Notes on Stationarity and extremum principles in mechanics (with applications to optimal design) Pauli Pedersen Department of Mechanical Engineering, Solid Mechanics Technical University of Denmark Nils Koppels Allè, Building 404, DK-2800 Kgs.Lyngby, Denmark email: [email protected] WORKING PRINT December 9, 2008 c ii Pauli Pedersen: Stationarity and extremum principles in mechanics Stationarity and extremum principles in mechanics c Copyright 2008 by Pauli Pedersen, ISBN 06 Preface The energy principles in mechanics play an important role, not only for analysis but also for design synthesis and optimization. However, when teaching mechanics it is mostly found difficult to communicate a basic understanding of these principles. What is the reason for this situation that so many teachers agree with? Should the reason be related to the students, to the teachers, or to the available textbooks? The present small book attempts to give an alternative nontraditionally presentation of the subject. The presentation in chapters 2 - 6 has earlier been used in a course on elasticity, anisotropy and laminates. A primary idea is to separate the mathematical derivation of an identity from the specific interpretations of this identity. Then also separate the stationarity principles from the extremum principles, and finally balance the physical interpretation of the non-physical variations, where also the aspect of infinitesimal variations is important. Hopefully, this alternative presentation will appeal to some readers. The chapters 7 - 9 with direct relation to optimal design use to a large extend the basic principles in mechanics, but also introduces new results from design variations, i.e., the sensitivity analysis for design. These chapters are influences by recently published papers, and here serve as examples to illustrate the simplicities that may results from using the basic principles in mechanics. Kgs. Lyngby, Winter 2008 Pauli Pedersen iii c iv Pauli Pedersen: Stationarity and extremum principles in mechanics Contents Preface iii Contents v 1 Introduction 1.1 Background . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 Layout of contents . . . . . . . . . . . . . . . . . . . . . . . . . . 1 1 1 2 Stationarity principles in mechanics 2.1 The work equation, an identity . . . . . . . . 2.2 Symbols and definitions . . . . . . . . . . . . 2.3 Real stress field and real displacement field . 2.4 Real stress field and virtual displacement field 2.5 Virtual stress field and real displacement field 2.6 Summing up . . . . . . . . . . . . . . . . . . . . . . . . 3 3 4 6 7 8 9 . . . . . . . . . . . . . 11 11 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14 14 16 . . . . . . . . 17 17 18 18 19 . . . . . . . . . . . . . . . . . . . 21 21 . . . . . . . . . . . . . . . . . . . 22 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24 26 3 Extremum principles in mechanics 3.1 Principle of minimum total potential energy 3.2 Principle of minimum total complementary(stress) potential energy . . 3.3 Overview of principles and their relations . 3.4 Summing up . . . . . . . . . . . . . . . . . 4 Potential relations and derivatives 4.1 Equilibrium . . . . . . . . . . . . 4.2 Relations with power law elasticity 4.3 Derivatives of elastic potentials . . 4.4 Summing up . . . . . . . . . . . . 5 Energy densities in matrix notation 5.1 Strain and stress energy densities 5.2 Energy densities in 1D non-linear elasticity . . . . . 5.3 Energy densities in 2D and 3D non-linear elasticity . 5.4 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . v c vi Pauli Pedersen: Stationarity and extremum principles in mechanics 6 Elastic energy in beam models 6.1 Elastic energy in a straight beam . . . . . . 6.2 Results for simple (Bernoulli-Euler) beams 6.3 Beam solutions by stress(complementary) principles . . . . . . . . . . . . . . . . . . 6.4 Summing up . . . . . . . . . . . . . . . . . 7 Some necessary conditions for optimality 7.1 Non-constrained problems . . . . . . . . . 7.2 Problems with a single constraint . . . . . . 7.3 Size optimization for stiffness and strength . 7.3.1 Size design with optimal stiffness . 7.3.2 Size design with optimal strength . 7.4 Shape optimization for stiffness and strength 7.4.1 Shape design with optimal stiffness 7.4.2 Shape design with optimal strength 7.5 Conditions with a simple shape parametrization . . . . . . . . 7.5.1 Possible iterative procedure . . . . 7.6 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27 27 29 . . . . . . . . . . . . . . . . . . . . . . . . . . 30 31 . . . . . . . . . . . . . . . . 33 33 34 34 34 35 35 36 37 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 38 40 41 . . . . . . . . . . . . . . . . 8 Analytical beam design 8.1 Optimality criterion for beam design . . . . . . . 8.1.1 Treated boundary conditions and loads . 8.1.2 Solutions in general . . . . . . . . . . . 8.2 Bernoulli-Euler cantilever beams . . . . . . . . . 8.2.1 Optimal compliances . . . . . . . . . . . 8.3 Timoshenko cantilever beams . . . . . . . . . . . 8.3.1 Design of beams with n = 1 cross sections 8.4 Examples of beam cross sections . . . . . . . . . 8.5 Summing up . . . . . . . . . . . . . . . . . . . . 9 The ultimate optimal material 9.1 The individual constitutive parameters 9.2 Sensitivity analysis . . . . . . . . . . 9.3 Final optimization . . . . . . . . . . . 9.4 Numerical aspects and comparison with isotropic material . . . . . . . . 9.5 Summing up . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 43 44 46 47 48 49 49 50 54 54 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57 57 57 58 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 59 59 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References 61 Index 63 Chapter 1 Introduction 1.1 Background The authors background for the present notes are years of teaching energy principles in a course on elasticity, based on (Pedersen 1998a). As stated in the preface it is mostly found difficult to communicate a basic understanding of these principles. However, when succeeding, the students has an important tool for many future applications. With a research background in optimal design, the unification that can be obtained using energy principles is found important, especially when the design sensitivity analysis is build on these principles. The obtained results are then valid for 1D, 2D and 3D problems, for anisotropy as well as for isotropy, often for non-linear as well as non linear elasticity, and for analytically research as well as for numerical research, see (Pedersen 1998b). With the present notes, the notions in earlier notes and papers are unified. Limiting these notes to mostly analytically derivations without numerically results, the total number of pages are few, but hopefully not too dry. 1.2 Layout of contents In (Pedersen 1998a) the present chapters 2 and 3 are combined in a single chapter, but it is found advantageously to separate the stationarity principles in chapter 2 with focus on the principle of virtual work. This principle is of most importance and is based on only a few assumptions, but with a rather abstract interpretation and therefore not immediately easy to communicate. Chapter 3 contains a graphical overview of the stationarity as well as the extremum principles. The extremum principles are mainly used as arguments for choosing approximations based on a stationarity principle. The main assumption behind chapter 4 is a proportionality relation between complementary energy (stress energy) and strain energy. From this follows proportionality relations between total potentials, strain energy, stress energy, external potential, and compliance. The names of stress energy and complementary energy are used synonymously, but in the notations only the superscript σ is applied, not the superscript C as in (Pedersen 1998a). Focus is thus put on the fact that complementary principles have stress/force as the primary variables, while strain principles has strain/displacement as the primary variable and superscript is applied for their related quantities. General simple first order design sensitivities is a result of derivatives on the 1 c 2 Pauli Pedersen: Stationarity and extremum principles in mechanics energy level. These results are obtained in fields of fixed strains or fixed stresses, and these results do not involve approximations. Chapter 4 is earlier published as an appendix in (Pedersen 2003). To be specific about the proportionality relation between strain energy and stress energy, chapter 5 presents a power law elasticity model, that is the most simple extension from linear elasticity. For this reversible model secant stiffnesses as well as tangent stiffnesses are derived for a three dimensional model, covering also anisotropic materials. Although being a one dimensional model, the beam is among the most important structural elements. It is therefore natural to include as an example in chapter 6, the use of energy principles in a linear elastic straight beam. These handbook formulas are applied in chapter 8 for design optimization. Before this design optimization a number of necessary conditions for design optimization are needed. The description in chapter 7 also focuses on the use of superelliptic description of shapes in two dimensional models, a description which has often been successfully applied. Chapter 8 contains results from recent research, see (Pedersen and Pedersen 2008), and demonstrates that optimal design based directly on an energy approach has promising aspects. It has been possible to find analytically described optimal beam designs, even for short beams where the Timoshenko beam theory must be used. The notes finish with chapter 9 on ultimate optimal material, related to compliance optimization,. i.e., again for an energy objective. The results in the reference research (Bendsøe, Guedes, Haber, Pedersen and Taylor 1994) are extended to be valid for non linear elastic materials as in (Pedersen 1998a). In active research on optimal design, this free material model is expected to play a major role. In the present notes this is taken as a further example on the importance of energy approaches, added design sensitivity analysis. Chapter 2 Stationarity principles in mechanics Energy principles play a central role in mechanics, but surprisingly few books treat the subject in a structured way. It is difficult to get an overview of the many different principles, and important questions are not presented, especially in relation to the necessary conditions for a certain principle. The present chapter is written as an alternative to the classical presentations, as by (Langhaar 1962) and (Washizu 1975). We shall show that all principles are specific interpretations of the same identity, and necessary conditions will not be introduced until absolutely needed. We shall refer only to a Cartesian 3-D coordinate system, and traditional tensor notation for summation and differentiation is applied. 2.1 Goal of the chapter Tensor notation The work equation, an identity Without physical interpretation of the quantities involved, we shall derive an important identity. The superscripts a and b are only part of the names (not powers) and their use will be explained later. From 1 b a b a b b a 1 b b σij ij = σij ij − (vi,j + vj,i ) + σij (v + vj,i ) (2.1) 2 2 i,j No physical interpretation and (2.2) a b a b a a b a b σij vj,i = σji vi,j = (σji − σij )vi,j + σij vi,j follows Z Z V V a b σij ij dV − Z V a σij bij 1 a a b (σ − σij )vi,j dV − 2 ji Z 1 b b − (vi,j + vj,i ) dV − 2 V (2.3) a b σij vi,j dV = 0 In (2.3) V is the volume of the domain of interest. Now let A be the surface that bounds this domain and nj the outward normal at a point of the surface. Then, using the theorem of divergence, the last part of (2.3) is rewritten to Z Z Z Z a b a b a a b a σij vi,j dV = (σij vi ),j − σij,j vib dV = σij vi nj dA − σij,j vib dV V V A V (2.4) 3 Using the theorem of divergence c 4 Pauli Pedersen: Stationarity and extremum principles in mechanics The identity Once more, by adding and subtracting the same quantities, we obtain the identity from which the stationarity principles of mechanics can be read without further calculations Z Z 1 b b a a b b σij ij − (vi,j + vj,i ) dV − σij ij dV − 2 V Z ZV Z 1 a a a b a b Tia vib dA+ (σji − σij )vi,j dV − (σij − Ti )vi dA − 2 A ZV Z A a a b a b (σij,j + pi )vi dV − pi vi dV = 0 (2.5) V V With the superscripts a and b we have indicated certain relations, and we shall now assume that all quantities with index a are related and that all quantities with index b are related. However, no relations exist between quantities with different index, and the equations therefore still have no physical interpretations. Let bij = bij (x) be a strain field derived from a displacement field vib = vib (x) with small strain assumption (engineering strains, linear strains, Cauchy strains) 1 b b bij = (vi,j + vj,i ) (small strain assumption) 2 (2.6) a = σ a (x) be a stress field with moment and force equilibrium with the and let σij ij field of volume forces pai = pai (x) and surface traction’s Tia = Tia (x) a a σij = σji a σij,j = −pai (2.7) a σij nj = Tia The work equation With (2.6) and (2.7) the identity (2.5) reduces to Z Z Z a b σij ij dV = Tia vib dA + pai vib dV V A (2.8) V which is often called the work equation, although it is merely an identity and does not express physical work when the a and b fields are not related. 