15) implies E$ = E,~. Following the same argument

Thus w is expressed in invariant form as (5.17) REMARKS 5.4. (1) $ depends on the relative measures E,@, 8, and paa. (2) The dependence of 3/ on the reference quantities aOup, A& and Ai, is discussed in detail in [16]. (3) The dependence of $ on Ai, could be eliminated as Ai, = -r,,A&. Cl
For a further discussion of Lie derivatives see [29, Section 1.61.

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

289

Using the Clausius-Duhem inequality and following standard arguments (see [34, Section 2.53 and references cited there), we write the properly invariant hyperelastic constitutive relations
(5.18)

5.3. ExampEe: Isotropic constitutive relations The simplest properly invariant isotropic constitutive relations for the effective membrane and shear stress resultants 6 and F, and for the stress couple resultant Gripis given by
nP* -

pEh
lY2
=

fpWe

YS
@aySPcy8) 7

&e-w

pEh3 12(1 - V)

(5.19)

where E is Youngs modulus, Y is Poissons ratio, K is the shear reduction coefficient, p = p/h is the three dimensional mass density taken to be independent of the through-the-thickness
Table 1 Basic kinematics
l

Definition

of the kinematics 4p: dc682+R3) t: .!2cc2-+s2 {E,,


E2,Q t

mid-surface ( director field inertial frame such that t = At(t # -t)


+ & (t x t)@J(tx t)

Orthogonal

transformation
[Gal

A = (t- t)1+
l

Exact relation between director and rotation vector 8 (exponential map) t = COSl~Gllf + ~ 8, where0*t=O, 8=@xt

Surface convected
a,, = PO,,

frame and
t a03 =

surface Jacobian
t0 T i, = Ita,, x a0211

a*

44a

a3

t,

i=

/I@,

%/I

Relative strain measures

290

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part 1

variable, and HPaYa=


REMARK 5.5.

+ a~%;y>> . 1 IU$u; + i(l - zJ><u,%z;

(5.20)

(1) The constitutive equation for the resultant couple given above is in terms of the symmetric part of m Mpa. It is assumed here and in the following developments that the skew-symmetric part of fiP is * zero, i.e.
&aPl = 0.

(5.21)

(2) Constitutive equations (5.19) can be justified on the basis of an asymptotic expansion of the three-dimensional Saint Venant-Kirchhoff model that employs definitions (AB4c) and (AB5) from Appendix B and retains terms up to order O(hlR), where h = h - h- is the thickness of the shell and R is the radius of curvature. Since terms involving products of aa, A& and Ai are of higher order, in classical shell theory the resulting constitutive equations are typically assumed to depend on the refence surface only though the first fundamental form (i.e., a;P); see [25; 35, p. 606; 37]:The general dependence of the constitutive equations on the reference surface is discussed in detail in [16]. 0 For convenience, and 2.
Table 2 Local momentum
l

the basic theory presented

in Sections l-5 is summarized

in Tables 1

equations.

Constitutive

relations

Balance of linear and director momentum

Stress resultants

and stress couples

na = npaap + qt

s = gBa, I=&+1 1= At + Ta,

+ fi3t

+ A,Gipa,

0 Effective stress resultants

$ = nPu _ @y +Q = qa - +a = q + A; ya6?
0 Constitutive

equations:

!P = +( 5, E,@, 8, , pap; aOap , Yo,, A$, Aim)

In what follows, with a slight

abuse in notations we write = (9, t) E %:, where $6: d-+ 11;8 defines the p~si~i*~ of the mid-surface, and 1: A$-+-_*defines the dire&x fiefd, We assume displacement boundary conditions of the fen-m

where CM is the b~u~da~ of & C R, The tangent space at Q = (gcr,b), denoted by T, 552, is the linear space af admissible variations defined as

Since &( 5 > E T,S2 for each & E &, we have the following alternative useful characterization of the variations 22: ti--+ T$ associated with the director field: fi) S~~~~~~ ~e~&~~p~~~~. Recalling the ane-to-one correspondence between T,S2 and TJt$-; we can write

REMARK

(6Sb) is pa~tie~la~l~ useful in a ~Qrnputati~~al context. The reasun for this lies in the relation 62 - E = 0. ~~~luw~~~ Remark 25 we can write

61. The rotated rep~ese~tati~~

292

f. C. Simo, D. D. Fox, Stress resultant geometrical& exact shell model. Part I

&T-E =0

6T=6Tl&

-+6T2E,.

