Sunteți pe pagina 1din 65
In. WV. VI. Theory of Buckling and Post-Buckling Behavior of Elastic Structures BERNARD BUDIANSKY Dieision of Engineering and Applied Physics Harvard Unitersity, Cambridge, Massachusetts Introduction . 6. 2 2 ee | Simple Models... ee ee A. Bifurcation and Post-Buckling Behavior... 6 6 ee ee et B. Initial Imperfections; Snap Buckling ©... +--+ 2 ee cree C. Imperfection Sensitivity ©... ee hortening Relation... ee ee Functional Notation, ional Calculus, and Frechet Derivatives ‘A. State Variables, Functions, and Functionals 5.2. 2 ee B. Norms, Linear and Multilinear Operators, Inner Products . . - - - + C. Variations; Gateaux and Frechet Derivatives... 6 2 ee D. Calculation of Frechet Derivatives 6... 22 22 ee Energy Approach 2. se ees A. Principle of Stationary.Pote B. Bifurcation Analysis... 0.0 ee eet C. Post-Buckling Analysi D. Initint Imperfections 2... 2 ee eee E. Load-"Shorening” Relation... 2s se F. Other Stationary Functionals 2... 2 eee G. Str A B. c . Stability. placement; Virtual-Work Approach Relation; Virtual-Work Equation... . . . Bifurcation and Post-Buckling Analysis»... 5 2s + > Special Cases. ee D. Shallow Shells; Donnell-Mushtari-Vlasov Shells... se + E, Initial Imperfections... 0 eee Mode Interaction... 20 ee ee ‘A. Simultaneous Buckling Modes; Post-Buckling Analysis... 5. + + + B, Initial Imperfections =... eee C. Nearly Simultaneous Modes 0. 2 ees References eee tte 2 Bernard Budiansky I. Introduction The general theory of buckling and post-buckling behavior of elastic structures enunciated by Koiter (1945) has spawned a considerable amount of research in this field, especially during the last ten years. A recent com- prehensive survey by Hutchinson and Koiter (1970) provides a very useful bibliography, together with an overview of the achievements, status, and goals of post-buckling theory. Their survey obviates the necessity of provid- ing a similarly comprehensive guide to the literature in the Present paper. Rather, the primary aim of this article is to provide a unified, general presen- tation of the basic theory in a form suitable for application to a wide variety of special problems. This will be done with the help of the succinct notation of functional analysis, which turns out to be remarkably appropriate for the purpose. In addition to Koiter’s original work, very many Papers on the general theory have emerged from the British school of post-buckling theorists (cg., Sewell, 1968; Thompson, 1969), almost exclusively in the language of finite- dimensional systems, In the United States, variations of the Koiter approach have usually been based on continuum concepts, with a bias toward virtual work (Budiansky and Hutchinson, 1964; Budiansky, 1965, 1969; Budiansky and Amazigo, 1968; Fitch, 1968; Cohen, 1968; Masur, 1973) rather than energy formulations (Seide, 1972). The derivations and results of the present exposition are equally applicable to continua and _finite-dimensional systems; the virtual work and energy approaches are given separate treat- ments, but their equivalence is made explicit. Throughout,the present paper, basic concepts of stability are relegated to a secondary role, in contrast to the central position they held in Koiter’s work. This, however, is largely a matter of taste, and will not affect essential conclusions concerning initial post-buckling behavior and imperfection sensitivity, II. Simple’ Models Very simple conceptual models can illustrate with remarkable verisimili- tude many of the essential characteristics of the buckling and post-buckling behavior of more complicated structural systems. Before undertaking a gen- eral analysis of arbitrary elastic structures, we will exploit such models in order to expose basic concepts of bifurcation buckling, snap buckling, imperfection-sensitivity, load-shortening relations, and stability. Buckling Behavior of Elastic Structures 3 A. BIFURCATION AND PosT-BUCKLING BEHAVIOR Consider the primitive modelt shown in Fig. 1, consisting of a vertical rigid rod of length L, fixed with respect to translation at its base, and elastically constrained against rotation € by a spring that supplies a restor- ing moment f (é). In the presence of a vertical load 2 applied as shown, the rod is in a state of static equilibrium only if aLsiné = f(2). (21) le a!" Fi. 1. Simpte model With f (é) given by SO=KiE+ KC + KE +e (Ki > 0) (2.2) the fundamental solution & = 0 of (2.1) is available for all 2, but in addition the buckled states represented by L=slOlLsing (23) exist for nonzero values of é. These states lie on an equilibrium path in the P-é plane (Fig. 2a) that is connected to the fundamental path ¢ = 0 at the critical load 2, given by h. = Ky/L. (2.4) + Thisis precisely the model used by Masur (1973) in his recent exposition of shell buckling, and by Koiter (1972) in his recent textbook. 4 Bernard Budiansky In the vicinity of € = 0 (indeed, for path < 7m) these buckled states lie on the ah thethyeaen, (25) where Ai/te =(Kx/Ky), Aa/2, = (Ky/Ky +4)... (2.6) Figure 2a illustrates a case for which K, < 0. » i Xe (0) (b) (ec) Fic. 2. Equilibrium paths. (a) 2, <0; (b) 2, =0, 22 > 0; (¢) 4, = 0,4; <0. The imagined transition at 2 = 2, of the state of the structure from the fundamental equilibrium path to the buckled path (in either direction) is called bifurcation buckling, because the fundamental path can naively be viewed as having split into several branches. For 4, £0, the bifurcation is asymmetric; symmetric bifurcations, illustrated in Fig. 2b,c, occur for 4, =0, 42 # 0. The phrase post-buckling behavior (or, more precisely, initial post- buckling behavior) refers to the character of the buckled equilibrium paths in the vicinity of the critical load 2,. The discovery of the equilibrium states in Fig. 2 does not, by itself, answer the question of what is the actual behavior to be expected of the structure modeled in Fig. ! when it is subjected to a load that is increased slowly from zero. At least part of the answer is provided by the introduction of initial imperfections. B. IntTIAL IMPERFECTIONS; SNAP BUCKLING Assume that in its unloaded state the rod must inevitably suffer some initial deviation € from a perfectly vertical position (Fig. 3); then, if € is the additional rotation produced by the load, equilibrium states must satisfy A=Sf (EYL sing + 2). (2.7) : Buckling Behavior of Elastic Structures 5 es Fic. 3. Initially imperfect simple model These are shown schematically (for the case 2, < 0) in Fig. 4a,b for positive and negative values of . The intersecting equilibrium paths of the perfect structures, shown dotted, have been deformed into disjoint branches by the presence of € ¢ 0, and with respect to loading that increases from 2 = 0 the branches that do not pass through the origin have evidently become irrelevant. For various values of @, the significant branches that do contain the origin make up the two familics shown in Fig. 5a,b; as TE] approaches zero, the members of cach family approach different pairs of branches of ‘the equilibrium paths associated with the perfect structure. For neither sign of& do paths from the origin ever approach the fundamental path above A= dh. _— I | | | 1 t | i (a) (o) Fic. 4, Equilibrium paths for imperfect model. (a) ¢ > 0; (b) ¢ <0. 6 Bernard Budiansky For a sufficiently small positive value of é, the associated equilibrium path (Fig. 5a) displays a local load maximum 4, < 4, ata rotation &, that depends on ¢. Obviously, if the actual loading is increased beyond 2,, something nasty will happen; for 2 > 4, static equilibrium is impossible at values of € in the vicinity of €, and soa dynamic process involving initial loadings must ensue. If velocity-dependent damping is invoked in a conceptual attempt to bring the structure eventually to rest, the final static configuration (if it exists) cannot be arbitrarily close to ¢,, no matter how slightly 2, has been exceeded. Accordingly, 2, is designated as the buckling load of the imperfect structure. This kind of buckling, associated with a local maximum in the static equilibrium path and the expected jump to a distant configuration when the maximum is exceeded, has variously been called snap buckling, snapping, limit-point buckling, and oil canning. {o) {b) E16. 5. Equilibrium paths for various values of inital impectection & (2, < 0) (a)2 > 0; (b)E <0. The curves of Fig. 5b, for Z < 0, do not suggest such a dramatic behavior. They do indicate that for loads in the vicinity of 2, significant increases occur in the rate of growth of rotation with load, but distinguished loads different from , that signal this increase are not usefully defined for various values of small €. Thus, for = < 0, 2, Still deserves to be called the buckling load of the structure, but this buckling is mild in comparison with the sharp, possibly catastrophic, snap buckling that occurs at 4, for E> 0. These observations, for 4/4, = K2/K, <0, are readily extended to other combinations of parameters in S(¢). If 2, > 0, small negative values of € provoke snapping (Fig. 6a); in the case of a symmetric bifurcation, with 4, = 0, snap buckling is induced by initial imperfections of either sign if 42 <0 (Fig. 6b); 22 > 0 produces only mild buckling (Fig, 6c). Clearly, it is Buckling Behavior of Elastic Structures 7 7? X Cc ds Z, _ LA €>0 Feo 7% é (a) (b) (c) Fic. 6, Equilibrium paths for perfect and imperfect models. (a) 4, > 0; (b) 2, = 0,23 < 05 (0) 4, = 0, 4, > 0. the initial post-buckling behavior of the idealized perfect structure—in par- ticular, whether the load increases or decreases after bifurcation—that deter- mines the kind of buckling to be expected in the imperfect one. C. IMPERFECTION SENSITIVITY If snap buckling does occur, the magnitude of 2, can be significantly below 2,. Maximizing 2 with respect to & in (2.7) requires that AL cos(E + €) = f'(é) (2.8) and this, together with (2.7) provides the asymptotic result 9g A = 2A (A4/2e)E]? (2.9) for small (€) and (2, €) < 0. Similarly, for 2, = 0, 2, < 0, Ale 1 = 3(—2,/2) GE (2.10) for small €. In both cases, small imperfections can make the snap-buckling load 2, substantially lower than the critical load A, of the ideally perfect structure. Structures in which this can occur are said to be imperfection sensitive. In the models studied, imperfection sensitivity is implied by 4, #0, or by 2; = 0, 22 < 0, and the seriousness of this sensitivity is then determined by the magnitudes of 2,/2, or 22/2,. [If 2, + 0 and 2, < 0, the approximation (2.9) may become inadequate for moderate values of @; 4, must then be found from a more accurate solution for the maximum value of 2 implied by (2.7), which will be lower than the separate values given by (2.9) and (2.10).] 8 Bernard Budiansky D. Loab-SHortENING RELATION Instead of prescribing the load 2 on the imperfect structure, we could impose a vertical displacement, or “shortening,” A at the point of load application (Fig. 3), given by A/L = cos(é + &) — cos (2.11) in terms of € and & In the limiting case f= 0 of the perfect model, A/L ® 427, and so, from (2.5) Hb, A+ (Ay/A)QA/L)"2— (E> 0) (2.12a) RL (A/2)QA/L)? — (E <0) (2.12b) & 1 + 2(2o/2,)(A/L) (2, = 0,2 #0) (2.13) for small A. The solid curves in Fig. 7 show typical load-shortening relations, together with associated 7-é relations. Note that asymmetric bifurcations é an é an O/L {e) (a) Fic. 7. Load-shortening relations for simple model. (a) 4, <0, & > 0; (b) 4, <0, (c) 4, = 0,2, < 0,2 2 0; (d) 4, =0,4,> 0,220. Buckling Behavior of Elastic Structures 9 give a vertical initial slope to the 2-A relation, whereas d7/dA is finite for symmetrical bifurcations. In the case of the imperfect structure, Eqs. (2.7) and (2.11), for a given €, supply the 4-A relation parametrically in terms of ¢, with typical results sketched as the dotted curves in Fig. 7. In those cases (Fig. 7a,c) that involve snap buckling when an increasing load 2 is imposed, the results indicate that prescription of a monotonically increasing shortening A would not provoke a violent response, but would simply induce a smooth peak in the magnitude of the associated load. However, a slight modification of the model can render it susceptible to snapping under an imposed shortening. Replace the rigid bottom support by an elastic spring that supplies a vertical re ing force ce, where ¢ is the vertical displacement of the bottom of the rod (Fig. 8a). Then A will, at any 1 » nl / Fic. 8. Load-shortening relations of (a) modified model. (b) 2, <0; (c) 2 = 0, 2; <0. load 2, exceed that found for the original model by the amount //c. For 4, < 0, and a sufficiently small € > 0, this leads to the possibility of a path in the 7-A plane (Fig. 8b) that displays a local maximum A,, in shortening. Under increasing A, snapping must occur as soon as A exceeds A,, , for then no neighboring static equilibrium state is to be found. The same thing can happen for 2, 40, 22 <0 (Fig. 8c) for small enough @ if, for = 0, the initial post-buckling value of d2/dA = (1/c) + (222/L/,) is positive—that is, if the initial post-buckling behavior of the perfect model shows a decrease in A as well as in 2. In all cases, when snapping occurs under increasing A, the load 2 will already have decreased from its peak value 7, . 