2.2 Real fields Virtual fields Symbols and definitions For a better overview, the stationarity principles of mechanics are divided into four groups covering the possible combinations of real fields (indexed by a superscript 0) and virtual fields (without index). The virtual fields are assumed to be sufficiently differentiable and admissible, but otherwise arbitrary and non-physical. An admissible displacement field must be kinematically admissible, i.e. it must satisfy the boundary conditions. An admissible stress field must be statically admissible, i.e., it must satisfy force equilibrium. A somewhat repeated definition of the quantities in the energy principles may be useful before the individual principles are stated and proved, all by specific interpretations of the work equation (2.8). Geometry variables: V = Volume of the continuum or structure A= Surface area that bounds the volume V (2.9) Stationarity principles 5 State variables at position x in the volume: ij = ij (x) = description of strain state at position x σij = σij (x) = description of stress state at position x Ti = Ti (x) = surface traction = force (per unit area) in direction i at position x of A pi = pi (x) = force (per unit volume) in direction i at position x in V (2.10) Work and energy quantities in strains and displacements: W = W (vi ) = Z Z A vi Ti (ṽi )dṽi dA + Z Z V 0 u = u (ij ) = U = Z Z Work of external loads vi pi (ṽi )dṽi dV 0 Strain energy density ij σij (˜ ij )d˜ ij 0 Total strain energy u dV V (2.11) Π = U − W Complementary work of external loads Complementary work and energy quantities in stresses and forces: Z Z pi Z Z Ti σ σ vi (p̃i )dp̃i dV vi (T̃i )dT̃i dA + W = W (Ti , pi ) = A V 0 uσ = uσ (σij ) = Uσ = Z Z 0 Stress energy density σij ij (σ̃ij )dσ̃ij 0 Total stress energy uσ dV V (2.12) Πσ = U σ − W σ Φ = W (dead loads, i.e.,W σ = 0) = Ti vi dA + A Z pi vi dV V Compliance from elastic energy Φ = U + Uσ = W + Wσ Φ = −Uext Total potential complementary energy Compliance of external loads Compliance relations: Z Total potential energy (2.13) Compliance from potential of external loads c 6 Pauli Pedersen: Stationarity and extremum principles in mechanics 2.3 Real stress field and real displacement field 0 As a first example of use of the work equation (2.8) we insert the real stress field σ ij 0 0 0 in equilibrium with the real load field Ti , pi , and the real displacement field vi from which the real strain field 0ij is derived. We get Z Z Z 0 0 0 0 p0i vi0 dV (2.14) Ti vi dA + σij ij dV = V A V For arbitrary constitutive relations we have Z Z 0 Z σ 0 0 ij ij ij 0 0 σij (˜ ij )d˜ ij + d(σ̃ij ˜ij ) = σij ij = 0 0 0 σij ij (σ̃ij )dσ̃ij (2.15) 0 which, together with the definitions in (2.11) - (2.12), gives 0 0 σij ij = u0 + uσ0 Z Zero sum of total potentials V 0 0 σij ij dV = U 0 + U σ0 Analogously for the external loads we get Z Z 0 0 Ti vi dA + p0i vi0 dV = W 0 + W σ0 A (2.16) (2.17) V and (2.14) can thus be written U 0 + U σ0 − (W 0 + W σ0 ) = Π0 + Πσ0 = 0 (2.18) i.e., the sum of the real total potentials is zero. Especially for a linear elastic material the definitions in (2.11) - (2.12) give Z 1 U 0 = U σ0 = σ 0 0 dV (2.19) 2 V ij ij In relation to the nature of the external forces, the concept of dead load is important. For dead loads the forces are independent of the displacement of their point of action, say a gravity load. For a dead load we get no complementary(stress) work, W σ0 = 0, and the work is thus Z Z 0 0 0 W = Ti vi dA + p0i vi0 dV A assuming W Clapeyron’s theorem V σ0 = 0 (dead load) (2.20) For a system with both linear elasticity and dead loads (2.14) gives 1 U 0 = W 0 2 (2.21) and thus External potential and compliance 1 Π0 := U 0 − W 0 = − W 0 = −U 0 (2.22) 2 which is often called Clapeyron’s theorem for linear elasticity. The ”missing” energy W 0 /2 is assumed to be dissipated before the static equilibrium with which we are concerned. Note, that if we by definition take −W 0 as given by (2.20) to be the external potential Uext , then the assumption of dead load is not necessary, but then again external potential is hardly a physical quantity. The quantity W 0 as given by (2.20) is also named the compliance, i.e., Φ as defined in (2.13). Stationarity principles 7 2.4 Real stress field and virtual displacement field Assume that vi is a kinematically admissible displacement field and that ij is the strain field derived from vi . Furthermore, as before, σ 0 , Ti0 , p0i are the real stress, surface traction and volume force fields. Then the work equation (2.8) reads Z Z Z 0 0 σij ij dV = Ti vi dA + p0i vi dV (2.23) V A V To distinguish the work by the external loads from the work of the reactions we divide the surface area A into (2.24) A = A T + Av where AT is the surface area without displacement control and Av is the surface area with given kinematic conditions. Furthermore, we describe the virtual field v i by a variation δvi relative to the real field vi0 vi = vi0 + δvi , ij = 0ij + δij (2.25) which with ij = (vi,j + vj,i )/2 gives 1 δij = (δvi,j + δvj,i ) 2 (2.26) Now, as vi is assumed to be kinematically admissible, we have δvi = 0 on the surface Av (but not necessarily vi = 0) , and thus (2.23) with (2.14) reduces to Z Z Z 0 Ti0 δvi dA + σij δij dV = p0i δvi dV (2.27) V AT V Virtual work principle This is called the virtual work principle or the principle of virtual displacements. Note that the virtual displacements and strains in (2.27) are infinitesimal and express energy and work variations without assumptions of linearity. Note also that the virtual work principle is a principle about a state, not a process. Often (2.23) and even (2.8) are also called the virtual work principle, but in this book we shall assume the virtual displacements and the virtual strains to be infinitesimal. Because stresses are fixed in the virtual work principle, a direct physical interpretation is not clear. However, it can be read as an energy balance which is valid for any kinematically admissible disturbance of the displacement field. Some specific cases of use of the virtual work principle lead us to specialized principles. Let us choose the very specific virtual displacement field δvi = ∆v corresponding to the single load Q δvi = 0 corresponding to all other external loads ∆ij derived from this field then (2.27) reduces to Z V 0 σij ∆ij dV = Q∆v or Q = (2.28) ∆U ∂U = ∆v ∂v (2.29) which is the first theorem of Castigliano. It is useful in determining stiffnesses. Note, that this theorem is valid independent of the specific constitutive behaviour (σ = Castigliano’s 1st theorem c 8 Pauli Pedersen: Stationarity and extremum principles in mechanics Unit displacement theorem for linear elasticity Stationary total potential energy σ()). The force Q should be interpreted as a generalized force; thus, if Q is an external moment, then v is the corresponding rotation. Now a much used theorem is obtained from (2.29) if we assume linear elasticity, because then we can set the displacement to ∆v = 1 and ∆ ij = 1ij for the resulting strains from this unit displacement field and get Z 0 1 σij ij dV (2.30) Q= V Returning to general non-linear elastic materials, we can interpret the virtual work principle as stationary potential energy. A potential is a scalar from which work can be derived. Let us assume that a material has a potential and the external loads has as well, then (2.29) states δU = δW or δΠ = 0 2.5 (2.31) Virtual stress field and real displacement field A virtual stress field σij is a statically admissible field, i.e., in equilibrium with the given external loads. Now, inserting also the real displacement field v i0 , and derived strain field 0ij in (2.27), we get Z Z Z Z 0 0 0 0 Ti vi dA + Ti vi dA + p0i vi0 dV (2.32) σij ij dV = V V Av AT On the surface Av , the surface traction’s Ti are the unknown reactions. Taking the virtual stress field as 0 σij = σij + δσij (2.33) 0 is the real stress field and δσ is an infinitesimal virtual stress field satiswhere σij ij fying δσij = δσji , δσij,j = 0 δσij nj = δTi where δTi = 0 on AT Stress virtual work principle (2.34) then using (2.14) we get Z V δσij 0ij dV = Z Av Ti vi0 dA (2.35) which expresses the principle of complementary(stress) virtual work, also called the principle of virtual stresses. Choosing a specific virtual field δσij = ∆σij where ∆σij nj = ∆Q corresponding to displacement v ∆T = ∆σij nj = 0 for all other places with prescribed vi 6= 0 (2.36) Stationarity principles 9 we get from (2.35) Z V ∆σij 0ij dV = ∆Qv or v = ∆U σ ∂U σ = ∆Q ∂Q (2.37) which is the second theorem named after Castigliano. It is valuable in determining flexibilities. Also, a unit theorem is obtained in complementary(stress) energies, read directly 1 from (2.37) when linear elasticity is assumed, i.e., ∆Q = 1 and ∆σ ij = σij v= Z V 1 0 σij ij dV (2.38) Finally, the parallel to stationary potential energy is the principle of stationary complementary(stress) potential energy δU σ = δW σ or δΠσ = 0 2.6 (2.39) Summing up • The identity (2.5) is obtained by rather simple mathematics and has no physical interpretation. • The work equation (2.8) involve two independent fields. A stress/force field in equilibrium, these fields are given super index a. A displacement field with derived strain field, these fields are given super index b. • Work as well as complementary work must in general be determined by integration and therefore depend on the force/displacement function. • Strain energy density as well as stress energy density is also determined by integration, applying the constitutive relation. • The notion of compliance is important in optimal design formulations and is therefore specifically defined. • With real stress field and virtual displacement field we get the virtual work principle, and from this a number of more specific principles. The virtual work principle is a principle about a state, not a process. • With virtual stress field and real displacement field we get the complementary virtual work principle (stress virtual work principle), and from this a number of more specific complementary principles. • Virtual displacements and virtual stresses are in general infinitesimal. Castigliano’s 2nd theorem Unit load theorem for linear elasticity Stationary total stress potential energy c 10 Pauli Pedersen: Stationarity and extremum principles in mechanics Chapter 3 Extremum principles in mechanics The stationarity principles of mechanics are based on very few assumptions. The principles of virtual work hold for any constitutive model and for any type of load, and for potential systems these virtual principles give stationary potential energies. Many approximation methods (like the finite element method) are based on and uniquely specified by these stationarity principles. However, this does not give us sufficient reason to choose an approximate solution that satisfies the same stationarity as the unknown real solution. Energy principles that in addition to stationarity give extremum can justify our choice. We choose the approximation for which the energy is closest to the real unknown energy. Furthermore, consistent approximation methods are a reasonable choice also for problems where an extremum cannot be proved. 3.1 Motivation for extremum Principle of minimum total potential energy We shall firstly prove the principle of minimum total potential energy δ 2 Π > 0, and for this we need assumptions concerning the constitutive model as well as for the load behaviour. Let us start with a single load Q (force or moment) and the corresponding displacement v (translation or rotation). For this force Q as a function of the corresponding displacement v, we will assume the following single load behaviour ∂v ∂Q ≥ 0, > 0, Q(v = 0) = 0 ∂v ∂Q (3.1) as illustrated in Figure 3.1a). We note that Q(v) is a function but, as ∂Q/∂v = 0 is a possibility, v(Q) is not strictly a function. Non-linearity and change of sign for curvature is possible. From the definition of W and W σ in (2.11) - (2.12) follows Z v ∂W Q(ṽ)dṽ ⇒ = Q = Q(v) W = ∂v 0 Z Q ∂W σ σ W = v(Q̃)dQ̃ ⇒ = v = v(Q) (3.2) ∂Q 0 As by (2.17) we have for this case of a single force W + W σ = Qv (3.3) 11 Assumptions Single force behaviour PSfrag replacements c 12 Pauli Pedersen: Stationarity and extremum principles in mechanics Q Q W Wσ b b (W ) W a v a) v a (W ) b) va W v W v = Qb = Qa v vb Figure 3.1: a): Illustration of a possible relation between force Q and corresponding displacement v. The work W and the complimentary work W σ are shown as areas. b): Work W of the force Q as a function of the displacement v that correspond to the force. Also the sign of the curvature of W = W (v) and of W σ = W σ (Q) is known from (3.1) and (3.2) ∂2W σ ∂Q ∂v ∂2W ≥ 0, >0 = = 2 2 ∂v ∂v ∂Q ∂Q Work function (3.4) From this it follows that we get a work function W = W (v), as illustrated in Figure 3.1b). From the tangents shown in Figure 3.1b) we read the inequalities (W )a + Qa (v b − v a ) ≤ (W )b (W )b − Qb (v b − v a ) ≤ (W )a (3.5) which together gives an inequality valid for v b > v a as well as for v a < v b , i.e., convexity Qa (v b − v a ) ≤ (W )b − (W )a (3.6) where the equality only holds for ∂Q/∂v ≡ 0 in all the actual interval from v a to v b and naturally for v a = v b . Same arguments hold for the stress quantities and thus we also have v a (Qb − Qa ) ≤ (W σ )b − (W σ )a Inequality for mixed product (3.7) with equality only for Qb = Qa because ∂v/∂Q > 0 is assumed in (3.1). Inserting Qa v a = (W )a + (W σ )a from (3.3) in (3.6) or (3.7) we get the inequality for the mixed products Qa v b ≤ (W )b + (W σ )a (3.8) By summation and/or integration we can extend the above results to a load system. Extremum principles 13 For a uniaxial stress/strain in terms of pure normal (σ, ) or, alternatively, pure shear (τ, γ), we assume a function very parallel to the load displacement function (3.1), ∂σ ∂ > 0, > 0, σ( = 0) = 0 ∂ ∂σ (3.9) i.e., well-defined functions for σ = σ() as well as for = (σ) because strict inequalities hold in (3.9). From the assumptions (3.9) follows in direct analogy to the load-work arguments which lead to (3.6)-(3.8) Uniaxial constitutive model σ a (b − a ) < ub − ua for b 6= a a (σ b − σ a ) < uσb − uσa for σ b 6= σ a σ a b < ub + uσa for a 6= b (3.10) with the definitions of energy densities in (2.11) - (2.12) and the previously discussed relation u + uσ = σ as stated in (2.16). For a multidimensional stress/strain state a direct generalization is not easy, and the assumption is therefore often stated directly as convexity of the energy density in the six-dimensional strain/stress spaces. With tensor symbols this is written a b σij (ij − aij ) < ub − ua for bij 6= aij b a b a aij (σij − σij ) < uσb − uσa for σij 6= σij a b σij ij < ub + uσa for a 6= b and σij ij = u + uσ for a = b (3.11) We now have the necessary inequalities to prove the extremum principles, and 0 and virtual disagain we start from the work equation (2.8). With real stresses σ ij placements, strains (vi − vi0 ), (ij − 0ij ) we get Z V 0 σij (ij − 0ij )dV = Z A Ti0 (vi − vi0 )dA + Z V p0i (vi − vi0 )dV From (3.11) follows that the left-hand side satisfies Z 0 σij (ij − 0ij )dV < U − U 0 for ij 6= 0ij General stress/strain state (3.12) (3.13) V and the right-hand side for dead loads (∂Ti0 /∂vi = 0, ∂p0i /∂vi = 0) gives Z Z p0i (vi − vi0 )dV = W − W 0 Ti0 (vi − vi0 )dA + A (3.14) V Using (3.13) as well as (3.14) in (3.12) we get the result U − U 0 > W − W 0 or Π > Π0 for ij 6= 0ij (3.15) i.e., the extremum principle for total potential energy. We note that the dead load assumption (3.14) is a necessary condition if we do not decide by definition to term the right-hand side of (3.12) as the negative external potential energy. Minimum potential c 14 Pauli Pedersen: Stationarity and extremum principles in mechanics 3.2 Principle of minimum total complementary(stress) potential energy We can directly establish the complementary(stress) principle because all the necessary inequalities were derived in the Section 3.1. In the work equation (2.8) we now insert vi0 , 0ij and the virtual stresses σij and get Z V 0ij (σij − 0 σij )dV = Z A vi0 (Ti − Ti0 )dA + Z V vi0 (pi − p0i )dV The left-hand side satisfies Z 0 0 0ij (σij − σij )dV < U σ − U σ0 for σij 6= σij (3.16) (3.17) V Minimum complementary (stress) potential The main part of the right-hand side is zero when σij is statically admissible because pi − p0i = 0 and Ti − Ti0 can only be different from zero at the reactions. For dead loads this part will be W σ − W σ0 , and with (3.17) in (3.16) we get 0 U σ − U σ0 > W σ − W σ0 or Πσ > Πσ0 for σij 6= σij i.e., the extremum principle for total complementary(stress) potential energy. Using also the earlier result (2.18) of Π0 + Πσ0 = 0 for only real fields we can with (3.15) and (3.18) set up two-sided bounds on approximate solutions. For the real solution we have (2.18), and by the sum of (3.15) and (3.18) we for an approximate solution get Π + Πσ > 0 or Π > −Πσ Two-sided bounds (3.18) (3.19) Furthermore, substitution of Π0 = −Πσ0 in (3.15) and (3.18) then gives the two-sided bounds Π > Π0 > −Πσ Πσ > Πσ0 > −Π 3.3 (3.20) Overview of principles and their relations Figure 5.1 illustrate the connections between the many different energy principles. The indicated horizontal dash line shows the division between the stationarity and the extremum principles. The indicated vertical dash line shows the division between the strain principles and the complementary stress principles. Extremum principles 15 b b a σija = σji ) + vj,i , bij = 12 (vi,j a σija nj = Tia , σij,j = −pai frag replacements 1st CASTIGLIANO R V σija bij dV = R T a v b dA + A i i VIRTUAL WORK PRINCIPLE R V pai vib dV 2nd CASTIGLIANO VIRTUAL COMPLEMENTARY (STRESS) PRINCIPLE ASSUMPTIONS ON LINEARITY LOADS- AND CONSTITUTIVE BEHAVIOUR STRAIN PRINCIPLES STRESS PRINCIPLES UNIT - LOAD UNIT - DISPLACEMENT PRINCIPLE LINEARITY STATIONARITY PRINCIPLES PRINCIPLE EXTREMUM PRINCIPLES MINIMUM OF TOTAL ELASTIC POTENTIAL MINIMUM OF TOTAL COMPLEMENTARY (STRESS) POTENTIAL TWO-SIDED BOUNDS MINIMUM OF INTERNAL MINIMUM OF INTERNAL ELASTIC COMPLEMENTARY STRAIN ENERGY ELASTIC STRESS ENERGY Figure 3.2: Overview of stationarity and extremum energy principles in mechanics. c 16 Pauli Pedersen: Stationarity and extremum principles in mechanics 3.4 Summing up • Extremum principles serve as argument for choosing approximate solutions that satisfy stationarity principles. • A basic assumption for extremum principles is that a force(moment) is a strict function of its corresponding translation(rotation). • For the constitutive behaviour a basic assumption for extremum principles is convexity of energy density in the six dimensional strain/stress space. • The principle of minimum total potential energy and the principle of minimum total complementary energy together give two-sided bounds for the real solution. • The graphical overview in Figure 5.1 shows stationary as well as extremum principles, strain principles as well as stress (complementary) principles. Chapter 4 Potential relations and derivatives This chapter shows some important potential relations and then the sensitivity analysis directly in energy terms, for linear as well as for non-linear power law elasticity. These results are often used as a basis for other formulations. For detail on the results in this chapter see (Pedersen 1998b) and (Masur 1970). Although not new, results like (4.10) and (4.14) are not so well known, and not intuitively understandable, even for the case of linear elasticity (p = 1). For optimal design these results give rise to important simplifications. 4.1 Not well known but important Equilibrium The general equation of energy equilibrium is (4.1) U + U σ + U ext = 0 with elastic strain energy U and elastic stress energy U σ (elastic complementary energy) from the corresponding densities u , uσ integrated over the structure/continuum volume V Z Z σ uσ dV (4.2) u dV and U = U = Energy equilibrium Internal potentials V V and the external potential U ext is defined by Z Z ext U := − Ti vi dA + pi vi dV A External potential (4.3) V with surface traction’s Ti , volume forces pi , corresponding displacements vi , and area A surrounding the volume V . The surface traction’s and the volume forces are assumed to be given. The displacements vi (displacement field) with resulting strains, stresses and energy densities are the solution for a given design, i.e. the solution for a given static problem of elasticity. Power law elastic materials resulting in the density relation uσ = pu everywhere in the continuum or structure give in total U σ = pU and the equilibrium (4.1) is then simplified to (1 + p)U = −U ext (4.4) with 0 < p ≤ 1 being a material power law constant. 17 Assumed power law elasticity c 18 Pauli Pedersen: Stationarity and extremum principles in mechanics 4.2 Total potentials Relations with power law elasticity Defining the total potential Π and the total complementary(stress) potential Πσ by (4.5) Π := U + U ext = −Πσ using Π + Πσ = 0, give from (4.4) the relations Π = −pU = −U σ = Potential relations p U ext = −Πσ 1+p (4.6) and by p > 0 and U > 0 get U σ > 0, Πσ > 0, Π < 0 and U ext < 0. From this follows that design for a number of differently stated extremum problems are equivalent and that their values at the extrema are related as max Π = −min Πσ = −min pU = −min U σ = max p U ext 1+p (4.7) with p = 1 for the specific case of linear elasticity. 4.3 Derivatives of elastic potentials The derivative of the total potential Π with respect to an arbitrary parameter, say a design parameter h, is dΠ /dh = (∂Π /∂h)fixed strains + (∂Π /∂) (d/dh) = (∂Π /∂h)fixed strains (4.8) Design independent loads because of stationary total potential ∂Π /∂ = 0 (virtual work principle) with respect to kinematically admissible strain variations. For design-independent external loads, (∂U ext /∂h)fixed strains = 0, the definition (4.5) then gives (∂Π /∂h)fixed strains = (∂U /∂h)fixed strains (4.9) and totally from (4.6), (4.8) and (4.9) get the result that is frequently used in design optimization Local design parameter 1 dU /dh = − (∂U /∂h)fixed strains p For a local design parameter he that only changes the design in the region e of the structure/continuum this gives the possibility of a localized determination of the sensitivity for the total elastic strain energy 1 dU /dhe = − (∂((ū )e Ve )/∂he )fixed strains p Remark (4.10) (4.11) where (ū )e is the mean strain energy density in the region of he and where Ve is the corresponding volume. Note that the only difference between linear (p = 1) and nonlinear material is the factor 1/p, and for a condition on stationarity dU /dhe = 0, p has no influence. Note, that the sensitivity is not physically localized, but still without approximation it is possible to determine the sensitivity localized. Potential relations 19 For the complementary potentials even more simple results are available dΠσ /dh = (∂Πσ /∂h)fixed stresses + (∂Πσ /∂σ) (dσ/dh) = (∂Πσ /∂h)fixed stresses (4.12) because of stationary total complementary potential ∂Πσ /∂σ = 0 (complementary virtual work principle) with respect to statically admissible stress variations. From U σ = Πσ then follows dU σ /dh = (∂U σ /∂h)fixed stresses (4.13) which with the relation p(dU /dh) = (dU σ /h) and (4.13), (4.10) gives (∂U σ /∂h)fixed stresses = − (∂U /∂h)fixed strains (4.14) valid also for p 6= 1. Formula (4.14) can also be found in the optimal design paper by (Masur 1970). 4.4 Summing up • With proportionality between strain energy and stress energy, simple proportional relations to external potential and to total potentials exist. • From this follows that sensitivity analyses are simplified, and alternative optimization objectives can be chosen. • First order sensitivities with fixed strain fields or fixed stress fields result from the virtual work principle or from the virtual complementary(stress) work principle. c 20 Pauli Pedersen: Stationarity and extremum principles in mechanics Chapter 5 Energy densities in matrix notation For linear strain models {}, {δ} are the vectors of strain and variational strain, and {σ}, {δσ} are the vectors of conjugated stress and conjugated variational stress. In this chapter energy densities is described by the linear strain quantities, but a similar description follows for non-linear strain notation. 5.1 Linear strain notation Strain and stress energy densities With {}, {δ} being the vectors of strain and variational strain, and {σ}, {δσ} being the vectors of conjugated stress and conjugated variational stress; then the variational strain energy density δu is defined by Z {} T {σ}T {d˜ } (5.1) δu = {σ} {δ} ⇒ u = Strain energy density 0 and thus having the same dimension as stress. The tilde ˜ indicate the difference between integration variable {˜ } and final strain state {}, i.e., the stress function is {σ} = {σ({˜ })}. The strain energy density is a function of strain, and from the strain energy density function u stress is obtained by differentiation {σ} = du = {σ({})} = [L̄]{} {d} and by its definition then the secant modulus [L̄]. The variational stress energy density δuσ is defined by Z {σ} σ T σ {}T {dσ̃} δu = {δσ} {} ⇒ u = (5.2) Stress energy density (5.3) 0 The stress energy density is a function of stress and from the stress energy density function uσ , and the secant compliance relation [L̄]−1 is obtained by differentiation {} = duσ = {({σ})} = [L̄]−1 {σ} {dσ} (5.4) Together the definitions of variational strain energy density and variational stress energy density gives δ({σ}T {}) = δu + δuσ (5.5) 21 c 22 Pauli Pedersen: Stationarity and extremum principles in mechanics and with integration from zero elastic energy density to final state follows (5.6) u + uσ = {σ}T {} For a one dimensional case Figure 5.1 illustrates the definitions (5.1) and (5.3) by areas, and the relation (5.6) is also directly recognized. σ uσ PSfrag replacements u Figure 5.1: Illustration of strain energy density u (5.1) and of stress energy density uσ (5.3) by areas. 5.2 Power law elasticity Energy densities in 1D non-linear elasticity The analysis is restricted to power law non-linear analysis, because for this model analytical explicit expressions for the constitutive matrices are obtained. Although the general case of 3D anisotropic behaviour is described for this nonlinearity, at first a 1D model is treated. The model of power law non-linear elasticity for a 1D model in terms of σ = σ() is σ = Ẽp or more general σ = | p−1 | E0 for ≥ 0 0 (5.7) with the positive material parameters E0 , p and the limit of linearity 0 measured in strain or in stress by σ0 . The material modulus for linear elasticity is E0 and thus σ0 = E0 0 . Figure 5.2 shows the drastic influence of the power p in the possible range 0 < p ≤ 1. Energy densities in matrix notation 23 σ E0 0.4 0.35 PSfrag replacements = p p−1 0 p = 1.0 p = 0.9 p = 0.8 p = 0.6 p = 0.4 p = 0.2 p = 0.1 0.3 0.25 0.2 0.15 0.1 0.05 0 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0 0.4 Figure 5.2: Illustration of the influence from power p in the applied power law non linear elasticity. For strains smaller than the limiting strain 0 a linear behavior is assumed. For convenience, here > 0 is assumed, and it follows that the secant modulus Es and the tangent modulus Et are σ = Ẽp−1 dσ = pẼp−1 = pEs Et := d Es := (5.8) Figure 5.3 illustrates the definition of the Et modulus by the tangent at an actual (, σ) value, and the Es modulus at the same point by the slope from the origin to the (, σ) value. σ Et PSfrag replacements actual (, σ) Es Figure 5.3: Illustration of the secant modulus Es and the tangent modulus Et . Secant and tangent modulus c 24 Pauli Pedersen: Stationarity and extremum principles in mechanics Inserting (5.7) in (5.1), the strain energy density for this material model is obtained Z 1 ˜p d˜ = Ẽp+1 (5.9) u = Ẽ p + 1 0 and from (5.6) with (5.21) get u + uσ = σ = Ẽp+1 Proportional relation ⇒ uσ = p Ẽp+1 p+1 (5.10) The simple proportional relation uσ = pu (5.11) between stress energy density and strain energy density often simplifies analysis (and especially sensitivity analysis for optimal design) to a large extent and give rise to a number of important general results. However, it should be kept in mind that the power law non-linear elasticity is a restrictive model. 