66)

Consequently, only two co~~o~e~~ (relative to the inertial frame {Ef}f=1,2,3) need to be considered. Thus, this representation automatically excludes the drill degree of freedom along E, = E. In the spatial description (6Sa), however, this constraint is only implicit in the condition 6t t = 0. q
l

6.2. Weak form of momentum ~fflu~ce By considering arbitrary variations 6c;D= (&4p,8t) E T, %, and integrating over the current surface q: ,s?+ lR3,the weak form of the momentum equations in Table 2 becomes G~~(~, 8Qi) := G(@, S@) + j-&[+
* 69 + I,& lit]

dp ,

(6.7)

where dp = jdtl ds2 is the (two-form) current surface measure and G(@, ;SQi) is the static weak form or virtual work expression defined in the following subsection. 6.2.1. Static weak form Using the divergence theorem for surfaces (integration lation gives the following expression:

by parts), a straightforward

manipu-

where G,,,(G@) is the virtual work of the external loading given by (6-9) Here rf-and 6 are the prescribed force and torque on the boundary 6 = &%, on a,&
,

of the shell; i.e. (6.10)

n= nv,

on d,&

and

where v = ~,a is the (one-form) a,&=@ and a,,,&r&aa,s42=0.

field normal to the boundary

+~(a&). Of course, a,& n

6.2.2. Component expression We now introduce the explicit expression for I and the component relations (4.20) and (4.22) into expression (6.8). Performing manipulations identical to those in the proof of Corollary 5.2, it can be shown that the weak form (6.8) is expressed in components as

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

293

6.3.

~ineuri~ution. Linearized struin measures

Next, we consider the linearization of the weak form (6.11) at a configuration Q) := (8, t) E % in the direction of an incremental tangent field A@ : = (Aq, At) ET, %. The resulting linearized weak form plays a central role in the numerical solution of boundary value problems considered in Part II of this work. To construct the linearization of (6.11), one considers a one-parameter family of configurations E M s>, = (F~, t,) E %Zwith the property that

(6.126) A systematic procedure for constructing GjEis to make use of the exponential map, as follows.

6.3. I. T~n~~n~ ~on~~~r~~iu~ and ex~~~~~~~~~ map Let (A+o,At) E T, (8. Define the one-parameter family of configurations setting v~:=~+EA~, and t,:=exp,[sAt],

(~7,t t,) E % by

(6.13) 2.4). Recafl

where exp,: T,S -tS is the exponential that one has the explicit expression exptfr At] :=cos]fsAt]ft+

map in the 5 sphere (see Proposition

sin]] s AtI] E At its AtI]

(6.14) of (6.12). To this

where At E T,S2. We now verify that (6.13) meets the required properties end, first observe that

Gil,=+, = P and exp,k Atll,,,

= 8,
derivative

(6.15)

so that (6.12a) holds. Next, by making use of the directional

formula we obtain

(6.16) so that (6.12b) also holds. Finally, we record a relation needed differentiating (6.16) it follows that in the next subsection. By

(6.17)

294

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

With these preliminary results at hand, we proceed to the linearization measures followed by the linearization of the weak form.

of the strain

6.3.2. Linearized strain measures be the kinematic quantities associated with the one-parameter family of Let (a& rt, K$) confIgurations E H @, = (pE, t,) defined by (6.13). For convenience, we introduce the notation (6.18) By making use of the results in Section 6.3.1, we readily obtain the expressions

tt*b,a + Aw,J Aq3 := A+Q~* t,, -t v,4n

AY, :=

,
At,, -

(6.19)

We now proceed to the linearization 6.4. Linearized weak form. Symmetry

of the weak form.