10 Bernard Budiansky E. Stapiuity The analysis of bifurcation buckling, together with consideration of initial imperfections, has apparently provided an understanding of the behavior of the simple model, but from a practical viewpoint the study is really not complete without some attention to stability in the dynamic sense. For example, the presumption that the static equilibrium paths of Fig. Sa will be followed up to 4 = J, involves the tacit assumption that if the prescribed loading 7 increases slowly enough—or is applied in small enough increments—then time-dependent deviations Ag(t) from the static equilib- rium path induced by the loading will be damped out whenever further loading ceases and 2 is held fixed. Also, with the load fixed, small enough accidental disturbances A in orientation (or Aé in angular velocity) produced by an external agency should similarly be damped out if the static equilibrium path is to have any useful physical validity. These requirements are just those of stability (more precisely, asymptotic stability) in the Lya- punoy sense. The same question of stability must be raised with respect to the curves for € < Oin Fig. 5b, in which 2 eventually exceeds the critical load 4, of the perfect model, and yet we presumed no practical difficulty in following the equilibrium paths by means of slow loading. Fortunately, the stability of finite-dimensional elastic systems subjected to conservative loads (i.¢., loads associated with an energy potential) can be assessed on the basis of the classical energy criterion. In static equilibrium, with the load fixed, the total potential energy of the system must be stationary with respect to its configurational arguments; but the equilibrium is stable if, and only if, this stationary point is actually a local minimum. The argument for sufficiency is simply that a proper minimum in the potential energy provides an energy barrier between the initial configuration of static equilibrium and any alternative ones; by keeping initial disturbances small enough the energy increment needed to surmount the barrier can be denied to the system, and damping may then be relied upon to return the structure—asymptotically—to its original configuration. Instability is implied by the absence of a minimum because a configurational disturbance of arbitrarily small size, and sufficiently small kinetic energy, can always be found that would place the system out of static equilibrium into a state of lower total energy (potential plus kinetic); then damping can only decrease the total energy of the ensuing states still more. Even if the damping should ultimately lead the system toward a state of static equilibrium with no kinetic energy, this state cannot possibly coincide with the original state of higher potential energy. This is a rough, abbreviated version of Koiter’s (1965a) proof.t 1 The writer is informed by Prof. Koiter that this proof is actually due to Jouquet (1930). Buckling Behavior of Elastic Structures 1 In the case of our simple model under prescribed load 2, the potential energy is given by =| Ede + Abfeos(e +2) — cos 2 and é¢/0§ = f(€) — ALsin(E + ) vanishes, by static equilibrium [Eq. (2.7)]. Stability is implied by a positive value for ag/0e? = f'(2) — AL cos(E + 2), (2.14) but along the equilibrium paths of Fig. 5a, 67/82? is positive at € = 0, and does not vanish until 2 reaches 2, [see Eq. (2.8)]. This verifies stability for 0 <2 2,, and that on the bifurcated paths the rules relating the sign of d2/| dé | to stability found for the imperfect case still apply. At 2 = 2,, 6/dé vanishes, and a decision on stability rests on the signs of higher derivatives; stability can thereby be verified at 2 = 4, in case (c) of Fig. 6 and instability in cases (a) and (b). It is no accident that the boundary between stability and instability on the fundamental path occurs at the bifurcation load. Given any analytic function $(é, 2), the vanishing of 46/aé on two intersecting paths of the 2-¢ plane implies that 67/0? (and also 67/é& 42) must vanish at the point of intersection. Ill. Functional Notation, Variational Calculus, and Frechet Derivatives A. STATE VARIABLES, FUNCTIONS, AND FUNCTIONALS In conventional continuum mechanics, the state of an elastic structure may be described by the spatial distributions of stresses, strains, and displacements. In the approximate analysis of particular structures on the basis of special theories, generalized stress, strain, and displacement varia- bles should always be identifiable. For example, in simple beam theory, the 12 Bernard Budiansky “stresses” are the bending moments, the generalized “strains” are the cur- vatures, and the lateral deflection along the beam is the pertinent displace- ment variable. Similarly, in a pin-jointed truss the member loads and clongations may appropriately be identified as the stress- and strainlike quantities, and the values of displacement at the joints suffice to constitute the set of appropriate displacement-state variables. In the context of any such particular theory it will be useful to denote a distribution of generalized stress by the single symbol ¢, and to imagine that ¢ represents a “point” ina space that contains all stress distributions that are mathematically (if not physically) possible. Similarly, we will use ¢ and u to denote points in strain and displacement spaces. These state variables will be related to each other in various ways by the ficld equations of the theory being used. Thus, the elastic constitutive equa- tions permit the calculation of in terms of e; the procedure for so doing may be regarded as a transformation from strain space to stress space. We can then say that o is a function of e, and it will be very convenient to use the single symbol ¢ to denote the name-of this function of e, as well as the value of stress the function produces. Thus we will write o = o[e], and not insist on the more respectable form ¢ = f[e] which carefully distinguishes the varia- _ ble o from the function f (The reader is warned, however, that our more casual notation is regarded with considerable distaste in certain circles.) In structural mechanics, global scalar quantities arise, such as the poten- tial energy $, which can be expressed as a function of the generalized displacement u. Such a scalar-valucd function is often called a functional and the value of ¢ for a particular choice of w may be expressed as = fu); this represents a transformation from displacement space to the one-dimensional space of real numbers. B. NorMs, LINeaR AND MULTILINEAR Operators, INNER PRODUCTS The generalization of length in the space of displacement is provided by the norm of u, written as ||uj|, which is a measure of the “size” of u. A norm should have the properties that |u|] > 0 for u #0, \jul| =0 for u=0, and Wael = || ul G.1a) for all scalars «. Also the triangle inequality Nea + tall S eu) + eal] (3.1b) Buckling Behavior of Elastic Structures 13 should be satisfied. Subject to these restrictions, the choice of a norm, in a particular theory, is often a matter of convenience; some choices make it easier to prove things than others, but in this paper we will not deal seriously with matters that require precise discrimination among various possible norms. A function (or functional) fis linear if Sau] = of [u) (3.2a) for all scalars «, and Sty + we] = fla] + Sle] (3.2b) for all uy, #2. For such a linear function, the notation Sf {u)] = Au may be introduced, and then A is called a linear operator. This operator may in turn depend on another variable v, and to emphasize this we could write f [osu] = Afe]u. Similarly, a function g that is linear in each of the variables u, and 1, can be expressed in terms of a bilinear operator B by an expression of the form glity, u2] = Buyu,. If Bu, = Buzu, for all 14, u,, then B is a symmetric bilinear operator. Finally, the expression Cu, 12 +++ u, may be used to denote a multilinear function of tty, t2,...,u, in terms of a multilinear operator C, and this operator is symmetric if the u's may be interchanged in any manner without affecting the result of the operation. If Bu, uy is a scalar and B is a symmetrical bilinear operator that satisfies Buju, = But > 0 for u, #0, then Bu, vz may be written in the inner product form tt, ty), and this may often be used to provide a convenient norm defined by Jul) = 0. But the two definitions are equivalent when f” is continuous in uv (Vainberg, 1964), and in this paper we will refer to f’ as the Frechet derivative, The relation (3.5) may be written conveniently as S'uy = (6/6e)f[u + ety] <0 (3.7) and this leads to a convenient definition for higher-order generalized deriva- tives of f/ Thus =0. (3.6) a Sy uy = gy lu + eur}iybe=o and, assuming sufficient continuity of the higher-order Frechet derivatives of J, this is the same as Sy, = (G7 /bey Oea)fu + et: + erttal yweyeo 6.8) wherein the order of differentiation is immaterial. It is evident, then, that Sty Uy =f" ty, So that f” is a symmetrical bilinear operator. Generalizing (3.8), we get the relation em Ce, G2 Sus uy ty ley=ea Buckling Behavior of Elastic Structures 15 satisfied by the nth-order Frechet derivative of f, which is a symmetrical multilinear operator. Note that identities like Sy thy = (Pfee2) fu + ert lesao and Suyugtiy = (CP /écy 2c3)f[u + cpu, + e2U2]k,-c.<0 follow from (3.9), and the convenient abbreviated notation f"u? = fu, u,, Sy US =f" uy uy Uy Will be used. This notation permits the succinct Taylor- series representation of an analytic function or functional in the famil looking form Sle] = Slo] +S To] tt = to) + AS "Tole = to? +. (3.10) D. CALCULATION oF FRECHET DERIVATIVES The formal definitions given for Frechet derivatives of any order are entirely equivalent to the familiar processes of “taking variations” in the calculus of variations. A few examples will illustrate such calculations. Consider the functional Fea i, [w? + 3(o'?? + xn] dx. The variable u is identified as the pair of functions w(x), c(x). Recall that 6F = [Be dw + 60" bv! + x Sw] dx. Thus, letting dw = wy, dv = v,, we have Fu, = {, lo. + x)wy + (6v'pv4] dx. Next, taking another variation, with dw = w , 5v = vy, leads to Fuji, = [ en + 60,03] dx, ‘o Continuing, 1 6wy ww de F'ty igus = to and then FO=0 for n>4. 16 Bernard Budiansky As another example, consider the nonlinear strain-displacement relation & = (Bu/dx) + 4[@u/dxy? + (@x/ex? + (éw/ex)?] of elasticity theory. We get, with U as the designation of all three displace- ment components u, », w, Guy | dude, , Gvev, | Bwewy ax ax ax 6,U, = and Ou, Guz | dv, Av, | Sw, aw. “UU 1 Oty, Ov, vy Siwy Big fee? Ox Ox | ox ox ax Ox Note, finally, that the conventional higher-order variations 5?F, 5°F, of the calculus of variations are simply SF =4F(OuP, OF “Qu, OOF = Posuy, so that the Taylor expansion (3.10) is the same as Su + du] = fu] + f'[uldu + Sf"[u] (Su)? + ; H=Sft f+ Ffto (11) Systematic, rigorous expositions of functional analysis may be found in Vainberg (1964) and Liusternik and Sobolev (1961); a very readable, elementary account has been given by Rall (1969). IV. Energy Approach A. PRINCIPLE OF STATIONARY POTENTIAL ENERGY Koiter’s (1945) general theory is based in large part on the principle of stationary potential energy, and the same approach is adopted in this section. With only a few exceptions, the results to be found are contained in Koiter’s early work, and the major differences here are of style, emphasis, and notation. The only state variable to be used in the energy approach is the gener- alized displacement u. The potential energy @ of the structural system under contemplation is considered to be a functional of u, but it will also be presumed to depend on a single scalar variable 2 which determines the magnitude (and possibly the distribution) of prescribed external loads on the system. Informally, we will refer to 2 as the “load.” It will be presumed that Buckling Behavior of Elastic Structures 17 certain geometric restrictions independent of 2 (typically, boundary or sup- port conditions) are placed on the displacements. These restrictions, together with any more general requirements concerning continuity that may be appropriate, constitute admissibility conditions that constrain u to lie in a well-defined subspace, from within which equilibrium states will be sought. This search will be based on the stationary energy principle which asserts that a particular w is an equilibrium state if, and only if, it is admis- sible and 56 [u; 2] = o'[u; Zou = 0 (4.1) for all admissible variations du. The admissibility conditions on du are hom- ogencous, and follow simply from the requirement that, for any scalar a, u + a du, as well as u, lie in the subspace of admissible displacements. Equa- tion (4.1) is usefully regarded as a variational equation of equilibrium. It is to be emphasized that the equation d¢ = 0, as the condition of equilibrium, is meant to reflect the use of a potential-energy functional appropriate to the special structural theory adopted for the analysis of the particular structure under consideration. Only those theories that enjoy such a principle of stationary potential energy are therefore susceptible to the present analysis. This is not, however, a serious restriction, since a good special theory should mimic general continuum mechanical theories in this respect. B. BIFURCATION ANALYSIS In order to discover conditions for bifurcation buckling, we assume first that there exists a findamental solution tt that varies smoothly with 7 as the load increases from zero. The variational equation of equilibrium (4.1) re- quires that $'[uo(2); A}ou = 0 (42) for all admissible 5u. Now suppose that, for some range of A, there is another solution = up(2) + e(2) (43) that intersects the fundamental one at 2,, in the sense that : lim v(2) = 0. (4.4) dake It will be further assumed that #9(2) exists for 2 greater than /, , so that a true bifurcation, rather than a limit point, is implied by (4.3) and (4.4). 