This content was uploaded by our users and we assume good faith they have the permission to share this book. If you own the copyright to this book and it is wrongfully on our website, we offer a simple DMCA procedure to remove your content from our site. Start by pressing the button below!
A(t) £ 50(3) with A(0) = /, A(l) = A. By a standard continuation argument, using ip and the connectedness of S3, we can "lift" this path to a (smooth) path t i—> p(t) £ 6'3 such that
50(3) of (3.15) induces a two-sheeted covering
between the tangent bundles. More precisely, we have $(p',r') = $(p, r) for (p,r) £ T53 if and only if (p1, r') = ±(p, r).
26
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Proof. The relation
(p',r') = $(p,r), then ip(p) = ip(p') and D
(po)- Since, by Lemma 3.2, if in (3.15) is a local diffeomorphism, the curve R(t) in SO(3) induces a unique curve p(t) in S3 such that R(t) = (p(p(t))(= R p(t)) and p(to) = PQ. Thus, the curve (u(t),p(t)) in R3 x S3 characterizes the motion in precisely the same manner as does the curve (u(t), R(t))- In other words, as announced at the beginning of this section, we can indeed use instead of R3 x 50(3), as the configuration space of the rigid body. Evidently, Co is a submanifold of R3 x R4, but, since by Theorem 3.3 the curve (u(t), — p(t)) also characterizes the same motion, it always has to be understood that both (u,p) € Co and (u, — p) € Co correspond to the same configuration of the body. With this, (t,x(t),p(t),x(t),p(t)) is the state of the body at time t and hence, by Definition A.4 of tangent bundles and Corollary 3.4, the state space is M. be a local parametrization of .A/I near XQ, which is defined on an open subset Vm C R m . Moreover, let £o € V m , £o € Rm be the unique points such that 0(£o) = ^o and Z>0(^o)^o = %o- 1fx(t) is any curve in M. with x(to) = XQ, x(tg) = x0, then f(t) := $~l(x(t)) is a curve in Vm with ^(t 0 ) = Co, Co(*o) = Co- Furthermore, we have x(t) = £>0(^(t))^"(t) + D2(j)(£(t))(£,(t))2,so that the system (7.1) may be rewritten as
where For x = >(£) the operator -D0(£) is a linear isomorphism from Rm onto TXM; whence D(j)(£)is a liXMtoRm.Therefore, the system (7.6)isnchangednear isomorphism ofT
76
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
if the first equation is multiplied by D0(£) T - In fact, recall that, by (7.3) and (7.7), andbecauseFisabundlemorphism,therightsideofthiquationXM,x =sinT 0(0- Since the term D^)1 D±G(t, 0(£), £>0(00T coincides with D^G'i>(t^,£.)T, where ^(t, £, 0 := G(t, 0(£), £0(0£), we obtain the system
which is now in Rm and, therefore, to which Theorem 6.1 potentially applies. In order to check that Theorem 6.1 does apply to (7.8), we must verify the three conditions
By (7.4(i)) we know that
if and only if D
0(£o) is a bijection from Rm to TXOM, we have by (7.4(iii)) (y(t))y(t), and y(0) = 0, which shows that with the e2-axis. The shaft is assumed to rotate with a constant velocity c. Then the translation of the rotation into a motion along the axis of the shaft can be modeled by the condition ur - ctany = 0. Since if is required to be small, this will be simplified to (U) is an open subset of M. (under the topology induced by the standard topology ofRn) and 0 is a homeomorphism oflA onto >(£•/) • (ii) 0 is an immersion on U. If XQ & M and (U, $) is a local d-dimensional Cp parametrization of M such that XQ € 4>(U), then (U,<j>) is called a local d-dimensional Cp parametrization of M near XQ. 133
Now all stated results follow directly from Theorem 6.1. Remark 7.1. The solution x ( t ) (but not A(£)) obtained in Theorem 7.1 is independent of the choice of G, as long as G -1 (0) is unchanged near (to,Xo,Xo) and condition (7,4(iii)) continues to hold. This follows from Theorem 6.3 after transformation of the problem using a parametrization. Theorem 7.1 is tailored to systems such as (5.18) (or its multibody generalization (5.49)). In that setting, the use of Theorem 7.1 requires the explicit knowledge of initial conditions (u(t0),p(tQ)) = (u0,p0) and (u(to), p(t0)) = (u 0 ,po). In this respect, we recall that Remark 4.2 explains how p$ and po can be recovered from two more frequently available data, namely, the rotation RQ specifying the initial position of the body in the reference frame and the initial angular velocity vector WQ. As in section 6.2, we want to consider now also the case when in (7.1) the full-rank condition (7.4(iii)) no longer holds. Then, as before, the constraint G = 0 has to be changed into a new constraint G = 0 and we have to consider the effects of that change.
CHAPTER 7.
7.2
DAE FORMULATION ON MANIFOLDS
77
A Projected Derivative Operator
As in the previous section, let M be a smooth m-dimensional manifold and consider a mapping G : R1 x TM —*• R fc , as in (7.2(iii)). For reasons that should be clear by now, we want to incorporate into our theory the case when for (t,x,x) € R1 x TM. the derivative DxG(t,x,x) with respect to the variable x has constant, yet not necessarily full, rank. For the case M = R" this was considered in section 6.2, where G = F, and there we also saw the relevance of the operator DxT(t,x,x) for these results. In fact, there is no hope of expecting to develop the desired extensions of the theory to manifolds without using the "derivative" DxG(t,x,x}. Unfortunately, similar to section 4.4, this derivative does not exist except when M = Rn or when G(t,x,x) is independent of ±, since x is the base variable of the bundle TM.. The most important occurrence of Dxr(t,x,x) in section 6.2 is in the term Q7x,x)DxT(t1x,x)x in the definition (6.18) of the mapping F. In this section we(t will show that the absence of a derivative DxG(t,x,x) in the manifold setting does not affect the existence of a well-defined operator replacing Q(t,x,x)DxG(t,x,x). In other words, it is still possible to make sense of the combined term QDXG, even though DXG alone is not defined. In addition, some consequences of this result, needed later, will also be examined. Let (to,xo,±o) € R1 x TM be fixed, so that XQ € M and XQ G TXOM. Moreover, let PO 6 £(R /C ) be a projection operator onto rge DxG(to, ZQ, ±o) and denote by Qo G £(R fe ) the operator Q0 := / - P0- Note that DxG(t0,XQ,xo) is unambiguously defined as the derivative of the mapping x e TXoM i—> G(to,xo,x) € R fc at x — XQ. Let Um c M be the domain of a chart 0"1 containing the point XQ, and (j>: V™ —> Um the corresponding parametrization of M defined on an open subset Vm C Rm. We denote by £o € Vm and £o € Rm the (unique) points for which <j>(£o) = x0, ^50(Co)Co = %$• Every mapping H on R1 x TM (or, more generally, on R1 x TUm) induces a mapping H^ on R1 x Vm x Rm via
L e m m a7.2 of the parametrization 4>. Proof. Let ip be another parametrization with VK^o) = X 0) D^>(r]o)r)o — XQi so that, with the change of coordinates B := (f>~o •!/>, we have G^(t, 77,77) = G^^.O^^B^fi). Differentiation with respect to r\ then implies that
Since, by equation (7.9), DfG^^o^o,^) = DxG(t0,xo,±o)D(f>(^0), QoD^^o^o^o) = 0, and, hence, by multiplying (7.10) by Q0, that
we obtain that
78
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
With
it then follows that
as had to be proved. Lemma 7.2 shows that for any point (t, x, x) £ R1 x TM. and any projection P(t, x, x) onto rge DxG(t,x,x), we may define, with Q := I — P, the operator [QDxG}(t,x,x) € C(TxM,Rk) by
arametrization of M. about x such that x = >(£) and x = -D0(£) whence Q*(£,£,£) = Q(t,x,x). The (bracketed) notation [QDXG] is meant to emphasize that, generally, Q and DXG cannot be separated, that is, that [QDxG](t,x,x) should be thought as, but is not equal to, Q(t,x,x}DxG(t,x,x), which does not exist since DxG(t, x, x) does not exist. Since in (7.11) we have x = £>>(£)£> it follows that
irrespective of the parametrization (/>. We are now in a position to generalize the definition of the set E(F) in (6.16) and begin with a preliminary remark. Lemma 7.3. The set
is independent of the choice of the projection Q(t, x, x) := / — P(t, x, x), where P(t, x, x) projects onto rge D±G(t,x,x). Proof. Since, for a given parametrization 0 with x = >(£) and x = £>>(£)£, we have
it follows from (7.11) that [QDxG](t,x,x)x + Q(t,x,x)DtG(t,x,x) = 0 if and only if <9*(*,&e)[-De G *(*.60€ + AG*(t,£,C)] = °- But <9*(*.£>0 = Q(*,x,x) projects onto a complement of the range of DxG(t, x, x), and, since DfG^(t, £, |) = DxG(t, x, ±)£>^>(£) and D
which is independent of Q(t, x,x). In view of Lemma 7.3 we may now define the set
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
79
which is independent of the choice of Q(t, x, x). In analogy to (6.17), suppose that
where V is an open neighborhood of E(G) in R1 x TM. This ensures that projection operators P(t,x,x) onto rge D±G(t,x,x) exist, which depend smoothly upon (t,x,x) e V. For instance, orthogonal projections have this property. As before, we set Q(t, x, x) := I - P(t,x,x) and define for (t,x,x) e V:
Obviously, since Q is smooth, the right side of (7.12) is a smooth function of (£,£,£). TherefoxG](t,x,x)x is a smooth function of (t,x,x) e V and also G in (7.15) isre, [QD smooth. A trivial but important consequence of (7.11) and (7.15) is that
which shows that the set G~ 1 (0) n G~l(Q) is independent of the choice of G, that is, ofQ. Remark 7.2. It is important to note that there is no ambiguity in this approach because the definition of [QDxG](t,x,x), and thus of [QDxG](t, x,x)x, requires no smoothness or even continuity of Q with respect to (t, x, x). In particular, Lemma 7.3 holds without any such continuity requirement. This is crucial to defining £(G) in (7.13) independently of Q(t,x,x), which in turn shows that the neighborhood V of £(G) on which the constant rank condition (7.14) holds is independent of Q(t,x,i), and finally permits the definition of [QDxG](t, x, x)x as a smooth function of (t, x, x) when Q itself is smooth. By using (7.12), we may rewrite (7.15) as G(t,x,x) = G^ (£,£,£), where x = <£(£), ± = D(t>(£)£, and & is given by (6.18) with T, P, and Q replaced by G*, P*, and Q*, respectively. By (7.9), this is Since (t,x,x) € G^O) (respectively, G"1^)) if and only if (£,£,£) £ (G*)- 1 ^) (respectively, (G*)- 1 (0)), it follows from (7.16) and (7.17) that
The relations (7.17) and (7.18), together with the arguments of Lemma 7.3, now allow us to recover important properties of G from their counterparts proved in section 6.2 for the case M = Rn. Lemma 7.4. For (t,x.x) 6 S(G) we have
Furthermore, the nullspace kerZ?iG(t,x,x) is independent of the choice of the projections P and Q used in the construction of G, Thus, in particular, rank DxG(t,x,x) is independent of P and Q.
80
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Proof. Let <j> again be a parametrization of M. and set x = 4>(£), x = £>^(£)£. By (7.9), and since D<j>(£) is an isomorphism of Rm onto TXM, we have ker D±G(t,x,x) = £>
Two different choices for the projections P and Q result in two different choices for the projections P^ and Q* used in constructing G*. By Lemma 6.5, the nullspace ker DfG^(t, £, £) is unaffected by these changes and hence ker D^G(t, x, x) is independent of P and Q. Clearly, this implies the independence of rank DxG(t,x,x). Unlike in the case M = R", the condition rank D±G(t,x,x) = k is not expressed by the (here meaningless) relation (6.27) for F = G. However, this relation continues to be available with F = G^, and this provides a way of checking the condition rank D±G(t,x,x) = k by using local parametrizations. Moreover, Lemma 6.4 is easily generalized to the manifold setting by means of (7.16), (7.17), and Lemma 6.4 itself with F = G*. Lemma 7.5. A smooth curve x(t) in M. defined on some open interval J C R1 satisfies G(t,x(t),x(t)) =0 on J if and only if it satisfies
at some point to G J.