The conceptual procedure is standard; see e.g. [34, Ch. 61. One substitutes the oneparameter family of configurations (6.13) into (6.11) and then one makes use of the relation between the Frechet directional derivative given by the standard formula

DG-A@ =2

E=OG(@$,

8~).

(6.20)

The linearization of (6.20) can be broken into two parts, D,G - A@ and D,G . A@, each of which possesses classical interpretations.
6.4.1. Operator expressions

For subsequent developments, it proves convenient to introduce a more compact notation. To this end, we define the following resultant stress and stress couple vectors. As is stated in Remark 5.5, we assume %ifpul = 0, i.e. 6 = G! We thus have

(6.21)

Furthermore,

utilizing expression

(6Sb) for the director variation field, set

(6.22)

.l. C. Simo, D.D.

Fox, Stress resultant geometrically exact shell model.

Part f

2%

We define the matrix differential V,l ag


t

operators

a
1

(623a)

(623b)

t
BbIn = t.2 2

t
Bbb =

ii

at

3x2

(6.23~)

where j is the (3 x 2) matrix obtain by deleting the third column of -4, i.e. (6.24) at hand, we can rewrite the weak form (6.11) as

4x2 = With this notation

G,t@Q,)
7

(6.25)

where dpO = jO de dt2 is the reference surface measure. Aiternatively , defining the total resultant stress vector as

R=Z
M L-1 and the differential
8X1

(6.26)

operator (6.27)

the weak form (6.25) can be written as G(@, SD) := /& B{ fiT) R d,u,, - G&j@)
l

(6.28)

296

J.C.

Simo, D.D.

Fox, Stress resultant geometrically exact shell model,

Part I

We emphasized once more that by employing the material representation of the director field variation, the weak form (628) is in a five degree uf freedom format with the constraint ~~~ditiu~ f 66 = 0 aut~rnati~~~y satisfied through the material refation ST * E3 = 0,
l