18 Bernard Budiansky The bifurcation buckling mode will be defined as ty = lim (e/lo)), (45) ante where |__|] represents a suitable norm; note that |Ju;|| = 1. Since the u given by (4.3) must satisfy equilibrium, $'[uo(4) + v(4); 2] Su = 0, and, under the assumption that ¢ is analytic in the vicinity of uo(2,), a Taylor-series expansion gives O'los A] Ou + $"[tgs Av bu + 46" [os Zo? Sue = 0 (4.6) for sufficiently small |2 — 4, |. But the first term vanishes by (4.2); dividing the rest of (4.6) by |v], and letting 2 > A, then gives "(uo (2); Aju, Su = 0 (4.7) as the variational equation governing the buckling mode m, and the critical load 2,. The notations UolZe) =e and — [ol ?,)5 are convenient; then (4.7) becomes iy du = 0. (4.8) Equation (4.8) constitutes the variational statement of a homogeneous eigenvalue problem. Admissibility conditions on v, and hence on uy, coincide with the homogeneous conditions on dx discussed earlier. This cigenvalue problem necd not, of course, have a unique solution for 4, but 4, will be defined as the lowest positive eigenvalue. This lowest critical load, in turn, need not be associated with a unique buckling mode «,, but, for the present, uniqueness (except for sign) will be assumed. Another derivation of (4.8) may be instructive. Since ¢'[u; 2] 6u = 0 all along any equilibrium path, differentiation of this equation with respect to time is permissible, and provides a variational equation for the displacement rate du/dt. Thus, along the fundamental path, vg yttto , 208 ltos AN) (¢ {ios AY + a) ou =0 (49) and on the bifurcated path a( ae “) nl deb [ug + $'[o + v: (4.10) ata aR Buckling Behavior of Elastic Structures 19 Letting 4 = A, in cach of these equations and subtracting gives Ge(dofdt), du = 0 (4.11) as the variational principle governing a nonzero displacement rate (dv/dt). that can be added to dup/dt on the fundamental path at 2, without violating equilibrium. This is simply another way to describe a bifurcation and (dv/dt), must be the same as the bifurcation mode u,, to within a multiplica- tive constant. Note that in the derivation of (4.11) there is no assumption that the loading rate d7/dt associated with (dv/dt), must vanish. Casual derivations in the literature of buckling equations often refer to “ buckling under constant load,” meaning, presumably, the existence of a nonzero displacement du/dt consistent with d//dt = 0. But formulation of the eigenvalue problem for bifurcation does not require an explicit assumption concerning d2/dt; and the eigenfunction u, is proportional to the total displacement rate du/dt = (du/d2)(d7/dt) + (dv/dt) only if d2/dt is zero. It may be noted that (4.11) also governs limit-point buckling; this follows from (4.9) for which a nontrivial solution duo/dt, with d2/dt = 0, becomes the limit-point-buckling mode. C. Post-BUCKLING ANALYSIS In order to proceed with a determination of vin (4.3) for 2 # 2, , introduce the scalar parameter é defined by g = Cyn), (4.12) where the bracket symbol represents any bilinear inner product; the only restriction on this inner product is that <1, u,> #0. Indeed, it is partic- ularly convenient to choose the norm |p|] = = 1, and this choice will be assumed henceforth. Then it follows that v= eu, +8, (4.13) where = 0, The parameter € is therefore a measure of the “amount” of buckling mode contained in the difference u(A) — 19(2) between the displacements on the bifurcated path and the fundamental path, at a given value of 2. This amount depends, of course, on the particular choice made for the inner-product norm, but for any choice, €0 for 22 ,. Furthermore, since 2 [el] = uy +h, where h-0 for 2>2,, it follows that jl] = E/(1 —4||hl?), and so v= (uy + A)/(1 — $I\h|?); hence, from (4.13), B= E{(uy +h) — 4]/Al|?) — ty} and therefore 3/0 for 2 2,. 20 Bernard Budiansky Each term in the Taylor expansion (4.6) of the variational equation of equilibrium will now, in turn, be expanded about 7,. We will use the notation HP = OM): EA ars 1 =A OWA HP = AP hase ce " " $= TOMA BP = Ibn (Id) and so on. Then, remembering that the first term of (4.6) vanishes, we have AGE + ~ Ae)OE + 2 — AYRE + Jo du EMO +L AMI 4 fe? ou + HY to ow 4 (4.15) The aim of the analysis is now to discover how u = uo(Z) + Eu, + & varies with 2 along the bifurcated path, and, with u9(2) considered known, this would appear to dictate a search for &(2) and (2). A small change in view- point is, however, very effective; consider € to be the i dependent variable, and look for 2 and 3 as functions of €. To this end, we will anticipate at least the asymptotic validity, for small é, of the perturbation expansions Led thet+ het, (4.16)¢ b= Su + Bry ten (4.17) The full asymptotic expansion for the displacements is then = uo(2) + y+ Cry + uy bo. (4.18) Although we know that &/é — 0 for € - 0, there is of course no guarantee, that these asymptotic expansions must involve only integral powers of €; but it will turn out that this prescription, with rare exceptions, is self-consistent,. in the sense that a systematic procedure can be developed for the calculation of the terms in (4.16) and (4.18). Substituting v = uv — up(Z) from (4.18) into (4.15), together with the ex- pansion (4.16) for 4, and making use of the fact that ds u, Su = 0, gives OP fbrg + Ay hl, + Apu?) du + PU ets + Ar Pha + a Pony + VF hLuy, a es=ssssee (4.19) It follows that the coefficients of é?, ¢°,... must vanish separately, for all tn many earlier papers, the notation 4 = 4,(1 + az + 62? + +++) has been used instead, Buckling Behavior of Elastic Structures 21 admissible variations in du. Set du = u,, and exploit the fact that eta, = Ply, =0 (4.20) by (4.8), wherein du can be chosen as uy. Similarly, #213 u, = 0, and so Ay = —$erdbend, (4.21) = Ala oihon + Halden + AO] MER yay $euy , : and expressions for the successive coefficients 2, 24,-.-in (4.16) can also be found in a straightforward way. It is seen that the post-buckling coefficient 4, depends only on the buckling mode u,, but the calculation of 2, would generally require the determination of 7, as well, and so on. It is now evident that the relations between 2 and é found for the bifurcated path are entirely similar to those illustrated in Fig. 2 for the 2-é relation of the simple model. Again, an asymmetric bifurcation corresponds to 2, # 0, and in symmetric bifurcation the load rises or falls during buckling according to the sign ofA. The results for 2, and 2, are, of course, not valid if $”u? = 0, but this occurrence, although possible, can only be accidental. In fact, as will be shown later when stability is discussed, it is generally to be expected that $12 <0. [Nevertheless, if ¢/u? should vanish, the post-buckling analysis can still proceed on the basis of (4.6), but nonintegral powers of € may then be needed in the expansions for 2 and u. Furthermore, even with a unique buckling mode #1, more than one post-buckling path for each sign of € could conceivably then intersect the fundamental path at 2,; on the other hand, there might not be a real bifurcation at all, despite the existence ofa solution to the eigenvalue problem.] A variational principle governing 1, can now be deduced from (4.19) by asserting that the coefficient of ¢? must vanish for all admissible Su: Pitty Su + Ay blu, du + 460? du = 0. (4.23) By virtue of (4.8) and (4.21) this is automatically satisfied for du = u,, Note that while (4.23) represents a nonhomogeneous problem for 1, there is an apparent lack of uniqueness in the solution since, by (4.8), any multiple of u, can be added to a particular solution of (4.23). But recall that since (i, u;) must vanish, it follows that = Oforn > 1. Consequently, ifti, isany particular solution of (4.23), u, must then be ty = i — Ciiy uy). (4.24) Once u, is found, 2, as given by (4.22) can be computed. It should be remarked that the use of (4.23) is not necessarily the best basis on which to calculate v2. The original variational statement ¢'[u] bu =0 22 Bernard Budiansky has, as its consequence, a set of field equations (be they algebraic or differen- tial) and possibly “natural” boundary conditions that, together with admis- sibility conditions on u (¢.g., geometrical boundary conditions) determine u. The expansions (4.16) and (4,18) could be substituted directly into these equations and a set of equations governing tg, v4, U2, ... could be found bya standard perturbation procedure, The resulting equations for u, must then coincide with those implied by the variational statement (4.23). In many (possibly most) buckling problems of practical interest, the bifur- cation is symmetric, the coefficient 2, vanishes, and the relations (4.22) and (4.23) for 2, and uv, then simplify considerably to da = —(baltut + guduaybeuh (422) Getty du + 4gtut du = 0. (4.23') In this case, since setting du = uy in (4.23') gives upuy = — 26203 the result for 2, can be written as Ja = = (Fut — WheuZyoeu;. (4.22") It should be noted that if 2, vanishes, then the value of 7, is independent of the particular choice made for the inner-product normalization in (4.24); by (4.20) the addition to uz of any multiple of u, leaves 12 in (4.22") unaffected. All of these results would assume slightly simpler forms if the inner product in (4.12) were chosen as Cu, with w, so normalized that fu} = —1; but this is not necessarily a conven- ient choice from the point of view of providing a transparent physical inter- pretation of é, and will not be made herein. = — dw D. IntriAL IMPERFECTIONS If the structure under analysis is not quite perfect, in that it contains a displacement tt before the application of load, its potential energy functional in terms of the additional displacement u that occurs during loading is ex- pressible in the form $ = ofu; 2] + Ye, ii: 4, (4.25) Buckling Behavior of Elastic Structures 23 where @ is the potential energy of the perfect structure, and wu, ti; A] is a functional of v and @ that may also depend on /. The additional contribution y to the energy may be always chosen to obey the conditions Wfu,0;2]=0 forall w (4.26) and v0.7; 7J]=0 forall i (4.27) ‘The variational equation of equilibrium is now $'[u, i; 2] du = 0, where the Frechet derivative is with respect to u. Hence glu; 2] ou + Y's, W; 2] du = 0, (4.28) and it is assumed that admissibility conditions on u remain the same as they were for the perfect structure. It is convenient to let a= i, (429) where |jiil] = 4, gives gov, Su = 0, (4.33) and this verifies that v, is indeed the same as tt. We now introduce € = (@, »,), as in the perfect case, and try to find a relationship among 2, 2, and € that is uniformly valid for suitably restricted small values of é, & and |2 — 4.|. Mote precisely, we will look for a dependence of 2 on é and € along curves € = af? in the ¢~ plane (Fig. 9), where a is a scalar parameter, and the exponent 7 > 0 will be chosen to suit our convenience. (For a given 7 any combination of € and € can be reached by an appropriate choice for a.) Expanding the operators in (4.32) about 2 = ,,and substituting = a2? gives 28 Sut (2 — ADD Su to + AGT? bu to + AGED? Gu te + OEY Su + EY EATE Su t+ OE YETH Ou 4 = 0, (4.34) where SUE lina: and so on. If the choice of an integer value for y > 2 is anticipated, it is appropriate to assume expansions in the form Aad tH E4R, CO He (4.35) and Gy + v2 + Boyt, (4.36) Buckling Behavior of Elastic Structures 25 where = 0, and 2,, 2), ..., as well as the v,, will depend on a. Then (4.34) becomes OP [bler + Ay bei, + AG ui] Su + Olhlvs + Ardliy + Gide, + Agha to Jou to $a yeti du t= 0 (4.37) wherein the leading term €#% 1, du = 0 has been dropped, and the omitted terms in the second bracket are all proportional to 2, or 2?. Now pick » = 2, and set du = wu; then, Ty bead + Age} + cali, = 0 whence [see (4.21)] 7, = 4, -aphes (4.38) where = Wing /2, he. (439) Hence, AB, +E (uphJE (4.40) and, having completed our exploitation of « = E/é?, we eliminate it from (4.40) to get Aw he + hye — (Eprsé) (441) for small enough values of € and & The approximation can be improved systematically, but for 2, + 0 appears adequate for asymptotic estimates of snapping loads 2, . If 2; < 0, and (Ep) > 0, then (4.41) represents a relation that does display a local maximum near 2 = 2,, which occurs at & = (Ep. Jay)? (4.42) giving Ald, % 1 — 2(-EpAy/2.)'