7.3
Nonmaximal Rank in the Manifold Case
We return again to the setting and notation of section 7.1 and assume now that the mapping G of (7.2(iii)) satisfies the constant, but not necessarily full, rank condition (7.14); that is,
where V is an open neighborhood of the set S(G) in R1 x TM. Recall that the set £(G) of (7.13) is independent of the projection Q appearing in the definition. The condition (7.14) ensures that the (smooth) mapping G in (7.15) is defined on V and, hence, if desired, can be extended to all of R1 x TM. Recall, moreover, that the smoothness of G depends upon the smoothness of P which cannot be continuous if (7.14) fails to hold. By Lemma 7.5 and Theorem 7.1 with G replaced by G, and since £(G) = G -1 (0) n G-1(0) by (7.16), we obtain at once the following generalization of Theorem 6.7. Theorem 7.6. Suppose that the constant rank condition (7.14) holds so that the mapping G in (7.15) is well defined and smooth. Let (to,xo,x0) € £(G) and assume that
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
81
rank DxG(t0,x0,xo) = k and ( A ( x 0 ) z , z ) > 0 for every z € kerDxG(t0,x0,x0), z ^ 0. ITien i/?,ere is an open interval J C R1 with to & J such that the problem (7.1) with G replaced by G and A, F satisfying the conditions of Theorem 7.1 has a unique solution (x(t),X(t)) defined on J and satisfying x(t0) = x0, ±o(to) = io- Furthermore, ( t , x ( t ) , ± ( t ) ) € £(G); whence G ( t , x ( t ) , x ( t ) ) = 0 for every t <E J. Lemma 7.4 shows that the hypotheses of Theorem 7.6 are independent of the choice of G defined by (7.15). As in the case M = R™, we then obtain the following extension of Theorem 6.9. Theorem 7.7. The solution x(t) (but not A(i)) obtained in Theorem 7.6 is independent of the choice of G. Proof. Once again let 0 : Vm —> M be a parametrization of A4 near XQ defined on some open subset Vm C R m and £0 6 Vm, £0 e Rm the unique points for which x o-fj 7-1), G is replaced by G, this system reduces to (7.8) with G* replaced by G*. Since G* = G* by (7.17), two different choices for G thus lead to two different choices for G*. But, by Theorem 6.9, the corresponding solution £(£) of the system (7.8) with either choice of G$ and £(to) = £o ; £(^o) = £o is independent of that particular choice. Since the solution obtained in Theorem 7.6 is x(t) = <j>(£(t)), it is independent of the choice of G. The significance of Theorem 7.6 regarding the constrained motion of a rigid body is easy to infer from the comments at the end of section 5.2, but there are some interesting subtleties that could not be mentioned before. Let B be a rigid body subjected to a constraint G = 0, where G : R1 x T(R3 x 3 5 ) —> R fc and rank DxG(to,XQ,Xo) = k at some point (£0,2:0, .TO) € G~ 1 (0). Then the generalized Gauss principle ensures that the motion of B beginning at the "initial" state (£o,£o,£o) is characterized by the system (7.1) with appropriate A(x), F(t.x,x) as in Theorem 5.1 (in particular, x — (u,p),x = (it,p)) and, of course, with the initial conditions x(to) = XQ, x(to) = XQ. Suppose now that rank D±G(t,x,x) = r < k for (t,x,x) in a neighborhood of (to,z 0 ) xo) and that (t0,x0,x0) € £(G). Recall that the set E(G) of (7.13) depends intrinsically upon G. Then G is well defined by (7.15) in the neighborhood of (to, XQ, XQ, and every motion of B that is compatible with the constraint G = 0 must also be compatible with the constraint G = 0. This is (i) of the "only if" part of Lemma 7.5. It also shows that there exists no motion of B originating from the state (to,x0,x0) that is compatible with G = 0 unless (t0,x0,Xo) € G- 1 (0)nG~ 1 (0). By (7.16), G-l(Q)riG-1(0) coincides with S(G), the very set in the vicinity of which G was constructed. This discussion shows that, when rank .D±G(£o,£o,io) — k, the motion of B starting at (to,xo,±0) 6 E(G) must be characterized by the system (7.1) with G replaced by G. Thus, the correct equations of motion, when rank D±G is constant but not full, are implied by what they are when rank D±G is full: No extra physical argument is needed and everything follows from the generalized Gauss principle. Now Theorem 7.6 asserts
82
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
that the solutions of the system (7.1) with G replaced by G and with the initial state (to,xo,xo) 6 £(G) do comply with the original constraint G —0. Theorem 7.7 settles the fundamental question of the uniqueness of the motion constrained by G = 0 and originating from the state (to,xo,xo). This is by no means an obvious issue, nor a usual one, since the equations of motion corresponding to the constraint G = 0 are not directly expressible in terms of G but, rather, in terms of G, and there is no canonical choice for G. By proving that the specific choice of G has no effect on the solution, Theorem 7.7 answers the uniqueness question in the only way possible for such a problem. Last, the fact that the hypotheses made in Theorem 7.6 are also independent of the choice of G (see Lemma 7.4) completes the justification of the procedure of replacing G by G in (7.1) to obtain the correct equations of motion whe rank D±G is constant but not full in the neighborhood of some point of £(G). While our discussion here focuses on the case of a single rigid body, everything extends verbatim to systems of rigid bodies or mass points or both: only the manifold M. changes. The case of a geometric constraint G(t,x*) = 0 considered in section 5.2 is certainly a simple example where the full-rank condition for D±G fails to hold since D±G = 0. But, as noted, more general examples are provided by systems of rigid bodies where some bodies are subjected to geometric constraints and others to kinematic ones. It was this wide range of examples that prompted our development here of the general theory for the rank-deficient case. Remark 7.3. Theorem 7.7 gives only a partial answer to the dependence of the solution in Theorem 7.6 upon G and G,j3ince it shows only that, for given G, the solution does not depend upon the choice of G. This does not address the dependence of the solution upon G and, in fact, this issue is not solved by Remark 7.1 unless r = k. However, it is true that the solutions of Theorem 7.6 do not depend upon G as long as S(G) is unchanged and the constant rank condition (7.14) continues to hold (with the same r). The problem can be reduced to the case M. = R" via parametrizations and then handled by the subimmersion theorem rather than the implicit function theorem used in the proof of Theorem 6.3 for r = k. The proof is technical and is omitted here.
7.4
Reduction to DAEs on Euclidian Space
The problems of section 7.1 were posed on a submanifold M of R" and, by means of local parametrizations, reduced to problems on R m , m = dim.M, of the form (7.8). This is appropriate for theoretical considerations but may. introduce complications when it comes to the numerical implementation. Accordingly, we describe now a procedure by which our systems on M. can be transformed into systems on Rn without the use of any local parametrization. This procedure represents a simple generalization of our approach in Chapter 5 when passing from the formulation (5.18) in Theorem 5.1 to (5.26) in Theorem 5.2. We use the same notation as in section 7.1. An m-dimensional submanifold M. C Rn is always described, at least locally, as the zero set of a submersion g : Rn —> R n ~ m . In
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
83
the case when M — R3 x S3 C R3 x R4, this is even true globally when one works with x — (u,p), g(u,p) : — \p\2 - 1. The same holds when M is a product of copies of R3 x 53. Hence, in the applications of interest here, it is not restrictive to assume that
where g is a submersion on M. = g~ 1 (0) and thus we have rank Dg(x) = n — m for every x e M. As before we continue to require that A(x) & £S(R") satisfies
but, in addition, we also assume that A(x) has a smooth extension AE(x) defined for every x € R n . Likewise, we assume that the tangent bundle morphism F : R1 x TM —> R extends as a smooth mapping FE : R1 x R" x Rn -> Rn and that G : R1 x TM -> R fc , k <m = dim.M, extends smoothly to GE : R1 x Rn x Rn -» R fe ; that is,
A solution (x(t), A(t)) of the system (7.1), with the curve x(t) lying in M, can also be viewed as a solution of the same system with x(t) as a curve in R n constrained by the condition g ( x ( t ) ) = 0. Thus, to solve the original problem, we may consider solving (6.1) in R" with A and F replaced by AE and F£, respectively, and with the constraint function F given by
However, things are not as simple since F in (7.22) never satisfies the full-rank condition DiF(i. x. x] = n—m+k at any point. In the previous section we developed a procedure for handling constraints that fail to satisfy the full-rank condition, and, in fact, for the case of R", the same but simpler procedure was discussed in section 6.2. For this discussion the simpler variant will suffice. Suppose that Since D±G(t, x, x) = D±GE(t, x, x)\T M , it follows that rank Dj..GF'(t, x, x) = k for every (t,x,x) € G-^O). It is then straightforward to check that the set E(F) of (6.16) (with F given by (7.22)) is because TM — {(x, x) € Rn x R" : g(x) = 0, Dg(x}± — 0}. Moreover, the relation D±T(t,x,x) = (DxGE(t,x,x),Q) shows that rank D±r(t,x,x) = k is constant on some open neighborhood V of S(F) in R1 x Rn x R n . The conditions now hold for constructing a mapping F via formula (6.18), and we may even choose P(t,x,x) independent of (t,x,x), that is, as the natural projection R n - m x R f c ^ R f c along the first factor. Then Q(t,x,x) = I-P(t,x,x) is the projection R n-m x R fc _j R n-m aiong ^ ne second factor and we obtain
84
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
An explicit example of such a mapping F is the formula (5.23) in the case when M = R3 x S\ x = (u,p), and g(u,p) := |(b|2 - 1). From (7.23) and (7.24), and because TXM = k e r D g ( x ) for every x e M, it follows that Let (t0,x0,±0) € £(F) be fixed. By (7.24), we have G(tQ,x0,±0) = 0. Moreover, (7.23) gives i&nk D±G(to,Xo,xo) = k. Therefore, conditions (7.4(ii)) and (7.4(iii)) hold with G. The remaining condition (7.4(i)) will be met by assuming that
With this, Theorem 7.1 applies and ensures that the system (7.1) with initial conditions x(t0) = x 0 (e M), x(t0) = i0(e TXOM) has a unique solution (x(t),X(t)) with x ( t ) lying in M.. By (7.25), the characterization TXOM = ker Dg(x0) and the relation
provide that kerDxG(to,xo,xo) = ker DxT(to,x0,±o). By (7.27), we obtain that
By (7.24), (7.26), and (7.28), and since rank D±T(t,x,x) = k on the open subset V, Theorem 6.7 ensures that the system (6.1) with A, F, and F replaced by AE, FE, and F, respectively, and with initial conditions x(t0) = x0 and X0(t0) = XQ has a unique solution (x(t), A(i)) with x(t) in R". The following theorem shows that the notation x(t) is consistent with that used above when calling (x(t), A(t)) the solution of (7.1) with x(t) lying in M.. Theorem 7.8. With the splitting A(t) = (A(*),//(*)) € Rfc x R n ~ m , the pair (x(t),\(t)) is the (unique) solution given by Theorem 7.1. Proof. Since (x(t),X(t)))),solves (6.1) with A, F, and F replaced by AB, FE, and f, respectively, we have, omitting t-dependence for notational convenience, that
andx(to) = x0, x(t0) = XQ, (t0,xo,xo) € S(F). Theorem 6.7 ensures that T(t,x(t),x(t)) + 0; whence g(x(t)) = 0 and therefore x(t) € M by (7.19) and G(t,x(t),x(t)} = 0 since GB 1 = G. Thus, x(t) lies in M. and satisfies the constraint (7 = 0. In particular, \R xTM
AE(x(t)) = A(x(t)) and FE(t,x(t),x(t)) = F(t,x(t),x(t)) by (7.21). For the completion of the proof, it remains to show that the first equation in (7.29) entails
since it is indeed the x-derivative of G that is involved in Theorem 7.1.
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
85
Thus, the first equation in (7.29) becomes
By the hypothesis (7.20), A(x) maps into TXM for x 6 M. Moreover, F is a bundle map; whenceF(t,x(t),x(tx(t}M.On the other hand,Dg(x)maps into(TXM)Lfor))£T x e MXM = kerDg(x) and rge Dg(x)T = [ker Dg(x)]-L. Thus, bymultiplyingsincimT (7.31) by the orthogonal projection onto TXM, we obtain
Here the last term is the tangential component (equal to the orthogonal projection) of D±GE(t,x,x)'rXontoTXM, discussion of the special case at the beginning of section 5.1). Thus, (7.29) holds and the proof is complete. A similar procedure can be used in the more general setting of Theorem 7.6, that is, when condition (7.23) fails. Here (7.15) and (7.11) indicate that the calculation of the mapping G in the manifold setting requires the use of parametrizatioiis and is rather complicated. Thus in this case a formulation of the problem in Euclidian space does offer a particular advantage. However, such a formulation in Euclidian space cannot be obtained with an arbitrary smooth extension GE of G. This contrasts with the case when condition (7.23). that is, condition (7.4(iii)) of Theorem 7.1, holds. In order to highlight this fact further, we proceed in reverse order and first describe a problem in Euclidian space that yields solutions to a problem on M essentially similar to the one resolved in Theorem 7.6. We assume that the mapping GE satisfies
on an open neighborhood V of E(G £ ) n (R1 x TM] in R 1 x R n x Rn (and not merely in R1 x TM). Here, T>(GE) is given by (6.16) with F replaced by GE, while F continues to be defined by (7.22). We thus have
so that V is also an open neighborhood of E(F) in R1 x Rn x Rn and
since rge Dxr(t,x,x) = rge DxGE(t,x,x) x {0} C Rk x R n ~ m . Then, it is obvious how a projection P(t,x,x) e £(Rk) onto rge D±GE(t,x,±) yields the linear projection P(t,x,x) x 0 from ~Rn~~m x R fc onto rge DxT(t, x, x), and, similarly, how Q(t,x,x) = I — P(t, x, x) provides the projection Q(t, x.x)xl onto some complement of rge DxT(t, x, x). Thus, we obtain a mapping F derived from F via formula (6.18) in the form
86
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
where GE is given by (6.18) with F replaced by GE. From (7.33) and the characterization TM = {(x,x) € R" x R™ : g(x) = 0, Dg(x}± = 0} it is clear that
which therefore leads to the assumption
Now let (t0,x0,x0) e R1 x TM. Since
it follows from Theorem 6.7 that, when
then the system (6.1), with A, F, and F replaced by AE, FE, and F, respectively, has a unique solution (x(t), A(t)) for which x(to) = XQ, x(t0) — ±o and which also satisfies (t,x(t),x(t)) € S(F). In particular, we have T(t,x(t),x(t)) = 0 and thus g(x(t)) = 0; that is, x(f) e M, as well as GE(t,x(t),x(t)) = G(t,x(t),x(t)) = 0. Below, we will show that, under our assumptions, the hypotheses of Theorem 7.6 are satisfied and that, when A(t) is split in the form A(t) = (A(<),//(*)) e R* x Rn~m, then (x(t), A(t)) is the (unique) solution provided by Theorem 7.6. In both steps, we need the following technical lemma whose proof will be postponed until the end of this section. Lemma 7.9. Under the stated assumptions the following two relations hold: (i) E(G) C E(GE) n (R1 x TM). (ii) I/rank DxGE(t,x,x) = rank DxG(t,x,x) for all (t,x,x) 6 V n (R1 x TM), and P(t,x,x) e £(R fe ) is a projection onto rge DxGE(t,x,±) for (t,x,x) £ V, then, with Q:=I-P,
Lemma 7.10. Under the stated assumptions the hypotheses of Theorem 7.6 are satisfied. Proof. First, we must show that rank D±G is constant on some open neighborhood of E(G) in R1 x TM and hence that (7.14) holds. By Lemma 7.9(i), it suffices to show that this is true on some open neighborhood of S(G£) n (R1 x TM} in R1 x TM. In the following all the mappings are assumed to be evaluated at a point (t,x,x) € S(GE) n (R1 x TM) = E(F) (see (7.33)). By (7.36) and (7.37) we haveAf = ra n—m+k and hence by Lemma 6.5 with Q replaced by Qx.1, rank (DiF + (Qx/)£> a; r) = n - m + k. Since D±T + (Q x I)DXT = (D±GE + QDXGE, Dg(x}} by (7.22), this means that rank (DxGE+QDxGE,Dg(x)) = n-m+k, i.e., dimkeT(D±GE+QDxGE,Dg(x)) = m-k. Since kerDg(x) = TXM, this also reads dimker^iGjf^+Q-DsGjf,^) = m-k.
