Article

Mathematics and Mechanics of Solids 000(00): 1–39 ©The Author(s) 2011 Nonlinear morphoelastic plates II: Reprints and permission: sagepub.co.uk/journalsPermissions.nav Exodus to buckled states DOI: 10.1177/1081286510387234 mms.sagepub.com

Joseph McMahon Program in Applied , University of Arizona, Tucson, USA

Alain Goriely OCCAM, Mathematical Institute, University of Oxford, UK

Michael Tabor Program in Applied Mathematics, University of Arizona, Tucson, USA Department of Mathematics, University of Arizona, Tucson, USA

Received May 7 2010; accepted July 15 2010

Abstract Morphoelasticity is the theory of growing elastic materials. The theory is based on the multiplicative decomposition of the deformation and provides a formulation of the deformation and stresses induced by growth. Following a companion paper, a general theory of growing non-linear elastic Kirchhoff plate is described. First, a complete geometric description of incompatibility with simple examples is given. Second, the stability of growing Kirchhoff plates is analyzed.

Keywords buckling, growth, Kirchhoff plates, non-linear elasticity, residual stress

1. Introduction In a companion paper [1] we demonstrated that residual stress can be generated through growth in a non- linear Kirchhoff plate. It has been established that under appropriate circumstances, residual stress induced through growth can be sufficient to create instabilities in an elastic three-dimensional material [2, 3]. It is therefore a natural question to investigate whether such stress fields can be sufficient to initiate buckling in plates. Surprisingly, the case of two-dimensional structures has not been analyzed within the framework of non-linear elasticity, although there in now a growing body of literature in the community, all related to extensions of the Föppl-von Kármán equations to include the effect of growth through extra terms [4, 5] (see McMahon et al. [1] for a general introduction and citations).

Corresponding author: Alain Goriely, OCCAM, Mathematical Institute, University of Oxford, 24-29 St Giles’, Oxford OX1 3LB, UK. Email: [email protected]

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 2 Mathematics and Mechanics of Solids 000(00)

Here, we combine the modeling of incompatible growth through a finite multiplicative decomposition with the global constraint principle. The global constraint principle has clarified and simplified the theory of finite deformations of rods and shells and allows for geometrically exact models to be constructed for a wide variety of [6, 7]. We demonstrate that such theories can be easily adapted to include growth. Further, we solve explicit examples of plate models and find residual stress and buckling caused by the elastic response to growth. Whereas, the notation and basic theory follows the discussion introduced by McMahon et al. [1], we further analyze the general theory of incompatibility based on differential before analyzing the buckling problem. The structure of the paper is as follows. In Section 2 we review the basic kinematics of finite deformations of solids and show its relation to the geometry of differentiable . Using the notation and tools of kinematics and , we construct concrete examples of continuously differentiable two-point fields with positive determinant that are not deformation . We expose three distinct ways in which such tensor fields can fail to meet the requirements of a deformation gradient. Such tensor fields are called incompatible growth fields. In Section 3 we introduce the multiplicative decomposition of the deformation gradient into growth and elastic components. The proper method for incorporating the multiplicative decomposition into constitutive relations is shown. Section 4 introduces the global constraint principle by which rod and shell theories are derived. In Section 5 we derive the kinematics of a buckled axisymmetric Kirchhoff plate, find the virtual displacements from such a configuration, and apply the global constraint principle to derive balance equations, which are simplified in Section 6. In Section 7 we introduce a non-linearly elastic constitutive relation, modify it with the multiplicative decomposition of the deformation gradient, and import the result into the balance equations. Section 8 explores the appearance of buckled solutions alongside un-buckled configurations. We find the impact of thickness on the onset of buckling in response to incompatible growth. In Section 9 we find large buckling in another example of elastic response to growth.

2. Geometry and kinematics of deformation and growth 2.1. Coordinates and E3 The space in which a solid body resides is three-dimensional Euclidean space E3, which consists of (spatial) points, each of which has an inner-product space of vectors attached to it. Here E3 has a fixed right- handed trio of orthogonal Cartesian axes, and each has unit vectors i, j,andk parallel to these axes. We also use cylindrical coordinates (R, , Z). The unit tangent vectors h1, h2,andk point in the directions of increasing R, ,andZ, respectively. They are related to the Cartesian tangent vectors by

h1() = cos i + sin j,

h2() =−sin i + cos j.(1)

Although E3 consists of spatial points, in most cases we refer to a point not by its coordinates, but by the anchored at the origin that points to the point.

2.2. with cylindrical coordinates We treat a solid body as a with boundary, which is a set of material points that can be assigned coordinates in an invertible manner, at least locally. In the case at hand, we consider a multiply connected manifold B that is assigned cylindrical coordinates (R, , Z). We may need to employ different branches of the azimuthal angle for situations such as those illustrated in Figures 1 and 2. The map from B to coordinate points is called a , and (B) will be called the space of coordinates.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 3

Figure 1. One chart for cylindrical coordinates assigns coordinates between 0 and 2π. The function X for this chart sends this portion of R3 into E3.

Figure 2. Some subsets of Bo (the open interior of B) may straddle the line that corresponds to = 0 = 2π in the chart in Figure 1. To avoid a disconnected image of the subset in R3, we choose a different chart, which assigns coordinates between π and 3π. The position function X for this chart sends this portion of R3 into E3.

Now that the body B is represented by a collection of coordinate points, we define the position function X : (B) → E3 by

X (R, , Z) = R cos i + R sin j + Z k = Rh1() + Zk.(2)

The map X ◦ takes each point p ∈ B and assigns it a position (X ◦ )(p) = X(R(p), (p), Z(p)) in E3 in an invertible manner. The set (X ◦ )(B) will often be called the ‘Euclidean body’. Note that the position function X maps to the same point in E3 regardless of whether the chart in Figure 1 or the chart in Figure 2 is used.

2.3. Tangent vectors 2.3.1. Tangent vectors in E3 To each point X in the Euclidean body there is attached a three-dimensional real whose elements are called tangent vectors. At X a for this tangent space is provided by the three partial of the position function X:

∂X E = = cos i + sin j,(3) R ∂R ∂X = X =− + E ∂ R sin i R cos j,(4) ∂X E = = k.(5) Z ∂Z

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 4 Mathematics and Mechanics of Solids 000(00)

Figure 3. Each tangent vector at a point X(R, , Z) in the Euclidean body corresponds to a unique tangent vector at (R, , Z) ∈ (B) − and to a unique abstract tangent vector at 1(R, , Z) ∈ B.