?. (4.43) This holds also for 7, > 0 and Ep < 0, and is seen to be similar to the result (2.9) for the simple model. For 2, Zp > 0, (4.41) does not display a local maximum in 2, but in the vicinity of 2 = A, gives -é relations like those shown in Fig, 5b for the simple model. It must be presumed that the appro- priate branches of the asymptotic result (4.41) can be matched smoothly to solutions A(Z; @) for small 2. If 2, = 0, the approximation (4.41) found is clearly inadequate (since for € #0 and € = 0, the post-buckling solution of the perfect structure is not obtained) and must be carricd one step further. The casiest procedure, 26 : Bernard Budiansky however, is to shift gears and set y equal to 3 instead of 2 [in (4.37)]. Then 2, = 0, and : bevy du = —Agiu? Su So that v2 is identified as w,, and with du = 1, the terms of order & in (4.37) give Jat aph,. (4.44) With « now replaced by &/%, (4.35) gives AR) + da? — (Ep /é), (4.45) and, for 2, < 0, the snapping load becomes Asfhg % 1 = 3(-22/2,)'? (REP, (4.46) at & = Ep2./-22,)"°. (447) Now the similarity to Eq. (2.10) is close, and with E replaced by Zp, the diagrams of Fig. 6b,c, for 2 near 2, show branches of (4.45) that should match into results for small 2. Essentially the same asymptotic result would be obtained if we stayed and continued with the perturbation analysis based on y = 2; furthermore, the results for 2, + Ocould have been found with y = 3, but ina less straightfor- ward fashion. All of the results found are susceptible to systematic improve. ment, but, as is in many asymptotic studies, the improvement may be illusory, or not worth the effort—or both. The imperfection sensitivity discussed is clearly dependent (via p) on the choice of it for the shape of the imperfection. Some general insight into the effect of this choice, and the consequent magnitude of the nondimensional parameter p, will be found later when the virtual work approach to the general theory is developed. E. LoaD-" SHORTENING” RELATION Returning to the perfect structure, we remark that in most problems the potential energy function @ can be expressed as oles 2] = 6 + Blu; 2], (4.48) where & is the strain energy of the system and B is the potential energy associated with the prescribed loading. Furthermore, it is usually true that B can be written in the form Blu; 2] = —2Au), (4.49) Buckling Behavior of Elastic Structures 27 where A[0] = 0, and by analogy with the simple model A[1] can be regarded as a generalized shortening. Assume further that 6 docs not depend ex- plicitly on 2 (only implicitly via 1), and it follows that A= —(0/62)d[u; 2), (4.50) but since (d/d2)g[u; 2] = [us A](du/d2) + (6/02) 6[u; 2] and ¢'[u; 2](du/d2) must vanish by (4.1), A = ~(d/dd)glu; 2] (4.51) Note that, by equilibrium along any path, 0 = (d/d2)b[u) du = (09/62) — f"(du/d2) bu = —A‘ du — $"(dufd2) bu. Now set u = u,, and du = ,; since @2(du/d2)u, = 0 by (4.8), we have Alu, = 0, (4.52) which says that at the critical load the shortening is stationary with respect to the buckling displacement. With the prebuckling shortening defined by Ao = —(W/d2)[uo(2); 2] = —b, we have A — Ag = ~(a/d2) Gls 2] — bo} = —(d/d2)(Glue + 052] — Ga} (da2\igou + Gu? + Lose +4 = —hov — Aboe? ++ — [Fo + d5v + App? + --Ji, where » = do/dd. But the bracketed expression is the Taylor expansion of $'[¥o + 52] about up, and, by equilibrium ¢'b must vanish; so does $% v, since ¢’ Su = - Su = 0 all along the fundamental path. Consequently, A Ay = 1650? and so, for small é, Aw Ag ® 4276217. (4.53) This gencral result has apparently not been exhibited hitherto. 28 Bernard Budiansky Accordingly, with 2% 2, + 4,€ + 2&7; APR A+ AVA) (2/b02)(A - Ao)!” for €>0 (4.54a) 1 AsAM-Q/b2j (A — Ad)? — for E< 0 (4.54b) 1+ {22/2 \f—bag [A — Ao] for 2,=0,2;40 (4.55) wherein, to be consistent with (4.53) and the condition #17 < 0 previously mentioned, A — Ay > 0. These results are, again, similar to those given in Eqs. (2.12) and (2.13) for the simple model (in which there was no prebuck- ling shortening Ag). It is worth emphasizing that they are quite general, including cases in which the prebuckling variation of Ag with Z is not linear. Some typical initial 2-A curves, together with post-buckling 2-¢ relations to which they correspond, are sketched in Fig. 10. _ is (0) 7 A & a rc) (e) Fic. 10. General load-shortening relations. (a) 2, <0,¢ > 0;(b)2, > 0,€ > 0; (c) 44 = 0, Jz <0; (d) Ay = 0, 2g > 0. For the imperfect structure Eq. (4.53) remains a valid first approximation for the relation between A-A, and €, and the variation of 2 with A-Ag is readily found from a cross-plotting elimination of € between (4.53) and either (4.41) or (4.45). Results similar to those given by the dotted curves in Fig. 10 would then be found. Buckling Behavior of Elastic Structures 29 In the case of symmetric bifurcation, it is useful to define the initial post- buckling stiffness of the perfect structure as K = d2/dA at 2 = 2,, and com- pare it to the corresponding prebuckling stiffness Ky at 2= A, on the fundamental path. [For asymmetric bifurcations, where A= Age —(2 ~ A)out/223, the two stiffnesses coincide.] We find YK = (1/Ko) + [(— dui /2A2] or, since dig) _ oo D7 ob K/Ko = [1 + (#29/2420.))"'. (4.56) This implies the variety of post-buckling behaviors indicated in Fig. 11. The initial post-buckling stiffness vanishes for 2, = 0, and for increasingly posi- tive value of 2, approaches the prebuckling stiffness. For negative 2), the load must drop during bifurcation, and at a critical negative value of A, given by dy = ~ bi 28. the 2-A curve becomes vertical. For 2, <2, da/dA again approaches Ko. As in the modified simple model, bifurcation 4; < 7, implies the pos- sibility of snap buckling (Fig. 8c) for an increasing a imposed on the struc- ture, whereas an increasing / leads to snap buckling for all 2, < 0. Values of K for 2, > 0 are useful in assessing the load-carrying capacity of structures that continue to carry loads after buckling without snapping. As in the simple model, snap buckling under imposed shortening for 4, # 0 is always possible for sufficiently small Z, when Zp/, < 0. Fic. 11. Initial post-buckling stiffnesses (2, = 0). 30 Bernard Budiansky F. OTHER STATIONARY FUNCTIONALS The preceding development can be applied on the basis of other station- ary principles that determine the state of the structure. For example, the Reissner-Hellinger principle (Reissner, 1953) states that a certain functional Q of stress, strain, and displacement must be stationary. If the state variable U is now used to denote an appropriate normed linear space of all the state variables contained in the argument of Q, the entire analysis gocs through essentially unchanged, with the starting point Q’ 5U = 0 as the governing variational statement. A similar remark concerns the development in terms of stationary poten- tial energy when there are constraints upon the displacement that are i: troduced by means of the method of Lagrangian multipliers. A modified functional FLU; 2) = fu; 2] + Ally involving the Lagrangian multiplier variable y is used instead of ¢, where now U is in the space of state variables w and v, and the linear operator A[u] is designed to have the property that the assertion Alu] dv = 0 implies the constraining relations. Again, with F'8U = fu; 2) du + vA'[u] du + Alu] dv = 0 as the governing variational statement of the problem, it follows that all of the preceding results remain valid with the replacement of ® by F, and wu by U. Example 1—Euler Column Although the post-buckling behavior of an Euler column has long been understood on the basis of the exact elastica solution, this problem usefully provides an elementary, yet nontrivial example of the application of the general theory. The usual uniform-column theory based on inextensional bending deformation gives the potential energy L L $= 4EI[ (dOfds)? ds — lt = J, cos 0 as}, (4.57) to where (see Fig. 12) 0 is the rotation of the column cross section, s is distance to the cross section measured along the centroidal axis, E/ is the bending stiffness, and P is the vertical load. By inextensionality, s and the total curved length L are invariant under deformation. Buckling Behavior of Elastic Structures 31 oe °f ® Fic. 12. Simply-supported Euler column. Successive Frechet differentiation of (4.57) gives L L ¢[0]0, = EIf (d0/ds)(d0,/ds) ds — P [ 0, sin 0 ds, 0 “0 L L $°[0}0,0 = ery (d0s/ds)(a0,/ds) ds — P [0,05 cos 0 ds, 0 0 Dy $°[0]0,0203 = P| 0,00 sin O ds, “0 L #[0]01 020302 = P{ 0,020,0, cos 0 ds. (4.58) to The prebuckling solution is clearly 0o(s) = 0, and the bifurcation equation (4.8) becomes $"[0]0, 50 = Er J * a0, fs) 5(d0/ds) ds — P, fo, 50 ds = 0" (4,59) 0 0 wherein 2 has been identified as P. For a simply supported column, any continuous 60 is admissible, and so, (4.59) implies El(d?0,/ds*) + P.0, = 0 and the natural boundary conditions d0,/ds = 0 at s=0,L. The lowest eigenvalue is the Euler critical load P, = WEI/L, and, with the use of the inner-product norm ||0|| = <0, 0>"/? based on L = (2/L) I, af ds we have 0, = cos(zs/L), where {|0,|| = 1. 32 Bernard Budiansky Since $/ = $”[0] = 0, the post-buckling coefficient 2, given by (4.21) vanishes, as it obviously should for the symmetrically bifurcating Euler column. From (4.58), L gO} = Pf Of ds = APL to and HO; L 0% ds = —4L, 0 Consequently, Eq. (4.22') gives 42/2. = 4 and so it is verified that the column is not imperfection sensitive. (This example is somewhat unusual in that the calculation of 2; did not require a determination of v3 .) During the buckling deformation 0 = & cos(ns/L) + PIP, © | + 30, (4.60) but a possibly more useful representation for P/P, would be in terms of the amplitude of displacement rather than rotation. Since the lateral deflection w is related to 0 by dw/ds = sin0 = 0 +--+, w = (EL/n)sin(as/L) + + and introduction into (4.60) of 7 = €L/z as the amplitude of the initial buckling displacement gives PIP, = 1 + (§n?)\(n/Ly. (4.61) Moderately large displacements are thus produced by small excursions beyond the initial load. The initial load-shortening relation, by (4.55), is now immediately found as PIP. = 1 + (A/2L), (4.62) where L A = ap/0P =L—[ cos Ods. *o The results of this analysis agree with those found from an expansion about P, of the classical exact solution in terms of elliptic integrals. Example 2—Inextensional Ring The uniform thin ring shown in Fig. 13 is presumed to be loaded by Pressure q that remains normal to the surface. Inextensionality requires that Buckling Behavior of Elastic Structures 33 Fic. 13. Thin ring, the nondimensional normal and tangential displacements shown in the figure satisfy doe dv\? (dw : v4 bea l(v+ +(E-9) =0. (4.63) The rotation @ of the ring cross section satisfies sin 0 = —(dw/dz) + v, and the curvature is therefore do dv dw) ieee a ee ee ee The strain energy is $E (3*(d0/ds)? ds, and the potential energy of the pres- sure is —q(AA), where AA is the reduction of enclosed area given exactly by _ AA = 4R? f [—2 — w? — v? + 2v(dw/dz)] dx. (4.65) 0 Consequently, the potential-energy functional F, modified to include a Lagrangian multiplier function v(x) to incorporate the constraining relation (4.63), is F = aialae. where de _ dul fav al _ del dx de do? ~ da fd ieee a [» oa tai tl dx =f" vf es Hw 7 Jo Fat? da, +} da : ‘ da, (4.66) 34 Bernard Budiansky where 2 = qR3/EI. With U representing the state variables w, v, and y, the variational equation Q’ 6U = 0 is satisfied in the fundamental state by Yo=h, Wo = =0 {as well as by rigid-body displacements, which may be suppressed). The bifurcation “oe OU, bu = 0 is now a {|e fw) 5 _ &w dw) dw, i, { cael Fep—v1 Ol 7 a ay,dw + ae . fr i [vt | iv =, afv + *) = (», + | a dz =0 (4.67) Standard operations of the calculus of variations then give, with vy, the differential equations dw, dbo Pw, dv 1 oy, 2 1 i] =0 on dex: dx Puy Poy dey. fader, Pos) _ aes eae a ge lee] wp + 2t=0. (4.68) dx The solution (again, ignoring rigid-body motion that can occur at any load) for the lowest buckling load is = 3 (4.69) with the mode wy, = cos 20, v, = —$sin 20, vy = 3. cos 20. (4.70) It is clear that the bifurcation is symmetric, and 2, = 0. For the purpose of calculating U, by means of Eq. (4.23'), the quantity $%’ U? SU is needed, and is found from (4.65) to be a QU? bu = a for(i + =) af» + *) + (» + a al da -f. “fo “( a) oft ~e) + (Gt-o) ol a (4.71) Buckling Behavior of Elastic Structures 35 wherein the first integral vanishes, by (4.70). The differential equations that result from (4.71) become yw, ,@w. dv, do, ae ta Te 7 ag 7 12 = 9608 40, Bry | dw, | dv. , dv, _ 9 dat Oa te tae = g8in dv, _ 9 “Wa = 6 (1 — cos 40), (4.72) and the solution for U, is w2 = — (9/16), v2 = (9/64)sin 40, vy = —(27/16)cos 40. (4.73) The terms in Eq. (4.22') for 2, can now be calculated; Q/’U} U, is ob- tained from (4.71), with 6U = U,; OF UF is = fd, wy)? div, \? nf” (2-2) (4 - Ba and Q2U? is 7 Csieg ao "4 7 I, [20.22 +» +f (Ft—o) | de The result for 2,//, works out to Azfd, = 27/32. (4.74) The generalized shortening A associated with the nondimensional func- tion (4.65) is related to the actual decrease in area (AA) by A = AA/R?. Accordingly, formula (4.55) gives the load-“ shortening” relation Afh, = 1 + (9/16)(AA/A), (4.75) where A = xR? is the original area enclosed by the ring. This result, together with Bd, = 1+ Q27/32)F2 +--+, w= E(cos 20) ~ (9/16)22 +--+, v = —(€/2)(sin 20) + ((9/64)sin 40)2? + and the equality 2, = 3, sum up the buckling and initial post-buckling beha- vior of the ring under normal hydrostatic pressure. (For moderate values of ¢, the results agree with those found by Carrier’s (1947) numerical evalua- tion of his exact, but implicit, analytical solution.) 