5.3 Effective strain/stress Energy densities in 2D and 3D non-linear elasticity Extension to 2D and 3D models is not trivial, and the definitions of effective strain and of effective stress must be chosen appropriately. In matrix notation the differential strain energy density du is similar to the variation in (5.1) du = {σ}T {d} (5.12) In analogy with (5.7) the constitutive secant modulus is p−1 p−1 e e {σ} = E0 [α]{} ⇒ {σ} = [L̄]{} with [L̄] = E0 [α] 0 0 (5.13) Constitutive secant modulus assuming linear elasticity for e ≤ 0 , and with the non-dimensional and constant matrix [α] describing the relative moduli (isotropy as well as non-isotropy). The reference strain is 0 and the corresponding reference modulus E0 . It follows from (5.13) that at the reference strain 0 , the scalar secant modulus is independent of the power p. The fact that the matrix [α] is constant means that the non-isotropic relations are unchanged, only the stiffness magnitude changes through the factor p−1 e . Inserting (5.13) in (5.12), with [α] being symmetric, give p−1 e {}T [α]{d} (5.14) du = E0 0 and the effective strain e must be defined so that {}T [α]{d} can be integrated. From the energy related definition 2e := {}T [α]{} (5.15) follows by differentiation with [α] constant and symmetric 2e de = 2{}T [α]{d} (5.16) Energy densities in matrix notation 25 and thus inserted in (5.14) the one dimensional result du = E0 pe de 0p−1 = Ẽpe de with Ẽ = E0 0p−1 (5.17) which is integrated to obtain the relations proved for the 1D case, i.e., u = 1 Ẽp+1 and uσ = pu p+1 e (5.18) The constitutive tangent modulus is obtained by differentiating (5.13) to get, with the use of (5.16), p−1 1−p e (5.19) E0 [α] [I] − 2 {}{}T [α] {d} = [L]{d} {dσ} = 0 e or alternatively written for the constitutive tangent modulus [L] = e 0 p−1 E0 1−p [α] − 2 {ζ}{ζ}T e with {ζ} = [α]{} (5.20) showing the influence of the dyadic product {ζ}{ζ}T . The model (5.13) with the definition (5.15) for non-linear elasticity may alternatively be derived from the strain energy potential u as determined in (5.21) u = Ẽ p+1 p+1 e (5.21) giving {σ} = Ẽ du de = (p + 1)pe d{} p+1 d{} (5.22) and from the definition of effective strain (5.15) de = −1 e [α]{} d{} (5.23) {σ} = Ẽep−1 [α]{} (5.24) that inserted in (5.22) give with Ẽ = E0 /0p−1 identical to the assumed model (5.13). Figure 5.4 shows the factor (e /0 )p−1 , that by the displacement iterations is determined for each element. It follows from Figure 5.4 that domains close to e = 0 are most sensitive. Constitutive tangent modulus c 26 Pauli Pedersen: Stationarity and extremum principles in mechanics p−1 e 0 1 p = 0.9 0.8 PSfrag replacements p = 0.8 0.6 0.4 p = 0.6 0.2 p = 0.4 p = 0.2 0 e p = 0.1 0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4 Figure 5.4: Illustration of the stiffness factor (e /0 )p−1 as function of the effective strain e . 5.4 Summing up • The notations u for strain energy density and uσ for stress energy density are chosen as alternative to the use of complementary energy density in order to focus on the relational dependence. • The sum of these energy densities is denoted φ and only for linear elasticity is u = uσ = 12 (u + uσ ) = 12 φ. • However, for all non linear elasticity is φ = u + uσ . • For power law non linear elasticity with power p is uσ = pu , and secant as well as tangent constitutive matrices are analytical available. • Note the misprint in the references (Pedersen 2005a), (Pedersen 2005b), and (Pedersen 2006), where in formulas corresponding to 5.19 the denominator 2e 2 and should be η 2 , as here and in for Green-Lagrange strains is written ηref ef f the early reference (Pedersen and Taylor 1993). Chapter 6 Elastic energy in beam models 6.1 Elastic energy in a straight beam In handbooks of strength of materials, like in (Sundstrøm 1998), we find the formula for elastic energy (based on linear, isotropic elasticity) in a straight beam of length b−a U = U σ = U σN + U σT + U σM + U σMx Z b 2 T2 M2 Mx2 N +β + + = dx 2EA 2GA 2EI 2GK a Handbook formulas (6.1) where the cross-sectional forces/moments are N = normal force, T = transverse(shear) force M = bending moment, Mx = torsional moment The material parameters of the assumed isotropic linear elastic behaviour are E = Young’s modulus, ν = Poisson’s ratio E G = shear modulus = 2(1 + ν) and the cross-sectional constants are A = area (tension/compression stiffness factor) I = moment of inertia (bending stiffness factor) K = torsional stiffness factor β = factor from the shear stress distribution Let us primarily prove the individual terms in (6.1) based on the definitions in Chapter 2. With only a normal force N and stresses uniformly distributed over the cross-section we get N σ σ σ2 2 E , = , u = u σ = = = A E 2 2E 2 2 N ⇒ u = uσ = 2A2 E σ= (6.2) 27 Normal force c 28 Pauli Pedersen: Stationarity and extremum principles in mechanics Bending moment With uniformly distributed energy density, the energy per length is u σ A = N 2 /(2EA) as stated in (6.1). With only a bending moment M the stresses vary linearly through the height h of the beam Mz −h h , for ≤z≤ I 2 2 2z2 M ⇒ u = uσ = (6.3) 2I 2 E R and the moment of inertia defined by I = A z 2 dA, the energy per length is R with σ 2 A u dA = M /(2EI) as stated in (6.1). After the two cases of pure normal stress let us analyze the case of only a transverse force T . The distribution of shear stresses τ = τ (z) (τ = σ12 ) will depend on the specific cross-sectional shape, and is here stated as σ= τ = τmax f (z) with f (z) = 0 for z = ± Transverse force h 2 (6.4) Then with engineering shear strain γ = 212 and µ as a cross-sectional constant, determined by a function f (z), the analog to (6.2) is τ τγ τ2 γ2G µT , τ = τmax f (z), γ = , u = uσ = = = A G 2 2G 2 2T 2 µ ⇒ u = uσ = f 2 (z) (6.5) 2A2 G R Integrating to energy per length A uσ dA, we see that the constant β in (6.1) must be defined by 2 Z µ 2 f (z)dA (6.6) β= A A τmax = Torsional moment For specific values of µ, β, see (Sundstrøm 1998) or an alternative handbook. Finally the case of only a torsional moment, here restricted to circular crosssections for which the stress distribution with outer radius R is τmax = τmax r Mx r Mx R , τ= = for rmin ≤ r ≤ R K R K (6.7) In analog to (6.5) we then get u = u σ = Decoupled energies τ2 M 2 r2 = x2 2G 2K G (6.8) R which with K = A r2 dA for circular cross-sections gives the energy per length as stated in (6.1). The non-circular cross-sections is not covered here. Energy is not linear in stresses as seen from (6.1), where N , T , M and M x are often termed generalized stresses. We therefore need to prove that the simple addition of the four energies is correct. For the normal stresses with both N and M we get σ= Mz N 2 M 2 z 2 2N M z N + ⇒ σ2 = 2 + + A I A I2 AI (6.9) Elastic energy in beam models 29 and the last R term will not give rise to energy because by definition of the beam axis we have A zdA = 0. For the shear stresses with both T and Mx it is more simple to look at the work of T and Mx instead of the elastic energy and then base the proof on the energy principles of Chapter 2. The beam displacements from T give no rotation around the beam length direction and therefore no work by Mx . Similar the beam cross-sectional rotation from Mx give no transverse displacement and therefore no work by T . Thus the decoupling’s in (6.1) are correct. frag replacements6.2 Results for simple (Bernoulli-Euler) beams M MB MA x Q x T a) b) Figure 6.1: Slender beam examples. Case a) for cantilever with concentrated force at the free end and case b) for a beam subjected to two end moments only. In Figure 6.1 is shown slender beams and although N 6= 0 and T 6= 0 the elastic energy from bending is often so dominating that we can simplify (6.1) to Uσ = Z b a or expressed in displacements v from M = U = Z b a EI 2 M2 dx 2EI (6.10) EId2 v/dx2 d2 v dx2 2 Bending energy only by dx (6.11) Let us in relation to the example in Figure 6.1a) discuss the error in neglecting the term with T 2 in (6.1). With load Q we have T = −Q and M = Qx giving U σT = Z L 0 Q2 Q2 L β dx = β , U σM = 2GA 2GA Z L 0 Q2 x 2 Q2 L 3 dx = 2EI 6EI (6.12) Neglected shear energy From this follows 3βEI 6β(1 + ν) U σT = = σM 2 U GAL Γ2 p with the slenderness ration Γ defined by Γ = L A/I (6.13) With β values of the order 1 and slenderness ratio mostly of the order 10-100 we see, how dominating the bending energy is. c 30 Pauli Pedersen: Stationarity and extremum principles in mechanics Elementary case The important case of linearly varying moment shown in Figure 6.1b) give M (x) = MA + (MB − MA )x/L and then from (6.10) with constant bending stiffness EI give Z L L 1 x 2 dx = Uσ = MA + (MB − MA ) (MA2 + MB2 + MA MB ) 2EI 0 L 6EI (6.14) which could have given (6.12) directly for MA = QL and MB = 0. 6.3 Beam solutions by stress(complementary) principles PSfrag replacements v0 x x=0 Q x=L Figure 6.2: Slender cantilever beam problem with end force only. The cantilever, slender beam problem is repeated in Figure 6.2 with displacement v0 = v(x = 0) corresponding to the force Q. We shall first list v0 from solving the differential equation (not shown) QL3 3EI How can we obtain this result with an energy principle? v0 = v(x = 0) = Four energy solutions (6.15) • The stress(complementary) virtual work principle states δU σ = δW σ , which with δW σ = v0 δQ and (6.12) gives (6.15) because v0 δQ = (2QL3 /(6EI))δQ must hold independently of δQ. • The 2nd Castigliano theorem states v0 = ∂U /∂Q and thus directly from (6.12) v0 = 2QL3 /(6EI). • The unit load theorem (2.38) for linear elasticity as here gives with σ 1 from Q = 1 and 0 from Q: Z LZ Z LZ Qxz 1xz 1 0 v0 = σ dAdx = dAdx = I IE 0 A 0 A Z L Qx2 QL3 dx = (6.16) EI 3EI 0 σ σ σ • The complementary total potential (6.12) RL 2 2 R Q energy is Π = U − W , i.e., from σ Π = 0 Q x /(2EI)dx - 0 v(Q̃)dQ̃ and thus stationarity of Πσ with respect to variation of Q (∂Πσ /∂Q) gives the result when the following differentiation is applied R Q d 0 v(Q̃)dQ̃ = v(Q) = v0 (6.17) dQ Minimum of Πσ (∂ 2 Πσ /∂Q2 > 0) gives L3 /(3EI) > 0, that is clearly satisfied. Elastic energy in beam models 31 6.4 Summing up • The elastic energy in a straight beam of linear, isotropic, homogeneous material is integrated along the beam axis, that contains the centers of cross sectional gravity. The formula 6.1 is taken from a handbook. • Although energy is not linear in the cross sectional forces and moments, the simple addition from each of these is proved. • The coefficient β to the shear force component must be determined by integration for a specific cross section. • As an example four different energy principles are shown to give the same result. c 32 Pauli Pedersen: Stationarity and extremum principles in mechanics Chapter 7 Some necessary conditions for optimality In this chapter we primarily present two necessary conditions for optimal solutions in general, i.e. not directly related to optimal design. The first condition is only valid for non-constrained problems and the second condition is valid for problems with only a single constraint in addition to the objective. After the two first sections we then determine, expressed in physical terms, the condition for solution of the most simple optimal design problems, for which sizes (or field of size) optimize stiffness as well as strength. The optimal designs to these two different problems are shown to be the same. In close relation to the analysis for size optimization, we treat optimization of shape, again to optimize stiffness as well as strength. The shape solution to these two different problems is shown in many cases also to be the same, very much in parallel to the result for size optimization. A simple parametrization for shape optimization is finally described. The goal of the present chapter is to obtain basic understanding of very simple optimal design problems, without involving extended numerical calculations. 7.1 General optimization Design optimization Size and shape Non-constrained problems The notion of Φ is often used for compliance as in (2.13), but in Sections 7.1 and 7.2 it is applied for a more general objective. A non-constrained optimization problem may be defined as Extremize Φ = Φ(he ) with variables he non-constrained (7.1) and a necessary condition for this is that the objective Φ is stationary with respect to all the independent variables he dΦ/dhe = 0 for all e (7.2) The optimization of material orientation is an important example of such a nonconstrained problem. Unfortunately, this problem is not simple to solve, because many local optima exist. Each variable (orientational angle θ e in a domain e) has several solutions to the stationarity condition (7.2). 33 Stationary objective c 34 Pauli Pedersen: Stationarity and extremum principles in mechanics 7.2 Problems with a single constraint An optimization problem with only a single constraint may be defined as Extremize Φ = Φ(he ) with variables he constrained by g = g(he ) = 0 Converted to nonconstrained To obtain an optimality condition we convert this problem to a non-constrained problem, using a Lagrangian function L = Φ − λg to be made stationary for arbitrary value of λ (the Lagrangian multiplier). It follows from (7.2) that a necessary optimality condition is dL/dhe = 0 for all e dΦ/dhe = λdg/dhe Proportional gradients (7.3) ⇒ (7.4) This general result we read as proportionality between the gradient of the objective and the gradient of the single constraint. The factor of proportionality, the Lagrangian multiplier λ, is determined by the constraint condition g(h e ) = 0, that follows from dL/dλ = −g = 0. The use of the optimality condition (7.4) for obtaining more general information about optimal design is very important. It should be noted that behind this result is the assumption that the constraint is active. Extensions to two and more constraints are possible, but the uncertainty about the active constraints is then often a limiting factor on the usefulness. An alternative look at the problem is to postulate (7.4) and then see that dg = 0 implies dΦ = 0, i.e. dΦ = 7.3 X ∂Φ X ∂g ∆he = λ ∆he = λdg = 0 ∂h ∂h e e e e (7.5) Size optimization for stiffness and strength The theoretical results for size optimization are more developed than those for shape optimization. Let us therefore start with some basis knowledge from size optimization, as it can be found in (Pedersen 1998b) for non-linear elasticity or in (Wasiutynski 1960) for linear elasticity. 7.3.1 Size design with optimal stiffness Homogeneous mass(volume) dependence If the objective is to minimize compliance (minimize elastic energy) for given total mass then we have (for optimal stiffness design with homogeneous assumptions and design independent loads): the ratio between sub-domain energy and sub-domain mass should be the same in all the design sub-domains. Let the P design parameters be he , then homogeneous mass relations are obtained P with M = e Me = e hm M̄ e , where M is the total mass, Me is the mass in doe main e, m is a given positive value, and M̄e is independent of P the design parameters. P The homogeneous energy relations are obtained with U = e Ue = e hne Ūe , where U is the total strain energy, Ue is the strain energy in domain e, n is a given positive value, and Ūe is explicitly independent of the design parameters. Restricted to problems with constant mass density we get, in all design domains, the same mean strain energy density. Furthermore, if the model has constant energy Some necessary conditions for optimality 35 density within a design domain, then the result for the optimal design is uniform strain energy density u∗ , i.e. u∗e = ū for all free design domains (7.6) where lower and upper size constraints are not reached. The symbolism here is a super-index ∗ related to the optimal design, and a overhead bar ¯ indicating a constant value for each domain e (mean value). Assume now that the necessary condition (7.6) give a global minimum solution, then for any other design the total strain energy U is larger (or equal to) X X X X X Ve = u∗e Ve (7.7) Ve∗ = ū U = ue Ve ≥ U∗ = u∗ Ve∗ = ū e e e e Stiffest design e where is the optimal volume of the design domain e. For an P alternativeP design with design volumes Ve we have the same total volume V V = e Ve = e Ve∗ . From (7.7) we get X (ue − u∗e )Ve ≥ 0 (7.8) Ve∗ e 7.3.2 Size design with optimal strength With positive volumes Ve we read from (7.8), that at least one ue is not less than u∗e . Thus if the strongest design is defined by minimum of maximum u e , then the stiffest design characterized by the optimality condition (7.6) is also the strongest design. We note that the strength may also be defined in relation to the von Mises stress or an alternative effective stress, and these measures are not always proportional to the energy density. For a detailed discussion of these aspects see (Pedersen 1998b). 