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
87
Thus rank xGfT M + QDxGfT M) = k at every point (t, x, x) 6 E(G£) n (R1 x TM)(D and hence on an open neighborhood VM of E(G£) (^(R1 x TM) in R1 x TM since k is thee maximum possible rank. Therefore we have rank (DXGE + QDXGE)\TTM — ^> where the mappings are evaluated at (t,x,x) € VM. This implies that rge DXGET^M = rge DXGE, for, given 2 e R n , there exists h 6 TX.M such that DxGEh+QDxGBh = DxGEz; whence DxGBh = DxGEz by multiplying both sides by P. Thus, rank DxGfT^M = rank DXGE on VM. But, since D^G^ M — D±GE, this amounts to rank DXG = rank DXGE on VM; whence rank DXG = r on V.M by (7.32), after shrinking VM if necessary, to ensure that VM C V. We now prove that
By the demonstrated constancy of rank DXG on V», G is well defined by the formula (7.15). The above proof also showed that rank DXG — r := rank DXGE on VM- Since rge DXG C rge DXGE, this implies that rge DXG = rge DXGE. so that Q|v^ is a (smooth) projection onto a complement of rge DXG and therefore can be used for the calculation of G. On VM, the conditions required in Lemma 7.9(ii) are satisfied. Therefore, by (7.15) and (7.39), and, since GE = G on R1 x TM, we have
which, of course, is simply the relation G = GE\vM since GE is given by (6.18) with F replaced by GE. By the openness of VM in R1 x TM, this implies that D±G — D±GE\TtM on VM and thus, because of E(G) C VM, that
Now (7.40) follows from (7.41) and the hypothesis (7.37) since, by Lemma 7.9(i), E(G) C S(G E ) n (R1 x TM). Actually, the relation G = GE\\>M, proved above, together with (7.33) and S(G) = G-1^) n G-HO) as well as S(G£) = (G^)- 1 ^) n G^'^O) (see (7.16) and (6.25) with F = GE, respectively) shows that
(However, here (7.32) and (7.37) are used via arguments involved earlier in the proof to justify (7.42). In general, only Lemma 7.9(i) is true.) By (7.42), the point (t 0 , x 0 ,x 0 ) 6 E(G £ )n(R 1 xTM) is actually in S(G). From (7.40) it also follows that rank D±G(t0, x0, XQ) = k. Finally, the last property to check for ensuring the applicability of Theorem 7.6 is that (A(xo)z.z) > 0 for z 6 ker DxG(t0,x0,±0), z -£ 0. But, due to (7.41), it is immediate that this is the earlier assumed condition (7.38).
88
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Remark 7.4. Conversely, the hypotheses of Theorem 7.6 do not imply the validity of the condition (7.32) except when r — k. Thus, condition (7.32) is the additional requirement necessary to transform the problem from one on M. to one on R n . It merely says that not every smooth extension GE of G can be used. Otherwise, such a transformation is not possible. Before proving Lemma 7.9, we complement Lemma 7.10 by generalizing Theorem 7.8 as follows. Theorem 7.11. With the splitting X(t) = ( \ ( t ) , f i ( t } ) £ R f e xR n ~ m , the curve (x(t),X(t)) is the unique solution given by Theorem 7.6. Proof. We proceed as in the proof of Theorem 7.8. As noted after (7.38), we have
whence x(t] lies in M and GE(t,x(t),x(t}) = 0, so that G(t,x(t),x(t)) = 0. Thus, it suffices to show that (x(t), A(t)) solves the equation
for then the uniqueness part of Theorem 7.6 (available by Lemma 7.10) proves our claim. The relation with F given by (7.35) yields
Next, by using AE(x(t)} = A(x(t}} and FE(t,x(t),±(t)) = F(t,x(t),x(t)} since x(t) lies in M, and by multiplication by the orthogonal projection onto TXM — kerDg(x), we obtain where [D±GE(t, x, x) T A]r is the tangential component of DxGE(t, z,i) T A and hence equals (D^GE(t,x,x)\TiM)r X. Now, by the relations (7.43), (7.42), and (7.41), we have D±GE(t,x,x)lTiM = D±G(t,x,x), and (7.44) follows. For completeness, we now prove Lemma 7.9. Proof of Lemma 7.9. (i) The characterization of S(G) in section 7.2 uses local parametrizations and a reduction to the case when M. is an open subset of Rm. Naturally, any suitable parametrization can be used. In particular, we can apply the parametrization induced by a local diffeomorphism of Rn = Rm x R n ~ m that maps (locally) R m x {0} onto M. This reduces the problem to the case when M. is an open subset of R m , GE has the form GE(t,xi,x2,xi,X2) relative to the splitting R" = Rm x R n ~ m , and G(t,x,x) = GE(t,xi,0,xi,0) with x = (n,0), x = (ii,0).
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
89
In that setting, to say that ( t , x , x ) € £(G) means that
and that i.e., that there exists hi & Rm for which
On the other hand, to say that (t,xi,X2,xi,X2) € T,(GE) means that G B (i,xi,x 2 l xi,x 2 ) = 0 arid that
i.e., that there exist (/ii,/i 2 ) e Rm x R n ~ m such that
Because
it is clear that when (7.45) holds, then (7.46) holds with x 2 = 0, £'2 = 0, and /i2 = 0. Thus {(t,x,x) € £(G),x = (x!,0),x = (x - i,0)} implies that (t,xi,0,xi,0) € £(G B ) and thus ( t , x , x ) € E(G E ). It is obvious that E(G) C R1 x TA1 and hence, as claimed, E(G) c S ( G £ ) n R 1 xTAi. (ii) For (t, x, x) e Vn(R J xTX) the condition rank D±GE(t, x, x) = rank D.iG(i, x, x) together with G = G^_ lxTjVf show that
Hence the projection P(t, x, x) can be viewed either as a projection onto rge D^GE(t, x, x) in Rfe or as a projection onto rge DxG(t,x,x) in R fc . Thus, the operator [QD xG](t,x,x..) makes sense for (t, x, x) e V n (R1 x TM). Since [QDXG] is defined via local parametrizations, we may proceed as in the proof of part (i) above, thereby reducing the problem to the case when M is an open subset of Rm C R™ and GE = G £ (i,xi,x 2 ,x 1 ,i 2 ), G(t,x,x) = GE(t,xi,0,±i,Q) where x = (xi,0), x = (xi,0). In that case, the definition of [QDxG](t,x,x) is
90
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
while, since TXM = Rm x {0} and x = (ii,0), x = (zi,0),
Thus [QDxG](t,x,x) = Q(t,x,x)DxGE(t,x,x)\TIM>
an
d the proof is complete.
Remark 7.5. The results of this section can be summarized as follows: Let M C R™ be characterized by M. = g~l(0), where g : Rn —» R ra ~ m satisfies the condition rank Dg(x) = n — m at every point x € M. Assume that for x € M, A(x) € £5(R") satisfies the condition rge A(x) C TXM of (7.20). Furthermore, assume that F — F(t,x,x) is a bundle morphism, that is, that F(t,x,x) £ TXM for (t,x,x) € R1 x TM. Finally let G : R1 x TM —> R fc . Now suppose that A, F, and G have (available) extensions to R1 x Rn x R", denoted by AE, FE, and GE', with AE(x) e £ 5 (R n ) for x e R". Assume that the condition (7.32), i.e.,
holds on some open neighborhood V of £(G £ ) in R1 x R™ x R". Then a mapping G~£ : R1 x R™ x Rn -> R fc can be defined by (6.18) and can be extended arbitrarily outside V. If the condition (7.37) holds, that is, if rank£»iGS(i,x,x) = k
V (t,x,x) € S(G£),
then the mapping T(t,x,x) := (GE(t,x,x),Dg(x)x) (7.36): rank Dxf(t,x,x)=n-m + k
satisfies the full-rank condition
V(t, x, x) e £(F) = T,(GE) n (R1 x TM) = S(G).
Now let (to,xo,io) £ S(F) = S(G) be any point such that (A^rro)-^ z) > 0 for every 2 e keri'iF^o^o^o) = kerZ)iG B (to>xo)^o)|T, M> z ^ 0- Then the problem (6.1) with ^4, F, and F replaced by AE, FE, and F, respectively, has a unique solution (x(t), A(t)) such that x(to) = XQ, i(io) — ^Oi and that, with the splitting A(£) = (A(t),/u(i)), the curve (x(t),X(t)) is the unique solution provided by Theorem 7.6, which involves the original mappings A, F, and G and not their extensions AE, FE, and GE. If r = k in (7.47), then Theorem 7.11 coincides with Theorem 7.8 and G^ =GE above. For r < k in (7.47) the hypotheses made in this summary are stronger than those needed for the validity of Theorem 7.6 since they put limitations upon the choice of the extension GE. But without them the solution (x(i),A(t)) of Theorem 7.6 can no longer be obtained by the above procedure of extending the problem to Euclidian space. Fortunately, this theoretical limitation does not appear to be a serious one in the applications to rigid body motion.
7.5
Characterization of State Spaces
The material developed in this chapter helps to clarify the concept of a state space for a system of rigid bodies and/or mass points. Let M. C R" be a submanifold representing
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
91
the configuration space of such a system, so that M is a product of copies of R3 x S1'3 and/or R 3 . Let the system be subjected to a constraint of the form G(t.x,x) = 0, where G : R1 x TM —> R fc , k < m — dim M. The essence of the results of this chapter is that, under general conditions, the only states (t,x,x) 6 R1 x TM compatible with the constraint are the points of (t,x.±) of the set £(G) introduced in (7.13), which depends intrinsically upon G. In fact, the motion of the system satisfying the constraint G = 0 under any external force F = F(t,x,x) can only occur from a state (to,xo,±o) € S(G), in which case the motion x ( t ) has the property that ( t , x ( f ) , x ( t ) ) € S(G) at all times. The conditions for E(G) to represent all the possible states of the system during the motion are simply that
where V is an open neighborhood of E(G) in R1 x TM., and that any mapping G denned from G by (7.15) satisfies the full-rank condition
Indeed, in Lemma 7.4 this condition was seen to be independent of the choice of G. Thus, when (7.48) and (7.49) hold, it makes sense to call S(G) the state space S of the system. Recall that (see (7.16))
and that, by (7.15) with Q = 0,
This definition of S coincides with the definitions introduced in Chapters 3 and 5 in the case of rather simple constraints G = 0. To corroborate this, we briefly review a few examples. (a) Mass-point systems. Here M = R3£. For r = k in (7.48) we have S(G) = G^O) by (7.50) and (7.51). The set G -1 (0) was indeed our choice for the state space S in (2.4) (where
which is S in (2.19). (b) Rigid body. In Chapter 3 we saw that the state space of a rigid body B is M — R3 x S3 provided that (u,p) and (u,—p) correspond to the same physical configuration. As a result, if G : R1 x TMm—> R* is used to constrain B, and if r — k in (7.48), then, as in (a) above, we find that
92
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
With x — (u,p), x = (u,p), this is the state space S of (5.7). A geometric constraint (r = 0) leads to
Note that in this case DXG exists because G is independent of x and, because Q = I, [QDXG] = [DXG] in (7.12) coincides with DXG. Clearly, the general procedure for the calculation of S(G), as well as the correct conditions, that is, (7.48) and (7.49), under which £(G) can legitimately be called the state space S of the system, make it possible to characterize S for much more complicated systems where intuition and common sense would be of limited help. It should also be observed that the "natural" choice of R1 x T(R3 x S3) for the state space of an unconstrained rigid body is also provided by this method in spite of the fact that the constraint involved in that case has no physical meaning. This is shown in the next example. (c) Unconstrained rigid body. The configuration space R3 x 53 of an unconstrained rigid body is characterized by {(u,p) G R3 x R4 : \p\ — 1 = 0}, and hence
may be viewed as a constraint on M := R3 x R 4 . Clearly, for this G we have r = 0 in (7.48) and G(t,u,p,u,p) := 2pTp by (6.18) with T = G. Since G~l(Q) = R1 x R3 x S3 and, obviously, rank D^,p}G(t,u,p,u,p) — k = 1 on G~ 1 (0), the condition (7.49) holds on G-^O) and hence on E(G) C G-^O). Thus
Since there is no "physical" body with configuration space R3 x R4, the constraint |p|2 — 1 = 0 has no physical meaning. Yet the result that R1 x T(R3 x 53) should be the state space of an unconstrained rigid body is the (physically) correct one.