This basis is called the coordinate basis. There is a corresponding dual basis of vectors Ej defined so that j · = δj E Ei i , the : ER = cos i + sin j,(6) E =−R−1 sin i + R−1 cos j,(7) EZ = k.(8)

The inner products Eij = Ei · Ej form a symmetric, positive-definite : ⎛ ⎞ ⎛ ⎞ ERR ER ERZ 100 ⎝ ⎠ ⎝ 2 ⎠ ER E EZ = 0 R 0 ;(9) EZR EZ EZZ 001 while the inner products Eij = Ei · Ej form the inverse matrix: ⎛ ⎞ ⎛ ⎞ ERR ER ERZ 100 ⎝ ER E EZ ⎠ = ⎝ 0 R−2 0 ⎠ . (10) EZR EZ EZZ 001

2.3.2. Tangent vectors in R3 For each tangent vector at a point X(R, , Z) in the Euclidean body there is a unique tangent vector at (R, , Z) ∈ (B). In the space of coordinates, the tangent vectors are represented by partial differential operators. The correspondence is ∂X ∂ E = at X(R, , Z) ←→ at (R, , Z), R ∂R ∂R ∂X ∂ = X ←→ E ∂ at X (R, , Z) ∂ at (R, , Z), ∂X ∂ E = at X(R, , Z) ←→ at (R, , Z). (11) Z ∂Z ∂Z There is also a one-to-one correspondence between the vectors Ei and the linear functionals on the vectors at (R, , Z): ER at X(R, , Z) ←→ dR at (R, , Z), E at X(R, , Z) ←→ d at (R, , Z), EZ at X(R, , Z) ←→ dZ at (R, , Z), (12) where the linear functionals are defined so that   ∂ dξ j = δj,whereξ R = R, ξ = , ξ Z = Z, (13) ∂ξi i

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 5 where we have used square brackets to indicate that ∂/∂ξi is the argument of dξ j. We need this notation for tangent vectors at (R, , Z) distinct from those at X(R, , Z) because we will be considering situations in which there is either no position function X from (B) or the position function is not globally one-to-one, so that two 3 distinct coordinate points (R1, 1, Z1)and(R2, 2, Z2) may be mapped to the same point in E .

2.4. Apriori, the tangent space at (R, , Z) ∈ (B) is just a real three-dimensional vector space, not an inner- product space. A Riemannian metric tensor is a smooth1, symmetric, positive-definite quadratic form on tangent vectors in (B). A metric tensor is used to assign sizes to vectors and, when used as a bilinear form, to compute the analogues of dot-products between vectors. A differentiable manifold equipped with a Riemannian metric tensor is called a Riemannian manifold. When a position function X is present, the tangent space at X(R, , Z) is an inner-product space, and we want the tangent space at (R, , Z) to ‘inherit’ the inner product. That is, we want an inner product  ,  such that

∂ ∂ = · = ∂ , ∂ ER ER 1, R R ∂ ∂ = · = 2 ∂, ∂ E E R ,

∂ ∂ , = E · E = 1, (14) ∂Z ∂Z Z Z with all other inner products of basis vectors equal to zero. Using the linear functionals in Equation (12), we can construct a bilinear form that satisfies these relations: M = dR ⊗ dR + R2 d ⊗ d + dZ ⊗ dZ. (15) Note the correspondence of M to the tensor M = ER ⊗ ER + R2E ⊗ E + EZ ⊗ EZ, (16) which can be used to define the inner products of ER, E,andEZ. If we treat the functions Eij = Ei · Ej as functions of (R, , Z) and ignore their original definition, M in Equation (15), which is known as a metric tensor, can be expressed as

i j M = Eij dξ ⊗ dξ . (17) Among other things, the metric tensor defines arclengths of . Let [a, b] s → γ (s) = (R(s), (s), Z(s)) be a continuously differentiable in (B). The length of γ is computed with the metric tensor as follows:

b ξ i ξ j b d d 2 2 2 Eij(γ (s)) ds = (R (s)) + (R(s) (s)) + (Z (s)) ds. (18) a ds ds a By design of M, the length in Equation (18) is the same as the Euclidean length of the Euclidean image X ◦ γ of the curve. Whenever there is a position function X, the metric tensor can be used to convert a triple of coordinate (dR/ds, d/ds, dZ/ds) into the squared norm of the corresponding Euclidean . First we write the triple of velocities as the tangent vector dR ∂ d ∂ dZ ∂ + + . (19) dS ∂R ds ∂ ds ∂Z Then we compute the inner product       dR ∂ d ∂ dZ ∂ dR ∂ d ∂ dZ ∂ dR 2 d 2 dZ 2 + + , + + = + R2 + . (20) dS ∂R ds ∂ ds ∂Z dS ∂R ds ∂ ds ∂Z ds ds ds

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 6 Mathematics and Mechanics of Solids 000(00)

θ

−1 Figure 4. A deformation of a solid body is computed in three steps. X maps each Euclidean point of the body to its correspond- ing coordinate point. f maps reference coordinate points to deformed coordinate points. The position function x maps deformed −1 coordinate points into E3. The full deformation is χ = x ◦ f ◦ X .

A tangent vector at (R, , Z) ∈ (B) corresponds to an abstract tangent vector at −1(R, , Z) ∈ B. Similarly, the metric tensor at (R, , Z) also corresponds to a symmetric, positive-definite quadratic form on tangent vectors at −1(R, , Z). Equipped with this metric tensor, which can also be used as a bilinear form, each tangent space at a point in B becomes an inner-product space, and B with the abstract metric tensor is an example of a Riemannian manifold, for which is a coordinate system and M is the metric tensor expressed in that coordinate system.

2.4.1. Deformation and change of metric tensor Consider a deformation χ in which the deformed configuration has coordinates (r, θ, z) and corresponding position function x(r, θ, z) = r cos θi + r sin θ j + zk. (21) Just as the position function X in Equation (2) induces the metric tensor in Equation (15) on the space of (R, , Z) points, the position functionx induces a metric tensor on the space of (r, θ, z) points: m = dr ⊗ dr + r2 dθ ⊗ dθ + dz ⊗ dz. (22) If we use the notation (ξ 1, ξ 2, ξ 3) = (R, , Z) for the reference coordinates and (ζ 1, ζ 2, ζ 3) = (r, θ, z)for the deformed coordinates, the deformation gradient can be written as ∂ζj F = e ⊗ Ei, (23) ∂ξi j where ∂x e = = cos θi + sin θj = h (θ), (24) r ∂r 1 ∂x = x =− θ + θ = θ eθ ∂θ r sin i r cos j rh2( ), (25) ∂x e = = k. (26) z ∂z

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 7

The deformation gradient maps tangent vectors at X ∈ (X ◦ )(B) to the corresponding, deformed tangent vectors at x = χ(X). The right Cauchy–Green tensor FT · F associated with F provides a quadratic form akin to the one in Equation (16):     ∂ζj ∂ζ T i k F · F = E ⊗ e · e ⊗ E ∂ξi j ∂ξk ∂ζj ∂ζ i k = e E ⊗ E (27) j ∂ξi ∂ξk

T where ej = ej · e .HereF · F allows the computation of deformed arclengths while using the reference coordinates (R, , Z). For example, if [a, b] s → γ (s) = (R(s), (s), Z(s)) is a continuously in (B), then the velocity of its Euclidean image is dR d dZ v = E + E + E , (28) ds R ds ds Z and its arclength after deformation is

b b  F · v ds = v · FT · F · v ds a a  b ∂ζj ∂ζ dξ i dξ k = e ds (29) j i k a ∂ξ ∂ξ ds ds

The metric tensor on (B) corresponding to the right Cauchy–Green tensor is

∂ζj ∂ζ i k e dξ ⊗ dξ , (30) j ∂ξi ∂ξk which is used as a quadratic form on velocity vectors in (B)oftheform

dξ i ∂ dR ∂ d ∂ dZ ∂ v = = + + . (31) ds ∂ξi ds ∂R ds ∂ ds ∂Z If v is the velocity vector of the curve s → γ (s)in(B), then the arclength assigned by the metric tensor in Equation (30) is       b ∂ζj ∂ζ b ∂ζj ∂ζ dξ m ∂ dξ n ∂ e (dξ i ⊗ dξ k)[v, v] ds = e dξ i dξ k ds j i k j i k m n a ∂ξ ∂ξ a ∂ξ ∂ξ ds ∂ξ ds ∂ξ  b ∂ζj ∂ζ dξ m dξ n = e δi δk ds j i k m n a ∂ξ ∂ξ ds ds  b ∂ζj ∂ζ dξ m dξ n = e ds. (32) j m n a ∂ξ ∂ξ ds ds

The metric tensor in Equation (30) is called the pull-back of the metric tensor m in Equation (17), under the map f (R, , Z) = (r, θ, z).