7.4 Also best strength Shape optimization for stiffness and strength In the following we use the same kind of reasoning to draw conclusions about shape optimization, without involving a solution to the actual stress problem. Thus we gain general knowledge, valuable for 3D and 2D-problems, for non-linear elastic as well as for linear problems, for non-isotropic or isotropic problems, for any external, design independent load. Also valid for non-homogeneous problems and independent of the solution procedure. In order to simplify the mathematics the design parametrization is chosen as illustrated in figure 7.1. An alternative parametrization with expansion in terms of shape design functions is formulated in (Dems and Mroz 1978), a paper closely related to this presentation. We assume a homogeneous state for the strain energy density u e within the volume Ve related to the shape parameter he , say a constant stress finite element. Let us now subject the shape to variation using only two parameters h i and hj . Furthermore, let the total volume V of the structure (continuum) be fixed, then ∆V = dVj dV dV dVi ∆hi + ∆hj = ∆hi + ∆hj = 0 dhi dhj dhi dhj (7.9) because we also assume the domain volumes to be depending only on one design parameter and with a positive gradient (to be used later) Ve = Ve (he ) and dVe /dhe > 0 General knowledge (7.10) Localized volume change c 36 Pauli Pedersen: Stationarity and extremum principles in mechanics Figure 7.1: Discretized design parametrization, showing two design domains i and j. 7.4.1 Shape design with optimal stiffness In shape optimization for extremum elastic strain energy the increment of the objective corresponding to increments ∆hi , ∆hj is ∆U = Design independent loads Localized energy change dU dU ∆hi + ∆hj dhi dhj which for power law non-linear elasticity σ = Ep can be written as 1 ∂U ∂U ∆U = − ∆hi + ∆hj p ∂hi ∂hj fixed strains (7.11) (7.12) This is proved in (Pedersen 1998b) for design independent loads, and follows from (4.10) in chapter 4. Therefore only the local energies U i = ui Vi and Uj = uj Vj are involved and the variations in the strain energy densities need not be determined, because the constitutive relations are unchanged. We have dVj dVi 1 ∆hi + uj ∆hj ) ∆U = − (ui p dhi dhj (7.13) and inserting (7.9) in (7.13) we obtain 1 dVi ∆U = − (ui − uj ) ∆hi p dhi (7.14) A necessary condition for optimality ∆U = 0 with dVi /dhi > 0 is therefore u i = u j . With all design parameters, eq. (7.9) and (7.13) are written X dVe ∆he ∆V = dhe e 1X dVe ∆U = − u e ∆he (7.15) p e dhe Constant energy density and we conclude that a necessary condition for optimality ∆U = 0 with constraint ∆V = 0 is constant strain energy density ue . Thus for the stiffest design the energy density along the shape(s) to be designed, here denoted u s , must be constant us = ū (7.16) Some necessary conditions for optimality 37 7.4.2 Shape design with optimal strength We now relate the stiffest design (minimum compliance) to the strongest design (minimum maximum strain energy density). Let us assume that the highest strain energy density is at the shape to be designed. With index s referring to shape design domains and index n referring to domains not subjected to design changes, this means that for the stiffest design we assume (7.17) us = ū > un A design domain that depends on design parameter is given index s (h s ) and a design domain which is not subjected to design change is given index n (h n ). For the total design domain we use index S and for the total domain not subjected to design, index N . The total elastic strain energy U is obtained from X X X X U = U S + U N = U s + U n = u s V s + un Vn ,i.e., s U = ū X Vs + s X n s n (7.18) u n V n n With unchanged domain N and for the stiffest design U > U∗ we obtain X X X X u s V s + u n V n > ū Vs∗ + u∗n Vn∗ ,i.e., s n X s (us − ū )Vs > s X n n (u∗n − un )Vn (7.19) as s ū Vs∗ = s ū Vs due to given total volume, and furthermore individual unchanged in the non-design domains Vn∗ = Vn . The right hand side might be negative, so we can not directly draw conclusions as from (7.8). However, in a complementary formulation with stress energies we can prove that the right hand side is non-negative and then the proof holds. The proof of increasing energy in the shape domain is as follows. We write the total stress energy Uσ as the sum of stress energy in the shape domain UσS and stress energy in the non-shape domain UσN and obtain P Basic assumption P Uσ = U σ S + U σ N ⇒ dUσS dUσN dUσ = + dh dh dh From the principle of complementary virtual work follows dUσ /dh = (∂Uσ /∂h)fixed stress field and we get dUσ ∂UσS ∂UσN = + dh ∂h ∂h fixed stress field (7.20) (7.21) where the last term is zero when h has no direct influence on the non-shape domain. Finally for the stiffest design we have dUσ /dh > 0 and from this we conclude ∂UσS dUσ 1 dUS = = >0 (7.22) ∂h fixed stress field dh p dh Summarizing the theoretical results of this section; we have for the general threedimensional case with non-isotropic, power law non-linear elastic material in an nonhomogeneous structure, and for any design independent single load case that: Detail of proof c 38 Pauli Pedersen: Stationarity and extremum principles in mechanics Design for stiffness and strength The minimum compliance shape design (stiffest shape design) has uniform energy density along the designed shape, as far as the geometrical constraints make this possible. If we furthermore assume that the highest energy densities are found at the designed shape, then the stiffest design is also the strongest design, as defined by a design which minimizes the maximum energy density. Note that these results are obtained without calculating the stress/strain fields and without specifying the constitutive behaviour. This behaviour need not be homogeneous and thus we can also include the multi-material case. 7.5 Good experience Conditions with a simple shape parametrization In the final conclusions in section 7.4.2, we have added the note ”as far as the geometrical constraints make this possible”. Also it was commented that normally the shape parametrization implies such a geometrical constraint. In this section we use a simple shape parametrization that makes a rather simple optimality condition possible. The limitations of using this simple parametrization can be evaluated by the possibility to obtain almost uniform energy density distribution along the shape to be designed. Many examples illustrate that the parametrization is in fact able to describe optimal shapes in many cases. Figure 7.2: A three parameter (α, β, η) description of an internal hole in a rectangular domain, specified by A, B. Figure 7.2 shows a single inclusion hole, where the shape of the boundary is modeled as a super-elliptic shape, described by only three non-dimensional parameters, relative axes α, β and power η x η y η + =1 αA βB Two or only one parameter (7.23) With known area of the hole we only have two parameters and if furthermore symmetry is enforced, say αA = βB, we only have one free parameter, which might be the power η. Figure 7.3 shows the great flexibility even for this one parameter description. This parametrization naturally has its limitation, but several examples Some necessary conditions for optimality 39 Figure 7.3: Shapes giving equal area of the hole, with powers of the super-elliptic shape being η = 0.75, 1.25. 1.75 and 3.00, respectively. show its usefulness, and furthermore it can easily be extended to 3D-problems by x η y η z η + =1 (7.24) + αA βB γC In the 2D-model (7.23) the area of the hole is Z αA x η 1/η 4 βB 1 − ( ) ) dx = 2αβABg(η) αA 0 with the function g = g(η) defined by η+1 2 1 Γ /Γ g(η) := Γ η η η (7.25) (7.26) where Γ is the Gamma-function. With the rectangular area being 4AB the relative area of the hole φ (relative to the area 4AB) and the relative area of the solid (relative density) ρ are 1 φ = αβg(η) = 1 − ρ 2 (7.27) Relative hole area or density An optimal design problem is formulated in order to extremize the elastic energy U for constant relative area Extremize U subject to φ(α, β, η) = φ̄ (7.28) Within the possibilities of the three parameters α, β, η this also minimizes energy concentration and returns constant energy density along the boundary of the hole, as discussed in section 7.4.2. Using the result (7.12) from sensitivity analysis we determine the differential of the elastic energy (p = 1 for linear elasticity) 1 ∂U ∂U ∂U dU = − dα + dβ + dη (7.29) p ∂α ∂β ∂η fixed strains and the differential of the constraint follows from (7.27) (using a formula manipulation program to differentiate the Gamma-functions) dα dβ p(η)dη + + dφ = φ (7.30) α β η2 Design problem c 40 Pauli Pedersen: Stationarity and extremum principles in mechanics Available functions with the function p = p(η) defined by 2 1 η+1 p(η) := Ψ −Ψ −Ψ η η η (7.31) where Ψ is the Psi-function. To illustrate that the functions g(η) and p(η) are wellbehaved functions we show in figure 7.4 these functions and there derivatives. These functions are available in many libraries of computer routines. Figure 7.4: Left g-function and right the p-function with their derivatives as a function of the shape power η. Optimality condition For model by the FEM The condition of dU = 0 when dφ = 0 is a necessary condition for optimality and thus (as in general with only a single constraint) we from (7.29) and (7.30) get the optimality condition by proportional gradients (7.4), i.e. ∂U ∂U η 2 ∂U (7.32) α =β = ∂α ∂β p(η) ∂η fixed strains In a fixed strain field the energy densities u are constant and only the volumes of domains (elements) connected to the hole boundary change. Thus in a finite element formulation the optimality condition (7.32) is written α X s Strength or stiffness us X ∂Vs ∂Vs η 2 X ∂Vs =β us = us ∂α ∂β p(η) s ∂η s (7.33) where index s refers to an element connected to the hole boundary. The only information needed in addition to the results from analysis is ∂Vs /∂α, ∂Vs /∂β, ∂Vs /∂η, i.e. only information from geometry. We note, P in agreement with section 7.4.2, that if us is constant along the hole boundary then s ∂Vs /∂α = ∂V /∂α = φ/α etc., and the optimality criterion (7.33) is satisfied by us φ = us φ = us φ. Thus a constant energy density along the boundary of the hole implies stationary total elastic energy. However, we can have stationary energy without constant energy density, if the possible designs are restricted. This is illustrated by the examples in chapter ??. 7.5.1 Possible iterative procedure The problem is how to find a boundary shape that satisfies (7.32) or in finite element formulation (7.33). The heuristic approach of successive iterations could be to estimate the Lagrange multiplier λ by the mean value ∂U η 2 ∂U 1 ∂U λestimated = +β + α (7.34) 3 ∂α ∂β p(η) ∂η Some necessary conditions for optimality 41 and then redefine α, β, η by αnew = λ/( ∂U η2 ∂U ∂U )old , βnew = λ/( )old , ( )new = λ/( )old ∂α ∂β p(η) ∂η (7.35) with iterations on λ to satisfy the constraint of (7.28) 1 φnew = αnew βnew g(ηnew ) = φ̄ 2 7.6 (7.36) Estimated multiplier Summing up In this chapter the important results to focus on are: • For non-constrained problems the necessary optimality condition is stationarity of the objective with respect to all the independent variables. • For problems with a single constraint the necessary optimality condition is proportionality between the gradient of the objective and the gradient of the single constraint. • For size optimization of stiffness and strength the stiffest design is characterized by the optimality condition of uniform energy density and this design is also the strongest design. • For shape optimization of stiffness, the minimum compliance shape design (stiffest shape design) has uniform energy density along the designed shape, as far as the geometrical constraints make this possible. • For shape optimization of strength, if we assume that the highest energy densities are found at the designed shape, then the strongest design, as defined by a design which minimizes the maximum energy density, is also the stiffest design. • For shape optimization a simple super-elliptic description makes it possible to design a wide spectrum of shapes, and the analytical treatment of this case is almost as simple as the classic elliptic case. • The super-elliptic description can be extended to include a skewness parameter, and then also describe triangular shapes, see (Pedersen 2004) for more detail. Collected results c 42 Pauli Pedersen: Stationarity and extremum principles in mechanics Chapter 8 Analytical beam design Beams are among the most important structural elements with classical results for analysis as well as for design. Especially the statically determinate cases constitute a basic knowledge in solid mechanics, treated in a one dimensional formulation. The present chapter originates from a study published in (Pedersen and Pedersen 2008) and describes the general aspects from an energy point of view, but only the most simple examples of the statically determinate beams are included. The analytical approach has two steps, first determining analytically the necessary optimality criterion as it is often found in the literature; for early reference on compliance minimization see (Huang 1968) and (Masur 1970). Second