7.6
ODE Formulations
The equations of constrained motion given in Chapter 7.1, or in more general, abstract form by the DAE (7.1), can sometimes be manipulated to yield new formulations. For instance, in the case of mass-point systems, we showed in section 2.1 that the Lagrange multiplier A can be explicitly calculated as a known function A = A(i, x, x) of the state of the system. This procedure leads to equations of motion represented by the ODE (2.12) instead of the DAE (2.11). This approach can be extended to systems of rigid bodies and, in fact, to general DAEs of the form (7.1) on a submanifold M of R n . As in section 7.1 we assume that A : M —> £S(R"), F : R1 x TM —* Rn is a tangent bundle morphism and G : R1 x TM —> R fc . We further assume that the full-rank condition
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
93
holds and that These assumptions, or at least minor variants, were already involved earlier, e.g., in Theorem 7.1. In addition, we require that A(x) is positive definite on TXA4,
which, together with (7.53), implies that
Since A(x] is symmetric, it follows from (7.55) that
Note that (7.54) is stronger than condition (7.4(i)) required in Theorem 7.1. From these conditions it follows that A(x)\TxM G GL(TXM] for every x G M. We spt
Suppose now that x — x(t), A = X(t) solves the system (7.1), where, of course, x(t) is a curve on M. By (7.56) we have A(x(t}}x(t] = A(x(t}}xT(t], where XT(£) e TX^M. is the tangential component of x(t). Thus, dropping the t-dependence for simplicity, we obtain and hence, with (7.57),
As a result, we have
By (7.52) and (7.57) the operator
is invertible; whence
94
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
As in Remark 5.1 we use a local extension GE of G to make sense of the partial derivative DxG(t,x(t),x(t))x(t). Then differentiation of the identity G(t,x(t),x(t)) = 0 yields
By writing x = XT + XN-, where, of course, XN represents the normal component of x, we obtain
But XN depends only upon x and x via the formula
where Vx denotes the second fundamental tensor of the manifold M at x 6 M. (see, e.g., Choquet-Bruhat, DeWitt-Morette, and Dillard-Bleick [7] and also Rabier and Rheinbold [18] for an elementary description and computational aspects). Hence, by substituting (7.61) into (7.60) and the result into (7.59), we obtain a representation
of A in terms of a known function A : R1 x TM —> R fc . With this equation (7.58) becomes As we shall see below, equation (7.63) is a second-order ODE on .A/I, which has a unique solution for given initial condition
Moreover, if x0 and ±0 in (7.64) are chosen so that G(t,xo,±0) — 0, then the solution of (7.63) satisfies the constraint G(t, x(t),x(t)) = 0 at all times. The simplest way to see this is to recall that the DAE (7.1) with the same initial condition has a unique solution by Theorem 7.6, and, because (7.62) gives a formula for A in (7.1), this solution must also solve the ODE (7.63). The uniqueness of the solution of (7.63) thus implies that it coincides with the solution of (7.1) and hence satisfies the constraint G = 0. In order to see that (7.63) is indeed an ODE on M with a unique solution satisfying the initial condition (7.64) we define
so that H : R1 x TM -+ R" is a bundle morphism. Let 4>: Vm -> M (m = dimM) be a local parametrization of M near x 0, and £o 6 V m , £o 6 Rm be such that >(£o) — ^o>
CHAPTER 7.
DAE FORMULATION ON MANIFOLDS
95
Z)0(Co)Co = x0. If x(t) is a curve on M with x(to) = XQ, x(to) = XQ, then £(£) := 0"1 (#(£)) is a curve in Vm with £(£o) = Co, C(^o) = Co- Furthermore, we have x ( t ) = D0(C(£))C(t) and x(t) = Zty(£(t))£(t) + D 2 0(C(£))(C(£)) 2 . Since £><£(£(*)) maps onto Ta;(t)Ai, the tangential component £T(£) is given by
where, for every x e A4, H(x) is the orthogonal projection onto TXM and depends smoothly upon x £ M. Thus, with the notation of (7.65), the equation (7.63) becomes
which, together with the initial condition £(to) = £o> £(£o) = Co, is a second-order ODE in RTO with a unique solution. Except perhaps for the full-rank condition (7.52), all the hypotheses listed at the beginning of this section that ensure the reducibility of the DAE (7.1) to the ODE (7.63) will hold when (7.1) accounts for the constrained motion of a rigid body or a system of rigid bodies and/or mass points. This can be readily verified from the explicit form of the corresponding operator A(x) (see sections 5.2 and 5.3, respectively). Naturally, when the constraint function G satisfies the constant rank condition
where V is an open neighborhood of E(G) in R1 x TM (with E(G) denned by (7.13)), and if where G is given by (7.15), then, as we already explained in section 7.3, the equations of constrained motion are given by a DAE of the form (7.1) with G replaced by G. Due to (7.66) this DAE reduces to a second-order ODE on A4 of the form (7.63). Consistent with Theorem 7.6, the relevant initial conditions for this ODE must satisfy the requirement (to,XQ,xo) 6 S(G); that is, by (7.16), G(t 0 ,x 0 ,x 0 ) = 0 and G(t0,xo,x0) = 0. Another alternate formulation is possible when, on some submanifold M. of R n , the DAE (7.1) has the form
and G : M. —> Rfc satisfies the full-rank condition
Naturally, (7.67) is obtained when in (7.1) G is replaced by G(t,x,x) = G(x,x) :— DG(x)x. The condition (7.68) ensures that the set C := G-1(0) is a submanifold of M.
96
PATRICK 3. RABIER AND WERNER C. RHEINBOLDT
of dimension m — k, m — dim M, usually called the constraint manifold of the problem.. The corresponding set S(G) in (7.13) is simply
As we saw, in order for (7.67) to account for the holonomic constraint G(x) = 0, the initial condition (£o,xo,xo) must be chosen in S(G) and hence such that G(XQ) — 0 and DG(x0)i0 = 0. In that case, the DAE (7.67) is simply
If x(f) is a curve in C, then the acceleration splits as x ( t ) = xr(t) +XAT(£), where XT and Xjv are the tangential and normal components of x. Although this notation is the same as the one used earlier, we emphasize that now "tangential" and "normal" are understood relative to the manifold C and not just to M. Since xw(t) — V£,t-,(x(t),x(t)), where V£ denotes the second fundamental tensor of C at x € C, we may rewrite (7.70) as
Since DG(x) maps into Rfe and TXC = ker DG(x) for every x £ C, observe that the operator DG(x)T e £(Rk,TxM) actually maps into (T^C)1- n TXM. Therefore, if Hc(x) € C(TXM) denotes, for any x e C, the orthogonal projection onto TXC, then multiplication of the first equation (7.71) by Uc(x) yields
Since XT € Tx(t)C, the left side of (7.72) coincides with Uc(x)A(x)Iic(x)xT. Hence, whenever the operator Hc(x)A(x)Hc(x) € L(T XC) is invertible, then (7.72) can be transformed into a second-order ODE on C. This invertibility property will hold for every x e C and every submanifold C of M if, as before, the operator A(x) 6 £s(Hn))ssatisfies rgej4(x) C TXM. and { A ( x ) z , z ) > 0 for every z € TXM, z ^ 0, and every x € M. As pointed out earlier, these conditions hold for problems arising from constrained rigid body motion (and also systems that possibly incorporate mass points). In summary, under the stated assumptions, the DAE (7.67) reduces to the secondorder ODE on C:
Note that the condition x e C need not be incorporated into (7.73) because it is redundant as soon as (7.73) is used with initial conditions x(
Chapter 8
Computational Methods Currently there appear to be no reliable general-purpose algorithms available on which production-level software may be based for solving the DAEs of the form (6.1) or, more generally. (7.1), arising as models of problems involving kinematic or mixed kinematic and geometric constraints. In this chapter we show that a class of algebraically explicit DAEs considered by Rheinboldt [23], based on local parametrizations, includes a method that can be adapted to the construction of an algorithm for solving the DAEs covered by Theorem 6.1. This algorithm is meant to show that it is indeed possible to construct an effective method for solving this class of DAEs. In section 8.3 we indicate some possible further directions of work that may lead to the development of other methods for the DAEs discussed in the previous chapters.
8.1
Computations on Manifolds
The DAE methods developed in [23] were based on a package of supporting algorithms, called MANPACK, for computing local parametrizations on submanifolds of R™ that are implicitly defined by local submersions (see Theorem A.I). This section provides a brief overview of some of the relevant MANPACK algorithms, as given in [22]. The presentation is independent of the earlier results. It should be noted that the computational tasks associated with implicitly defined manifolds differ considerably from those arising in connection with manifolds defined in explicit, parametric form as in, e.g., computational graphics. In fact, unlike in the latter case, for implicitly defined manifolds the algorithms for determining local parametrizations and their derivatives still need to be made available. Throughout this section it is assumed that F : R" -» R K , 0 < K < n, is a smooth mapping and a submersion on its nonempty zero set
Thus £ is a d — (n - K)-dimensional submanifold of Rn. Definition A.2 recalls the concept of a local parametrization of such submanifolds. The following result provides 97
98
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
a basis for the computational evaluation of such a local parametrization. We henceforth denote by the canonical injection that maps Rd isomorphically to {0} x Rd. Theorem 8.1. Under the stated assumption about F, let U e £(R d ,R") be an isomorphism from R6* onto a d-dimensional linear subspace T C R™. Then the mapping
is a local diffeomorphism
on an open neighborhood of xc € £ in R" if and only if
If (8.3) holds at xc, then there exists an open set Ud ofRd such that the pair (Ud,$), defined with the mapping (j) = H~1 o j : Ud —> R", is a local parametrization ofS near xc. Proof. If DH(xc)h = 0 for some h e R n , then DT(xc)h = 0 and U^h = 0; whence h £ TxcS and, because
also h € T x . By (8.3) this implies that h = 0. Conversely, for any nonzero h € TxcSnT-1 we have DT(xc)h = 0, and (8.4) requires that UTh = 0, which together imply that DH(xc)h = 0. Hence, if (8.3) holds, then there is some open neighborhood Un of xc in R" such that H is a diffeomorphism from Un onto the open set H(Un] in R™. Evidently, the set # ( £ n W n ) = H(Un) n ({0} x R d ) is open in {0} x Rd and Ud = j-lH(Y.r\Un) is an open subset of Rd. This shows that <j> = H~l o j maps Ud onto the open subset £ nUn of £. Since H maps £ r\Un into {0} x R d , the inverse j~l o H^nu» of 4> is well defined and both
CHAPTER 8.
COMPUTATIONAL METHODS
99
By definition of 0 we have
Thus the evaluation of x = 0(y) for given y 6 Ud requires finding zeros of the nonlinear mapping
For this a chord Newton method
works well in practice. Here we used the fact that, because (8.3) is assumed to hold at x° e S, the derivative DH(x°} is nonsingular. By standard results, it follows that there exists a 6 > 0 such that for ||y||2 < 6/2 the process (8.8) converges to a unique zero x* of Hy in some neighborhood of x° and thus gives x* = (f>(y). For the special choice x° = xc + Ucy, ||y||2 < <5/2, the iterates satisfy 0 = C/J (xk — x °] ~ V — Uc (%k — ^o)) which implies that the process can be applied in the form
with a y-dependence only at the starting point. This shows that, for any y near the origin of R d , the algorithm of Table 8.1 produces the point x = >(y) in the local parametrization (Ud,
100
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
computation of the nullspace of Z. A simple one is based on the LQ-factorization (with row pivoting)
Here P is a K x K permutation matrix; L is a K x re, nonsingular, lower triangular matrix; and Q is an n x n orthogonal matrix partitioned such that Qi and Qi are n x K and n x d matrices, respectively. Then, clearly, the d columns of Q^ form the desired orthonormal basis of T. This justifies the algorithm in Table 8.2. Table 8.2: Algorithm COBAS. Input: {Z} compute row-pivoted LQ-factorization (8.10) of Z; for j = l , 2 , . . . , d do V? :=Q2ei; return: Obviously, when the tangential coordinate system is used at zc, then COBAS can be applied with the Jacobian matrix DT(xc) as the matrix Z. In that case, GPHI simplifies considerably if the LQ-factorization (8.10) of DT(xc) is also applied for solving the corrector equation DH(x°)w = q. Then, for each step, only a K x K, instead of an n x n, lower-triangular system has to be solved. For details we refer to [22]. In general, the matrix Z used in Table 8.2 is constructed from the Jacobian matrix DT(xc). For example, it is frequently important to work with coordinate subspaces T that contain a specific canonical basis vector, say, e", of R™. Often the reason for this is that the independent variable xv represented by this vector is of a special nature, as, for instance, in our setting here, when xu corresponds to the time variable in a nonautoiiomous DAE. Evidently, in order to ensure that ev 6 T we have to replace the I'th column of Dr(xc) by a zero column. This leads to the algorithm in Table 8.3. Table 8.3: Algorithm GNBAS. Input: form Z by zeroing column v of DT(x°); by COBAS compute orthonormal basis Uc of ker W; Output: Of course, we require that the constructed matrix Z still has maximal rank K — dim T^-. In order to verify this during the computation of the basis of the nullspace, we may replace the LQ-factorization of Z in COBAS by the singular value decomposition (SVD) and then apply scaling and standard rank tests. The details should be evident. Let (W d ,0) be a local parametrization of S near x° e S. Then by Theorem A.7 the pair (U x R d , ($>, D0)) is a local parametrization of the tangent bundle T£ near (x0,v) for any v € TXoT,. For the evaluation of this parametrization of T£ we need now an algorithm for the computation of the derivative D<j> of 0.
CHAPTER 8.