36 Bernard Budiansky Similar studies can be made for rings under other kinds of normal loading, such as dead loading, constant centrally directed load, and centrally directed loading governed by an inverse-square law. Plates were analyzed by this theory (Koiter, 1945) and an early applica- tion to a shell was executed (Koiter, 1956). G. StapiLity As in the study of the simple model, the general analysis is incomplete without some indication that imperfection-sensitive structures enjoy stable equilibrium states up to the snapping load; the stability of imperfection- insensitive structures should also be assessed. As discussed earlier in Section II, a proper minimum in the potential energy implies stability of finite-dimensional systems, In turn, this minimum exists if the second varia- tion $”(6u)? is positive, in the sense that $”(6u)? > 0 for all admissible du # 057 for, if ¢’ du = 0 for the system in equilibrium, [u + du] — o[u] = 36’(5u)? + R, (4.76) where, in finite-dimensional systems, the remainder R is dominated by $"(6uy? for sufficiently small 6u; therefore Ad > 0 in some neighborhood of us; hence a positive second variation ensures stability. Similarly, if 6" (6u)? is negative for any particular éu, A¢[e du] < 0 for sufficiently small ¢, @ cannot have a minimum, and [with the help ofa little damping (Section 11)], instabi- lity is implied. Unfortunately, this simple reliance on the sign of the second variation as a stability criterion is not easy to justify rigorously in the study of continuous systems (Koiter, 1963a, 1965b, 1966). While a non-negative second variation remains necessary for stability, a general proof of sufficiency is not available. Part of the trouble (related to distinctions between strong and weak minima in the calculus of variations) is that in continua the remainder R in Eq. (4.76) may no longer be doniinated by a positive second variation during every possible approach to zero of du. Further, mutually contradictory decisions concerning stability can be reached on the basis of various norms adopted, in the definition of stability, for the sizes of disturbances and the responses they produce. (In finite-dimensional systems, ail such norms are equivalent to each other.) It appears necessary to study each continuous system separately, with physically appropriate, as well as analytically congenial, definitions of stability introduced in each case. The outcome of several such studies (Koiter, 1965a, 1967) has been that the second variation remains a + If there exists a constant ® > 0 such that $°(5u)? > Sul}? for all Su # 0, then o"(5u)? is said to be positive definite with respect to the norm | Positive and positive definite are equivalent in finite dimensional systems, but not necessarily in continua. Buckling Behavior of Elastic Structures 37 sensible indicator of stability, With the viewpoint that this will probably continue to be true in general, the present general analysis of stability will be restricted to an examination of the second variation. Consider first the equilibrium states along the fundamental path. If stabi- lity is presumed for sufficiently low 2, it must follow that any transition to instability as 2 increases cannot occur below 2 = 2,. For below the transi- tion load, say, 7, 4 w? > 0 for allay, and above J, $f w? < 0 for some w’s; we will assume that at 2 = 7, jw? = 0 vanishes for at least one w = iv,f re- mains non-negative for all other w, and, therefore. as a function of 1, is stationary at w = ih; thus the eigenvalue equation % i dw = 0 must be satisfied at 2 = 7; and therefore, since 2 is the lowest possible cigenvalue, 2 = },. Furthermore, if , is unique, ii is proportional to 1. Since $% 3 is positive for 2 < 2,, and out = Oat 2 = 2,, it follows that bu? < 0. However, the special case $n? = 0 will not be considered ex- Plicitly, although a separate treatment can be executed without essential difficulty. (It is worth noting that for fu} <0, bifurcation at J, implies instability on the fundamental path for 2 > 2,, and conversely, such insta- bility implies the bifurcation path discovered in the earlier analysis. On the other hand, if $707 = 0 instability need not necessarily occur for 2 > 4, and there may not even be a real bifurcation cither.) With this case dis- regarded, the earlier presumption that $71? < 0 is thus justified on the basis of stability considerations. Consider now the equilibrium paths for imperfection-sensitive structures that are traversed as 2 increases from zero up to the snapping load 4, (Fig. 9a,c). Along such a path we can write u = to(2) + 4, 2), (4.77) where now 4(, 2), defined for 2 < 4,() is the single-valued branch of 5 in Eq. (4.30) that satisfies lim uf, 2) = 0. (4.78) é-0 Along the considered path G"(5u)? = $"[uy + 45 ASul? + YW" [uo + v1, tH; 2](5u)? (4.79) which, expanded about u = uo and i = 0, provides # (Gu)? = GslSuy? + Hou)? + ABBUP(SuP? + > + Botidduy? + (4.80) We will not worry about pathological cases in which no such i exists, although an example can be contrived. As a simple example, if, for continuous w, o"w? = f§ (1 — Ax)? dx, we have $"»* positive for 2< 1, and indefinite for 4 > 1, but there is no i # 0 for which Jo (1 — x)? dx vanishes. Such situations are unlikely to arise in well-formulated problems of mechanics. 38 Bernard Budiansky Consequently, for sufficiently small €, ¢”(6u)? must have the same sign as o(5u)?, which is positive for 2 < 7,. Thus, stability up to snapping has been established, for small enoygh imperfections. In the case of an imperfection-insensitive structure (Fig. 9b,d) the same conclusion clearly applies for 4 < 2,. For 2 in some range 2’ > 2 > 2,, the expression (4.77) continues to apply, but the limit in (4.78) is replaced by lim v2) = o(2), (481) where v is the displacement increment along the bifurcated path originally defined in Eq. (4.3). The appropriate expansion of (4.79) for 2 > 2, is now about u = uo + v and ti = 0, and this gives 8" (6u)? = "Lt + e(2);AM(5u)? + 8" [tg + (2); Aer = v)(Suy? + + + EW'*[tug + v, 0; Adu)? + + = 0. (4.82) Accordingly, for 2 > ., $"(6u)? will share the sign of 6"[tio + v3 2](5u)? for sufficiently small €. This is twice the second variation along the bifurcated path of the perfect structure, the sign of which will be discovered by means of a search for the minimum value of "Uo + 0; A](Su)?/A(Su)?, (4.83) where A(éu)? = du, du) = [|u|]? and Au? = 1, with respect to an admis- sible (5u). If this minimum valuc is positive, then the numerator is positive for all du. With the assumption that A has been so chosen that the mini- mum, denoted by f, actually exists for 6u = w, the consequent requirement that (4.83) be stationary with respect to variations in w implies that wand are solutions of the eigenvalue problem o'w dw — PAw dw = 0 (4.84) on the bifurcated path. For 2 = ,, v = 0, and the solution is clearly ~ = 0 and w = w,. Hence, for 7 > 2,, the expansions wen tarp + Fwy te, B= oh +OBr+s+ (485) appear appropriate, where, without loss of generality, <1,,,> =0 for n= 2. Recalling that r = Gu, + €u, + +++, expanding (4.84) about 2 = A, and r = 0, and writing Aww dw w, dw), gives UPewe + Ahly + Pluz — B, Aus} dw + Gens + Ay ows + Arey + VI Guy + Pum, + Pouy we + Ay blud + $pitud — By Aw, — Br Ay} dw + = T In fact, $"(5u)? is then positive-definite with respect to || ||- Ps Buckling Behavior of Elastic Structures 39 Setting div = u, in the coefficient of € gives Br = Ar dey + Gen, where use has been made of the fact that Au} = = ||, ||? = 1. But since @/u} = —2A, bf u? [Eq. (4.21)] it follows that Br =~ ben. Since A= A, + 2,€ + + on the bifurcated path, it follows from (4.85) that P= —(2—A)bUF+00-—2) (n= 2) (4.87) and therefore, since fu? < 0, f is positive for some range of 2 > 2,, which verifies stability on the bifurcated path, as well as on the associated imperfect-structure path, for sufficiently small | @ |. If 4, = 0, then f, = 0, and it follows from (4.86) that dev, Sw = — Gu? Sw, (4.88) and comparison with Eq. (4.23’) then shows that _ (4.89) From the vanishing of the coefficient of € in (4.86) with éw = u,, and with the use of Eq. (4.22’), the result Br = - 220, bit (4.90) is found, so that B= -22-2)¢02 +002-—2)" (nz. (4.91) Again, f > 0 for sufficiently small 2 — 4, > 0, and stability follows for the bifurcated path of the perfect structure, as well as for the imperfect structure with sufficiently small imperfections. The special case 2 = 2, has so far been exempt from the analysis, but is easily handled separately. The minimum of [lo + 145 AJ(Su)? _ Glo + 145 A] (Ou)? + YW Tuo + 115 Ae(Ou? A(su)? - A(Sul is now sought. For 4, + 0, Eq. (4.40) shows that @ = O(22) for = 2,,and, from (4.36), ty = uy + €2v) +--+. Series of the form (4.85) may again be chosen for the minimum f and the corresponding minimizing du = w, and then (4.92) Bw dw — Bw dw =0 expands into ews + Peuz — Py Ary} Sw + E4Gewwy + Gelo2u, + Guy w, + 468} — By Awe — Boy} dw + O(@) ++ = 0. (4.93) 40 Bernard Budiansky The choice 6u = u,, now gives By = Goa} = — 2A, bout and so. B= (—25ud)Ag + Ol) — (n= 2). (494) Hence, at 2 = 2, stability holds, for sufficiently small € (or @), if 2, € > 0. If 2, = 0, Eq. (4.45) shows that € = O(é*) at 2 = 2,, and we have f, by = ty; then (4.93) leads to We = uy and Bo = 302 upu, + $pout = —322 ben}. Consequently, B = (—32, fu3)e? + O(2") (n= 7 at 2=2,, and since 2, > 0, we have stability for sufficiently small € > 0. To sum up, for sufficiently small imperfections, stability, as assessed by the sign of the second variation, has been verified along imperfect-structure equilibrium paths emanating from 2 = 0, up to the snapping load J, in imperfection-sensitive structures; and, up to some load higher than A, in imperfection-insensitive structures. [Similar conditions hold for the imperfection-insensitive perfect structure, except that the second variation is inadequate to determine stability at 2 = 2,. A study of higher-order varia- tions (Koiter, 1945) indicates the same results as for the simple model: instability for 2, + 0, stability for 2, = 0, 2, > 0.] It is now easy to show that along the post-snapping branches of paths (Fig. 10a,c) where 2 is decreasing from 4,, instability must hold, although this fact is not of too much practical interest. It is only necessary to contemplate u = tio() + tulZ, 2), where t is the branch of B that, for 2 < 7, < 4,, approaches the perfect- structure bifurcated solution as — 0. The analysis following Eq. (4.81) is then applicable, the only change being that 2 is less than 2, instead of being, greater. The conclusions from (4.87) and (4.91) are consequently reversed, and instability is deduced for sufficiently small @ and (2 — 2.) < 0. A few final remarks on stability: It is tempting (and common) simply to look at a curve of load versus a deflection and conclude that there is stability if the load is “rising,” instability otherwise. But this can have pitfalls. Ob- viously, an undetected bifurcation load on a rising curve will invalidate the conclusion at loads above the critical load. This explains why the calcula- tions just made concerning stability are limited to small € and small neigh- borhoods of A,. Furthermore, equilibrium on a “falling” 2-é curve can be Buckling Behavior of Elastic Structures al stable; again, a bifurcation on such a curve could conceivably cleanse the contemplated falling path of its instability. There is, however, at least one pictorial stability theorem of some generality that may be worth mentioning. It is always true that a negative slope on a load-shortening curve can only be associated with an unstable equilibrium state; that is, d2/dA <0 implies instability under prescribed loading. To prove this, let ¢[u; 4] be the poten- tial energy of the system, perfect or not, and note that along any (smooth) equilibrium path u(A), $'[u(2); 2](du/d2) = 0, by equilibrium; then d du dg’ du a ales) ~ at oR a. But (?u/d22) = 0, again by equilibrium, and so (dp'fa?)