analytical step is to determine an explicit analytical solution to the optimality criterion and this step is limited to statically determinate beams as the optimality criterion can then be written independent of design. The problems of the present chapter are chosen so that a full analytical approach is reasonable. For long beams (Bernoulli-Euler beams) a number of optimal designs are analytically described as refereed in (Save, Prager and Sacchi 1985) and (Rozvany 1989), with most focus on designs based on plastic collapse. These problems may be formulated to maximize stiffness or maximize eigenfrequency for a given amount of material volume or mass, see the review (Olhoff and Taylor 1983). The solutions depend on the actual boundary conditions and load conditions in addition to the chosen design type, i.e., the chosen relation between cross sectional area and cross sectional moment of inertia. Bernoulli-Euler beam theory is applied in most treated cases , and little analytical attention is given to short beams for which Timoshenko beam theory is necessary. In short beam structures the elastic energy from shear forces can not be neglected. From a practical point-of-view more realistic designs are obtained when this additional energy is taken into account, but only a few optimal designs with analytical description are then available. The goal of the present chapter is to present such results from (Pedersen and Pedersen 2008). Primarily, the statics of some statically determinate cases are presented with one dimensional distribution of shear force T = T (x) and of bending moment M = M (x). Material is assumed isotropic and when the beam has cross sectional distribution of area A = A(x) and cross sectional moment of inertia I = I(x), then the distribution of elastic energy per unit length is given. In beams the cross sectional moment of inertia I = I(x) plays a major role and thus we need a relation between area and moment of inertia. Three possibilities are treated in the present chapter with I(x) proportional to A(x), to A2 (x), or to A3 (x). The cross sectional types 43 Statically determinate beams Analytical steps Timoshenko beam theory c 44 Pauli Pedersen: Stationarity and extremum principles in mechanics Cross sectional size Beam length Use of optimal design results specifically treated in the present chapter are limited to those presented in Section 8.4. To illustrate the influence of beam length on the design, the common expression I(x) = CAn (x) is further specified as C = γb2(2−n) with γ as a non-dimensional constant in the range 0.08 ≤ γ ≤ 0.25 for the treated cases. Beam length L is given relative to a cross sectional length parameter b, i.e., L = ηb, treating η as a non-dimensional length parameter. In optimal design formulation for minimum compliance with point wise design variables, a necessary optimality criterion is uniform elastic energy density. In beam design with the cross sectional area A = A(x) as design parameter the optimality criterion must be stated for this two dimensional design domain. The gradient of the elastic energy per unit length with respect to area change, must as a necessary condition be the same for all areas, i.e., at all positions x. For the simplest cases, i.e., all Bernoulli-Euler beams and Timoshenko beams with I(x) linearly depending on A(x) this means that the mean value of elastic energy density should be the same for all areas. However, this later statement of the optimality criterion is not valid for all Timoshenko beams. Four cases of statically determinate beams are analyzed in (Pedersen and Pedersen 2008): Bernoulli-Euler cantilevers, Bernoulli-Euler simply supported, Timoshenko cantilevers, and Timoshenko simply supported. For each of these four beam models three load cases are applied, and in addition to the cross sectional parameter n = 1, 2, 3 in An (x), this adds up to 36 individual cases. As examples in the present chapter are only presented the cantilever beams. In two and three dimensional design of beam like structures, this knowledge on A∗ (x) can be used to compare with numerically obtained results or as initial designs. The presented values of compliance decrease also give possibility for comparing with alternatively obtained results. 8.1 Optimality criterion for beam design The optimization problem is stated as Minimize compliance Φ for a given volume V = Compliance objective Z L A(x)dx (8.1) 0 where the cross sectional area A(x) is integrated along the beam axis x from 0 to length L. The length is specified relative to a cross sectional reference length b as L = ηb with the influence of the non-dimensional parameter η, for short beams showed in the range 1 ≤ η ≤ 5. The area function A(x) is the design function to be optimized, then denoted A∗ (x). The compliance Φ is the work of external dead loads and may be evaluated as the sum of internal elastic energies (stress energy + strain energy) Φ = U σ + U . Using beam theory under the assumption of zero normal force and zero torsional moment (with linear elasticity Φ = 2Uσ = 2U ) the compliance is Φ= Z L 0 φ(x)dx with φ(x) = β M 2 (x) T 2 (x) + GA(x) EI(x) (8.2) where φ(x) is twice the stress elastic energy per unit length, T (x) the shear force, M (x) the bending moment, I(x) the cross sectional moment of inertia, G the shear Analytical beam designs 45 modulus, and E Young’s modulus. For isotropic material we have G determined by Poisson’s ratio ν as E = 2(1 + ν)G. The factor from the cross sectional distribution of shear stresses is β, here approximated to be unchanged along the beam, and being in the range 1 < β < 2. In Section 8.4 different models of type I(x) = CAn (x) are shown, with values of the power n = 1, 2, 3 and correspondingly different dimensions of the quantity C that does not depend on the actual design variable. Introducing the model C = γb 2(2−n) give I(x) = γb2(2−n) An (x) for n = 1, 2, or 3 (8.3) Model for moment of inertia where b is a cross sectional reference length, assumed constant for the optimization. Section 8.4 shows examples where for n = 1 the size b is the width w or the height 1 ≤ γ ≤ 41 . For n = 2 the size b disappear in (8.3) and as reference value h with 12 √ p 1 b a kind of mean value V /L may be chosen, for this case 4π ≤ γ ≤ 183 . Finally 1 for n = 3 the size b is again the width w and γ = 12 . Totally the non-dimensional parameter γ is in the range 0.08 ≤ γ ≤ 0.25 for the cases shown in Section 8.4. Twice the stress elastic energy per unit length, (8.2) and (8.3) is 2 T (x) M 2 (x)b2n 1 α + 4 n φ(x) = γE A(x) b A (x) with α = 2γβ(1 + ν) (8.4) For practical cases the defined non-dimensional parameter α is of the order 1. With a non-dimensional parenthesis and a design independent factor, then (8.4) is written Q2 (T (x)/Q)2 (M (x)/(Qb))2 + φ(x) = α γEb2 A(x)/b2 (A(x)/b2 )n with α = 2γβ(1 + ν) (8.5) Compliance per unit length where Q is the total external force. The shear force distribution T (x) and the moment distribution M (x) are independent of design for statically determinate cases. The optimality condition for an objective with a single active constraint is given by proportional gradients, as stated in (7.4). For the compliance problem (8.1) with A(x) as design parameter and λ̄ as a positive constant this give Z L Z L dΦ dφ(x) dV 1dx or = dx = −λ̄ = −λ̄ dA(x) dA(x) 0 dA(x) 0 Z L dφ(x) + λ̄ dx = 0 (8.6) dA(x) 0 to hold for any variation of A(x). For the statically determinate cases only A(x) in (8.4) is varying and differentiation gives dφ(x) 1 T 2 (x) M 2 (x)b2n = −α 2 − n 4 n+1 (8.7) dA(x) γE A (x) b A (x) and the necessary condition (8.6) to hold at all x may be expressed by a new nondimensional positive constant λ, defined by λ=α (M (x)/(Qb))2 (T (x)/Q)2 + n (A∗ (x)/b2 )2 (A∗ (x)/b2 )n+1 (8.8) Optimality criterion c 46 Pauli Pedersen: Stationarity and extremum principles in mechanics where A∗ (x) is the optimal area distribution. The optimality criterion (8.8) is valid in general for the studied cases, and the optimal area function can be determined from (8.8). For Bernoulli-Euler beams (α = 0) and for Timoshenko beams with n = 1 the constant λ is proportional to the mean value of elastic energy density, as seen from (8.5), i.e., φ(x)/A(x). 8.1.1 Treated boundary conditions and loads Figure 8.1 shows the elementary cases that are all treated in (Pedersen and Pedersen 2008), all being statically determinate. The cantilever cases 1), 2) and 3) give most simple force/moment distributions (8.9) and case 4) is in reality identical to case 1) when half the length is designed. The other simply supported cases 5) and 6) have additional force/moment components as stated in (8.10). PSfrag replacements x x T (x) T (x) M (x) M (x) L/2 Q 1) Q 4) Q 2) 5) Q 3) L Q Q 6) L Figure 8.1: Boundary conditions and loads for 6 beam cases. Cantilever static distributions For the cantilever cases with 0 ≤ x ≤ L = ηb or 0 ≤ x̃ ≤ η, with x̃ = x/b 1) T (x) = Q ⇒ M (x) = −Qx ⇒ 2) 3) Q x⇒ L Q M (x) = − x2 ⇒ 2L Q T (x) = 2 x2 ⇒ L Q M (x) = − 2 x3 ⇒ 3L T (x) = T (x) =1 Q M (x) = −x̃ Qb T (x) x̃ = Q η M (x) x̃2 =− Qb 2η T (x) x̃2 = 2 Q η x̃3 M (x) =− 2 Qb 3η (8.9) Analytical beam designs 47 For the simply supported cases with 0 ≤ x ≤ L = ηb or 0 ≤ x̃ ≤ η, omitting case 4 5) 6) Q Q + x⇒ 2 L Q 2 Q x ⇒ M (x) = x − 2 2L Q Q T (x) = − + 2 x2 ⇒ 3 L Q Q 3 M (x) = x − x ⇒ 3 3L2 T (x) = − T (x) 1 x̃ =− + Q 2 η M (x) x̃ x̃2 = − Qb 2 2η T (x) 1 x̃2 =− + 2 Q 3 η M (x) x̃ x̃3 = − 2 Qb 3 3η Simply supported static distributions (8.10) 8.1.2 Solutions in general For the Bernoulli-Euler beams the shear force term in the optimality criterion (8.8) is omitted and the solution is (proportional to) s M (x) 2 n+1 ∗ ( ) (8.11) [A (x)]B ∝ Qb The influence of the shear force in Timoshenko beams complicates the solution for n = 2 and 3, but for n = 1 the simplicity remains s M (x) 2 T (x) 2 ∗ α( [A (x)]T,n=1 ∝ ) +( ) (8.12) Q Qb although then depending on the value of α (of the order 1). For n = 2 and 3 the optimality criterion (8.8) is rewritten to polynomial form λ( (A∗ (x) n+1 T (x) 2 (A∗ (x) n−1 M (x) 2 ) ( ) =0 ) − α( ) − n( 2 2 b Q b Qb A∗ (x) b2 = T,n=3 v s u u α T (x) α T (x) 2 2 3 M (x) 2 t 2 + ( ( ) + ( ) ) 2λ Q 2λ Q λ Qb Timoshenko design formula for n = 1 (8.13) with solutions that depend on the constant λ, and iterations are necessary. Solutions are obtained in the inverse sense, that a specified λ directly gives a corresponding volume. For n = 3 a second order polynomial in the squared (A∗ (x)/b2 ) with the solution BernoulliEuler design formula Timoshenko design formula for n = 3 (8.14) For n = 2 a third order polynomial in (A∗ (x)/b2 ), also with a simplified solution because the component of second order is zero ∗ A (x) = b2 T,n=2 v s u u 3 −α T (x) 2 3 1 M (x) 1 M (x) 2 2 t 2 + ( ) + ( ) ) + ( ) λ Qb λ Qb 3λ Q v s u u 3 −α T (x) 2 3 M (x) 1 1 M (x) 2 2 t 2 (8.15) ( ) − ( ) ) + ( ) λ Qb λ Qb 3λ Q Timoshenko design formula for n = 2 c 48 Pauli Pedersen: Stationarity and extremum principles in mechanics In the following sections, specific solutions for cantilever beams are presented with focus on the optimal area function A∗ (x), on the integration to satisfy the volume constraint and on the values of decrease in compliance relative to the compliance for a uniform beam. 8.2 Bernoulli-Euler cantilever beams For the case of Bernoulli-Euler cantilever beams the optimal solution is not depending on the length of the beam, and the elastic energy per unit length (8.4) simplifies to x2m M 2 (x) = C̃ n EI(x) A (x) Q2 with C̃ = 2 m γEL2(m−1) b2(2−n) φ(x) = (8.16) using the cross sectional modeling (8.3) for n = 1, 2 and 3, and the load cases 1), 2), 3) correspond to m = 1, 2, 3, respectively. The optimality criterion then state that φ(x)/A(x) should be constant and simplifies to 2m (8.17) A∗ (x) = Kx n+1 The constant K is determined by the volume constraint (8.1) V = Z L A∗ (x)dx = K 0 Z L 2m x n+1 dx 0 2m+n+1 n+1 =K L n+1 2m + n + 1 2m + n + 1 V giving K = 2m+n+1 n+1 L n+1 BernoulliEuler cantilever designs (8.18) The optimal designs are illustrated in Figure 8.2, corresponding to combinations of m = 1, 2, 3, and n = 1, 2, 3, with the non-dimensional length coordinate 0 ≤ x/L ≤ 1 in order to focus on the form rather than satisfying a common volume constraint. 