COMPUTATIONAL METHODS
101
Let T be a coordinate subspace of E at x° and, as before, suppose that on Rn and R the canonical bases are chosen and that the columns of the n x d matrix Uc form an orthonormal basis of the given coordinate subspace T at x°. Then by differentiation of (8.6), it follows that d
Since the Jacobian of H is nonsingular at x c , this shows that at any the derivative D(j)(y) can be computed as follows in Table 8.4. Table 8.4: Algorithm DGPHI. Input: compute LU-factorization of A : — for j — 1 , . . . , d do: Output:
solve Azj = ed+j;
By further differentiation of (8.11) we obtain
which, because of the nonsingularity of D#(0(y)), defines D24>(y)(v1,v2) uniquely. The algorithm of Table 8.5 for computing D2(j> assumes that the vector w — D2r(x)(u1,u2) has already been computed for given x = 0(y) and ul = D4>(y)vl, i = 1, 2. Table 8.5: Algorithm D2GPHI. Input:
compute the LU-factorization of A := solve Az Output:
8.2
A DAE Solver for Kinematic Constraints
In this section we utilize the algorithms of section 8.1 in the development of the mentioned algorithm for solving DAEs of the form (6.1); that is,
102
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
under the conditions of Theorem 6.1 and with N = In + 1. Thus, in particular, by (6.7(iii)) the submersion theorem, Theorem A.I, ensures that E := r -1 (0) is a submanifold of Rn of dimension d = N — K. The proof of Theorem 6.1 showed that the DAE can be locally reduced to an equivalent ODE. This suggests that by applying a standard ODE solver to a suitable sequence of these local ODEs it should be feasible to compute a solution of the original DA However, while the proof of Theorem 6.1 is constructive, it Involves the use of certain projections that are not readily computable. In fact, it is easily seen that these constructions require, in essence, calculations with certain local parametrizations of the manifold E defined by the constraint function. Instead of attempting to translate the earlier proof into a numerical procedure, we sketch here a different proof of Theorem 6.1 that exhibits directly the steps of the desired computational algorithm and that, in particular, reflects the use of the parametrization algorithms discussed in section 8.1. For simplification we introduce the notation
where as before x and later also x are not derivatives but denote vectors in R n . Then (8.13) can be written as the first-order DAE
For any solution (v(t),X(t)) of (8.15) we noted repeatedly that the curve v = v(t) also satisfies the differentiated constraint DT(v(i))v(t) = 0 and therefore that the solution must remain in the set
Evidently, for any (vc,w°, A c ) 6 II we have
as well as A(xc)xc = F(vc) + £>iI>c)TAc. For given (vc,wc, A c ) e Ef let (Ud,(f)) be a local parametrization of E near vc. Then, because of (v c , wc) E TS, Theorem A.7 asserts that (Ud x Rd, (<^>, D
CHAPTER 8.
COMPUTATIONAL METHODS
103
For any fixed y e Ud consider the ./V-dimensional linear system
The operator K(0) is invertible. In fact, by (6.7(i)), K(0)h = 0 for some (hl,h2) e Rd x RK implies that D(j)(Q)hl = 0 and h2 — 0; whence h1 = 0 by the injectivity o D$(Q), and therefore h = 0. Thus, after shrinking Ud if needed, the operator K(y) of (8.17) will be nonsingular for any y € Ud and hence the mapping
is well defined. Consider now the local initial value problem
where the prime denotes the derivative operator d/ds and the initial condition corresponds to our assumption that 0(0) = vc. By standard ODE theory, (8.19) has a unique solution y : J >—> Ud on an open interval J containing the origin. By definition of <j> and the construction (8.17), (8.18) of * we obtain
Now set v(s) —
Comparing this with the notation (8.14) we see that the first component equation (corresponding to the original variable t) has the form t'(s) = 1. The initial condition j/(0) — 0 together with 0(0) = vc gives -u(O) — v°; whence, in particular, 4(0) = tc. Thus the first equation of (8.21) requires that t(s) = tc + s and hence that the operators d/dt and d/ds are the same. But then (8.21) is exactly the DAE (8.15). Moreover, it follows readily that, by our choice of (vc, wc, A c ) G II, this solution must satisfy the initial conditions
In other words, we proved that for any (vc, wc, A c ) 6 II there exists a local solution of the DAE (8.15) satisfying (8.22). Conversely, let t 6 J i-> (v(t},X(t)), 0 6 J be any solution of (8.15). Then as noted, we necessarily have (v(t),i>(t),X(t)) € II for t e J, and, hence, the solution can satisfy the initial condition (8.22) only if (vc,wc, A c ) e II. For any t0 € J write v° = v(t0), w° = w(ta), A° = A(t 0 ), and let (Ud,(j>) be a local parametrization of S near v°. Then the mapping ^ of (8.18) is well defined on Ud
104
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
(possibly after shrinking this neighborhood). For the local curve y : J i—» Ud, y = <j> we have v(t) =
1
ov,
Since the matrix on the left equals K ( y ( t ) ) and hence is nonsingular for y(t) 6 Ud, this implies that ^>(y) = (y(t),X(t)) and, hence, that the local curve is a solution of the initial value problem (8.19). Thus the DAE initial value problem (8.15), (8.22) with (vc,wc, Xc) € n is locally equivalent to the initial value problem (8.19). This shows that we can solve the local ODE to obtain the solution of the (global) DAE. Of course, for this we have to work, in general, with several, appropriately constructed, local parametrizations of E to obtain the solution of the DAE on a prescribed interval. Any standard ODE solver requires an algorithm for evaluating, at any y e Ud, the right side of (8.19), that is, the derivative vector y. This is accomplished here by the algorithm of Table 8.6. It assumes that the current local parametrization of S near xc is available and uses as input, besides the local vector y, the corresponding global point v = 4>(y). Note that as byproducts we also obtain the derivative v and the algebraic variable A. Table 8.6: Algorithm DYNH. Input:
use GPHI with y, xc, Uc, DH(xc) to get v = (j>(y); evaluate use DGPHI to compute D4>(y)', evaluate solve the linear system Output:
For the overall solution algorithm recall that only the initial conditions x(t0) = x0, x(to) = XQ satisfying the consistency condition T(to,xo,±o) = 0 are required and that consistent conditions for x(to] and A(io) will be determined by the process. With this, our solution algorithm DAENH assumes the form given in Table 8.7 below. It retains, where possible, a computed local parametrization over several steps of the ODE solver. The local basis at v G S is constructed by means of the algorithm GNBAS of Table 8.3 with index v — 1. Hence, because v contains, by (8.14), the time t as first component, GNBAS preserves the t variable. This explicit availability of t allows for a test of whether a prescribed terminal time has been reached. In Table 8.7 we simply work with a generic, user-supplied termination test, called t-test, which is assumed to return t-test := 'true' when the process is to be terminated. For the solution of the local ODEs (8.19) we use a user-prescribed ODE solver denoted by @[DYNH], which calls on the algorithm DYNH of Table 8.6 for the evaluation of the right side of the local ODE. More specifically, 6[DYNH](y, h) e Rd is the "next" point generated by this solver for the current point
CHAPTER 8.
COMPUTATIONAL METHODS
105
Table 8.7: Algorithm DAENH. Input: {t,x,p,h,km&x}; mode = init, k = 0; while mode = init then vc := (t,x,p);
evaluate use GNBAS with v — 1 to construct basis Uc at v°; set up DH(xc) y = (0,0); mode = step; while mode = step then use DYNH with y, vc, Uc, DH(xc) to get y, v, A; if DYNH failed then mode = init; break; Output: if t-test = 'true' then return success; if k > /cmax then return failure; take local ODE step: y = " [DYNH](y, h); if step unacceptable then mode = init; break; fc:=fc+l; endwhile; endwhile. y e Ud and stepsize h. The method is expected to generate a local error estimate for testing whether the step from y to 0[DYNH](y, h) is accepted or has to be repeated with a smaller h. A certain number of such step reductions causes the mode indicator to be set to init. In addition, we also set mode = init when the iteration methods in DYNH or GPHI converge poorly or not at all or when in GPHI, in the case of acceptable convergence, the distance between the computed point and the starting point of the process becomes too large. The algorithm DAENH represents an adaptation of the algorithms DAEN2 and DAEQ2 of [23]. As the latter two algorithms, DAENH has been implemented effectively with various local ODE solvers 0[DYNH] of different order. In particular, for the numerical experiments discussed in Chapter 9 we used the explicit Dormand-Prince Runge-Kutta DOPRI 5(4). In other work we also used several stiff methods such as RADAU and SDIRK.
8.3
Outlook on Computational Approaches
As noted earlier, much work is still needed in the computational analysis of the DAEs arising as models of rigid multibody problems with general kinematic or mixed kinematic and geometric constraints. Standard DAE software, such as the widely used code DASSL (see Brenan, Campbell,
106
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
and Petzold [5]) are designed for index-one DAEs and are certainly not useable for the production solution of the DAEs (6.1), which, typically, have index two,13 not to mention (7.1). There exists a large amount of literature on the study of computational algorithms for holonomic problems and we cite here only the books by Brenan, Campbell, and Petzold [5] and Hairer and Wanner [13], where other references can be found. Much of this literature concentrates strictly on the computational solution of the Euler-Lagrange equations (of index three)
In other words, except possibly for illustrative examples, there is little connection with the underlying mechanical problems. While this may be reasonably acceptable for most planar problems, in the case of systems of three-dimensional rigid bodies our discussion in the previous chapters shows that the details of the modeling aspects have to be considered carefully. In other words one cannot presume simply that such questions as, for instance, the handling of the condition p 6 S3, have been resolved already. Broadly speaking, the methods for solving (8.23) may be subdivided into two classes. The methods of the first class are characterized by their use of some local parametrization of the constraint manifold r -1 (0). Wehage and Haug [32] appear to have been the first to propose the use of certain specific parametrizations for the computational solution of (8.23), and, since then, this has been taken up by several authors. The extension of the approach to certain general classes of algebraically explicit DAEs by Rheinboldt [23] provided the basis of our algorithm in section 8.2 for the DAEs (6.1). The methods of the second class work with (8.23) or the DAEs obtained from that system by replacing the geometric constraint F(x) = 0 by its induced kinematic constraint DT(x)x = 0 or, alternately, by the acceleration constraint
As we noted before, the resulting DAEs are mathematically equivalent provided consistent initial conditions are used. But, in general, this equivalence is not preserved by the computational solution procedures. For the systems involving the induced kinematic (respectively, the acceleration) constraint, this exhibits itself in a drift of the computed approximations away from the geometric constraint manifold F"1(0) (respectively, from the tangent bundle Tr-1(0)). The computational methods of the first class work, in essence, with these modified DAEs but introduce various techniques to ameliorate or suppress the drift phenomenon. For details we refer again to the cited books and the original literature. It appears to be entirely feasible to apply analogous approaches also for the solution of DAEs of the general form (6.1). In contrast, it appears to be highly unlikely that the techniques underlying the methods of the second class for (8.23) can be applied to the DAE on a manifold M. of the 13 We shall not enter into any discussion of the index definition since it plays no further role in this presentation.
CHAPTER 8.
COMPUTATIONAL METHODS
107
form (7.1). Under the conditions of Theorem 7.1, it is clearly feasible to reduce this DAE locally to the form (6.1) via local parametrizations of the manifold A4. Moreover, since in practice, this manifold is usually a product of several copies of R3 x 513, such loca parametrizations can be computed relatively easily and without recourse to the MANPACK algorithms. The situation becomes more complex in the case of mixed kinematic and geometric constraints when, a priori, the full-rank condition of Theorem 7.1 fails to hold. Yet, again, as shown in section 7.4 a local parametrization approach still remains feasible, at least, in theory. However, in all these cases it is to be expected that any efficient software implementation has to be based directly on the details of the mechanical formulation rather than on a general DAE in the form (7.1). In other words, the software has to incorporate much of the mechanical modeling process. An example for this is the DADS system of Haug [14] for multibody systems subjected to geometric constraints. On the basis of a detailed description of the parts of the mechanical problem and their interconnections, DADS constructs the DAE corresponding to (5.58) in terms of the variables u and p. For details of the solution algorithms used in the system we refer to [14]. In particular, in [14] procedures are also indicated that apply a solver for the computation of approximations in (u,p) but allow work during each step in terms of the variables u and u; and their derivatives. The approach is similar to that discussed in the first half of section 7.6, and hence it appears to be feasible to adapt these algorithms also to systems of the forms (5.51) and (5.58).
This page intentionally left blank
Chapter 9
Computational Examples This chapter presents several computational examples of systems with kinematic or mixed kinematic and geometric constraints. The main intent is to show that the algorithm DAENH of section 8.2 provides an accurate and effective tool for solving such problems. Of course, examples with pure geometric constraints are not considered since they are widely available in the literature and, as we noted before, this case is not of central concern in this monograph. For ease of presentation only fairly small model problems were chosen. By nature this resulted in a selection of problems which are either planar or for which simple models in terms of special variables, such as certain angles, etc., are readily available. Since, in the latter case, the forces involved in these problems are generally conservative, it appeared to us artificial, if not pedantic, to rewrite each of them in terms of the ( u , p ) variables. But for several of these three-dimensional problems it is shown that there are some natural techniques that can often be applied to arrive at a model in terms of the (u,p) variables. Generally, of course, for larger and more complex problems than those chosen here it is rarely possible to use special variables, and then a consistent use of the (u,p) formulation becomes a necessity. However, as noted before and discussed in [14], for the computation it then becomes essential to develop the models directly from a description of the mechanical problem, that is, from a characterization of the various bodies and their interconnections. The development of a software system for that purpose is certainly outside the framework of this monograph. In all examples, the canonical basis {el, e2,e3} of R3 is chosen as the reference frame.
9.1
Some Introductory Planar Examples
In the context of the theory presented here, planar problems are of little interest since they are essentially of the same type as the classical mass-point problems discussed in section 2.1. However, from a computational viewpoint these planar problems are certainly not at all trivial and, in fact, they can and do exhibit most of the difficulties encountered in the 109
110
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
three-dimensional cases. Accordingly, we present in this section three planar examples that provide an initial illustration of the behavior of our computational method. In particular, the first two choices represent special cases of a simple single body problem for which exact solutions are known that allow for a direct comparison with our computed results. The third example involves a practically interesting multibody problem, which we present here in a simplified, planar version. While it would not be difficult to expand this example into a more realistic three-dimensional model, it appeared to us that the added complexity hardly justified the small gains in numerical insight. The first two examples concern the classical model problem of the motion of a rigid body that "skates" on a tilted plane. This supporting plane is chosen to be span {e1, e2} and the direction of the gravitational force is — sin /Je1 — cos /3e3. In other words, we consider a plane that is tilted around the e2-axis by a fixed angle 0 from the horizontal. At any time, the (single) sharp-edged skate is assumed to point in a specific direction and to be in contact with the plane at exactly one point. The problem is planar in the sense of section 3.3 if we assume that the point of contact is also the body's center of gravity. Let (ui, 112)T be the point of contact and
As shown in Neimark and Fufaev [17], it is readily verified that the Lagrangian has the form where m is the mass and J is the moment of inertia of the body with respect to e3. Thus as discussed in sections 2.3 and 4.1 it is here possible to work directly with the variables HI, u^, and (p. The problem becomes dimensionless with the substitutions
where r is a reference length. Then we obtain, in line with (4.42), the DAE
For our first example we assume that the supporting plane is horizontal, that is, that /? = 0. Then it is easily verified that under any initial conditions
that are consistent with the constraint (9.1), the solution of (9.3) with /3 = 0 and under
CHAPTER 9.