response with properly invariant stored energy ft.mtion tlr(S, q+, a,, Pap; aomPTTo,, &L Al)> G(@, S@), as given by (6.28), is the first variation of the functionat (629) (6.30)

~~~A~~

62,

For the case of hyperelastic

is the

~tentia~

energy of the external goading (we assume dead ~uading~~)_ iI

We ~CSW proceed to the ~inea~~ati~~ of the weak form (6.28). 6.42. Material part The material part, D,G A@, of the tangent operator arises as a resuit of Iinearizing the substitutive equation at fixed ~~~rnet~. For illustrative purposes, we consider elastic response . Accordingly,
l

(6.31) where

aP*243%

f3P128P22

aPdPl2

is the symme~~

tangent elasticity tensor.

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

297

6.4.3. Geometric part The geometric part, D,G A@, of the tangent operator arises from linearizing the geometric part of the static weak form with the material stress resultants held constant. With straight forward manipulations, it can be shown that
l

{D~{~~}.A~}*~=Y{~~}.~~*{~~}~,

(6.33)

where

KG :=

and

I
-3 at2 03 03 03

Pl,

Pl,

03

n1213 n221, 6T21, -(ljia~K,4 + 436,)1, -iih;;1, $1, q1, 221, -?2ya13 -6iz2ya13
03

FPl, _ 12

rii 1213 ?ii**1, -&22y*13

(6.34) 15x1s

riP13

m1213 riiZ213
03
03

03

03

a .3 at'
a

Y=

13

(6.35)

13 $i

13 I$

Note that K,, the geometric

tangent stiffness, is symmetric. process (i.e. the variations to be process) are (89, Se), where St = Stl X t. Thus A(W) = 0,

REMARK

6.3. The primary variables in the linearization

held constant in the linearization where A(st) ZO. In fact,

=88xAt = -(WAt)t The tangent operator . Cl as (6.36)

is now expressed

(6.37)

298

J.C.

Simo, D.D.

Fox, Stress resultant geometrically exact shell model. Part I

7. Closure and overview of subsequent research We have presented a careful development of the continuum basis of a geomet~cally exact shell theory obtained by reduction of the three-dimensional theory by means of the oYte inextensible director kinematic assumption. Our subsequent work will be concerned with numerical analysis and computational aspects involved in the implementation of the present formulation within the framework of the finite element method. In particular, we plan to focus on the following aspects: (a) The linearized theory. The relevant field equations are obtained simply by particularizing the results developed above to the reference configuration. The linear theory provides an ideal framework to address, in detail, issues concerned with interpolation and convergence of the finite element procedure. We provide a detailed account of the proposed interpolation, develop explicit matrix expressions, and assess the performance of the model in standard benchmark problems, as well as new test problems. Our results indicate that the present theory exactly matches results obtained with the degenerated solid approach. (b) Computational aspects of the fully nonlinear theory. The central topic to be addressed is the development of discrete update procedure for the director field which is exact and singularity free regardless of the magnitude of the rotational increments. In contrast with current procedures, often only second-order accurate, the proposed approach is exact, and preserves objectivity for arbitrary large displacement and rotation increments. Computationally, the need for using quaternion parameters to avoid singularities is bypassed. Detailed account of the interpolation procedure, and expressions for the residual and consistent tangent matrix are given. The formulation is assessed through numerical simulations involving instability and bi~rcation phenomena. (c) Enhunced models including through the thickness stretch. The model considered in this paper can be easily extended to accommodate thickness changes in the shell by means of a one extensible director kinematic assumption. It is shown that this additional effect adds very little complexity to the model while allowing for an entirely new class of problems to be addressed. The advantages of such a model include: exact recovery of the plane stress constitutive equations in the thin shell limit, explicit account of the through-the-thickness-stress for problems dominated by behavior such as delamination and avoidance of numerical problems associated with the undetermined pressure-like inextensible constraint. (d) Inelasticity in stress resultants at finite struins. Elastoplastic constitutive models at finite strains that incorporate classical kinematic and isotropic hardening rules are formulated entirely in stress resultants. The formulation completely bypasses the need for the costly integration-through-the-thickness procedure in the traditional degenerated solid formulation. The integration of the elastoplastic equations is performed by means of recently proposed general return mapping algorithms that can accommodate non-smooth yield surfaces. (e) Nonlinear dynamics. Computational aspects. The present geometric setting, and the associated Hamiltonian structure considered in [39] can be exploited to implement time stepping algorithms that exactly preserve the constants of motions. Recent work indicates that one can develop time stepping algorithms which are, in fact, discrete canonical transformations, thus preserving the discrete Hamiltonian structure of the model. We plan to address these and related issues in our subsequent work.

299

Appendix A. &dance equatims For completeness, we include in this appendix a derivation of the resultant form of the balance of linear and angular momentum from the three-dimensional integral balance laws. For a detailed discussion in the general context of Cosserat surfaces, see [35,36]. An alternative derivation is given in [32].

The three-dimensional

integral equation

of balance of hnear momentum

is written

where ii is the unit body force per unit ri d,Y = jV@-I$ dT change of variables

normal to the arbitrary spatial volume T with boundary a? and B is the mass. By using the kinematic assumption (3.3), the normal map relation (fi is the normal to a x [h-, h], with surface element dT), and the relation, (A.1) becomes

We now decompose the surface integral into an integral over the lateral surfaces (fi = v,E) and an integral over the top and bottom surfaces ($ = +E3). Using the definitions of the reciprocal basis (3.12) and the mid-surface (3.22) we have

a3

Recalling the definition of the stress resultant (4.5) and by using the divergence surfaces, the first term in (A.3) can be written

theorem

for

I II
ad

h-

h+ jagd&dy=

j j(in),,d&. &

(A4

Defining the surface loading term

300

J.C.

Simo, D.l).

Fox, Stress resuhrtt geome3ri&~lly exact shell model. Part I

and using definitun (3.231, we have

(A-6)
Since the choice of Y (or equivalently &) is arbitrary, assuming enough smoothness so that the standard locahzation result holds; see [26, p. 383, we obtain the resultant form of the local balance of linear momentum (4.12a)

The three-dimensional @xoiidY+

integral equation of balance of angular momentum @xpBdCP=$ Q,xp&dclr,

is written (A-8)

SI aqr