2.5. Incompatible growth Each fixed two-point tensor with positive determinant can be the value of a deformation gradient at a point, but not every two-point tensor field with positive determinant is equal to a deformation gradient. A continuously

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 8 Mathematics and Mechanics of Solids 000(00) differentiable two-point tensor field G with positive determinant will be called an incompatible growth field if it is not equal to a deformation gradient. Even if G is an incompatible growth field, it can be used to create a new metric tensor in the same fashion that a deformation gradient can. Let = i ⊗ j G Gjei E . (33) Then the analogue of the right Cauchy–Green tensor is

T · = j i ⊗ k G G ej GiGkE E . (34)

This is a symmetric, positive-definite quadratic form on tangent vectors at X ∈ (X ◦ )(B), and it corresponds to the metric tensor j ξ i ⊗ ξ k ej GiGk d d (35) on (B). Before we can demonstrate the ways in which G can fail to be a deformation gradient, we must delve a little deeper into differential geometry.

2.6. Immersibility and embeddability 2.6.1. Immersions, embeddings, and isometry An immersion of a set ⊂ R3 of coordinate points (ξ 1, ξ 2, ξ 3)isa differentiable function Y : → E3 such that the gradient of Y has positive determinant throughout :   ∂Y ∂Y ∂Y det > 0. (36) ∂ξ1 ∂ξ2 ∂ξ3

This ensures that no volumes in are collapsed to areas in E3, and that no volumes undergo a reversal of i j orientation. If is equipped with a metric tensor Eij dξ ⊗ dξ , an immersion Y is called isometric (with the metric tensor) if ∂Y ∂Y · = E , i, j = 1, 2, 3, in . (37) ∂ξi ∂ξj ij An immersion is locally invertible, but it need not be globally so. An immersion Y of is called an embedding if it is invertible throughout . An embedding is called isometric if it satisfies Equation (37).

2.6.2. Riemann–Christoffel curvature tensor Not every Riemannian metric has an isometric immersion into E3.There is a test for the existence of such a map from coordinates to E3. The test involves the Riemann–Christoffel curvature tensor, whose computation we now show. i j If Eij dξ ⊗ dξ is a Riemannian metric, then the of the first kind are defined by   1 ∂E ∂E ∂E = ik + jk − ij , (38) ijk 2 ∂ξj ∂ξi ∂ξk and the Christoffel symbols of the second kind are defined by

= k ij E ijk, (39)

ij −1 where [E ] = [Eij] . In classical tensor analysis, the Christoffel symbols of the second kind are defined as the coefficients used to express the rates of change of coordinate tangent vectors in Euclidean space

∂Ej = E , (40) ∂ξi ij

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 9

and the Christoffel symbols of the first kind are defined by ‘lowering’ the raised index in ij: = ijk Ek ij. (41) The Riemann–Christoffel curvature tensor R is a four-indexed tensor whose components satisfy ∂ ∂ ik ij m m R = − + − . (42) ijk ∂ξj ∂ξk ij k m ik j m

The test is embodied in the three-dimensional case of the ‘Fundamental Theorem of ’, as described in Theorem 1.6-1 of Ciarlet [8], which we paraphrase as follows.

3 2 Theorem 2.1. Let be an open, simply connected set in R and let C = [Eij] be a C symmetric positive- definite matrix-valued function on that satisfies

R ≡ O in ,

3 where R ijk is defined by Equation (42). Then there exists a C immersion YYof such that

∂Y ∂Y E = · in . ij ∂ξi ∂ξj

For a deeper exploration, see Ciarlet [8] or McMahon [9]. We consider simple cases in which isometric embeddability fails.

2.7. Example: immersion without embedding We construct an example by considering two manifolds with cylindrical coordinate systems, metric , and position functions isometric with the given metric tensors:

manifold B coordinate system = (R, , Z) metric tensor M = dR ⊗ dR + R2 d ⊗ d + dZ ⊗ dZ M = ER ⊗ ER + R2 E ⊗ E + EZ ⊗ EZ isometric position function X(R, , Z) = R cos() i + R sin() j + Zk manifold M coordinate system ψ = (r, θ, z) metric tensor m = dr ⊗ dr + r2 dθ ⊗ dθ + dz ⊗ dz m = er ⊗ er + r2 eθ ⊗ eθ + ez ⊗ ez isometric position function x(r, θ, z) = r cos(θ) i + r sin(θ) j + z k

In each case we have also included the ‘Euclidean image’ of the metric tensor under the isometric position functions, where

er = cos(θ) i + sin(θ) j, eθ =−r−1 sin(θ) i + r−1 cos(θ) j, ez = k. (43)

We consider a two-point tensor field of the form γ R 2 Z G = γ1 er ⊗ E + eθ ⊗ E + ez ⊗ E , (44) γ1

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 10 Mathematics and Mechanics of Solids 000(00)

where γ1 and γ2 are positive constants. If this were a deformation gradient, the reference coordinates-to- deformed coordinates part of the deformation would have the form

γ2 r = γ1R, θ = , z = Z. (45) γ1

The analogue of the right Cauchy–Green tensor is     γ γ T · = γ R ⊗ + 2 ⊗ + Z ⊗ · γ ⊗ R + 2 ⊗ + ⊗ Z G G 1 E er γ E eθ E ez 1 er E γ eθ E ez E 1  1 γ 2 2 R R 2 Z Z = γ e E ⊗ E + eθθ E ⊗ E + e E ⊗ E 1 rr γ zz  1 γ 2 = γ 2 R ⊗ R + 2 2 ⊗ + Z ⊗ Z 1 E E r E E E E γ1 = γ 2 R ⊗ R + γ 2 2 ⊗ + Z ⊗ Z 1 E E 2 R E E E E . (46)

This corresponds to the metric tensor

γ 2 ⊗ + γ 2 2 ⊗ + ⊗ 1 dR dR 2 R d d dZ dZ (47) on (B). Applying Equations (38), (39), and (42) to the coefficients of this metric yields an identically zero Riemann–Christoffel curvature tensor. Theorem 2.1 ensures the existence of an isometric immersion of any simply connected subset of (B). But we will now show that any such isometric immersion is not an isometric embedding. Consider the mapsx ◦ f : (B) → E3 shown in Figures 5 and 6:   γ x ◦ = x γ 2 (x f )(R, , Z) x 1R, γ , Z 1    γ2 γ2 = γ1R cos i + γ1R sin j + Z k. (48) γ1 γ1

The coordinate basis vectors with respect to (R, , Z)are     ∂ x ◦ γ γ (x f ) = γ 2 + γ 2 ∂ 1 cos γ i 1 sin γ j R 1  1   ∂(x ◦ f ) γ2 γ2 =−γ2R sin i + γ2R cos j ∂ γ1 γ1 ∂(x ◦ f ) = k, (49) ∂Z whose dot products form a symmetric positive-definite matrix: ⎛ ⎞ γ 2 1 00 ⎝ 2 ⎠ 0 γ2R 0 . (50) 001

Thus,x ◦ f maps (B)intoE3 in a fashion isometric with the metric in Equation (47). As illustrated in Figures 5 and 6, however, this isometric immersion is not an embedding if γ1 = γ2. A theorem on the ‘rigidity’ of the class of isometric immersions shows that the problems found inx ◦ f will be found in any isometric immersion of (B) with the metric tensor in Equation (47).