(du/dd) = 0. (4.95) Next, since A[u] = ae 4/62 does not depend explicitly on A, at 1 -(%) du 06" du a ad _(e 5) ay d dd} dd. and so, with (4.95), dAfdd = "(dufddy. (4.96) This shows that if dA/dd (or d2/dA) is negative, so is $”(6u) for du = du/d?, which proves instability. V. Stress, Strain, Displacement; Virtual-Work Approach A. StRESS-STRAIN RELATION; VIRTUAL-Work EQUATION Recall the Eqs. (4.48) and (4.49) concerning the form of the potential energy of applied loads, and write the total potential energy functional as = 6[e] — Ad[u] (5.1) in terms of the shortening A[u], and of the strain energy 6, now considered to be a functional of the generalized strains e. The strains, in turn are given bya strain-displacement function e[u] in terms of the displacements. The varia- tional equation of equilibrium is now bu = Se] de — AA'u] du = 0, (5.2) 42 Bernard Budiansky where 6e = e[u] Ou. (5.3) If we introduce the notation o =6{e), (5.4) Eq. (5.2) becomes o dc — Au] bu = 0. (5.5) This has the form of a principle of virtual work, stating that, in an equili- brium state, the change of the potential energy of the loads associated with “virtual” displacements du must equal the internal virtual work of stress ¢ acting through strains de that are compatible, via (5.3), with the displace- ments du. [We have thus endowed the symbol ¢ with three meanings; it is the stress state a; it is a linear operator on strains, making ¢ de the total work of the stresses acting through dc; and it is a function o[e] of strain consistent with (5.4). In earlier work, the notation @ - de has been used for the inner product representing internal virtual work, but it has become clear that the dot is really not needed. The appropriate interpretation for the symbol ¢ will always be clear from the context in which it is used, and, with good will on the part of the reader, the fluency afforded by the notation will soon be apparent.] We can, if we wish, assert Eq. (5.5) as the variational equation of equilibrium ab initio, without reference to the total potential energy; indeed, in plasticity problems, there may be no relations of the form (5.4), and no potential energy functional. However, in the present elastic case, the link to energy is desirable, in order to derive and exploit certain symmetry proper- ties embedded in the stress-strain relations. Thus, since 6[«] de, = ofe] de, it follows that 6"[e] be, ez = a'[e] de dey, and since &” is a symmetric bilinear operator, so must o’ be. That is, O'b 18, = G62, (5.6) for all €,, €2. The field equations governing the problem of finding a, ¢, and u may now be collected as o bc — 2A’ Su =0 (equilibrium), (5.7a) o=ol[e] (stress-strain), (5.7b) e = 2, gives oye, bu + opel Su — AAtu, du = 0, (5.13) 44 Bernard Budiansky where the subscript c denotes evaluation at 2 = 7, on the fundamental path. The stress-strain relation (5.7b) may be expanded as oy + 5 = ole +] = 40 + oon + 40gy? +7 (5.14) and so cancellation of oo, division by ||v||, and the approach to 2, gives oy = 068). (5.15) Similarly, the strain-displacement relation (5.7c) yields by = ebay. (5.16) Equations (5.13), (5.15), and (5.16) constitute an eigenvalue problem for 7, and the eigenfunctions 1,, 6, and &. It seems like a good idea now to check the consistency of these governing equations with Eq. (4.8) for the same eigenvalue problem in terms of ¢. We have ole) = Efele fu] - AAT] = ofele'[u] — 24°): Dropping explicit indication of argument of functions, taking another Frechet derivative, and evaluating it at 2 = 2, gives 2 = a2 (60) + Geel — DAL, Then td, Ou = heey e uF opeluy Su + Aly du. But since e,u, =e, by (5.16), and ajc, =o, by (5.15), we get that Guu, u = 0 is precisely Eq. (5.13). To study the bifurcated path, expand Egs. (5.8) by writing veu—u(d) = oy + Cm te, = 0 — 66(A) = fo, + a, + (5.17) nS —b0(4) = Se + Pepto, and substitute into Eq. (5.14), as well as into a similarly expanded version of (5.7c). Further, introduce the expansions O82) = of + (2 = 2) +4 = 20)'3P + (0) = e+ (= 2,)AP + A= 2.) + AIP) = AM + (2 = 2)AO + A= 2 FAM Fo where the dots represent total derivatives with respect to /, as in Eqs. (4.14). Finally, introduce the familiar relation 2 = 2, + 2,€ 4+ 2,@? +++, and then the following results are found by systematic collection of like powers of €: (5.18) G1 = 028, Fp = Oley + AyOLey + Fave} (5.19) Buckling Behavior of Elastic Structures 45 and by =, by = Oy + Apel + dele. (5.20) A similar process of substituting into the expansion (5.12) of the remain- ing field equation (5.7a) gives a fairly complicated expression that will not be written out in its full glory, but pieces, as needed and occasionally simplified, will be extracted. The terms of order € in this expression simply reproduce the buckling equation (5.13). The terms of order é? give the result afo28 + o.0ft, — AAluy + 6; 00, + 40,800? — YA FAG cety + ott, + 0,8 ~ Alu, — 2, Agu,}} du = 0. (5.21) Set du = u, and note that the first three terms of (5.21) may then be reduced to O28, — 6,00 by exploiting Eq. (5.13), with du = u,. In turn, with the help of (5.19) and (5.20), 0, and u, may be eliminated to transform this to Ay(oy Gu — Gpe}) + Ao, elu} + So%e2. Consequently, the coefficient 2, can be found as _ Ba yee + ocef'u} + ofc} — 2rd oa (5.22) A= where D = {o.0lu} + oeilud + 2oyéiuy + Gee? — Ate? — 2, Atul}. (5.23) Since none of the functions o[¢], 2[u], and A[u] depend explicitly on 2, we.can write og = o¢é., é2 = efi, and AY = Az'i,. Hence D can be expressed in the alternative form D = G25 + acl udit, + 2a eltty te + ofe2é, — Mud — Audie . (5.24) Once again, as an exercise in manipulation, let us verify that (5.22) agrees with the earlier result given by Eq. (4.21) in terms of }. We have, keeping the notation concise, and using dummy arguments 2, v2, ..-, $y = ae'v, — 2A'v,, "0,02 = Ge", 02 + 0'e'v, £02 — 2A"D, 02, $0, V2 Vs = G8"v, Vyv5 + O'e'UZe"Uy My (5.25) +o’ £"b205 + o'e'v28"0, 05 + o"e've'v, — JA" vy 0703, so that bout = aqelu} + Boleiuy elu? + offer) — AA" 46 Bernard Budiansky With ec. uy = €,, this is the numerator of (5.22). A similar calculation shows that #21} =D, and so the desired check is accomplished. In fact, it now hardly seems necessary to continue a direct systematic derivation of addi- tional results in terms of o, ¢, and w, but rather obtain them by transforma- tion of those already found by the energy approach. Thus, if 2, = 0, the result (4.22’) for 2, turns to be equivalent to Dy = -VDQoyeluyuy + o2eud + 3a, etd + oleh es iy + oeetuty + hove} + doe ut (5.26) =A ACU, — 2, diut. In deriving this result, use was made of the field equations governing a2, £2, and uz when 2, = 0, which, from (5.19) and (5.20), are GO, = O66, +4ored, by = eh, + 4etut (5.27) and, from (5.21), O26, OU + Feely OU — 2_Agy Su + oy ecuy du + docu? du — $2, At"u? 6u=0, — (5.28) where, as before, 1) is rendered unique by making = 0. But it is expected that the variational equation (5.28) will rarely be used to set up the problem, preference usually being given to a direct perturbation expansion, via (5.17), of the governing local equations (usually nonlinear differential equations). It may be remarked that the relation Aju, = 0 found in (4.52) has the consequence, from (5.10), that Ge EMy = -€, = 0, which means that the stresses in the structure at bifurcation do no work through the initial strains of the buckling mode. C. Speciat Ca: The results found are very general, far more so than one would usually require, and in various particular problems one or more of the following simplifications can be invoked: (i) linear stress-strain relations: o” = 0 for n > 2; (ii) quadratic strain-displacement relations: e” = 0 for n > 3; iii) linear shortening-displacement relations: A” = 0 for n > 2. g-Gisp: Buckling Behavior of Elastic Structures 47 In very many problems all three of these conditions are met. Limitation to elastic strains makes (i) valid in all but rubberlike materials. A quadratic strain-displacement (ii) relation is, of course, obtained when the Lagrangian strain tensor, or a simplified variant thereof, is employed. (On the other hand, if a nonlinear curvature-displacement relation is invoked, as in the ting problem, it may be cubic.) The linearity of Afu] occurs in dead-loading situations (though not in hydrostatic loading). In any event, with (i)-(iii) valid we get D = Gud = Gelut + 2o elu, in, , (5.29) and the results for 2, and 2 simplify considerably to Jy = 40, eluG[(o-60u3 + 20,02, ite) (5.22') and, if 2, = 0, Jy = —Qoyetuyug + o26tuZ)(Gclut + 2oyeluyil,). 5.26") Also, 62 = 05 €2, € = 6 uz + 4elu?, and the variational equation (5.28) col- lapses to 026, SU + Getty Ou + oeetuy Su = 0. (5.28") The buckling equation (5.13) also simplifies slightly. These results agree with those given by Fitch (1968). Another simplifying assumption, independent of (i)-(iii) above, is that ofa (iv) linear fundamental state. This assumption provides the common situation associated with buckling problems in which the prebuckling displacement, stresses, and strains vary linearly with 2, and the load 7 appears linearly in the bifurcation eigenvalue problem. We will define assumption (iv) to mean a special kind of linearity associated with the functions o[e], e[u], and A[u], and the fundamental solu- tions t%, Gp, and ég, in which not only do the identities olkco] = Kole}. efKtto] = Keltig}, Alu] = Kua}, (5.30) hold, for any constant &, but, in addition, the operators o$”, , AY? are all independent of 2 for n > 1.7 This guarantees that the fundamental solution of the ficld equations (5.7) will be linear in 7, and it also makes (5.13) a linear eigenvalue problem, Furthermore, with 2 = 69 = A® = 0 forn > 1, and the relation , = (I/2,)o,, D in (5.23) reduces to D = (I/A,Jocetud — A, Atul] = —(1/2.)o 101 (531) ¥This means that even though ¢[x] may be nonlinear, eftig + t] = [tg] + e[e] for all c, and similarly for o[e] and Alu], For example, ife, = (ée/éx) + 4(éw/ex)?, and the fundamental state is ty = 2X, Wo = 0, this condition is satisfied. 48 Bernard Budiansky where the last equality follows from Eq. (5.13). The numerators of (5.22) and (5.26) remain unchanged under just assumption (iv), but if all the assump- tions (i)-(iv) can be invoked D= $i = (1/A.)oce"u, (5.32) and the results for 4, and 2, become Ayla, = —40, cB /o-ctu = Ao. e2/o ey (5.22") and, for 2, = 0, Igfdg = —2o ety My + o2eUF/o,e0u} = 2oy ely ty + 0, 80u2/o ey. (5.26") The results are essentially those given by Budiansky (1965) and Budiansky and Amazigo (1968). All of the formulas of the preceding section concerning the generalized shortening and the post-buckling stiffness [Eqs. (4.54)-(4.56)] are im- mediately transformable to the notation ore the present section simply by replacing du? by D. Note that ¢.= = —Ati= —(I//,)o, &,, by (5.10). If the assumptions (i)-(iii) apply, then Se (4.56) and (5.29) give ise (1- Be Goch + esta) (5.33) Ko 2h Geb. Under assumption (iv) [and not necessarily any of (i)-(iii)] $= —(Ui.Jo.8 = — (INA? )oe8e » and since we also have (Eq. (5.32)] fu? = —(1/2.)o18; for this case, we get K_ 4, \ 018, ]7 Ke [+ (35) 2 A, oe8guy — A Age a w (1 — fe sei = ed) (2) 9 If all the assumptions (i)-(iv) hold, K 2, o,ctu\ ~? ao (ase) | a Example 3—Flat plate Consider a square, simply supported, flat, plate of uniform thickness com- pressed normally along all edges as shown in Fig. 14, by means of rigid, Buckling Behavior of Elastic Structures 49 lubricated loading heads. The problem will be treated on the basis of the familiar von Karman equations eraw otra fy a (SESW BE A DVtw = ( ay? oe? to a (5.36) é? . a _ VIF + Et i a = (5.37) where D? = E15/12(1 — v?), in terms of the normal displacement w and the Airy function F giving stress resultants N, = 0?F/0)?, N, = ?F/0x?, Ny, = —@F/éx éy. The generalized stresses of this problem are the bending y APY ” =] : a x APY Fic. 14. Simply-supported fiat plate. moments M,, M,, and M,, and the stress resultants N,, N,, and N,,; the displacements are u, v, and w; the generalized strain state consists of the mean in-plane strains ¢, , ¢,, and 7,, and the curvatures K,, K,,and K,,. The stress-strain relation c[u] is represented by éu 1 fdw\? 7 x ales) : év , 1féw\? 7 ee 3) ; K.= (5.38) ou ov éGwéw aw ay tat oar Ke = 7 agaye 50 Bernard Budiansky The stress-strain relation o[e] is My= DIK, + °K) 0, = [EM — v)lley + v6) M,=D[Ky+ 1K.) 0, =[E/(1~vJ]fe, + ve, (5.39) My =D(l~v)Ky, ty = [E/2(1 + ry and the virtual-work statement (5.7a), in the present context, is OF OF OF t ai =z 08, — ——— 5y J [i Sta + aE Os — Bay Bian + M, 6K, + M, 6K, + 2Myy ox.| dx dy = [external virtual work] (5.40) wherein the relations between stress resultants and stress function have been introduced. The differential equations (5.36) and (5.37) are consistent with Egs. (5.38)-(5.40) which are the realizations, for plates, of Eqs. (5.7). This shows that the various results of the general theory can be used. In the special plate problem under consideration, the boundary conditions of simple support are w= 2wax?=0 at x=0,a, w= dway?=0 at y=0,a, (5.41) - and the conditions of zero shear and uniform normal displacement along the edges are guaranteed by @F/ox dy = F/ax3 = at x=0,4, ‘PF/ox dy = OFF =0 at y=Oa (5.42) The prescribed force conditions require that (6F/6y)(0, a) — (@F/ay)(0, 0) = —AP,/a, (6F/éx)(a, 0) — (6F/ax)(0, 0) = —AP,/a. (5.43) The expansions for w and in the general theory are now represented adequately by [oles eetepes ean where the fundamental solution is simply Wo =0, Fo = —(2/2a)[P,y? + P,x?]. (5.45) Buckling Behavior of Elastic Structures st Assuming that P, and P, are fixed in magnitude, we look for the critical value of the scalar multiplier 2. The bifurcation equations, most casily found by substituting (5.44) into (5.36) and (5.37), and retaining terms of order ¢, are just DV‘, + (Ae/a)[P(@w/0x?) + P,(w4/03?)] = 0, VIF, = 0. (5.46) Since Fo satisfies the boundary conditions on F, those on F; become hom- ogeneous and the solution is w, = tsin(xx/a)sin(xy/a), FF, = 0,4 2, = 4n?Dja(P, + Py), (5.47) where we have made (Wy)max The bifurcation is clearly symmetric, so 4, = 0. To calculate 2,, we will need w,, F,, and by collecting terms of order €? in (5.36) and (5.37) we get oe Pe Vis + 7 [P. 2 +p, =, ; i ew, 8 2)? yp ow, Fwy (wy ' . VtF, + Et (ee By? ( xd 3) | . (5.48) Using w, as found in (5.47) gives WF, = (n*E8/2a*)[cos(2nx/a) + cos(2xy/a)] and the solution to (5.48), for any orthogonality condition = 0, is w2=0, — F, =(Et?