COMPUTATIONAL EXAMPLES
111
the assumption
For the computations we used u° = 0, (p° — 0, ii° = (—4,0) T , <^° = 2; whence a — 2 and the solution is
In this case the contact point of the skate describes a circle centered at (0, — 2) T with radius two. The computations with the algorithm DAENH of Table 8.7 were performed in double precision using the floating-point processor of a Pentium-Pro chip. A relative tolerance of 10~8 was chosen for the ODE solver. The maximum absolute error of the seven components of the computed solution after 10 revolutions was 0.4 x 10~7. For the second example we consider (9.3) with /? ^ 0 and under the assumption that at t = 0 there is only an angular velocity. In other words we have the initial conditions
Then it is easily verified that (9.3) has the exact solution
For the computations we used a — 1 and f3 = arcsin 1. Then as shown in Figure 9.1 the point of contact traces a cycloid in the {e1,e2}-plane with its cusps at the points (0, fc-7r/2)T, k = 0,1, 2 , . . . . At these cusps the velocity becomes zero, the skate comes to a stop, and the motion reverses itself, that is, changes the sign of its direction. The accuracy of the computed solution is of the same order of magnitude as in the first case. More specifically, again with the relative tolerance of 10~8 for the ODE solver, the maximum absolute error of the seven components of u ( t ) , u(t), X(t) equaled about 0.2 x 10~7 over about 10 loops of the cycloid. Thus in both examples the performance of DAENH is certainly highly satisfactory. As the third example we consider a simplified model of a commercially available transmission (see [15]), which translates the constant rotational motion of a smooth driveshaft into a motion along the axis of the shaft. The effect is similar to that of a worm drive, i.e., of a nut moving along a screw-threaded shaft, but here, in effect, the thread can be continuously varied. The transmission consists of a box containing an assembly of three coupled rings placed around the shaft. The inner surface of these rings is prepared to ensure good frictional contact with the shaft. The rings are arranged such that their symmetry axes form an angle ±
112
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.1: Path of skate on inclined (iti,U2)-plane. of a lever,
CHAPTER 9.
COMPUTATIONAL EXAMPLES
113
Now we attach the aforementioned three bodies. In practice, the transmission is in a box to which the other equipment is then connected. Irrespective of the specific physical arrangement, we simply assume that the shock absorber of mass ms has its center of gravity at (us, 0)T and is connected to the ring body by means of an elastic spring with spring constant 7 and a damper with damping constant 6.
Figure 9.2: Schematic of the roll-ring model. In Figure 9.2 the box containing the transmission ring is not shown and the spring and damper simply connect the shock absorber with a vertical line through the center of gravity of the ring body. From the center of gravity of the shock absorber hangs a planar double pendulum consisting of two rigid bars. The pivot pin of the first bar is at the center of gravity of the shock absorber, and the second bar swings around a pivot at the other end of the first bar. The bars are characterized by their lengths i\, 1% and masses m\, m^- Then their moments of inertia are
The arrangement of the four bodies is schematically shown in Figure 9.2. Since the model is planar, we can work with the configuration vector
At a given time, let v,w € R2 be the location of the centers of gravity of the two bars, and a\ and a? the angles formed by the axes of the bars with the direction —e2.
114
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Then the two pivots are modeled by the holonomic constraints
For the complete model this has to be combined with the nonholonomic constraint (9.8); in other words the complete model involves mixed constraints. In accordance with the analysis of section 6.2 we introduce the kinematic constraint
It is now straightforward to see that the equations of motion are of the form
where A — diag (mr,Jr,ms,mi,mi,Ji,m2,m2,J2) and and N is the external moment applied to the ring body when its axis is tilted. For the computation we use the data
and construct a moment function N(t) for driving the transmission as follows: In the piecewise constant function
each one of the eight discontinuities at to = 0.5,1.0,1.5,2.0,5.5,6.0,6.5,7.0 is replaced by a cubic spline that is nonzero for \t — t0\ < 0.1, and then the moment is computed as N(t) = av(t) with a scale factor a = 1.0 x 10~3. Now for the consistent initial conditions
CHAPTER 9.
COMPUTATIONAL EXAMPLES
115
Figure 9.3: Dynamic, behavior of roll-ring model. the computed behavior of the variables
9.2
Three-Dimensional Problems
In this section we turn now to problems in three dimensions, which require the full theory of the previous chapters. As a first example we consider a homogenous ball rolling without sliding on a horizontal platter that rotates with a constant angular velocity around a normal direction. The supporting platter is supposed to be contained in the plane span {e1, e 2 } with e3
116
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
as its axis of rotation. Let Q be the (constant) angular velocity of rotation of the platter and assume that the ball has radius r. We derive the equations of motion in terms of the variables (u,p), that is, in the form of the system (5.26). In the reference frame, let u = (ui, U2j u 3) T be the ball's center of gravity. Then the condition of contact is uz = r and, if it holds, x — (ui, u^, 0)T is the point of contact. Let u be the angular velocity of the ball in the body-fixed frame; then by (4.4) the velocity of x, in the reference frame, is given by The no-slip condition requires that, in a frame fixed in the platter, the velocity of the contact point in the plane of the platter must be zero. Since any given point y = (yi,y2;0) T of the platter has the velocityfi(—2/2,2/1, 0)T in the reference frame, it follows from (9-14) that
By (4.21) and (4.30) we obtain (dropping the t-dependence)
where we used the fact that, because of pTp = 0, the projection H(p)TH(p) onto {p}~L maps p into itself. Now recall that the matrix representation of K(p) (in the canonical bases) is given by the last three rows of (4.29). Thus the no-slip constraints (9.15) can be written as
For the equations of motion (5.26) we have in this case Fr = 0, N = 0, M — mis, and J = J(,/3, where m and J;, are the mass and moment of inertia of the ball. With the obvious extensions for Fr, N, and G we then obtain
I
The term H(p)T JH(p)p on the right of the second equation of (5.26) is here proportional to p and was simply subsumed in the term p,p. The problem becomes dimensionless with the following substitutions:
where, of course, r is the radius of the ball and T a reference time. For the computations we used fl = 2, J = 0.6 and the consistent initial conditions
CHAPTER 9,
COMPUTATIONAL EXAMPLES
117
The initial spin ensures that the trajectory of the ball is an ellipse with focal point at the origin. The computation proceeded through slightly more than four orbits and the resulting (overlapping) trajectory in the (e1, e2)-plane is shown in Figure 9.4. Of course, under different initial conditions the ball can fly out along a hyperbolic arc.
Figure 9.4: Trajectory of ball on rotating (m, U2)-plane. There are numerous other rolling-ball examples in the literature. Among these we select here a ball of radius r rolling, without sliding, on the inside of a cone, which in the reference frame is defined by
The opening angle 7, 0 < 7 < ^TT, of the cone is fixed. Here it is natural to introduce the polar coordinates (p, a) so that any point on the cone is specified by
When the Eulerian angles (p, ip, 0 are used to characterize the rotation of the ball, the configuration vector becomes y = (p, a, ,#) T and, as shown by Neimark and Fufaev [17], the problem has the Lagrangian
At the point of contact consider a tangential coordinate system that is transformed into
118
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
the reference frame by the rotation
Let (jj denote the angular velocity vector of the ball in this tangential frame. Then the conditions of rolling without sliding have the form
By using the well-known relations between w and the derivatives of the Eulerian angles (see, e.g., again [17]) and transforming the results into the reference frame by means of (9.22), the constraints (9.23) become
The formulation becomes dimensionless by the substitutions
Then, in accordance with the discussion of section 4.5, it follows from (9.21) and (9.24) that the equations of motion have the form
with
and
Let 7 = 1; then the initial conditions
CHAPTER 9.
COMPUTATIONAL EXAMPLES
119
are consistent and provide the ball with an initial spin. The computational results in Figure 9.5 show that the resulting trajectory is similar to that of the skate in Figure 9.1. The spin causes the ball to roll up briefly until the upward component of the velocity becomes zero, and the direction reverses itself. The ball then rolls downward in an arc until the position of the spin axis causes the trajectory to move upward again and the cycle repeats itself.
Figure 9.5: Path of ball inside cone in (a, p)-coordinates. We used in this example the special variables p, a, (p, •(/>, 9. In analogy to the previous example it is fairly straightforward to model the problem in the (u, p) variables, as shown below. Let u be the center of gravity of the ball in the reference frame and u = (u 1 ,u 2 ,0) T be its projection onto the {e1,e2}-plane. In analogy to (9.22) we introduce the vectors
forming an orthonormal basis of R3. Since u T 6 2 = 0, we have u = (u T 6 1 )6 1 + (w T 6 3 )6 3 , where x — (u T fe 3 )6 3 lies on the cone and u — x = (UTbl)bl is a normal vector. Thus the ball is in contact with the cone exactly if
When this condition holds then
120
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.6: Three rolling cylinders. is the point of contact and, as in the previous example, we can use (4.4) and (9.16) to obtain the velocity of the point of contact in the reference frame as since x — u = —rbl by (9.26). The no-slip condition now requires that the component of x(t] parallel to the tangent plane span {b2, b3} at x must be zero, that is, omitting the t-dependence,
With F® = (0,0, — 77M7COS7)T and N = 0, the equations of motion (5.26) are now fully determined. Note that in this formulation we have, besides the two no-slip conditions (9.29), also the contact condition (9.26) as an explicit constraint. In the previous form (9.25) the contact was implicit in the choice of the variables. As a third three-dimensional example we consider a system of three cylinders as shown in Figure 9.6. The two lower cylinders have identical radius and roll on a horizontal plane, while the third one rolls on top of them. All cylinders are assumed to be sufficiently long to ensure that in the time interval under consideration the top cylinder does not slip off the lower ones. Let span {el,e2} again be the supporting plane and denote by r\ and r2 the radii of the lower two and the upper cylinder, respectively. Because the cylinders are rolling without slipping, the angle a between the axes of the lower cylinders has to remain fixed. Thus, as shown in Figure 9.6, it is no loss of generality to assume that the axis of one of the lower cylinders is parallel to the e1-axis. Of course, the angle 9 between el and the axis of the upper cylinder will vary with time. Let x = (xi,X2,X3) T denote the center of gravity of the upper cylinder. Then the condition of contact with the supporting plane requires that £3 — 1r\ + r^. We follow here Neimark and Fufaev [17] and use as special variables the coordinates xi, #2,
CHAPTER 9.
COMPUTATIONAL EXAMPLES
121
the angle 0, and the angles of rotation fa, fa<, and tp of the two lower and the upper cylinder, respectively. In other words the configuration is characterized by the vector
j/ = (xi,x 2,0,ip,fa,ip2)T-
Then the Lagrangian of the system is
where m and J", J% are the mass and the principal central moments of inertia, respectively, of the upper cylinder, and J1, J2 are the moments of inertia of the lower cylinders about some generator line. For the points of contact between the upper and lower cylinders we obtain
and vl = (0,2ri^i,0) T and v2 = (-2risinaV>2,2ricosm/>i,0) T , respectively, are the velocities of these two points of contact. Evidently the angular velocity of the upper cylinder in the reference frame is uj = (<£> cos $,<,:> sin #,$) T . Thus, in agreement with (4.4), and (4.3), the no-slip conditions must be
which gives the constraint
Note that there should have been six scalar no-slip conditions but two of them are trivial zero-relations. With (2.29) it is now straightforward to generate from (9.30) and (9.31) the equations of motion in the form of the DAE
For the computations we chose the constants
122
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.7: Paths of the contact points and of the center of gravity of the upper cylinder. and the consistent initial conditions
Figure 9.7 shows the computed projections onto the (x^a^-plane of the paths zl(t), z2(t) of the contact points, and of x(t). In addition, the axis-line of the upper cylinder at times t = 0.0,2.7,3.72,4.8 are shown. The upper cylinder first rolls in the direction of the positive £i-axis with decreasing axis angle 9, then at t = 2.7 it reverses itself and rolls backward until t = 3.7 where it begins again to roll in the positive Zi-direction. From t = 2.7 on, B increases monotonically. It is not difficult to model the three-cylinder problem in terms of the (u,p)-variables. Before sketching this, we briefly discuss the constraint arising from one cylinder with radius r rolling, without sliding, on a horizontal plane span {e1, e2}. For this let (u,p) € R3 x S3 denote the configuration variable of this cylinder. For simplicity we assume that the cylinder is homogenous, so that u(t) remains on the cylinder-axis at all times. Contact of the cylinder with the plane implies that ^3 = r and the condition of rolling without sliding requires the axis of the cylinder to have a constant direction. With no loss
CHAPTER 9.