2m x n+1 1 2m n+1 0.8 = 1 2 2 3 1 4 3 3 2 2 3 0.6 PSfrag replacements 0.4 0.2 0 0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1 x L Figure 8.2: For Bernoulli-Euler cantilever beams, illustration of optimal cross sectional area distributions. Analytical beam designs 49 8.2.1 Optimal compliances For a uniform beam (cylindrical) the size of the area is A(x) = V /L and the total energy from (8.16) with C̃ inserted is L Z L n L 2m φ(x)dx = C̃( ) Φ= x dx V 0 0 1 b2(n−2) Ln+3 Q2 = 2 m (2m + 1) γEV n Z (8.19) For the optimal designs in (8.17) the area distributions give ∗ Φ = C̃ Z L 0 Z L ∗ φ (x)dx = C̃ 0 x2m Z C̃ 2mn dx = Kn K n x n+1 L x2m dx = (A∗ (x))n Z L 0 2m x n+1 dx = 0 C̃V K n+1 (8.20) R L 2m using (8.18) for the integral 0 x n+1 dx. Inserting the obtained value for K (also in (8.18)) and C̃ from (8.16) finally give ∗ Φ = Z L 0 φ∗ (x)dx = b2(n−2) Ln+3 Q2 (n + 1)n+1 m2 (2m + n + 1)n+1 γEV n (8.21) The ratio of the compliances (8.21) and (8.19) is Φ∗ (2m + 1)(n + 1)n+1 = Φ (2m + n + 1)n+1 (8.22) with resulting values for the different combinations of load case power m = 1, 2, 3 and cross section parameterization power n = 1, 2, 3 in Table 8.1. As seen from this table we can for unchanged volume have considerable decrease in the compliance by using the optimal forms in Figure 8.2. The notion ”form” is here preferred, because the present problems can not be characterized as shape design problems (the beam axis is unchanged). Ratios Φ∗ /Φ n=1 n=2 n=3 m=1 m=2 0.750 0.556 0.648 0.394 0.593 0.313 m=3 0.438 0.259 0.179 Table 8.1: For Bernoulli-Euler cantilever beams, values for the ratios Φ∗ /Φ for different combinations of load case m and design parameterization power n. 8.3 Timoshenko cantilever beams Short beams are beams where the energy from shear forces must be taken into account (Timoshenko beams). Thus the influence of the T (x) distributions in (8.9) and (8.10) must be taken into account for obtaining optimal designs. Note that the singular designs, which for Bernoulli-Euler beams follows from A ∗ (x) = 0 implied by Optimally obtained compliance c 50 Pauli Pedersen: Stationarity and extremum principles in mechanics M (x) = 0, are then treated more properly if T (x) 6= 0 at the same position. For more extended discussion on singularities of optimal beams, see (Olhoff and Niordson 1979). The non-dimensional moment distribution for the cases 1), 2) and 3), corresponding to m = 1, 2, 3, is M (x) x̃m = for m = 1, 2, or 3 Qb mη m−1 (8.23) and the non-dimensional shear force distribution for these cases is T (x) x̃m−1 = m−1 for m = 1, 2, or 3 Q η (8.24) Inserting specifically these distributions, twice the stress elastic energy per unit length (8.5) is ! x̃2(m−1) /η 2(m−1) x̃2m /(m2 η 2(m−1) ) Q2 α + (8.25) φ(x) = γEb2 A(x)/b2 (A(x)/b2 )n 8.3.1 Design of beams with n = 1 cross sections Beams with n = 1 cross sections are for the cases in Section 8.4 related to thin walled beams or to pure width design. For these cases the general solution is (8.12). Inserting the distributions (8.24) and (8.23) the solution, keeping m as parameter, is A∗ (x) ∝ q αx̃2(m−1) + x̃2m /m2 (8.26) valid for the load cases 1), 2), and 3). For a single load at the free end, load case 1) with m = 1, we get A∗ (x) = KT 1 p α + x̃2 (8.27) The constant KT 1 is determined by the volume constraint, i.e. by integration of (8.27) Z ηp α + x̃2 dx̃ with x̃ = x/b ⇒ V = KT 1 b 0 KT 1 p 4V p / 2η α + η 2 − α ln α + 2α ln (η + α + η 2 ) = b (8.28) Figure 8.3 shows the optimal area distributions in (8.27) for a number of chosen values of α = 0.4, 0.6, 0.8, 1, 1.5, 2, 3, and the beam length is chosen as L = 4b, i.e., η = 4. As expected the effect of taking the shear force into account (the parameter α 6= 0) is mainly noted at the tip end of the cantilever beam, where the optimal area √ is KT 1 α. Further along the beam, a linear change of design area is seen, like for the Bernoulli-Euler beams with 2m/(n + 1) = 1, as seen in Figure 8.2. For a uniform load, load case 2) with m = 2, the solution is A∗ (x) = KT 2 x̃ p 4α + x̃2 (8.29) Analytical beam designs 51 A∗ /KT 1 5 4 L = 4b (η = 4) 3 2 PSfrag replacements 1 α = 3, 2, 1.5, 1 0.8, 0.6, 0.4 0 0 0.5 1 1.5 2 2.5 3 3.5 x/b 4 Figure 8.3: The optimal design distributions for cross sectional areas, corresponding to load case 1) are illustrated for a Timoshenko cantilever beam which account for the shear force. The cross sectional moment of inertia is assumed proportional to the cross sectional area. Note, that KT 1 is depending on α as given in (8.28). and the volume constraint gives KT 2 Z η p V 2 x̃ 4α + x̃ dx̃ = / b 0 3V / (4α + η 2 )3/2 − 8α3/2 = b (8.30) For a triangular load distribution, load case 3) with m = 3, the solution is p A∗ (x) = KT 3 x̃2 9α + x̃2 (8.31) and the volume constraint gives Z η p V x̃2 9α + x̃2 dx̃) KT 3 = /( b 0 p p 8V /(η(9α + 2η 2 ) 9α + η 2 + 81α2 ln (3 α)− = b p 81α2 ln (η + 9α + η 2 )) (8.32) The optimal area distributions in (8.29) and (8.31) are presented in Figure 8.4. These designs are very similar to the optimal designs for Bernoulli-Euler beams with 2m/(n + 1) = 2 and 3 in Figure 8.2, but here the influence of the parameter α value is shown. Optimal compliances To determine how much is obtained by the designs (8.27), (8.29) and (8.31) relative to the compliance for a uniform design, the three energies per length are written Timoshenko cantilever designs c 52 Pauli Pedersen: Stationarity and extremum principles in mechanics A∗ /KT 3 for upper bunch of curves A∗ /KT 2 for lower bunch of curves 100 80 L = 4b (η = 4) 60 PSfrag replacements α = 3, 2, 1.5, 1, 0.8, 0.6, 0.4 40 20 0 0 0.5 1 1.5 2 2.5 3 3.5 4 x/b Figure 8.4: The optimal design distributions for cross sectional areas, are illustrated for a Timoshenko cantilever beam which account for the shear force. The lower bunch of curves correspond to load case 2) and the upper bunch of curves correspond to load case 3). The cross sectional moment of inertia is assumed proportional to the cross sectional area. Note, that KT 2 and KT 3 are depending on α as given in (8.30) and (8.32). together, i.e., (8.25) for n = 1 φ(x) = η 2(1−m) Q2 2(m−1) αx̃ + x̃2m /m2 γEA(x) (8.33) For the design of uniform area A(x) = V /(ηb) the integrated compliance is Z η 3−2m b2 Q2 η 2(m−1) Φ= x̃ (α + x̃2 /m2 )dx̃ γEV 0 Q2 b2 η 2 (αm2 (2m + 1) + η 2 (2m − 1)) = 2 (8.34) m γE V (4m2 − 1) For the optimal designs (8.27), (8.29) and (8.31) the integrated compliance is for n = 1 determined directly by the constants KT m , as presented in (8.28), (8.30) and (8.32). Z ηb V Q2 ∗ (8.35) Φ = φ∗ (x)dx = 2 m γE KT2 m 0 The resulting ratios Φ∗ /Φ are then only a function of the non-dimensional parameters α, η and m. Φ∗ V2 4m2 − 1 = 2 2 2 2 Φ b KT m η (αm (2m + 1) + η 2 (2m − 1)) (8.36) Table 8.2 list, for a number of combinations of values α = 1 and 2, η = 1, 3 and 5, the obtained relative decrease in compliance, corresponding to the load cases 1), 2), and 3) (m = 1, 2 and 3). Analytical beam designs 53 α, η m=1 m=2 m=3 1, 1 0.988 0.733 0.549 1, 3 0.888 0.662 0.515 1, 5 0.828 0.616 0.486 2, 1 0.996 0.741 0.552 2, 3 0.931 0.692 0.531 2, 5 0.867 0.646 0.506 Table 8.2: Values for the ratios Φ∗ /Φ for different combinations of shear parameter α and load case m. The cross sectional moment of inertia is assumed proportional to the cross sectional area (n = 1). In relation to the obtained decrease in compliance, we note that the shear parameter α do not have a significant influence, neither do the length parameter, i.e., the value of the non-dimensional parameter η. Table 8.2 shows that not so much is obtained for shorter beams and for higher values of α, all as expected and as it is directly seen from (8.36). The load distribution has most influence on this relative decrease of compliance. To compare with the resulting decrease for Bernoulli-Euler beams , the case of α, η = 1, 5 is chosen, and for the three load cases (m = 1, 2, 3) Table 8.2 shows for the Timoshenko beam model that the compliances by optimization decreased to 82.8%, 61.6%, and 48.6%, respectively. With the Bernoulli-Euler beam model the corresponding compliances by optimization decreased to 75%, 55.6%, and 43.8%, according to Table 8.1. w Sfrag replacements I(x) = w w = 2r 1 A2 (x) 12 I(x) = w 1 A2 (x) 4π I(x) = h √ 3 2 A (x) 18 I(x) = I(x) = I(x) ' w2 A(x) 6 I(x) ' h2 A(x) 12 1 A3 (x) 12w2 w2 A(x) 8 I(x) ' Optimally obtained compliance w2 A(x) 12 I(x) ' h2 A(x) 4 Figure 8.5: Moment of inertia expressed by area for different cross sections. Solid square with side length w and below thin walled. Solid circular with radius r = w/2 and below thin walled. Solid equal sided triangular with side length w and below thin walled. Finally rectangular with width w and height h and below a pure flange approximation. For the thin walled cross sections the wall thickness is applied as design parameter. c 54 Pauli Pedersen: Stationarity and extremum principles in mechanics 8.4 Examples of beam cross sections Figure 8.5 gives for a number of beam cross sections, the moment of inertia expressed in terms of the cross sectional area. For thin walled beams the assumed uniform thickness t is also assumed relatively small and thus approximated directly by the areas. For the square 4wt(x) = A(x), for the circular πwt(x) = A(x), for the triangular 3wt(x) = A(x) and for the ”flange or sandwich” beam 2wt(x) = A(x). For the circular and the equal sided triangular cross sections I(x) is relative to any direction through the center of gravity. For the rectangular cross sections I(x) is relative to the symmetry lines (also through the center of gravity). 8.5 Summing up • A large number of analytically (mostly explicit) optimal area distributions are derived for compliance minimization with a given structural volume. This is possible only for statically determinate cases, where the shear force distribution T (x) and the moment distribution M (x) are independent of design. The graphical displays of these designs give the background for basic understanding of the influence from boundary conditions, from load distribution, and from the cross sectional modeling. Dealing with both long beams (BernoulliEuler model) and with short beams (Timoshenko beam model), the influence from shear force is made clear. All these different models are designed based on an energy approach to directly obtain optimality criteria analytically. • The direct energy approach has enabled a unified analysis for the specific cases, i.e., for three problems of cantilever beams and three problems of simple supported beams, each with three cross sectional types, i.e., 18 cases for Bernoulli-Euler beams and 18 cases for Timoshenko beams. • The effect of optimization relative to uniform cross sectional design (cylindrical beam) also depends on these aspects, with optimal compliance values often half the original or even less, but also almost unchanged for specific cases. • Taking the area distribution as design variable, then a numerical two or three dimensional model of a beam like structure may be compared. The obtained optimal volume distribution may be compared with A∗ (x)dx from the present study or A∗ (x) may be used in an initial design for a similar or more complicated problem. Many cases are studied and the main parameters are related to the integer values m = 1, 2, 3 and n = 1, 2, 3. The load distribution determines m and the chosen cross sectional type is modeled with the power n. • For long (Bernoulli-Euler) √ cantilever beams the optimal designs (8.17) are n+1 given by the function x̃2m with Figure 8.2 showing these different forms. Values for the obtained ratio of decreased compliance, relative to a uniform beam (cylindrical beam), are listed in Table 8.1. • For long (Bernoulli-Euler) simply supported beams the optimal beam designs also depend on the length, expressed by the ratio of length over the chosen cross sectional length parameter η = L/b. For these results see (Pedersen and Pedersen 2008), that also show results for Timoshenko simply supported beams. Analytical beam designs 55 • For short (Timoshenko) cantilever beams the optimal designs, furthermore, are depending on the parameter α that describes the relative influence from the shear force. This parameter α = 2γβ(1 + ν) is depending on Poisson’s ratio ν for the assumed isotropic material, on the shear stress distribution by β, and on a factor from cross sectional type γ. A practical range of 0.5 < α < 2 is included in the chapter. For the ”thin walled” modeling (or pure width design) of n = 1 the optimal design functions (8.27), (8.29) and (8.31) are still simple. Figure 8.3 shows optimal area functions for a load concentrated at the free end and Figure 8.4 shows results for distributed loads. Details on constraint scaling are presented and the obtained ratio of decreased compliances for chosen values of α, η and m are listed in Table 8.2. • The cross sectional modeling cases of n = 2 and 3 give rise to a change from direct scaling in order to satisfy the volume constraint for the two solutions (8.14) and (8.15) that depend implicitly on the non-dimensional Lagrangian multiplier λ. This constant is determined by simple bisection. c 56 Pauli Pedersen: Stationarity and extremum principles in mechanics Chapter 9 The ultimate optimal material 9.1 The individual constitutive parameters In ultimate optimal material design, also named free material design, we represent the material properties in the most general form possible for an elastic continuum, namely the unrestricted set of components in positive semi-definite constitutive matrices. For a given material (given constitutive relations), we normally measure cost by the amount of material, say by thickness or density. With the free material we need a measure of the ”amount of a matrix”, and cost is then measured on the basis of invariants of these matrices. With reference to the paper by (Bendsøe et al. 1994), we extend the results obtained in that paper to be valid also for power law non-linear elasticity, as done in (Pedersen 1998b). If we choose as cost constraint the Frobenius norm (length of a matrix) of the constitutive matrix, then the analytical proof of the optimal constitutive matrix, even for 3D-problems, is rather direct. 9.2 Free material Frobenius norm Sensitivity analysis With localized sensitivity analysis as shown in chapter 4, and also given specifically by (4.10), we have du 1 ∂u 1 ∂u (9.1) =− = −Ve dh p ∂h fixed strains p ∂he fixed strains where Ve is the volume of the domain of the localized design variable h e , (here a component of the constitutive matrix). Thus minimum total strain energy u implies maximum strain energy density u in a fixed strain field. (In domains of non-constant strain energy density, the notion of mean value ū should be used). The strain energy density depends homogeneously on the squared effective strain e , see (5.15) in chapter 5. The problem formulation can therefore be stated as Maximize 2e := {}T [α]{} subject to Frobenius([α]) = F([α]) = 1 (9.2) where the matrix [α] describes the non-dimensional part of the constitutive matrix in the secant formulation (??). In the invariant formulation for [α] we can choose the coordinate system of principal strains {}T = {{1 2 3 } {0 0 0}} (9.3) 57 Localized determined sensitivities Design problem c 58 Pauli Pedersen: Stationarity and extremum principles in mechanics Principal strains only Direct conclusions and obtain α1111 α1122 α1133 1 2 2e = {}T [α]{} = {1 2 3 } α1122 α2222 α2233 α1133 α2233 α3333 3 (9.4) Now, the Frobenius norm of a matrix is defined as the square root of the sum of the squares of all the elements of the matrix (equal to the squared length of the contracted vector). It thus follows directly that for optimality, the matrix elements not involved in (9.4) must be zero. This means directly that also for the non-linear, power law materials we have: • the optimal material is orthotropic • principal directions of material, strain and stress are aligned • there is no shear stiffness This result for linear elastic material is proved in (Bendsøe et al. 1994), based also on a constraint on the trace of the constitutive matrix. Here, the extension to non-linear elastic material follows directly from the localized sensitivity result (9.1). For simplicity of proof we have chosen the Frobenius norm as the constraint. 