/32)[cos(2nx/a) + cos(2ny/a)]. (5.49) To calculate 2,, we note that all of the simplifying assumptions (i)-iv) discussed earlier are satisfied, and we can use Eqs. (5.26"). Note that e”1, Uy is, from (5.38), just a set of three in-plane strains given by aw, dw, dw, dw, w, Ew dy oy ox dy ~ dy Ox’ corresponding to ¢,, €,, and 7,,. Since w, = 0, we need only concern our- selves with 2 eu? and o, gu? in (5.26"), and by (5.40) these become @F,\ (aw,\?_ (62F2) (aw,\? AR) (3) m) |! > and o2euy = ff _ mE 32% TA possible inconsequential linear contribution to F, is ignored. 52 Bernard Budiansky (since °F ,/dx dy = 0) and 2 éwy\? ee ; accent = 2 ff [r.( my + a no aps ON ag = - 7) |: 120 — v)a? The result for 43/2, = —360 13/0, e212 is simply dafd, = 31 — ¥?) (5.50) independent of the ratio of P, and P,. Note that wya. © (Et) for small é, so that we have the result Whe 1 + BL = ¥7)(Wnalt)? +o (5.51) Finally, since oot = [A2(1 — v*)ar/E][PE + P2 — 20P,P,) Eq. (5.35) gives K_ (FE Reo('ae ae aee) “) For uniaxial loading, P, = 0, and K/Ky = 4, independent of y. For P, = P, and v = $, K/Ko = 4; this shows that, for balanced biaxial compression, the stiffness associated with change in area of the plate is reduced by a factor of 4 after buckling. D. SHALLOW SHELLS; DONNELL-MUSHTARI-VLASOV SHELLS In shallow-shell theory, the basic equations (5.7) assume the forms, in tensor notation, with respect to generalized coordinates €', €?: J[ Wr? 5Kyp +N 3B,5] dS = [external virtual work], (5.53a) s 2p Et oop Se Mit = Tq all ~ VIR + vKG0"4 a __&t ARO N# = [1 — VE" + VEL), (5.53b) Exp = [Us,p + Up.a] + bag W +3WL Wy, Ky = —Way- (5.53c) Here M** and N** are stress couples and stress resultants; U, and W are the surface and normal displacements; E,, and K,, are stretching and bending Buckling Behavior of Elastic Structures 33 strains; byp = ag iS an appropriate curvature tensor, where z is the normal distance to a reference plane; and the metric tensor g’? is calculated in the flat space of the reference plane. In Donnell-Mushtari-Vlasov theory, the same equations are used, but by, and g,y are curvature and metric tensors computed on the surface. [See Sanders (1963) for further details.] The expansions (5.17) may be written as ©) a ey U; UA), -PUs| oo |Us « nl of tehaf te yal tos (5.54) Ww) VW, Ww. Sy ap nr ) _ _ nv) _ fare fuel) wre) wre ape, ( |, iE (2) E,(2. Ey 22) Ey tee (at Bel aE] on gs Kay) Kup Kyp, where now 2 is a scalar multiplier of a prescribed external edge and surface loading distribution. For dead loading, conditions (i)-(iii) are satisfied, but not necessarily (iv), and the applicable formulas (5.22') and (5.26’), give ry a) 3 NOW, W. dS Ae 3 ee (5.57) 2 IPLAC Wp + NW We, as and, for 24 = 0, ree 2 @ wm a 1 2N*W Wop + NOW Wg] dS a | a (5.58) SP INPW Wp + 2NIV Wey] dS Also, Eq. (5.33) gives . i) mM a K A, Sf [NEW . Wp + INP W,, Wey] dS|~ >= - a 7 (5.59) Ko 22; JP NPE, dS Here the dot, as usual, means total derivative with respect to 2 along the fundamental path, and the index c—placed wherever convenient—denotes evaluation at 2 Similarly, if (iv) is also satisfied, formulas (5.22”), (5.26"), and (5.35) gi Tee NOW W5 dS ow (5.60) JP NPY IW, g ds 54 Bernard Budiansky and, for 2, = 0, we Q) Ray a) } QNW Wg + NOW. IV. g] dS = —JERNTW Wag NOW a ol 5 (5.61) SIs NOW. ,1V,» ds wy a) c 1 WwW. K _ [2 SENPW. Wp ds (502) Ko |) 73h, NPE, aS In the absence of tangential surface loads, the actual solution of particular problems is usually most easily executed by exploitation of the Airy stress function F giving the stress resultants as NY = COOK, (5.63) where €*° is the alternative tensor. The displacement U, may then be elim- inated from explicit consideration and the governing differential equations for W and F become the Karman-Donnell equations: DVEW + 82 Ogg Foy + P= BOPP oy Wesg (ENUF = 2b Wo, = — 42°27 Wap Woo, (5.64) where p is the normal surface loading. Expansion of F into a series like (5.55) thus leads to the equations [assuming (iv) does not necessarily hold] qa (yy a a) rr—~—~—~—~—~—~—~—~—~—~—~—”r”r—C=C“ EE a a ay (YEQVS F = 8° !byg Wo, + E°P™Wag(2e)W 0 = 0, (5.65) a) a for the eigenvalue 2, and the buckling mode HW, F.1f 7, = 0, the equations 42) for 1, and F; [which provide N’* via (5.63)] are Q) @) @) Q) DVAW + 2°S "bag Foy — NPQ), op — 2PPTE oy Wag w = SF Wap 2) Q Q) (EQVAF = &°° bag Woy + OPW ag 2)IV, oy ay a) CO (0) or It is implicit here that if N**(2) and W,,,(4) are nonlinear in 2, they can nevertheless be found without too much difficulty from a particular solution of the full nonlinear equations (5.64). Typically, this occurs in problems of Buckling Behavior of Elastic Structures 55 shells of revolution, in which a nonlinear, prebuckling axisymmetric state precedes nonaxisymmetric bifurcation. If assumption (iv) holds, and the fundamental state is linear, then Eqs. (5.65) and (5.66) are simplified to the extent that terms in $¥*(2,) drop out, and N?(2_) becomes 7, Nz, where Nz is independent of 2.. The initial post-buckling behavior of many shells has been studied on the basis of this theory. Examples involving linear prebuckling states are con- tained in Budiansky and Amazigo (1968), Budiansky (1969), Hutchinson (19672, 1968), Danielson (1969), Hutchinson and Amazigo (1967), and Hutchinson and Frauenthal (1969), problems in which nonlinear prebuckling states are taken into consideration are contained in Fitch (1968), Hutchin- son and Amazigo (1967), Hutchinson and Frauenthal (1969), Fitch and Budiansky (1970), and Budiansky and Hutchinson (1972). Imperfection sen- sitivity is discovered in most of these shell problems. A much earlier post- buckling shell analysis was made by Koiter (1956) on the basis of his general formulation (Koiter, 1945) in terms of displacements. The alternative stress— strain-displacement approach appears to make shell calculations somewhat easier. E. INITIAL IMPERFECTIONS The results of Eqs. (4.43) and (4.46) for the effects of initial imperfections remain valid, but now some insight into the magnitude of the parameter p in these formulas will be obtained. It is appropriate, in many problems, to assert that in the presence of an initial imperfection 1 = éii, the strain— displacement relations are Gx e{u +H] — efi] (5.67) giving the strain Z of the imperfect structure in terms of the additional displacement and the old function ¢[u] of the perfect structure. Similarly, the generalized shortening is modified to A= Alu + 0] — Ala. (5.68) The stress-strain relation is ¢ = ofé], and so the Frechet derivative with respect to u of the modified potential energy ¢ is hao tw =olelele + W]— 7A ii). Since @' = ofé|e'[u] — 2A‘[u], where ¢ = e{u], this identifies ’ du as W' du = oféle'[u + i] ou — ofele'[u] ou — 2A'[u + ti] + 2Afu). Taking the Frechet derivative with respect to i then gives wt du = o'fe)fe[u +H) — e[ulfiete + i] ou + oféle’[u + ti du — 24H du, (5.69) 56 Bernard Budiansky and evaluation at u = u,, t= 0, 6u = 1, gives a —™rmr™—é—s—s—OSO_OSCisC_seés|.# = ole, — Ee + o-ctim, — A Atiny, (5.70) where é = e'[u] |,-9. But now note that, by (5.13) oyegit + oat th — AAU, = if 6u = it, which is an acceptable choice for du if fi is assumed to be an admissible displacement shape. Hence (5.70) reduces to Fim, = 08h and, from (4.39), we find p= ~oytii/2, $end. (5.71) With ¢2ui = D given by (5.23) or (5.24) we can compute p directly in terms of the generalized stress, strain, and displacement quantities of the present development, Results equivalent to (5.71) were derived by Fitch (1968) under assumptions (i)-(iii), which are not necessary for its validity. But now a substantial simplification results if we have a linear prebuckling state [assumption (iv)], and take i = u,—that is, we consider the influence of an imperfection in the shape of the buckling mode. For then, by (5.31), 2. beu} = —,0,, and, since &=0, we have #u, = chu, =. Con- sequently, p = 1! Accordingly, we have the remarkable result that, quite generally for structures having linear fundamental paths and imperfections in the shape of the buckling mode, the results for imperfection sensitivity given by Eqs. (4.43) and (4.46) become precisely those found for the simple model in Eqs. (2.9) and (2.10). Let us now return to (5.69) and calculate ett du for a linear prebuckling state, with it = 1). Setting u = ug, 1 = 0 gives Wott, Su = oi — Eu, £5 du + Goehu, du ~ IASu, du, and since, for a linear prebuckling state, cp = ot, # =e) =e, Go = (A/2-)o., Ag = AZ, this reduces to Witu, du = —(2/2,)o ye du (5.72 with the help of (5.13). But note also that, from (5.25), eu, Ou = Geely Su — Alay Su = (1/A )oceguy du — A, Agi, du] = —(1/2)o, 65 du. (5.73) Buckling Behavior of Elastic Structures 57 Consequently, Stuy du = Abluy, du (5.74) and the equation of equilibrium (4.32) for the imperfect structure can be expanded into iB OU + (2 — Ae)GED Our to + AGED? Su to + Aght? bu to + AEGlu, Su+ += 0, — (5.75) where now u = (2/2,)u. + ¥. It is tempting to execute an approximate solu- tion of (5.75) for a relation among 2, v, and é that is uniformly valid for both 2 and (2 — 2.) small. Note that for € and 2 small, linearization with respect to band € gives GED bu + (2 — A)GLE du + AEhlu, du = 0, and it is evident that the solution for b is v = €u,, where = f(r. — 2). (5.76) So, for Ay # 0, use = Eu, in a Galerkin solution of (5.75), setting du = 1, and keeping terms of order é, €?, and 2. This gives G.- E+E = 6.77) which, with p = 1, agrees with (4.41) for 2 near 2, ,and with (5.76) for 2, and hence €, small. One might therefore expect that a calculation of 7, on the basis of (5.77) would be more reliable than one based on (4.41). The maxi- mization of 2 by use of (5.77) leads to [= AJP + AAr/)(A/A JE = 0. (5.78) This is in asymptotic agreement with (4.43) for p = | and sufficiently small ind, furthermore, provides the mathematically palatable result 7, > 0 for €— 00, as opposed to the obvious breakdown of (4.43) [and (2.9)] for finite values of @. Nevertheless, it must be remarked that the superiority of (5.78) over (2.9) cannot be established rigorously without a study of the effects of higher-order terms missing from (5.77). If 2,=0, an approximate Galerkin solution of (5.75) based on Eu, + Eu, gives (2. — AE + LG = 28, (5.79) where terms of order higher than & and € are dropped. The calculation for 2, now gives, for 2, < 0, (1 = se)? — 3 3(—Aa/de)(4/2e) 2 | = 0. (5.80) Again, this result has a more reliable look about it than Eq. (2.10), with which it agrees asymptotically, but. no proof is available. 58 Bernard Budiansky VI. Mode Interaction A. SIMULTANEOUS BUCKLING Mopes; PosT-BUCKLING ANALYSIS Suppose that the solution of the eigenvalue problem (4.8) for bifurcation _ yields several linearly independent eigenmodes u{?, u'?, ..., uS*, all asso- ciated with the lowest eigenvalue 2,. Without loss of generality, these modes canbe orthogonalized to each other in the form puPuP = for ix} (6.1) Assume, in generalization of the earlier condition $%u? <0, that PAu) < 0 for each i. It will, of course, also be true that bul? = 0 (6.2) for all i, j, since the bifurcation equation #1 du = 0 must hold for each i. It is appropriate now to expand the displacement along a particular post- buckling path into x 1 Hold) + EY yu? + Buy to, (63) fet where =0 for nz 2, (6.4) - and the y, are fixed numbers, satisfying Y}_, v? = 1, that describe the rela- tive initial contributions of the various modes. It continues to be convenient to normalize each u{) according to [etl = may (but need not) be chosen to coincide with —$fg. The magnitudes of the v, cannot be chosen arbitrarily; as will be shown, they must conform to tequirements imposed by equilibrium. Thus, replacing 1, in Eq, (4.19) by 38, y,1u4? gives _ x ET Chur + AGED vjul + ral x oe) jeu 71 int +OC-}du+ =0. (6.6) Setting du successively equal to u{? (i = 1, 2,..., N) in the coefficient of € Buckling Behavior of Elastic Structures 59 provides the simultaneous nonlinear equations voN DY Vaiaryye = AA (6.7) Gat k= where iy. = SPC ual (68) ~ and Ap= —be(uPY. (6.9) These equations must be solved for the v/s (subject to the constraint > v? = 1) and the associated 2,. There may be many essentially different sets of solutions (in addition to trivial duplications obtained by changing the signs of 2, and the ¥/s), and they would correspond to a variety of possible starting behaviors of different bifurcated paths. It is easily shown that unless all the a,,’s vanish there is:at least one solution with 2, # 0. Ifa; = 0 for each i, there are no uncoupled solutions corresponding to nonzero ,, but unless all the other a’s vanish too, there will nevertheless be one or more asymmetric bifurcations, involving two or more modes, with attendant imperfection sensitivity. This kind of mode interaction occurs in approximate solutions for buckling of cylinders under axial compression (Koiter, 1945) and of spheres under external pressure (Hutchinson, 1967b) in both of which problems many simultancous buck- ling modes are available. These structures are notorious for their extreme imperfection sensitivity, having snapping loads that, in practice, are usually less than half the bifurcation load. In the case of a nonzero 2,, the generalized shortening that occurs during a coupled-mode bifurcation can still be computed on the basis of Eq. (4.53), except that i, is, again, replaced by Y. v;1{?. Since $2(¥ yu)? = —Y A,v, this gives (6.10) and since 2 — 7, = 2,é we get de % 1 + (Ay/A_)[2(A — Ao/¥, Ayv7]'? (sign 2¢) (6.11) as the proper generalization of Eq, (4.54) for the single-mode case, If the assumptions (i)-(iv) of the previous section hold, it may be conven- ient to formulate Eqs. (6.7) in terms of stress, strain, and displacement. We TA theorem of Bezout is quoted in Sewell (1968) and Johns and Chilver (1971) to theeffect that if (6.7) is not degenerate (with infinitely many solutions) there are at most 2° — 1 essen- tially different real solutions, and at least one solution. 