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 11

3 3 Figure 5. If γ2 >γ1,thenx ◦ f : (B) → E double-covers part of E .

3 Figure 6. If γ2 <γ1,thenx ◦f does not map (B) to a complete annulus in E . This map is not continuous on the original Euclidean body; it represents a ripping of the annulus.

Theorem 2.2. Let be an open, connected subset of R3, and let Y and ZZbetwoC1 immersions such that their associated metric tensors satisfy ∂Y ∂Y ∂Z ∂Z · = · , in . ∂ξi ∂ξj ∂ξi ∂ξj

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 12 Mathematics and Mechanics of Solids 000(00)

Then there exist a constant vector c and a constant orthogonal matrix Q such that

Z(ξ 1, ξ 2, ξ 3) = Q · Y (ξ 1, ξ 2, ξ 3) + c for each (ξ 1, ξ 2, ξ 3) ∈ .

See Theorem 1.7-1 of Ciarlet [8]. No combination of constant rotation, reflection, and translation of x ◦ f will solve the problems displayed 3 in Figures 5 and 6. If γ1 = γ2, the annulus cannot be embedded in E in a fashion that agrees with the metric −1 in Equation (47). The map χ = x ◦ f ◦ X described above is a local deformation on subsets of the annulus, but it is not a global deformation. Hence, G is not equal to a global deformation gradient; it is an incompatible growth field.

2.8. Lack of immersion Consider next a growth field of the form γ R 2R Z G = γ e ⊗ E +  eθ ⊗ E + e ⊗ E , (51) 1 r R γ z 0 1(s)ds where γ1 and γ2 are positive, continuously differentiable functions of R,and

R γ (R)R r(R) = γ (s) ds, θ(R, ) =  2 , z = Z. (52) 1 R γ 0 0 1(s) ds The analogue of the right Cauchy–Green tensor is   γ 2 T 2 R R 2R Z Z · = γ ⊗ +  θθ ⊗ + ⊗ G G 1 err E E R e E E ezz E E γ1(s) ds  0  2 γ R = γ 2 ER ⊗ ER +  2 r2 E ⊗ E + EZ ⊗ EZ 1 R γ 0 1(s) ds = γ 2 R ⊗ R + γ 2 2 ⊗ + Z ⊗ Z 1 E E 2 R E E E E , (53) which corresponds to the metric tensor γ 2 ⊗ + γ 2 2 ⊗ + ⊗ 1 dR dR 2 R d d dZ dZ (54) on (B). Applying Equations (38), (39), and (42) to the coefficients of the metric in Equation (54) yields a Riemann–Christoffel curvature tensor with four entries that may be non-zero: γ    R =−R = R =−R = R 2 γ γ + γ γ − γ γ + γ RR RR RR RR 2 1 R 1 2 1 2 2 R 2 . (55) γ1 Unless γ + Rγ 2 2 = constant, (56) γ1 the two-point tensor field G in Equation (51) is not even locally a deformation gradient. It is another example of an incompatible growth field.

2.9. Non-metric complications Note that computing the Riemann–Christoffel curvature tensor tests properties of GT ·G and not of G alone. We provide here an example of a two-point tensor field with positive determinant and identically zero associated Riemann–Christoffel curvature that is not a deformation gradient, even locally.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 13

Let R Z G = (cos φ er − sin φ ez) ⊗ E + (− sin φ er + sin φ ez) ⊗ E + eθ ⊗ E , (57) with r = R, θ = ,andz = Z.Hereφ is an R-dependent angle. This corresponds to a rotation of the vectors er and ez in the θ = constant . The analogue of the right Cauchy–Green tensor is     T 2 2 R R 2 2 Z Z G · G = err cos φ + ezz sin φ E ⊗ E + eθθ E ⊗ E + err sin φ + ezz cos φ E ⊗ E = ER ⊗ ER + R2 E ⊗ E + EZ ⊗ EZ, (58) which is the same as the Euclidean metric tensor induced by the standard position function X from cylindrical coordinates to E3. The associated Riemann–Christoffel curvature tensor is identically zero. This tells us that GT · G is equal to the right Cauchy–Green tensor for some isometric immersion. It does not imply, however, that G is itself the gradient of an isometric immersion. Consider the form of a deformation gradient using cylindrical coordinates (R, , Z)and(r, θ, Z):     ∂r ∂r ∂r ∂θ ∂θ ∂θ = ⊗ R + + Z + ⊗ R + + Z F er ∂ E ∂ E ∂ E eθ ∂ E ∂ E ∂ E  R  Z   R  Z   GRAD r  GRAD θ ∂z ∂z ∂z + ⊗ R + + Z ez ∂ E ∂ E ∂ E . (59)  R  Z  GRAD z

If G in Equation (57) is a deformation gradient, then

GRAD r = cos φ ER − sin φ EZ, (60) and the vector field on the right-hand side should have identically zero . However,   CURL cos φ ER − sin φ EZ = Rφ cos φ E, (61) so for non-constant φ the two-point tensor field G in Equation (57) is not a deformation gradient. All information about orientation is absent from GT ·G.HereG fails to be even a local deformation gradient because it describes a re-orientation of material fibers without any bending, stretching, or compression of the body. If the material fibers were re-oriented in this fashion in a deformation, the result of this microscopic re-orientation would be a macroscopic change in the shape of the body. Considering incompatible growth such as that in Equation (57) requires more advanced differential geom- etry, including the theory of linear connections. Incompatible growth of this kind will not be considered here.

3. Deformation and hyperelasticity 3.1. The multiplicative decomposition We imagine incompatible growth to be non-elastic, i.e. it generates no elastic strain energy. A Riemannian manifold that undergoes incompatible growth has intrinsic arclengths that preclude it from being isometrically embedded in E3. Since a solid body must be embedded isometrically in E3, the transformation of an incom- patibly grown body is incomplete without some distortion that changes the ‘grown’ metric tensor to one that allows isometric embedding. We assume that the body has an elastic response to the incompatible growth, and we assume that the full deformation gradient F can be decomposed as in Rodriguez et al. [10] and related works:

F = A · G, (62)

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 14 Mathematics and Mechanics of Solids 000(00) where A is a two-point tensor field with positive determinant that describes the distortion caused by the body’s elastic response to the incompatible growth field G. A rough description is that the incompatible growth, described by G, transforms infinitesimal neighborhoods in such a way that they cannot fit together in E3,and the elastic response, described by A, transforms these neighborhoods so that they again form a solid body in E3, without overlaps or voids.