COMPUTATIONAL EXAMPLES
123
of generality, suppose that the axis-direction coincides with e1 at time t = 0 and that the reference frame {e^e^e 3 } and the body-fixed frame coincide at that time. At times t the direction of the axis is given by Rplel = -Rje1; whence the constraint Rpel = e1. By (3.13) (with Xi — pi, 1 < i < 4) this is equivalent to y>2 = PJ, — 0. But these conditions do not prevent the cylinder from sliding in the axis direction. To keep this from happening we must require that the velocity of one contact point, say, x = (w 1 ,u 2 ,0) T , vanishes at all times. Then, arguing as in the example of a ball rolling on a platter (but now with ft — 0), we find that ui = 0, and u^ = 2r(pip(> - popi) — 0. (Compare with (9.17).and note that p2 = p3 = Q because of p2 = Pz — 0.) Thus altogether we obtain the five constraints
It is informative to observe that, in view of \p\2 = 1 and p% — p% = 0. we may set po = cosijj, pi — sinip, in which case the last constraint of (9.33) becomes u\ when considering the first lower cylinder (so that r = r\). The reason why here the "natural" variable is lip and not tj) is consistent with Remark 3.3. With Po = cos-)/;, pi = sin?/;, the rotation Rp is the rotation of angle 2i/> in the {e1,e2}-plane. Obviously, the first three constraints of (9.33) are geometric and hence must be differentiated. This gives the kinematic constraints
Remark 9.1. The case of a nonhomogenous cylinder is slightly more complicated. If the center of mass is not on the axis, let WQ denote the coordinate of some material point on the axis at time t — 0 in the cylinder-fixed frame. Then v(t) = u(t] + Rp(t)w0 are the coordinates of the same material point at time t in the reference frame. Thus the constraint (9.34) holds with v in place of u. Since v — u + Rpwo and Rp = K(p)H(p)T (see (4.30)) we obtain
in place of the first equation in (9.34). The other two equations are unchanged. We return now to the modelling of the three-cylinder problem in terms of the variables (u ,pl) € R3 x 53, i = 1,2,3. Here the indices 1,2,3 and the terminology "first" and "second" cylinder refers to the two lower ones, while the "third" is the upper cylinder. Let T :-- {e^e^e3} denote a reference frame such that span {el,e2} is the horizontal plane. As in Figure 9.6 suppose that el is parallel to the axis of the first cylinder. Then, as discussed above, the condition of rolling without sliding for that cylinder is accounted for by the constraint (9.34) with u = u1, p — pl. In order to apply the same result to i
124
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
the second cylinder, which, in general, is not parallel to the first one, let Q be a rotation of the reference frame around e3 such that Qre1 is in the direction of the axis of the second cylinder. The corresponding change of coordinates affects only the first equation in (9.34) but it has to be kept in mind that under this coordinate transformation Rpz and K(p2)H(p2)~r are changed to QrRp2Q and QT K(p2)H(p2)TQ,m.respectively. With this we obtain the constraints expressing the rolling conditions for the two lower cylinders in the reference frame f in the form of ten scalar equations. Suppose now that a € R3 represents the direction of the axis of the upper cylinder at time t = 0; whence Rl3a is its direction at time t. The condition that the upper cylinder rests on the lower two is accounted for by the requirement that two distinct points on its axis, say u3 and u3 + R\a, remain at the distance 1r\ + TI of the horizontal plane (assuming homogeneity). This yields two scalar geometric constraints which, of course, need to be differentiated. It remains to obtain the constraint that the upper cylinder rolls without sliding on the lower two. As before, this requires the calculation of the two contact points in the reference frame J-, which shall be denoted again by z1 and z2. Their third coordinates equal 1r\ while their first two coordinates are the points of intersection of the projected axis of the upper cylinder with the projected axes of the lower two. These points are easily calculated since the three axes have the parametric equations
The velocity of the material point in the upper cylinder that coincides with z1 at time t is (9.35) while the velocity of the corresponding material point in the first cylinder is (9.36) The no-slip condition of z1 requires equating the velocities in (9.35) and (9.36) where it should be noted that the third components of both of these equations vanish; whence we obtain two, and not three, scalar constraints. In fact, when the material point in the upper cylinder reaches z1, its third coordinate achieves a minimum and hence its velocity component in the direction e3 vanishes. In the case of (9.36), the third coordinate of the material point achieves a maximum. Naturally, similar considerations apply to z 2 , of course, with Q T R p iQ in place of Rpz. As a result the (u, p)-formulation of the problem comprises 16 scalar constraints and 18 variables. If we view p1, p2, p3 as vectors in R4, the number of variables becomes 21 and now, with the three added constraints (p4)Tp* = 0, i = 1,2,3, (see (5.51)), we have 19 constraints. Since there are no external forces or moments, the constrained system is given by the DAE (5.51) with FE = 0 and NE — 0. Given the size of the system, the full-rank condition for the constraints can only be checked numerically. For the next problem, we consider a single sharp-edged disk (or wheel) rolling on a rough plane where by "sharp-edged" we mean that the disk cannot slide sideways. This
CHAPTER 9.
COMPUTATIONAL EXAMPLES
125
is a frequently addressed problem in the literature; in fact, it is often used as the prime example for showing that a kinematic constraint need not imply any restrictions on the configurations (see again Neimark and Fufaev [17]). As before we use span {e1, e 2 } as the supporting plane and —e3 as the direction of the gravitational force. The body-fixed frame {a1, a 2 , a3}, with origin at the center of gravity of the disk, is chosen such that a1 is the rotation axis, and a 2 and a3 span the plane of the disk. A critical assumption is that the disk is never parallel to the supporting plane, that is, that the planes spanja^a 2 } and spanfe^e 2 } intersect in a unique line spanned by 6 = (e3 x al}/\\e* x a 1 ^. If this holds and the disk is in contact with the plane, then there is a unique point of contact x = (xi,X2,0) T . The position of the disk can now be characterized by x\, x^, and three angles, namely, the angle (p between e1 and b, the angle 9 between e3 and the plane of the disk, and the angle ip between a? and the radius from the center of gravity to the point of contact. In other words, the configuration is characterized by the vector y — (xi,X2,tp,d,ij))~r. Here we assume that\0\ < ir/2/to ensure that the disk is not flat. on the supporting plane. Since the disk is assumed to have a sharp edge, the velocity x of the point of contact must always be in the direction of the vector £>; that is, we have the nonholonomic constraints
where r is the radius of the disk. The center of gravity is at the point
With this we see that the system has the Lagrangian (see [17, p. 102])
Here m denotes the mass of the disk and Ji, J% its principal moment of inertia, respectively. Once again, for the computation it is useful to substitute the dimensionless quantities
Since the system is conservative, we obtain then the equations of motion in the form
126
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
where a straightforward calculation shows that
with
and
As in the first two examples of this section, it is easy to formulate this problem in the (u,p)-variables. Let P e £(R3) denote the orthogonal projection onto spanfe1^2}. By considering the symmetry-plane spanje3^1} of the system as shown in Figure 9.8 we observe that the condition of contact is
whence the condition of contact becomes
If the contact condition (9.39) holds, then it also follows easily from the figure that the point of contact is
Now let (j be the angular velocity of the disk in the body-fixed frame. Then, as before, in the reference frame, the velocity of the point of contact x is, by (4.4),
where, as in (9.16), RpU = 2K(p)p. The no-slip condition requires that Px(t) = 0. Hence using (9.40) we obtain the no-slip constraints
CHAPTER 9.
COMPUTATIONAL EXAMPLES
127
Figure 9.8: Symmetry plane of the disk system. Finally, again using Figure 9.8 we note that gravity induces the moment
With this, most of the information is available for obtaining the DAE (5.26) for the disk system in the (u,p)-variables. For this observe that by (4.30), in terms of (u,p), the vector a1 becomes Similarly the moment N can be expressed in the body-fixed frame; we will omit the details. Finally, note that the geometric constraint (9.39) (in terms of ( u , p ) variables) must still be differentiated. This problem shows that the case of mixed constraints in the (u,p)-formulation may arise even when a single rigid body is considered. In the study of equations (9.38), an important aspect is, of course, the stability of the solutions. It is well known, for instance, that the motion of a rolling disk can be stable only when the velocity is sufficiently large. Stability studies are outside the topic of this monograph. We only note that, typically (see again [17]), these studies address the stability of two classes of special solutions of (9.38) which are a good source of computational examples. The first of these two classes is that of the trivial rectilinear motions of the contact point with constant velocity
for which it is interesting to study the effect of small perturbations. As an example, we used Ji = 0.35, J2 = 0.5, 00 = 0.3 and introduced a small tilt 0° = 1.0 x 10~5; in other
128
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.9: Terminating disk path in (iri,X2)-plane. words we solved (9.38) with the initial conditions
Figure 9.9 shows the resulting path traced by the contact point in the supporting plane. As expected, it is initially almost a straight line, but the tilt of the disk is increasing. Due to the moment induced by gravity this tilting motion accelerates, which in turn causes the trajectory to curve away from the straight path. Shortly thereafter the tilt angle passes through 7r/2 and accordingly we terminated the calculation at that point. Note that (9.38) does not reflect the condition \Q\ < ir/2, which was essential for the derivation of the system. Yet it turns out that, as in our example, it is not rare to encounter points along a solution where \9\ = n/2 and hence where the solution becomes physically meaningless. Of course, the computational algorithm will not "see" any of these points and will continue beyond them although often with very erratic results. Thus, during the computation we have to monitor the value of the tilt angle 6 and terminate the calculation when \8\ has passed through Tr/2. The second class of special solutions arising often in stability studies is that of the circular motions
which exist under the assumption that the equation
has a solution q, 00 £ (-7r/2,7r/2). Once again we consider the effect of certain perturbations on these solutions. For this let Ji := 0.35, J2 '•— 0.5, #o : = 0.25. Then (9.43) is a quadratic equation in q that has a
CHAPTER 9.
COMPUTATIONAL EXAMPLES
129
positive solution q :— q° = 2.17080452. We solve (9.38) with various perturbed values of g, that is, with the initial data
For q — q° the resulting trajectory in the supporting plane is a circle of radius l/q° centered at the origin. It is interesting that the solution behavior is rather stable under changes of q. For q = q° + A with A = 0 , 1 0 ~ 7 , . . . , 10~2,0.08, Table 9.1 shows, besides A and q, the deviation from circularity
where the maximum is taken over all computed points during a ^-interval of about 10 orbits. The computed trajectories spiral away from the circle fairly slowly. Only when beginning with q = q° + 0.08 did we encounter a more erratic trajectory, namely, that shown in Figure 9.10. Table 9.1: Perturbation of circular disk motion.
A 0.0 1.0(-7) 1.0(-6) 1.0(-5) 1.0(-4) 1.0(-3) 1.0(-2) 0.08
q
£
2.17080452
6.660(-8)
2.1708046 2.1708055 2.1708145 2.1709045 2.1718045 2.1808045 2.2508045
5.948(-7) 7.223(-6) 7.198(-5) 7.160(-4) 6.834(-3) 5.239(-2) 3.743(-l)
As the final problem we consider the motion of a two-wheeled cart on an inclined plane in a slight generalization of a related problem given by Neimark and Fufaev [17]. As before, we use span {e^e2} as the supporting plane and (sin/3,0, -cos/3)T as the direction of the gravitational force. Both wheels have the same radius r > 0 and are assumed to have sharp edges. As sketched in Figure 9.11, the following variables will be used for the specification of the current configuration of the cart: (i) The point of intersection (xi, £2,r ) T between the symmetry axis of the cart with the axle on which the wheels are mounted, (ii) the angle 6 between e1 and the symmetry axis of the cart, and (iii) the angles of rotation (j)f and <pr of the left and right wheel, respectively (each measured from a fixed radius vector). Thus the configuration is characterized by the vector y = (xi,X2,d,
130
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.10: Perturbed circular disk motion in (x1, x2)-plane.
from (xi, X2, r) T and is to remain always at a distance r from the supporting plane. This can be ensured, for example, by introducing a support under the pole of the cart that slides freely on the supporting plane. No vertical displacements of the cart body are allowed. Let a > 0 be the half-length of the axis. Then the conditions of rolling without sliding require for the left and right wheels the following preservation conditions for the velocities:
In addition, the absence of lateral sliding demands that
Clearly, the constrained system is conservative and has the Lagrangian
Here m is the mass of the entire cart, Jc is the moment of inertia of the cart when it rotates as a whole about the axis with direction e3 passing through (xi,X2,r) T , and Jw is the axial moment of inertia of each wheel. For the computation it is useful to work with the dimensionless configuration vector
CHAPTER 9.
COMPUTATIONAL EXAMPLES
131
Figure 9.11: Two-wheeled cart. Correspondingly we introduce the dimensionless quantities
Hence, in accordance with (4.42), we obtain the equations of motion
where
For the computation we used a = 0.5, Jc = 1, Jw — 0.5, g — 0.3, and the following consistent initial conditions
The computed path of the point (xi,£2) T in the supporting plane is shown in Figure 9.12. By (9.52) the initial velocity vector (ii,£2) T has a positive first component and hence points upward. The cart begins to move in that direction, but then gravity takes over and forces it to roll backward down the slope. Since here the turning velocity Q is positive, the path turns gradually upward again until, just slightly above
132
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
Figure 9.12: Motion of two-wheeled cart in (o;i,a:2)-plane. the e2-axis, it reaches a cusp point where the motion stops and the cart begins to roll downward once more. Thus the motion of the cart is similar to that of the skating body in Figure 9.1. But since <^;(0) and (^r(0) are distinct, the two wheels will retain different angular velocities throughout the motion. This causes the downward swings to have different depths depending on whether the cart is rolling forward or backward.