60 Bernard Budiansky have [see Eq. (5.25)] uw = o'e'ue"vw + o'e've” my + oe'we"uv so that, with ¢u{? = ef, ace" = of, the coupling coefficient (6.8) is Gi = bin + ay + Oni (6.12) where bye = ofall, (6.13) Also, Eq. (6.9) for A; gives [see (5.31)] Ay = —(UA)ocec(u? — AAUP?) = (1/A.JoPe?. (6.14) In addition to (or in the absence of) asymmetric bifurcations, there may also exist symmetric bifurcations involving several simultaneous modes. The expansion (6.3) would remain appropriate, but then the admissible combina- tions of v,’s would have to be determined from equations governing u, and 4, that follow from Eq. (6.6). So far, the need for such calculations has not been apparent in any specific problems of practical interest, and this procedure will not be pursued here. [See Johns and Chilver (1971) and Sewell (1970) for some recent studies for finite-dimensional systems] B. INITIAL IMPERFECTIONS A thoroughgoing discussion of the effects of initial imperfections in the case of simultaneous bifurcation modes would involve the exploration of a variety of imperfection shapes that provoke responses from the separate coupled modes in different ways. Indeed, the possibility exists that certain special imperfections would not lead to limit point snapping, but may in- stead induce bifurcations at loads less than the critical load of the perfect structure (Koiter, 1963b). It must be expected, however, that this situation, while it may sometimes be attractive analytically, would be exceptional, and limit point snapping would generally occur. Furthermore, as a demonstra- tion of imperfection sensitivity, it will suffice here to study the simplified situation contemplated in the previous section in which assumption (iv) ofa linear prebuckling state is met. Ii this case, let us contemplate the effect of an initial imperfection in the shape of the initial coupled bifurcation of the perfect structure. It then becomes apparent that the situation is indis- tinguishable from the single-mode case if only the bifurcation mode 1, is replaced by }' v,u{?. That is, if we choose the imperfection as (6.15) Buckling Behavior of Elastic Structures 61 the old asymptotic result (2.9), namely Ald 1 ZA"? (44% <0) remains valid. Furthermore, the modified, possibly better, equation (5.78) is also applicable. CC. NeaRLy SIMULTANEOUS Moves It may happen that the eigenvalue problem for bifurcation has several solutions u{? corresponding to critical loads 2 (i = 1, 2,..., N) that, while not coincident, are close to each other. This occurs in the sphere under external pressure (Koiter, 1969) [when simplifications rendering the modes coincident (Hutchinson, 1967b) are not made]. Much interest has also arisen in two-mode problems associated with nearly coincident eigenvalues asso- ciated with local plate buckling and overall column or panel buckling (van der Neut, 1968; Koiter and Kuiken, 1971; Tvergaard, 1972). Clearly, a single-mode analysis for post-buckling behavior and imperfection sensitivity based on the lowest critical load would not generally provide the coupled- mode results in the limit as these loads do become coincident, A reasonable way to handle the problem is to execute a Galerkin solution based on the use of the approximation u = tio(4) + S31 &u?, obtaining algebraic equations governing the és, 2, and initial imperfections. Thus we demand that thine Seis ‘| du + W'[u, 7; 2] du = 0 (6.16) A for du = uf) (j = 1,2,..., N). Seti fi, and expand (6.16) as follows: Po Su + $5 Y Eul? du + AGO(Y Eu)? Su +--+ + WO, 0; 2] du + Edgtir du + --- = 0. (6.17) The first terms of each line in (6.17) vanish, and the omitted terms are of order €}, 2, and &. Now ghoul = bul) + (2 — A) buy + Of — 29), (6.18) where or = Polos PL Be = (MAYES br-a,- Similarly HGuPuP = GetiPuP + O(2 — 2) 62 Bernard Budiausky for any m, and so (6.17) becomes Da~ we : tee cu ou +5 YY emg i= j= x Pent Pup du + + Eysindu+---=0 (6.19) where the vanishing terms (7,1) du have already been dropped and the omitted terms are of order g2)¢. (2— ARE, (AMES, (2 — 22), and It is now best to drop back to the special case of a linear fundamental state, and write Me ar] cu? (6.20) thus restricting the imperfection to a linear combination of the buckling modes. Note that by (5.25) 0 Pegu? du — ALP du + aes du, (621) ty) Ou = and prebuckling linearity gives Hou? Su = (1/2 oct? Su — Azul? Su, (6.22) where no indices appear on eZ, AZ, and ¢ since they do not depend on 4, Since cit uP = PZ uPuP = 0 and gD = jg, it follows from (6.21) that oe = 0 (6.23) and (OAM eLUUP — AruUD = hulu? = 0. (6.24) Equations (6.23) and (6.24) supply various forms of the orthogonality rela- tions that must hold among bifurcation modes associated with different critical loads when the prebuckling state is linear. Now by (5.74) and (6.20): . Syn Su = AYE, htu? du. (6.25) i=t Hence, putting du = uf, and choosing m=k gives the simultaneous equations ae Y= Aaaes DY i=1 jet +N), (6.26) Buckling Behavior of Elastic Structures 63 where 4, = —#%(u\»)? is independent of 2, and Gin = bau PuPup. (6.27) [If the additional conditions o’ = ¢ = A” = 0 are met—and they will be if (i)-(ii) hold—¢¢ becomes independent of 2, and Gi; is the same as the a; given by (6.12).] It is now a matter of solving (6.26), perhaps numerically, for equilibrium paths relating 2 to the ,’s for various assumed @,’s. This setup is uniformly valid for vanishingly small differences among the 20, giving results for snap- ping loads that approach those implied by the analysis for coincident loads. It may sometimes be desirable to depend on a somewhat improved set of equations that incorporate individual u contributions associated with some of the modes. To this end, the following procedure is suggested. Calcu- late uv) for a given mode, treating the problem as in the previous sections for isolated bifurcation. Then use the expression w= u)+ > x Eu? + Y ePuY i= in a Galerkin solution of ¢' du + y’ du = 0, making approximations similar to those in the above analysis, but keeping terms of order ¢,¢,¢,. Further details will not be explored here. ACKNOWLEDGMENTS This work was supported in part by the Air Force Office of Scientific Research under Grant No. AFOSR-73-2476, in part by the National Aeronautics and Space Administration under Grant No. NGL 22-007-012. and by the Division of Engineering and Applicd Physics, Harvard University. Helpful comments from J. W. Hutchinson and W. T. ‘Koiter are gratefully acknowledged. REFERENCES Bupiansky, B. (1965). Dynamic buckling of elastic structures: criteria and estimates. Proc. Int, Conf. Dynamic Stability of Structures, Northwestern Uniy., Evanston, linois, pp. 83-106. Bupiansky, B. (1969). Postbuckling behavior of cylinders in torsion. Proc, IUTAM Symp. Theory Thin Shells, Copenhagen, 2nd pp. 212-233, Bupiansky, B, and Awazico, J. C. (1968). Initial post-buckling behavior of cylindrical shells under external pressure. J. Math. Phys, 47, 223-23: Bupiaxsky, B, and Hurcntxsox, J. W. (1964). Dynamic buckling of imperiection-sensitive structures, Proc. Int. Congr. Appl. Mech., Munich, X1, pp. 636-651 Bupiaxsky, B,, and Hutcuixsos, J. W. (1972). Buckling of circular cylindrical shells under axial compression. In“ Contributions to the Theory of Aircraft Structures” (van der Neut Anniversary Volume). pp. 239-260, Delft Univ. Press, Dellt. ‘Carrier, G. F. (1947), On the buckling of elastic rings. J. Math. Phys. (Cambridge, Mass.) 26 94-103. 64 Bernard Budiansky Cones, G. A. (1968). Effect of a nonlinear prebuckling state on the postbuckling behavior and imperfection-sensitivity of elastic structures. AIAA J., 6, 1616-1620. See also Cohen, G. A. (1969). ALAA J. 7, 1407-1408. DANIELSON, D. A. (1969). Buckling and initial postbuckling behavior of spheroidal shells under Pressure. AIAA J. 7, 936-944, Firctt, J. R. (1968). The buckling and postbuckling behavior of spherical caps under con centrated load. Int. J. Solids Struct. 4, 421-446, Firctt, J. R., and Buptaxsky, B. (1970). The buckling and postbuckling of spherical caps under axisymmetric load. AIAA J. 8, 686-692, Hurctxson, J. W. (1967a). Initial postbuckling behavior of toroidal shell segments. Int. J. Solids Struct. 3, 97-115. Hurcixsox, J. W. (1967b). Imperfection sensi J. Appl. Mech. 34, 49-55. Hurcutxson, J. W. (1968). Buckling and initial postbuckling behavior of oval cylindrical shells under axial compression. J. Appl. Mech. 38, 66-72. Hurcuixsos, J. W., and Astazico, J. C. (1967). Imperfection sensitivity of eccentrically stiffened cylindrical shells. ALAA J. 8, 392-401. Hurcinxson, J. W., and FRAUENTHAL, J. C. (1969). Elastic postbuckling behavior of stiffened and barreled cylindrical shells. J. Appl, Mech. 36, 784-790. Hurctuxsox, J. W., and Korrer, W. T. (1970). Postbuckling theory. appl. Mech. Rev. 23, 1353-1366. Jouxs, K. C.. and Cryer, A. H. (1971). Multiple path generation at coincident branching points. Int, J. Mech. Sci. 13, 899-910. Jovquer, Emite (1930). Amortisement des oscillations et stabilité séculaire. Proc. Int. Congr. Appl. Mech., 3rd Stockholm, Vol. 3, pp. 111-1 Korrer, W. T. (1945). On the Stability of Elastic Equilibrium (in Dutch). Thesis, Delft Uni H. J. Paris, Amsterdam; English transl, (4) NASA TT-F10, 833 (1967), (b) AFFDL-TR-70- 25 (1970). Korrer, W. T. (1956). Buckling and postbuckling behavior of a cylindrical panel under axial compression. NLR Rep. No. $476. Rep. Trans, Nat, Acro. Res. Inst., Vol. 20. Nat. Aero. Res. Inst., Amsterdam, Korrer, W. T. (1963a) On the concept of stability of equilibrium for continuous bodies. Proc. Kon. Ned. Akad. Wetensch. B66, 173-177 (1963a). Kotter, W. T. (1963b). The effect of axisymmetric imperfections on the buckling of cylindrical shells under axial compression. Proc. Kon, Ned. Akad. Wetensch. B66, 265. Korter, W. T. (1965a). On the instability of equilibrium in the absence of a minimum of the potential energy. Proc. Kon. Ned. Akad. Wetensch. B68, 107-113. Korter, W. T. (1965b). The energy criterion of stability for continuous elastic bodies. Proc. Kon. Ned. Akad. Wetensch, B68, 178-202. Korrer, W. T. (1966). Purpose and achievements of research in elastic stabili Conf, Soc. Eng. Sci, 41h, North Carolina State Univ., Raleigh, N.C. Korrer, W. T. (1967). A sufficient condition for the stability of shallow shells. Proc. Kon. Akad. Wetensch. B70, 367-375. Korrtr, W. T. (1969). The nonlinear buckling problem of a complete spherical shells under uniform external pressure: I, II, II] and FV. Proc. Kon. Ned. Akad. Wetensch. B72. 40-123, Korrrr, W. T. (1972). “Stijfheid en sterke 1-Grondslagen.” Scheltema & Holkema, Haarlem. Korrer, W, T., and Kuiken, G. D. C. (1971). The interaction between local buckling and overall buckling on the behavior of built-up columns, WIHD 23, Laboratorium voor Technische Mechanica, Dellt. ity of externally pressurized spherical shells, y. Proc. Tech. Buckling Behavior of Elastic Structures 65 Liustersik, L. A, and Sosotev, V, J. (1961). “Elements of Functional Analysis. Ungar, New York. Masur, E. F. (1973) Buckling of shells—general introduction and review. ASCE National Structural Engineering Meeting, 1973, San Francisco, California, Preprint 2000. RaLt, Louts B. (1969). “Computational solution of nonlinear operator equations.” Wiley, New York. Retssxer, Eric (1953). On a variational theorem for finite elastic deformations. J. Math. Phys. (Cambridge, Mass.) 32, Nos. 2-3, 129-135, Sanpers, J. L. (1963). Nonlinear Theories for Thin Shells, Quart. Appl. Math. 21, 21-36. Sewe, P. (1972). A reexamination of Koiter’s theory of initial postbuckling behavior and imperfection sensitivity of structures. Symp. Thin Shell Structures. California Institute of Technology, June 1972, Iu “Thin-Shell Structures” (Y. C. Fung and E. E. Scchler, eds.) Prentice-Hall, Englewood Cliffs, New Jersey, 1974. SeweLt, M. J. (1968). A general theory of equilibrium paths through critical points, 1, Il. Proc. Roy. Soc. A 306, 201-223, 225-238, SeweLt, M. J. (1970), On the branching of equilibrium paths. Proc. Roy. Soc. A 315, 499-517. Tuomrson, J. M. T. (1969). A general theory for the equilibrium and stability of discrete conservative systems, Z, Math. Phys, 20, 797-846. See also J. M. T. THOMPSON AND G. W. Hunr, (1973). “A General Theory of Elastic Stability.” Wiley, New York. TVERGAARD, V1GGo (1972). Influence of post-buckling behavior on optimum design of stiffened panels, Rep. No. 35. Danish Center for Applied Mathematics, Technical University of Denmark. VainnerG, M. M. (1964). “Variational Methods for the Study of Nonlinear Operators.” Holden-Day, San Francisco, California, VAN DER NEUT, A, (1968). The interaction of local buckling and column failure of thin-walled compression members. Proc. Int. Congr. Appl. Mech. 12th, Stanford Unit, 1968. rederich

S-ar putea să vă placă și