~~~ gr

iii I

where the definitions in (A.11 apply. Again employing the kinematic assumption (J.3), the normal map relation ii dY = jV@ -&dI, and the change of variables relation, (A.8) becomes

IRecal! that the normals to the lateral, top, and bottom surfaces are given by & = v,E, fi = E3, and IV = -E3, respectively. Thus, with the definition of the reciprocal basis (3.12) and the definition of the mid-surface (3.22) we can write

(A.10) Using the stress and stress couple resultant defining the external couple as definitions (4.5), (AJ), (3.23), and (3.24) and

J. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

301

(A.ll) we can rephrase. (A.10) as (A.12) .a


J

where the divergence theorem, balance of linear momentum (A.7), and the relation ti, = t X t were used. Since V (or equivalently L$) can be chosen arbitrarily, assuming enough smoothness we obtain the local resultant form of the balance of angular momentum (4.12b) as 1 TLI T(,rn ),p+~,llXnn+m=zPi,. (A.13)

Appendix B. Stress resultant

components

In this appendix we derive explicit component expressions for the stress and stress couple resultants from the three-dimensional theory. First, we present some preliminary results that are helpful in the following development:
g, =

@,,I I$ g, = a, + O,, ,
AEa, + Ait,
ag

g, = t

t,, =

(B.1)

= aSgp + u39.

From definitions 1 no1= Substitution


na

(4.7) and (4.11) we recall that


h+

I I h-

agjdt,

fi=:

1
I I

:_ Sagj d5 .

(B-2)

of (B.l)
= -

into (B.2) yields


h+

h+

h_

aPa(ap + [hga,)jd[

+ L
j

I h-

(cT~~+ &iiaP)tj

dl

From component
p= 1 j

expressions
h+

(4.22) relative to the surface convected basis {a,, t} we obtain eAEaPa)j dt (B.4a)

I h-

(f?

(03a

tAiaP)jd[,

(B.4b)

302

.I. C. Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

(B .4c)

(B.4d) From (4.25) and (4.27) we now obtain component shear stress resultants as relations for the effective membrane and

Acknowledgment This work was performed under the auspices of the Air Force Office of Scientific Research. Support was provided by grant nos. 2-DJA-544 and 2-DJA-771 with Stanford University. This support is gratefully acknowledged. We wish to thank Professors S. Antman, T. Hughes, P. Krishnaprassad, 3. Marsden, C. Steele and R. Taylor for many helpful discussions.

References PI R. Abraham, J.E. Marsden and T. Ratiu, Manifolds, Tensor Analysis, and Applications
Springer, New York, 2nd ed., 1987).