3.2. Constitutive relation of an elastic body with growth 3.2.1. Traditional hyperelasticity A growthless hyperelastic body has a strain-energy density W(F) (a function of the Cauchy–Green tensor FT · F in an isotropic material) with units elastic energy [W(F)] = . (63) reference volume The nominal stress tensor, which is the of the first Piola–Kirchhoff stress tensor, is ∂ S = W(F), (64) ∂F so the units of S are [W(F)] elastic energy / reference volume elastic energy / deformed length [S] = = = . (65) [F] deformed length / reference length reference area If we recognize elastic energy per final length as elastic force in the deformed configuration, then the units of S reflect the role of the first Piola-Kirchhoff stress as a used to convert area in the reference configuration into force in the deformed configuration. The T is related to S by

− T = (det F) 1 F · S, (66) so the units of T are

[T] = [det F]−1 [F][S]   −1 = deformed volume deformed length · elastic energy / deformed length reference volume reference length reference area elastic energy / deformed length = . (67) deformed area These units reflect T’s role as a linear map used to convert area in the deformed configuration to force in the deformed configuration.

3.2.2. Hyperelasticity with growth In a body with growth included via the multiplicative decomposition in Equation (62), we distinguish three states of the body. The pre-growth state is the traditional reference configuration; the post-growth, pre-elastic response state (described by a non-isometrically embeddable Riemannian manifold) is called intermediate; and the final (Euclidean) configuration is found after the elastic response to the growth. The lengths, areas, and volumes of these three states will be considered distinct. In this model, the elastic strain-energy density’s argument is A, the sole (first-order) descriptor of the elastic response. Just as F maps from a tangent space in the reference configuration to a tangent space in the deformed configuration in traditional hyperelasticity, A maps from a tangent space in the intermediate configuration to a tangent space in the final configuration. Considering the units of W(F) in Equation (63), we conclude that the units of W(A)are elastic energy [W(A)] = . (68) intermediate volume

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 15

The analogue of the nominal stress has units   ∂ elastic energy / intermediate volume elastic energy / final length W(A) = = , (69) ∂A final length / intermediate length intermediate area while the analogue of the Cauchy stress has units   − ∂ (det A) 1 A · W(A) ∂A   −1 = final volume · final length · elastic energy / final length intermediate volume intermediate length intermediate area elastic energy / final length = . (70) final area A value of this analogue of the Cauchy stress is a linear map that is used to convert area in the final config- uration to force in the final configuration. Since the elastic strain-energy depends only on the elastic response A, this is the true Cauchy stress for a body in which the deformation gradient is decomposed as F = A · G. The formula above provides the form of the (Eulerian) Cauchy stress tensor for a hyperelastic body with growth, but we will need the first Piola–Kirchhoff stress tensor P to express balance laws. In hyperelasticity with growth, P remains a linear map from tangent spaces in the reference configuration to tangent spaces in the final (post-growth, post-elastic response) configuration, and it is related to the Cauchy stress tensor in the traditional manner: = · −T P (det F) T F   det F ∂ = · · −T A ∂ W(A) F det A  A  ∂ = (det G) A · W(A) · F−T. (71) ∂A A more complete argument for importing incompatible growth into elasticity in this fashion can be found in Chen and Hoger [11].

4. Reduced plate theory In a reduced theory of shells or rods, the deformations allowed are restricted to a specific class. In the case of plates (a special type of shell), for example, the body is allowed to deform to any in a family of ‘plate-like’ configurations. This is different from a constraint such as incompressibility in that restriction to such a class of deformations cannot be expressed as a set of local constraints. With global constraints in place, weak forms of balance laws are expressed in terms of virtual displacements tangent to the (infinite-dimensional) manifold of allowed deformations. The theory is developed in Antman and Marlow [6], Marlow [7], and Chapter XIV of Antman [12]. Further, it is assumed that the first Piola–Kirchhoff stress tensor can be decomposed into the sum of an active, constitutively defined portion plus a reactive or latent portion:

P = Pact + Plat. (72)

Here Plat is the portion of the stress that keeps the body plate-like, and it is assumed to do no work in the sense that  ∂x Plat : dV = 0 (73) (X◦)(B) ∂X  for each virtual displacement x tangent to the manifold of allowed configurations, where (X ◦ )(B)isthe (Euclidean) reference configuration of the body. This combination of assumptions forms the global constraint

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 16 Mathematics and Mechanics of Solids 000(00)

principle. We use Plat to denote that portion of the stress that constrains the body to its plate-like configuration but performs no work in doing so. The equations of equilibrium can be expressed in weak form as

  ∂x  (Pact + Plat) : dV = τ · x dS, (74) (X◦)(B) ∂X ∂(X◦)(B) where ∂(X ◦ )(B) is the radial boundary of the reference configuration and τ is the applied traction vector at a point on this boundary. By Equation (73), the equations of equilibrium of the constrained body have the weak form   ∂x  Pact : dV = τ · x dS, (75) (X◦)(B) ∂X ∂(X◦)(B)  where x is again a virtual displacement tangent to the manifold of allowed configurations. As pointed out in Section 4 of Marlow [7], Equation (75) will amount to an condition, as opposed to a stronger pointwise condition, that the active stress Pact will be required to satisfy on the boundary. It is through simplifications such as this that the so-called global constraint principle makes non-linear rod and shell problems manageable.

5. The Kirchhoff plate 5.1. Kirchhoff constraints In a Kirchhoff plate, the final configuration is assumed to have the form

(x ◦ f )(R, , Z) = r(R, ) + Zd(R, ), (76) where the function r describes a two-dimensional ‘middle surface’ of the plate, and the director field d is a unit vector-valued function that satisfies      ∂r ∂r  ∂r ∂r  d = ×  ×  . (77) ∂R ∂ ∂R ∂

Thus, d can be viewed as a function of r and its partial derivatives. The three-dimensional Kirchhoff plate consists of a two-dimensional surface from which one-dimensional material fibers sprout in a perpendicular direction. See Figure 7. Momentarily viewingx as a function of r and its derivatives for notational convenience, we find that a virtual displacement from this class of configurations has the form

        1  ∂r ∂ r ∂r ∂ r ∂r ∂r x = lim x r + r, +  , +  −x r, , →0  ∂R ∂R ∂ ∂ ∂R ∂       1  (r, +r, ) × (r, +r, ) r, ×r, = lim r + r + Z R R − r + Z R →   0  r,R ×r, (r,R +r,R ) × (r, +r, )    r, ×r, +r, ×r, = r + Z R R , (78) r,R ×r,

 where r is an as-regular-as-needed function of R and . We could derive the vector equation of equilibrium by considering the weak form of the method of virtual work with this virtual displacement, but there is an easier but less direct method. The director field must satisfy the (local) constraints d · d = 1andd · r,α = 0, α = R, . (79)

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 17

Figure 7. A radial slice of an axisymmetric Kirchhoff plate. ζ is the height of a point of the middle surface, and a material point’s height above the middle surface is measured along the unit vector d, which is perpendicular to the middle surface. When measured from the middle surface, the heights of the top and bottom surfaces are, respectively, H1 and H2.