9.3 Positive definite Final optimization The further analysis relates only to the sub-matrix in (9.4). To fulfill the condition of being positive definite, we have as necessary conditions α1111 > 0, α2222 > 0, α3333 > 0 2 2 2 α1111 α2222 > α1122 , α1111 α3333 > α1133 , α2222 α3333 > α2233 (9.5) The problem formulation (9.1) can now be written as Maximize 2e = α1111 21 + α2222 22 + α3333 23 + 2α1122 1 2 + 2α1133 1 3 + 2α2233 2 3 (9.6) New design problem constrained by (9.5) and by given Frobenius norm F Optimality condition The general necessary condition for optimality is proportional gradients (see (7.4) in chapter 7), i.e. for this specific case 2 2 2 2 2 2 F 2 − 1 = α1111 + α2222 + α3333 + 2α1122 + 2α1133 + 2α2233 −1=0 d(2e )/dαiijj = λd(F 2 )/dαiijj (9.7) (9.8) with the same λ for all αiijj . This gives the result 21 Optimal modulus matrix 22 23 1 2 1 3 2 3 = = (9.9) α1111 α2222 α3333 α1122 α1133 α2233 and we can finally write the resulting constitutive matrix in the directions of principal strains/stresses (evaluating λ to satisfy (9.7): 2 1 1 2 1 3 0 0 0 1 2 22 2 3 0 0 0 1 3 2 3 23 0 0 0 1 [α]optimal = (9.10) 0 0 0 0 0 0 (1 + 2 + 3 )2 0 0 0 0 0 0 0 0 0 0 0 0 = = = Ultimate optimal material 59 9.4 Numerical aspects and comparison with isotropic material The result (9.10) is valid also for power law, non-linear elastic materials. We note that the matrix in (9.10) has only one non-zero eigenvalue and that the material therefore only has stiffness in relation to the specified strain condition. For the ultimate optimal material, the effective strain e , the strain energy density u , and the Frobenius norm F are 2e = 21 + 22 + 23 1 u = E (2 + 22 + 23 )(p+1)/2 p+1 1 F = 1 (9.11) We can obtain the same effective strain and strain energy density with an isotropic, zero Poisson’s ratio material [α] = [I], but then the corresponding Frobenius norm is F = 6, i.e. the material cost is six times greater. As shown in (Bendsøe et al. 1994), the zero Poisson’s ratio material may be valuable in numerical calculation, because of the degeneracy of the ultimate optimal material. 9.5 Summing up In this chapter the important results to focus on are: • The ultimate optimal material is very degenerate and is only stable in relation to the specific strain state for which it is designed. • The obtained solution is also valid for power law non-linear elastic materials, and simple arguments lead to the obtained analytical solution. • The direct comparison with isotropic, zero Poisson’s ratio material is most interesting, and can be used for obtaining numerical solutions to specific problems. Collected results c 60 Pauli Pedersen: Stationarity and extremum principles in mechanics References Bendsøe, M. P., Guedes, J. M., Haber, R. B., Pedersen, P. and Taylor, J. E. (1994), ‘An analytical model to predict optimal material properties in the context of optimal structural design’, J. Applied Mechanics 61, 930–937. Dems, K. and Mroz, Z. (1978), ‘Multiparameter structural shape optimization by finite element method’, Int. J. Numer. Meth. Engng. 13, 247–263. Huang, N. C. (1968), ‘Optimal design of elastic structures for maximum stiffness’, Int. J. Solids Structures 4, 689–700. Langhaar, H. L. (1962), Energy methods in applied mechanics, John Wiley and Sons, Inc. Masur, E. F. (1970), ‘Optimum stiffness and strength of elastic structures’, J. Eng. Mech. Div., ASCE EM5, 621–649. Olhoff, N. and Niordson, F. I. (1979), ‘Some problems concerning singularities of optimal beams and columns’, Z A M M 59, T16–T26. Olhoff, N. and Taylor, J. E. (1983), ‘On structural optimization’, Journal of Applied Mechanics 50, 1139–1151. 58 references. Pedersen, N. L. (2004), ‘Optimization of holes in plates for control of eigenfrequencies’, Struct. Multidisc. Optim. 28(1), 1–10. Pedersen, P. (1998a), Elasticity - Anisotropy - Laminates with Matrix Formulation, Finite Element and Index to Matrices, Solid Mechanics, DTU, Kgs. Lyngby, Denmark. 320 pages - also available at http://www.fam.web.mek.dtu.dk/html/pp.html. Pedersen, P. (1998b), ‘Some general optimal design results using anisotropic power law nonlinear elasticity’, Structural Optimization 15, 73–80. Pedersen, P. (2003), ‘A note on design of fiber-nets for maximum stiffness’, J. of Elasticity 73, 127–145. Pedersen, P. (2005a), ‘Analytical stiffness matrices with Green-Lagrange strain measure’, Int. J. Numer. Meth. Engng. 62, 334–352. Pedersen, P. (2005b), ‘Axisymmetric analytical stiffness matrices with Green-Lagrange strains’, Computational Mechanics 35, 227–235. Pedersen, P. (2006), ‘Analytical stiffness matrices for tetrahedral elements’, Computer Methods in Applied Mechanics and Engineering 196, 261–278. Pedersen, P. and Pedersen, N. L. (2008), ‘Analytical optimal designs for long and short statically determinate beam structures’, Struct. Multidisc. Optim. pp. 1–15. on line. Pedersen, P. and Taylor, J. E. (1993), Optimal design based on power-law non-linear elasticity, in P. Pedersen, ed., ‘Optimal Design with Advanced Materials,’, Elsevier, pp. 51–66. Rozvany, G. N. I. (1989), Structural Design via Optimality Criteria, Kluwer, Dordrecht, The Netherlands. Save, M., Prager, W. and Sacchi, G. (1985), Structural optimization, optimality criteria, Vol. 1, Plenum Press. 61 c 62 Pauli Pedersen: Stationarity and extremum principles in mechanics Sundstrøm, B., ed. (1998), Handbok och Formelsamling i Hållfasthetslaera (in Swedish), KTH, Stockholm. 398 pages. Washizu, K. (1975), Variational methods in elasticity and plasticity, Pergamon Press, Ltd. Wasiutynski, Z. (1960), ‘On the congruency of the forming according to the minimum potential energy with that according to equal strength’, Bull. de l’Academie Polonaise des Sciences, Serie des Sciences Techniques 8(6), 259–268. Index admissible displacement field, 4 stress field, 4 Also best strength, 35 Analytical steps, 43 Assumed power law elasticity, 17 Assumptions, 11 Available functions, 40 Basic assumption, 37 beam cross sections rectangular, 53 solid circular, 53 solid equal sided triangular, 53 solid square, 53 thin walled, 53 Beam length, 44 Beam solutions by stress(complementary) principles, 30 Bending energy only, 29 Bending moment, 28 Bernoulli-Euler cantilever beams, 48 Bernoulli-Euler cantilever designs, 48 Bernoulli-Euler design formula, 47 cantilever, slender beam, 30 Castigliano’s 1st theorem, 7 Castigliano’s 2nd theorem, 9 Cauchy strains, 4 Clapeyron’s theorem, 6 Collected results, 41, 59 complementary energy density, 26 Complementary work of external loads, 5 complementary(stress) virtual work, 8 compliance decrease, 44 Compliance from elastic energy, 5 Compliance from potential of external loads, 5 compliance minimization, 43 Compliance objective, 44 Compliance of external loads, 5 Compliance per unit length, 45 Conditions with a simple shape parametrization, 38 Constant energy density, 36 constitutive individual parameter, 57 matrix non-dimensional, 57 positive semi-definite matrix, 57 Constitutive secant modulus, 24 Constitutive tangent modulus, 25 Converted to non-constrained, 34 corresponding displacement, 11 cross sectional reference length, 45 Cross sectional size, 44 cross-sectional constants, 27 cross-sectional forces/moments, 27 curvature, 11 dead load, 6, 13 Decoupled energies, 28 Derivatives of elastic potentials, 18 Design for stiffness and strength, 38 Design independent loads, 18, 36 Design optimization, 33 Design problem, 39, 57 Detail of proof, 37 differential strain energy density, 24 Direct conclusions, 58 direction principal strain/stress, 58 displacement control, 7 displacement field, 4 distribution of shear stresses, 45 domain non-shape, 37 shape, 37 effective strain, 24, 57 stress, 24 Effective strain/stress, 24 Elastic energy in a straight beam, 27 elastic strain energy, 17 elastic stress energy, 17 elasticity power law non-linear, 36, 57 Elementary case, 30 Energy densities in 1D non-linear elasticity, 22 Energy densities in 2D and 3D non-linear elasticity, 24 Energy equilibrium, 17 engineering strains, 4 Equilibrium, 17 equilibrium simplified, 17 Estimated multiplier, 41 Examples of beam cross sections, 54 explicit analytical solution, 43 External potential, 17 External potential and compliance, 6 extrema relations, 18 extremum principle for total complementary(stress)potential energy, 14 for total potential energy, 13 Final optimization, 58 finite element, 40 fixed stress field, 37 For model by the FEM, 40 force equilibrium, 4 Four energy solutions, 30 Free material, 57 Frobenius norm, 57 gamma-function, 39 General knowledge, 35 General optimization, 33 General stress/strain state, 13 63 c 64 Pauli Pedersen: Stationarity and extremum principles in mechanics geometrical constraint, 38 given kinematic conditions, 7 Goal of the chapter, 3 Good experience, 38 gradient of the elastic energy, 44 Frobenius, 58 Normal force, 27 Not well known but important, 17 Numerical aspects and comparison with isotropic material, 59 Handbook formulas, 27 heuristic approach successive iteration, 40 homogeneous energy relation, 34 mass relation, 34 Homogeneous mass(volume) dependence, 34 optimal design formulation, 44 optimal designs, 48 optimal material aligned, 58 degenerate, 59 no shear stiffness, 58 non-zero eigenvalue, 59 orthotropic, 58 Optimal modulus matrix, 58 Optimality condition, 40, 58 optimality condition, 34 Optimality criterion, 45 optimality criterion, 46 Optimality criterion for beam design, 44 Optimally obtained compliance, 49, 53 optimize material orientation, 33 stiffness, 33 strength, 33 outward normal, 3 Overview of principles and their relations, 14 identity, 4 Inequality for mixed product, 12 influence of beam length, 44 Internal potentials, 17 invariant matrix, 57 matrix length, 57 kinematically admissible, 4 Lagrange function, 34 multiplier, 34, 40 linear elastic material, 6 linear elasticity and dead loads, 6 Linear strain notation, 21 linear strains, 4 Local design parameter, 18 Localized determined sensitivities, 57 Localized energy change, 36 Localized volume change, 35 material design free material, 57 ultimate design, 57 material parameters, 27 matrix invariant, 57 maximize eigenfrequency, 43 stiffness, 43 maximum strain energy density, 57 minimize compliance, 44 minimum compliance, 37 maximum strain energy density, 37 total strain energy, 57 Minimum complementary (stress) potential, 14 Minimum potential, 13 mixed products, 12 Model for moment of inertia, 45 moment distribution, 45 moment equilibrium, 4 Motivation for extremum, 11 multidimensional stress/strain state, 13 necessary condition, 45 a single constraint, 34 non-constrained, 33 positive definite, 58 proportional gradient, 58 proportionality, 34 necessary optimality criterion, 43 Neglected shear energy, 29 New design problem, 58 No physical interpretation, 3 Non-constrained problems, 33 norm parametrization boundary shape, 38 super-ellipse, 38 Poisson’s ratio, 45 zero, 59 Positive definite, 58 Potential relations, 18 Power law elasticity, 22 Principal strains only, 58 principle of minimum total potential energy, 11 overview, 14 Principle of minimum total complementary(stress) potential energy, 14 Principle of minimum total potential energy, 11 principle of virtual displacements, 7 principle of virtual stresses, 8 Problems with a single constraint, 34 Proportional gradients, 34 Proportional relation, 24 psi-function, 40 ratio of the compliances, 49 real displacement field, 6 load field, 6 strain field, 6 stress field, 6 Real fields, 4 Real stress field and real displacement field, 6 Real stress field and virtual displacement field, 7 Relations with power law elasticity, 18 relative decrease in compliance, 52 Relative hole area or density, 39 Remark, 18 Results for simple (Bernoulli-Euler) beams, 29 Secant and tangent modulus, 23 secant formulation, 57 Sensitivity analysis, 57 sensitivity analysis localized, 57 shape optimization References 65 stiffness, 36 Shape optimization for stiffness and strength, 35 shear force distribution, 45 shear modulus, 45 shear stresses, 28 Simply supported static distributions, 47 Single force behaviour, 11 single load at the free end, 50 singularities of optimal beams, 50 six-dimensional strain/stress spaces, 13 Size and shape, 33 size optimization stiffness, 34 strength, 35 Size optimization for stiffness and strength, 34 specialized principles, 7 statically admissible, 4 Statically determinate beams, 43 statically determinate cases, 45 Stationary objective, 33 stationary total complementary potential, 19 stationary total potential, 18 Stationary total potential energy, 8 Stationary total stress potential energy, 9 Stiffest design, 35 stiffest design shape, 36 size, 35 strain effective, 57 principal, 57 Strain and stress energy densities, 21 strain by differentiation, 21 Strain energy density, 5, 21 strain field, 4 strength von Mises, 35 Strength or stiffness, 40 stress by differentiation, 21 stress energy complementary formulation, 37 Stress energy density, 5, 21 stress field, 4 Stress virtual work principle, 8 strongest design shape, 38 size, 35 successive iteration heuristic approach, 40 Summing up chapter 8, 54 chapter 6, 31 chapter 5, 26 chapter 3, 16 chapter 7, 41 chapter 4, 19 chapter 2, 9 chapter 9, 59 super-elliptic shape, 38 surface area, 7 surface traction’s, 4, 17 Symbols and definitions, 4 tangent modulus, 23 Tensor notation, 3 The identity, 4 The individual constitutive parameters, 57 The work equation, 4 The work equation, an identity, 3 theorem stiffest shape design, 37 strongest shape design, 38 theorem of divergence, 3 Timoshenko beam theory, 43 Timoshenko cantilever beams, 49 Timoshenko cantilever designs, 51 Timoshenko design formula for n = 1, 47 Timoshenko design formula for n = 2, 47 Torsional moment, 28 total external force, 45 Total potential complementary energy, 5 Total potential energy, 5 Total potentials, 18 Total strain energy, 5 Total stress energy, 5 Transverse force, 28 triangular load distribution, 51 Two or only one parameter, 38 Two-sided bounds, 14 Uniaxial constitutive model, 13 uniform load, 50 uniform strain energy density, 35 unit displacement field, 8 Unit displacement theorem for linear elasticity, 8 Unit load theorem for linear elasticity, 9 Use of optimal design results, 44 Using the theorem of divergence, 3 variational strain, 21 variational strain energy density, 21 variational stress energy density, 21 Virtual fields, 4 Virtual stress field and real displacement field, 8 Virtual work principle, 7 volume forces, 4 work equation, 4 Work function, 12 Work of external loads, 5 Young’s modulus, 45 Zero sum of total potentials, 6
© Copyright 2026 Paperzz