PI R. Abraham and J.E. Marsden, Foundations

of Mechanics (Addison-Wesley,

Reading, MA, 2nd ed., 1978). (Addison-Wesley,

t31 S. Ahmad, B.M. Irons and O.C. Zienkiewicz,

Analysis of thick and thin shell structures by curved finite elements, Internat. J. Numer. Meths. Engrg. 2 (1970) 419-451. 141S.S. Antman, Ordinary differential equations of nonlinear elasticity I: Foundations of the theories of non-linearly elastic rods and shell, Arch. Rat. Mech. Anal. 61 (4) (1976) 307-351. PI S.S. Antman, Ordinary differential equations of nonlinear elasticity II: Existency and regularity theory for conservative boundary value problems, Arch. Rat. Mech. Anal. 61 (2) (1976) 353-393. J.H. Argyris and D.W. Scharpf, The SHEBA family of shell elements for the matrix displacement method. bl Part I. Natural definition of geometry and strains, Aeron. J. Roy. Aeron. Sot. 72 (1968) 873-878. [71 J.H. Argyris and D.W. Scharpf, The SHEBA family of shell elements for the matrix displacement method, Part II. Interpolation scheme and stiffness matrix, Aeron. J. Roy. Aeron. Sot. 71 (1968) 878-883. J.H. Argyris and D.W. Scharpf, The SHEBA family of shell elements for the matrix displacement method, PI Part IIl. Large displacements, Aeron. J. Roy. Aeron. Sot. 73 (1969) 423-426. f91 J.H. Argyris, M. Haase and G.A. Malejannakis, Natural geometry of surfaces with specific reference to the matrix dispIacement analysis of shells, Proc. Kon. Nederl. Akad. Wet. Ser. B 76 (5) (1973).

J.C.

Simo, D. D. Fox, Stress resultant geometrically exact shell model. Part I

303

[lo] J.H. Argyris and P.C. Dunne, A simple theory of geometrical stiffness with applications to beam and shell problems, Second Internationaf Symposium on Computing Methods in Appfied Sciences and Engineering, Versaifles, France, 1975. [ll] J.H. Argyris, An excursion into large rotations, Comput. Meths. Appl. Mech. Engrg. 32 (1982) 85-155. [12] K.J. Bathe, Finite Efement Procedures in Engineering Analysis (Prentice-Hall, Engiewood Cliffs, NJ, 1982). f13] K.J. Bathe and E.N. Dvorkin, A continuum mechanics based four-node shell element for general non-linear analysis, Internat. J. Computer-Aided Engrg. Software 1 (1984). [14] M.F. Beatty, Vector analysis of finite rigid rotations, J. Appl, Mech. 44 (1977) 501-502. [lS] M. Bernadou and J.M. Boisserie, The Finite Element Method in Thin Shell Theory: Application to Arch Dam Simulations (Birkhauser, Boston, MA, 1982). [16] M.M. Carroll and P.M. Naghdi, The influence of the reference geometry on the response of elastic shells, Arch. Rat. Mech. Anal. 48 (1972) 302-318. [17] Y. Choquet-Bruhat and C. Dewitt-Morette, Analysis, Manifolds and Physics (North-Holland, Amsterdam, 2nd ed., 1984). 1181 H. Cohen and C.N. Defilva, Nonlinear theory of elastic directed surface, J. Math. Phys. 7 (6) (1966) 960-966, [l9] E. Cosserar and F. Cosserat, Theorie des corps deformables, in: Chwolson, Trait6 de Physique (Paris, 2nd ed., 1909) 953-1173. [20] M. Crisfield, Finite Elements on Solution Procedures for Structural Analysis, I, Linear Analysis (Pineridge Press, Swansea, U.K., 1986). [21] J.L. Ericksen and C. Truesdell, Exact theory of stress and strain in rods and shells, Arch. Rat. Mech. Anal. 1 (4) (1958) 295-323. [22] H. Goldstein, Classical Mechanics {Addison-Wesley, Reading, MA, 2nd ed., 1981). [23] A.E. Green and N. Laws, A general theory of rods, Proc. Roy. Sot. London Ser. A 293 (1966) 145-155. [24] A.E. Green and P.M. Naghdi, On the derivation of shell theories by direct approach, J. Appl. Mech. (1974) 173-176. 1251 A.E. Green and W. Zerna, Theoretical Elasticity (Clarendon Press, Oxford, 2nd ed., 196Q). [26] M. Gurtin, An introduction to Continuum Mechanics (Academic Press, New York. 1981). [27] J.O. Hallquist, D.J. Benson and G.L. Goudreau, Implementation of a modified Hughes-Liu she11 into a fully vectorized explicit finite element code, in: P. Bergan et al., eds., Finite Element Methods for Nonlinear Problems (Springer, Berlin, 1986) 283-297. [28] T.J.R. Hughes and W.K. Liu, Nonlinear finite element analysis of shells: Part I. Three-dimensional shells, Comput. Meths. Appl. Mech. Engrg. 26 (1981) 331-362. [29] T.J.R. Hughes and W.K. Liu, Nonlinear finite element analysis of shells: Part II. Two-dimensional shells, Comput. Meths. Appl. Mech. Engrg. 27 (1981) 167-182. [30] T.J.R. Hughes and E. Carnoy, Nonlinear finite element shell formulation accounting for large membrane strains, Comput. Meths. Appl. Mech. Engrg. 39 (1983) 69-82. [31] T.J.R. Hughes, The Finite Element Method (Prentice-Hall, Englewood Cliffs, NJ, 1987). [32] A. Libai and J.G. Simmonds, Nonlinear elastic shell theory, in: J. Hutchinson and T.Y. Wu, eds., Advances in Applied Mechanics (Academic Press, New York, 1983). [33] W.K. Liu, ES. Law, D. Lam and T. Belytschko, Resultant-stress degenerated-shell eiement, Cornput. Meths. Appl. Mech. Engrg. 55 (1986) 259-300, [34] J.E. Marsden and T.J.R. Hughes, Mathematical Foundations of Elasticity (Prentice-Hail, Englewood Cliffs, NJ, 1983). [35] P.M. Naghdi, The theory of shells, in: C. Truesdell, ed., Handbuch der Physik, Vol Via/Z, Mechanics of Solids II (Springer, Berlin, 1972). [36] P.M. Naghdi, Finite deformations of elastic rods and shells, in: Proceedings IUTAM S~~sium on Finite Elasticity, Lehigh University, Bethlehem, PA, 1980. [37] F.I. Niordson, Shell Theory, North-Holland Series in Applied Mathematics and Mechanics (North-Holland, Amsterdam, 1985). [3X] K.C. Parks and G.M. Stanley, A curved Co shell element based on assumed natural-coordinate strains, J. Appl. Mech. 53 (2) (1986) 278-290. [39] H. Parish, Nonlinear analysis of shells using isoparamefric elements, in: T.J.R. Hughes et al., eds,, ASME 48 (ASME, New York, 1981).