 Encoding this into the expression Pact : ∂x/∂X is complicated, even in the axisymmetric case. If we introduce Lagrange multipliers λ, βR,andβ, we can consider instead the integral of the following:          k α τ · r + Zd ,k −λd · d + β d · r,α + d · r,α      k α k α = τ · r,k +β d · r,α + τ · Zd ,k + (β r,α −λd) · d

     α α α Z α = (τ + β d) · r,α + τ · Zd,α + τ + β r,α −λd · d, (80)

  where k = R, , Z and α = R, . Thanks to the Lagrange multipliers, we can treat r and d as though they were independent virtual displacements. We will be considering a plate with no body forces and no applied tractions, so the weak form of the equations of equilibrium will be !      α α α Z α (τ + β d) · r,α + τ · Zd,α + τ + β r,α −λd · d dV = 0 (81) (X◦)(B)

  for all admissible virtual displacements r and d. Note that         ∂   τ R · + = · R · + r Zd , R Pact E ∂ r Zd R  ∂   = ( · ) · + Pact h1 ∂ r Zd (82)   R       ∂   τ · + = · · + r Zd , Pact E ∂ r Zd     h ∂   = P · 2 · r + Zd . (83) act R ∂

5.2. Kinematics of an axisymmetric kirchhoff plate We consider axisymmetric deformations of an axisymmetric Kirchhoff plate. The reference configuration has the form X (R, , Z) = R cos() i + R sin() j + Z k = R h1() + Z k, (84) where h1() = cos() i + sin() j. (85)

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 18 Mathematics and Mechanics of Solids 000(00)

As a function of (R, , Z), the final configuration will have the form ◦ = + ζ + (x f )(R, , Z) r(R) h1( ) (R) k Z d(R, ). (86) r(R,)

Here r is the cylindrical radius of a point in the middle surface, and ζ is the height of a point in the middle surface. See Figure 7. An outward-pointing radial tangent vector to the middle surface is d (r(R) h () + ζ (R) k) = r (R) h () + ζ (R) k. (87) dR 1 1 We define a to be the corresponding unit vector:

r h1 + ζ k a = = cos φ h1 + sin φ k, (88) (r )2 + (ζ )2 where φ is the angle formed by a and the radial unit vector h1:   ζ φ = arctan . (89) r

The director d is the ‘upward’-pointing unit vector orthogonal to a:

d =−sin φ h1 + cos φ k. (90)

The azimuthal unit vector h2 is the third member of the local orthonormal basis. To construct the deformation gradient, we appeal to the relation ∂ (χ ◦ X)(ξ 1, ξ 2, ξ 3) = F · E , (91) ∂ξi i

1 2 3 where F and the Ei are evaluated at X (ξ , ξ , ξ ), ∂ ∂d (r(R) h () + ζ (R) k + Z d(R, )) = r (R) h () + ζ (R) k + Z (R, ) ∂R 1 1 ∂R

= r h1 + ζ k − Zφ (cos φ h1 + sin φ k)

= r h1 + ζ k − Zφ a. (92)

∂ ∂d ( + ζ + ) = + ∂ r(R) h1( ) (R) k Z d(R, ) r(R) h2( ) Z ∂(R, ) ∂ = − φ h1 r h2 Z sin ∂ = (r − Z sin φ) h2, (93) ∂ (r(R) h () + ζ (R) k + Z d(R, )) = d(R, ). (94) ∂Z 1 These results show that   F = r h + ζ k − Zφ a ⊗ ER + (r − Z sin φ) h ⊗ E + d ⊗ EZ 1   2   r − Z sin φ = r h + ζ k − Zφ a ⊗ h + h ⊗ h + d ⊗ k, (95) 1 1 R 2 2

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 19 where it must be understood that the vectors on the right-hand sides of the tensor products are anchored at the reference point, while those on the left-hand sides are anchored at the deformed point. We will find it convenient to use the orthonormal basis {a, h2, d} in the deformed configuration, so we will re-write the left-hand side of the first in F. Note that           a cos φ sin φ h h cos φ − sin φ a = 1 =⇒ 1 = . (96) d − sin φ cos φ k k sin φ cos φ d

As a result, + ζ = φ − φ + ζ φ + φ r h1 k r (cos a sin d)  (sin a cos d) = r cos φ + ζ sin φ a + ζ cos φ − r sin φ d. (97)

By definition of φ, though, ζ = r tan φ,so φ + ζ φ = φ + φ φ r cos sin r (cos tan sin ) sin2 φ = r cos φ + cos φ = r sec φ, (98) ζ cos φ − r sin φ = r (tan φ cos φ − sin φ) = 0. (99)

The deformation gradient can thus be written as     r − Z sin φ F = r sec φ − Zφ a ⊗ h + h ⊗ h + d ⊗ k. (100) 1 R 2 2

We will need the following expression for later computations:   −1  − r − Z sin φ F−T = r sec φ − Zφ 1 a ⊗ h + h ⊗ h + d ⊗ k. (101) 1 R 2 2

We will use incompatible growth tensors of the form presented in Equations (51) and (52): γ R 2R Z G = γ e ⊗ E +  eθ ⊗ E + e ⊗ E , 1 r R γ z 0 1(s)ds = γ1 h1 ⊗ h1 + γ2 h2 ⊗ h2 + k ⊗ k, (102)

It should be noted that in Equation (102) the vectors on the right-hand side of each tensor product are anchored at a point in the reference configuration and each vector on the left-hand side of a tensor product is anchored at the image under the incompatible growth discussed above. It should also be noted that we may be abusing notation by using Euclidean tangent vectors on the left-hand sides of tensor products in G and on the right-hand sides of tensor products in A. If the post-growth, pre-elastic response state is not immersible, then there is no position function to map this intermediate state into Euclidean space, even locally, and thus the tangent vectors in this state should only be expressed as partial differential operators with respect to r, θ,andz or as the duals to those operators. However, the forms of GT · G (analogous to the right Cauchy-Green tensor) and A · AT (analogous to the left Cauchy-Green tensor) are correct when derived from our expressions. With F and G so defined, the tensor describing the elastic response to growth is

−1 A = F · G    r sec φ − Zφ r − Z sin φ = a ⊗ h1 + h2 ⊗ h2 + d ⊗ k. (103) γ1 γ2R

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 20 Mathematics and Mechanics of Solids 000(00)

The vector on the right-hand side of each tensor product is anchored at a point in the incompatibly grown state, and each vector on the left-hand side of a tensor product is anchored at a point in the final configuration. The most important tensor for the constitutive relation will be the Eulerian tensor     φ − φ 2 − φ 2 T r sec Z r Z sin A · A = a ⊗ a + h2 ⊗ h2 + d ⊗ d. (104) γ1 γ2R

 We now consider variations with r. We start with the following expression

        ∂ ∂ α α R r h2 r (τ + β d) · r,α dV = Pact · h1 + β d · + Pact · + β d · dV plate plate ∂R R ∂      ∂   ∂ r R r h2 = · Pact · h1 + β d + · Pact · + β d dV. plate ∂R ∂ R

It will be helpful to consider βR and β as components of a vector β:

R R β = β ER + β E = β h1 + β R h2. (105)