Appendix
Submanifolds This appendix collects some basic definitions and results on submanifolds of finitedimensional spaces in a form that is readily applicable for computations. For proofs we refer, e.g., to Abraham, Marsden, and Ratiu [2]. Throughout this appendix, the dimension n of the ambient space is given, d is an integer with 0 < d < n, and p denotes a positive integer or oo. Recall that, when G : E —> Rm is of class Cp on an open subset E C R n , then G is called an immersion or submersion at a point z0 € E if its first derivative DG(xo) € £(R n ,R m ) is a one to one mapping or a mapping onto R m , respectively. More generally, G is an immersion or submersion on a subset S C E if it has that property at each point of 5. Note that these definitions require n < m for G to be an immersion and n > m for it to be a submersion. Clearly, if n > m and DG(xo) has maximal rank m, then G is a submersion at XQ £ E. Definition A.I. A subset M C R" is a d-dimensional Cp-submanifold o/R™ if M is nonempty and for every point XQ 6 M there, exists an open neighborhood U of XQ in Rn and a submersion G : U i-> Rn~d of class Cp such that M n U = G~1(Q). The following result is frequently used. Theorem A.I (Submersion Theorem). Suppose that for the Cp mapping G : E C R" H+ R m , n > m > 0, on the open set E, the set M = G^O) = {x & E : G(x) = 0} is not empty and G is a submersion on M. Then M. is an (n — m)-dimensional Cp-submanifold o/R". An essential property of manifolds is the concept of a local parametrization. Definition A.2. Let M be a nonempty subset o/R n . A local d-dimensional Cp parametrization of M is a pair (U,
134
PATRICK J. RABIER AND WERNER C. RHEINBOLDT
The close relationship between submanifolds and local parametrizations is provided by the following result. Theorem A.2. A nonempty subset M C R" is a d-dimensional Cp-submanifold o/R™ if and only if for every XQ € M. there exists a local d-dimensional Cp parametrization of M. near XQ. The following result shows that, when M is a d-dimensional C^-submanifold of R™, then any local Cp parametrization of M is necessarily d-dimensional and hence we may suppress for them the specification of the dimension. Theorem A.3. Let M be a d-dimensional Cp-submanifold o/R". // (14,(f>) is a local k-dimensional Cp parametrization of M, then k = d. Moreover, if(Ki,(f>i) and (^2,^2) are two local Cp parametrizations of A4, then ip^1 ° (f>i is a Cp-diffeomorphism ofV\ := ^OM^i) n <^(W2)) onto V2 := ^(^(Wi) n 0 2 (W 2 )). If (U, <j>) is a local Cp parametrization of a d-dimensional Ctp-submanifold M of R", the pair (^(W),^" 1 ) is called a local chart of M. Moreover, if (Ui,
(A.I) where (L(, 0) is any local Cp parametrization of M. near XQ. Theorem A.4. If V is an open neighborhood of XQ in R™ and G : V i-> Ttn~d is a submersion at XQ such that G~l(0) = M. n V, then TXOM — ker DG(XQ). Let / : M —> R1 be a function on a given submanifold M. of R n . Then a point XQ & M is a critical point of / if for some local parametrization (U, (/>) of M. near XQ, the function / o <j> is differentiable at £0 = ^~ I (XQ) and D(f o <^>)(£o) = 0. It is easily seen that this condition is independent of the choice of the local parametrization (U, >) of M. near XQ. Critical points include local extrema of / on M.. If E C R" is an open subset and h : E —» R1 is differentiable, then the gradient V/i(x) of h at x e E is the unique vector in Rn such that (V/i(x), v} = Dh(x)v for every v £ R n , where {-, •} denotes the Euclidian inner product on R n . With these definitions, Theorem A.4 together with the submersion theorem, Theorem A.I, leads to the frequently used Lagrange multiplier theorem.
APPENDIX.
SUBMANIFOLDS
135
Theorem A.5 (Lagrange Multiplier Theorem). Assume that for the C1 mapping G : E C Rn >—> R771, n > m > 0. on the open set E, the set M = G~l(0) is not empty and G is a submersion on M.. Let h : E C Rn >—»• R1 be a Cl functional on E and XQ 6 M. a critical point of h. Then there exists AQ G Rm such that (XQ, AQ) is a critical point of the functional h : E x Rm >—> R 1 , h(x, A) = h(x) + (G(x), A); that is, (XQ, AQ) is a solution of the system
Definition A.4. Let M be a d-dimensional Cp-submanifold o/R™. The subset TM — Uxe.MKx} x TXM]}].of R™ x R™ is the tangent bundle of M and, for every x € M, {x} x TXM is the fiber ofTM above x. Since TxRn = R" for every x € R n , we have TR™ = R" x R™. Thus, the tangent bundle of a submanifold M. of R™ appears as a subset of TRn. The following result shows that when M. is sufficiently smooth, then this subset is a submanifold of TRn. Theorem A.6. Let M be a d-dimensional Cp-submanifold o/R n and p>2. Then TM is a Id-dimensional Cp'1-submanifold o/TR" = R" x Rn = R2n. The next result addresses the construction of local parametrizations of TM. from local parametrizations of M • Theorem A.7. Let M be a d-dimensional Cp-submanifold of R" with p > 2, and x0 e M. Then for any local Cp parametrization (U,<j>) of M near x0, the pair (U x R rf , (0,D>)) is a local Cp~l parametrization ofTM near (XQ,V) and any v & TXOM. The inverses of the local parametrizations (U x Rr, ($, D<j>)) of TM obtained in Theorem A.7 are called (tangent) bundle charts.
This page intentionally left blank
References [1] ABRAHAM, R. AND MARSDEN, J. E., jamin/Cummings, Reading, MA, 1978.
Foundations of Mechanics, 2nd ed.,
Ben-
[2] ABRAHAM, R., MARSDEN, J. E., AND RATIU, T., Manifolds, Tensor Analysis, and Applications. 2nd ed., Springer-Verlag, New York, 1988. [3] ARNOLD, V. I., Mathematical Methods of Classical Mechanics. Springer-Verlag, New York, 1978. [4] BREDON, G. E., Topology and Geometry, Springer-Verlag, New York, 1993. [5] BRENAN. K. E., CAMPBELL, S. L., AND PETZOLD, L. R., Numerical Solution of InitialValue Problems in Differential-Algebraic Equations, North-Holland, New York, 1989; reprinted as Classics in Applied Mathematics 14, SIAM, Philadelphia, 1996. [6] CARATHEODORY, C., Der Schlitten, Z. Angew. Math. Mech., 13 (1933), 71-76. [7] CHOQUET-BRUHAT, Y., DE\VITT-MORETTE, C., AND DILLARD-BLEICK, M., Manifolds, and Physics, Vol. I, North-Holland, Amsterdam, 1982.
Analysis,
[8] CONDON, E. U., Kinematics. In Handbook of Physics, E. U. Condon and H. Odishaw, eds.. McGraw-Hill, New York, 1958. [9] DIRAC, P. M. A., Generalized Hamiltonian dynamics, Canad. J. Math., 2 (1950), 129-148. [10] DIRAC, P. M. A., Generalized Hamiltonian dynamics, Proc. Roy. Soc. London Ser. A, 2462 (1958), 326-332. [11] GAUSS, K. F., Uber ein neues allgemeines Grundgesetz der Mechanik, J. Reine Angew. Math. (Crelle), 4 (1829), 25-28. [12] GOLUB, G. H. AND VAN LOAN, C. F., Matrix Computations, 2nd ed., Johns Hopkins Univ. Press, Baltimore, MD, 1989. [13] HAIRER, E. AND WANNER, G., Solving Ordinary Differential Equations II: Stiff Differential-Algebraic Problems, 2nd ed., Springer-Verlag, New York, 1996.
and
[14] HAUG, E. J., Computer Aided Kinematics and Dynamics of Mechanical Systems, Vol. I, Allyn and Bacon, Boston, MA, 1989. [15] JOACHIM UHING, KG, GMBH AND Co. Uhing-Linearantrieb, Firmenprospekt 07d 03.91, 1991. [16] LEI, X., Singularities of a Sheet Metal Stretching Problem and Quasilinear Second Order Ordinary Differential Equations, Ph.D. thesis, University of Pittsburgh, Pittsburgh, PA, 1996. 137
138
REFERENCES
[17] NEIMARK, J. I. AND FUFAEV, N. A., Dynamics of Nonholonomic Systems, Transl. Math. Monogr. 33, Amer. Math. Society, Providence, RI, 1972. [18] RABIER, P. J. AND RHEINBOLDT, W. C., On a computational method for the second fundamental tensor and its application ta bifurcation problems, Numer. Math., 57 (1990), 681-694. [19] RABIER, P. J. AND RHEINBOLDT, W. C., On impasse points of quasilinear differentialalgebraic equations, J. Math. Anal. Appl., 181 (1994), 429-454. [20] RABIER, P. J. AND RHEINBOLDT, W. C., On the computation of impasse points of quasilinear differential-algebraic equations, Math. Comp., 62 (1994), 133-154. [21] RABIER, P. J. AND RHEINBOLDT, W. C., On the numerical solution of the Euler-Lagrange equations, SI AM J. Numer. Anal, 32 (1995), 318-329. [22] RHEINBOLDT, W. C., MANPACK: A set of algorithms for computations on implicitly denned manifolds, Comput. Math. Appl, 27 (1996), 15-28. [23] RHEINBOLDT, W. C., Solving algebraically explicit DAE's with the MANPACK manifold algorithms, Comput. Math. Appl, 27 (1997), 31-43. [24] ROBERSON, R. E. AND SCHWERTASSEK, R., Dynamics of Multibody Systems, SpringerVerlag, Berlin, 1988. [25] RUND, H., The Hamilton-Jacobi Theory in the Calculus of Variations, Van Nostrand, Toronto, Canada, 1966. [26] SCHMIDT, T. AND Hou, M., Rollringgetriebe. Tech. report, Regel. und Messtechnik, Bergische Univ. GH, Wuppertal, Germany, 1992. [27] SOMMERFELD, A., Mechanik, 3rd ed., Leipzig, Germany, 1949.
Akadem. Verlagsgesellschaft Geest and Portig,
[28] SPARSCHUH, S. AND HAGEDORN, P., On the Gauss principle in the numerical integration of mechanical systems. In Real-Time Integration Methods for Mechanical Systems Simulation, E. J. Haug and R. C. Deyo, eds., NATO ASI Series F, Vol. 69, Springer-Verlag, New York, 1990, 293-300. [29] SYNGB, J. L., On the geometry of dynamics, Philos. Trans. Roy. Soc. London Ser. A, 226 (1926), 31-106. [30] SYNGE, J. L., Geodesies in non-holonomic geometry, Math. Ann., 99 (1928), 738-751. [31] SYNGE, J. L., Geometrical mechanics and de Broglie waves, Cambridge Univ. Press., Cambridge, UK, 1954. [32] WEHAGE, R. A. AND HAUG, E. J., Generalized coordinate partitioning for dimension reduction in analysis of constrained dynamic sytems, J. Mech. Design, 104 (1982), 247255. [33] WHITNEY, H., The self-intersections of a smooth n-manifold in 2n-space, Ann. of Math., 45 (1944), 220-248. [34] ZEIDLER, E., Nonlinear Functional Analysis and Its Applications, Springer-Verlag, New York, 1985.
Index DeWitt-Morette, C., 94 differential principles, 9 Dillard-Bleick, M., 94 Dirac, P. M. A., 2
Abraham, R., 3, 29, 133 angular velocity, 2, 30 Arnold, V. I, 31 Bredon, G. E., 20 Brenan, K. E., 105, 106 bundle morphism, 94
embedding, 3, 20 energy norm, 48, 52 Euler angles, 20 parameters, 23 Euler-Jacobi operator, 37, 38 Euler-Lagrange equation, 13 external force, 17, 18, 31
Campbell, S. L., 105, 106 Caratheodory, C., 16 Cayley-Klein parameters, 23 center of mass, 19, 30 Choquet-Brubat, Y., 94 computational drift, 14 Condon, E. U., 23 configuration space, 3, 10, 20, 26, 29, 34, 55, 91 conservative force, 42, 43, 130 constant rank condition, 66, 70, 79-81, 91, 95 constrained motion, 49 constraining force, 15 constraint geometric, 13, 50, 53, 54, 65 holonomic, 1, 13, 15-17, 96 ideal, 15 kinematic, 2, 4, 10-12, 48, 50, 53, 56, 65 manifold, 96 mixed, 65, 114 nonholonomic, 1, 4, 10, 16 state, 10 coordinate subspace, 98-101 covering, 25, 26 critical point, 134
frame, 19 friction, 16 Fufaev, N. A., 1, 2, 16, 110, 117, 120, 125, 129 full-rank condition, 11, 13, 16, 48, 56, 65, 71, 72, 75, 91, 92, 95 Gauss principle, 3, 4, 9, 11, 14, 47 Gauss, K. P., 9 general linear group, 19 generalized Gauss principle, 4, 46, 48, 49, 53, 81 Golub, G. H., 21 gradient, 134 Hagedorn, P., 3 Hairer, E., 106 Hamilton's principle, 9, 16, 17, 40 Haug, E. J., 1, 4, 5, 23, 24, 29, 106, 107 Hou, M., 112 immersion, 133 inertia tensor, 31, 55 initial condition, 12, 14, 51, 62, 71, 75, 96, 103
DAE
existence, 62, 75, 102 formulation, 12-14, 17, 49, 52, 53, 56, 61, 95, 101 139
INDEX
140 kinetic energy, 17, 41, 43, 52 Lagrange multiplier, 12, 17, 46, 48, 51, 56, 92 multiplier theorem, 3, 12, 14, 48, 49, 134 Lagrangian, 16, 18, 43, 110, 117, 121, 125, 130 Lei, X., 72 local parametrization, 6, 49, 75, 97, 106, 133, 135 Marsden, J. E., 3, 29, 133 mass matrix, 10, 11, 30, 55 mass-point system, 10 moment, 31 multibody systems, 55 Neimark, J. I., 1, 2, 16, 110, 117, 120, 125, 129 Newton's law, 4, 9-11, 14, 31, 36 method, 99 orthogonal group, 19 Pauli spin matrices, 21 Petzold, L. R., 105, 106 planar motion, 3, 27, 43, 58 potential, 16, 18, 42, 52 principle of virtual work, 14 quasi coordinates, 32 quaternion algebra, 4, 21 conjugation, 22 modulus, 22 multiplication, 21 pure. 22 Rabier, P. J., 72, 94, 96 Ratiu, T., 133 Rheinboldt, W. C., 72, 94, 96, 97, 106 Riesz representation theorem, 46, 74 rigid body, 19, 29 Roberson, R. E., 2 rotation group, 20 Rund, H, 2 Schmidt, T., 112
Schwertassek, R., 2 second fundamental tensor, 55, 94, 96 Sommerfeld, A., 16 Sparschuh, S., 3 special orthogonal group , see rotation group 20 state space, 10, 26, 34, 56, 90 subimmersion theorem, 82 submanifold, 133 submersion, 133 submersion theorem, 16, 20, 97, 102, 133 Synge, J. L., 2 tangent bundle, 16, 26, 32, 135 bundle morphism, 74 space, 134 tangential coordinate subspace, 98 total energy, 52 trivialization, 32 Uhing transmission, 111 unconstrained motion, 2, 4, 10, 15, 16, 29 Van Loan, C. P., 21 variational principles, 9 virtual displacement, 15, 16 velocities, 14 Wanner, G., 106 Wehage, R. A., 106 Whitney, H., 20 Zeidler, E., 11