304
[40] E. Ramm,

J.C. Simo, D.D. Fox, Stress resultant geometrically exact shell model. Part I

A plate/shell element for large deflection and rotations, in: K.J. Bathe, J.T. Oden and W. Wunderlich, eds. (Formulations and Computational Algorithms in Finite Element Analysis) (MIT Press, Cambridge, MA, 1977). [41] J.G. Simmonds and D.A. Danielson, Nonlinear shell theory with a finite rotation vector, Kon. Ned. Ak. Wet. B73 (1970) 460-478. [42] J.G. Simmonds and D.A. Danielson, Nonlinear shell theory with finite rotation and stress-function vectors, J. Appl. Mech. 39 (1972) 1085-1090. [43] J.C. Simo, A finite strain beam formulation. The three-dimensional dynamic problem. Part I, Comput. Meths. Appl. Mech. Engrg. 49 (1985) 55-70. [44] J.C. Simo, J.E. Marsden and P.S. Krishnaprassad, The Hamiltonian structure of elasticity. The convective representation of solids, rods and plates, Arch. Rat. Mech. Anal. (1987) (to appear). [45] J.C. Simo and L.V. Quoc, A three-dimensional finite-strain rod model. Part II: Geometric and computational aspects, Comput. Meths. Appl. Mech. Engrg. 58 (1986) 79-116. [46] J.C. Simo and L. Vu-Quoc, A beam model including shear and torsional warping distortions based on an exact geometric description of nonlinear deformations, Internat. J. Solids and Structures (1987) (to appear). 1471 J.C. Simo and L. Vu-Quoc, On the dyn~i~ in Space of rods undergoing large motions--A geometri~lly exact approach, Comput. Meths. Appl. Mech. Engrg. (to appear). [48] G. Stanley, Continuum-based shell elements, Ph.D. Dissertation, Applied Mechanics Division, Stanford University, Stanford, CA, 1985.

Potrebbero piacerti anche