With β so defined, we can consider the two-point tensor field Pact + d ⊗ β.  Consider the planar portion of r · (Pact + d ⊗ β): " " # # " " # #      · + ⊗ β = · · + · β · + · · + · β · r (Pact d )planar r Pact h1 r d ( h1) h1 r Pact h2 r d ( h2) h2.(106)

In polar coordinates, the of a planar vector field is

1 ∂ 1 ∂g Div (f h + g h ) = (Rf ) + , (107) 1 2 R ∂R R ∂

 · + ⊗ β so the divergence of r (Pact d )planar is " # " " # #  1 ∂   · ( + ⊗ β) = · · + · (β · ) Div r Pact d planar ∂ R r Pact h1 R r d h1 R R " " # # 1 ∂   + r · P · h + r · d (β · h ) R ∂ act 2 2  ∂ r    1 ∂   = · P · h + βRd + r · R P · h + R βRd ∂R act 1 R ∂R act 1  1 ∂ r  1 ∂ + · (P · h + (β · h ) d) + r · (P · h + (β · h ) d) R ∂ act 2 2 R ∂ act 2 2  ∂ r    1 ∂   = · P · h + βRd + r · R P · h + R βRd ∂R act 1 R ∂R act 1  1 ∂ r    1 ∂   + · P · h + βR d + r · P · h + βR d R ∂ act 2 R ∂ act 2  ∂ r    1 ∂   = · P · h + βRd + r · R P · h + R βRd ∂R act 1 R ∂R act 1      ∂ r h  ∂ h + · P · 2 + β d + r · P · 2 + β d . (108) ∂ act R ∂ act R

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 21

 The r portion of Equation (81) can be written as      ∂   ∂ r R r h2 · Pact · h1 + β d + · Pact · + β d dV plate ∂R ∂ R " #  = · + ⊗ β Div r (Pact d )planar dV plate  !  ∂   ∂ 1 R h2 − r · R Pact · h1 + R β d + Pact · + β d dV. (109) plate R ∂R ∂ R By the in the plane, the integral of the first integrand can be re-written " #  · + ⊗ β Div r (Pact d )planar dv plate " # ! H2  = · + ⊗ β Div r (Pact d )planar RdRd dZ H slice 1  " # ! H2  = · + ⊗ β · ν r (Pact d )planar d dZ ∂ H1 (slice) !  H2 = · + ⊗ β · ν r (Pact d )planar dZ d , (110) ∂(slice) H1 where the ‘slice’ mentioned is a planar slice of the reference configuration of the plate, ∂(slice) is the curve that forms the planar boundary of the slice, ν is the outward-pointing unit on ∂(slice), and d is arclength measure on ∂(slice). In order to balance the boundary term in the weak form of the principle of virtual work, we must specify the value of H2 H2 + ⊗ β · ν = + ⊗ β · (Pact d )planar dZ (Pact d ) h1 dZ (111) H1 H1 on the boundary of the middle surface. For an unloaded plate, this integral is zero at each point on the boundary of the middle surface. Now that boundary terms are balanced, we have  !  ∂   ∂ 1 R h2 r · R Pact · h1 + R β d + Pact · + β d dV = 0 (112) plate R ∂R ∂ R  for each as-smooth-as-needed function r of R and .Since  !  1 ∂   ∂ h r · R P · h + R βRd + P · 2 + βd dv R ∂R act 1 ∂ act R plate  ! H2 1 ∂   ∂ h  = · + βR + · 2 + β · ∂ R Pact h1 R d ∂ Pact d dZ rRdRd (113) slice H1 R R R = 0,  for such a large class of r, we conclude that the integral in Z is identically zero:  ! H2 1 ∂   ∂ h · + βR + · 2 + β = ∂ R Pact h1 R d ∂ Pact d dZ 0. (114) H1 R R R

  Next, we consider the variation in d. We apply the same procedure as for the variation in r. The result will be a weak version of the balance of in the constrained plate with zero applied moments.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 22 Mathematics and Mechanics of Solids 000(00) !     α Z α τ · Zd,α + τ + β r,α −λd · d dV plate ⎧ ⎫ ⎨   ⎬ ∂d ∂d    = τ R · + τ · + τ Z + βα −λ · ⎩ Z Z r,α d d⎭ dV plate ∂R ∂    = Z Div d · Pact dV plate planar  !    ∂ ∂ Z α 1 h2 + d · τ + β r,α −λd − Z (R P · h ) − Z P · dV R ∂R act 1 ∂ act R  plate    H2 = d · Z Pact dZ · ν d ∂ (slice) H1  !  H2 ∂ ∂ Z α 1 h2 + · τ + β α −λ − ( · ) − · θ d r, d Z ∂ R Pact h1 Z ∂ Pact dZRdRd . (115) slice H1 R R R

As before, ν = h1 is the outward-pointing unit normal to the planar boundary ∂(slice) of the slice in the reference configuration, and d is the arclength measure on ∂(slice). To balance the boundary terms, we must set values of the moment

H2 Z (Pact · h1) dZ (116) H1 on the planar boundary of the middle surface. Since we assume no applied moments in addition to no applied tractions, we set this integral equal to zero at each point on the boundary of the middle surface. We have found that

 !  H2 ∂ ∂ Z α 1 h2 · τ + β α −λ − ( · ) − · θ = d r, d Z ∂ R Pact h1 Z ∂ Pact dZRdRd 0 (117) slice H1 R R R

 for all as-smooth-as-needed functions d of R and . We conclude that

 ! H2 ∂ ∂ Z α 1 h2 τ + β α −λ − ( · ) − · = r, d Z ∂ R Pact h1 Z ∂ Pact dZ 0. (118) H1 R R R

5.3. Addressing Lagrange multipliers We are not yet ready to introduce the constitutive relation to close the system of equations. The quantities λ, βR,andβ are not constitutively defined. However, there is enough information in the equations above to get around this complication. First we claim that if we take the cross-product with d,thetermλd will vanish. By the form of F−T found in Equation (101), we can be sure that F−T · k = d. If we use an isotropic hyperelastic constitutive relation, then the form of A = F · G−1 in such a relation will guarantee that T · d points in the direction d. As a result,

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 23

H2 H2 Z τ dZ = (Pact · k) dZ H1 H1 H2   = (det F) T · F−T · k dZ H1 H2 = (det F)(T · d) dZ is a scalar multiple of d, (119) H1 which implies H2 d × τ Z dZ = 0, (120) H1 the elimination we sought. When we take the cross-product of d with Equation (118), we have     1 ∂ H2 ∂ H2 h d × R Z (P · h ) dZ + d × Z P · 2 dZ R ∂R act 1 ∂ act R H1 H1 H2 α −d × r,α β dZ = 0. (121) H1 We consider all of these terms in detail to find expressions for βR and β in terms of kinematic and constitutively defined quantities. By the symmetry of the deformation, the vector field Pact · h1 has zero projection onto h2. It is determined solely by its projections onto h1 and k:

(Pact · h1) = (h1 · Pact · h1) h1 + (k · Pact · h1) k (122)

Since h1 and k are R-independent, 1 ∂ 1 ∂ 1 ∂ R (P · h ) = h R (h · P · h ) + k R (k · P · h ) . (123) R ∂R act 1 1 R ∂R 1 act 1 R ∂R act 1 In particular, 1 ∂ 1 ∂ h · R (P · h ) = R (h · P · h ) , (124) 1 R ∂R act 1 R ∂R 1 act 1 1 ∂ 1 ∂ k · R (P · h ) = R (k · P · h ) . (125) R ∂R act 1 R ∂R act 1 We have   1 ∂ H2 d × R Z (P · h ) dZ R ∂R act 1 H1   1 ∂ H2 = (cos φ k − sin φ h ) × h R Z (h · P · h ) dZ 1 1 R ∂R 1 act 1  !H1 1 ∂ H2 + k R Z (k · P · h ) dZ R ∂R act 1  H 1    cos φ ∂ H2 sin φ ∂ H2 = h R Z (h · P · h ) dZ + h R Z (k · P · h ) dZ 2 R ∂R 1 act 1 2 R ∂R act 1  H 1   H1 ! cos φ ∂ H2 sin φ ∂ H2 = ( · · ) + ( · · ) h2 ∂ R Z h1 Pact h1 dZ ∂ R Z k Pact h1 dZ . (126) R R H1 R R H1

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 24 Mathematics and Mechanics of Solids 000(00)

The symmetry of the deformation also ensures that Pact · h2 point along h2:

(Pact · h2) = (h2 · Pact · h2) h2. (127)

Further, the coefficient (h2 · Pact · h2) is -independent, so the - of (Pact · h2) is ∂ ∂ ( · ) = ( · · ) ∂ Pact h2 ∂ h2 Pact h2 h2 ∂h = ( · · ) 2 h2 Pact h2 ∂ =−(h2 · Pact · h2) h1. (128) The cross-product with d is       ∂ H2 h H2 h d × Z P · 2 dZ = (cos φ k − sin φ h ) × −h Z h · P · 2 dZ ∂ act R 1 1 2 act R H1 H1 cos φ H2 =−h2 Z (h2 · Pact · h2) dZ. (129) R H1 The remaining terms are

H2 ∂ H2 ∂ H2 R r R r d × r,α β dZ = d × β dZ + d × β dZ ∂R ∂ H1 H1 H1 !   H2 H2 R = (cos φ k − sin φ h1) × r h1 + ζ k β dZ + r h2 β dZ    H1  H1 H2 H2 R R = r cos φ β dZ h2 + ζ sin φ β dZ h2  H 1   H1  H2 H2 − r cos φ β h1 − r sin φ β k. (130) H1 H1 What we have shown is     cos φ ∂ H2 sin φ ∂ H2 R Z (h · P · h ) dZ + R Z (k · P · h ) dZ R ∂R 1 act 1 R ∂R act 1 H1 ! H1 cos φ H2 − Z (h · P · h ) dZ h R 2 act 2 2 H1     H2 H2 R = r cos φ + ζ sin φ β dZ h2 − r (cos φ h1 − sin φ k) β dZ. (131) H1 H1  H2 β Note that the left-hand side is a scalar multiple of h2,andthat H dZ on the right-hand side is the coefficient 1 of a linear combination of h1 and k, both of which are orthogonal to h2. Since this holds for each (R, ), we H2 see that β dZ = 0. Computing the h2-projection reveals H1   cos φ ∂ H2 R Z (h · P · h ) dZ R ∂R 1 act 1  H1  sin φ ∂ H2 + R Z (k · P · h ) dZ R ∂R act 1 H1   φ H2   H2 cos R − Z (h2 · Pact · h2) dZ = r cos φ + ζ sin φ β dZ . (132) R H1 H1

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 25

If we multiply by R anddividebycosφ,wehave   ∂ H2 R Z (h · P · h ) dZ ∂R 1 act 1  H1  ∂ H2 + tan φ R Z (k · P · h ) dZ ∂R act 1 H1   H2   H2 R − Z (h2 · Pact · h2) dZ = R r + ζ tan φ β dZ H1  H1    H2 = R r + r tan2 φ βR dZ  H1  H2 = Rr sec2 φ βR dZ . (133) H1 In the next section we will find another expression for the integral of βR, and this equation will become a differential equation in terms of kinematical and constitutively defined quantities.

6. Formulation of the problem as a system of ordinary differential equations Since we have eliminated all differentiation with respect to in the differential equations, we can view them as a system of ordinary differential equations with independent variable R. Thanks to the discoveries of the last section, Equation (114) can now be written as ! 1 d H2 H2 H2 h R (h · P · h ) dZ + k R (k · P · h ) dZ + R d βR dZ R dR 1 1 act 1 act 1 H1  H1 H1 H2 h2 −h1 h2 · Pact · dZ = 0. (134) H1 R The projection of this equation onto k is ! H2 H2 1 d R R (k · Pact · h1) dZ + R cos φ β dZ = 0, (135) R dR H1 H1 so that H2 H2 R R (k · Pact · h1) dZ + R cos φ β dZ = constant. (136) H1 H1 Recall that we must assign the value of the stress

H2   R Pact · h1 + β d dZ (137) H1 on the boundary (R = Rmax) of the middle surface. If this stress is zero, then in particular its k-projection is zero. This implies that the constant on the right-hand side of Equation (136) is zero, and

H2 H2 R β dZ =−sec φ (k · Pact · h1) dZ. (138) H1 H1 We will see later (Equation (163)) that the isotropic hyperelastic constitutive relation we choose ensures that Pact · h1 points in the direction a = cos φ h1 + sin φ k. As a result of the direction of Pact · h1,

Pact · h1 = (a · Pact · h1) a, (139)

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 26 Mathematics and Mechanics of Solids 000(00)

so (h1 · Pact · h1) and (k · Pact · h1) satisfy

k · Pact · h1 = (a · Pact · h1) k · a = sin φ (a · Pact · h1) , (140)

h1 · Pact · h1 = (a · Pact · h1) h1 · a = cos φ (a · Pact · h1) , (141)

(k · Pact · h1) = tan φ (h1 · Pact · h1) . (142) and H2 H2 R β dZ =−sec φ tan φ (h1 · Pact · h1) dZ. (143) H1 H1 The h1-projection of Equation (134) becomes !   1 d H2 H2 H2 h R (h · P · h ) dZ − R sin φ βR dZ = h · P · 2 dZ R dR 1 act 1 2 act R H1 H1 ! H1   1 d H2 H2 H2 h R (h · P · h ) dZ + R tan2 φ (h · P · h ) dZ = h · P · 2 dZ R dR 1 act 1 1 act 1 2 act R H1 H1 ! H1 H2 H2 d 2 R sec φ (h1 · Pact · h1) dZ = (h2 · Pact · h2) dZ. (144) dR H1 H1 We have one more scalar ordinary differential equation from the previous section:   d H2 R Z (h · P · h ) dZ dR 1 act 1  H1  d H2 + tan φ R Z (k · P · h ) dZ dR act 1 H1   H2 H2 2 R − Z (h2 · Pact · h2) dZ = Rr sec φ β dZ H1 H1 H2 3 =−Rr tan φ sec φ (h1 · Pact · h1) dZ, (145) H1 where we have used Equation (143) to re-write the right-hand side. Note that   d H2 tan φ R Z (k · P · h ) dZ dR act 1 H1  d H2 = tan φ R tan φ Z (h · P · h ) dZ dR 1 act 1 H1   H2 H2 2 2 d = φ tan φ sec φR Z (h1 · Pact · h1) dZ + tan φ R Z (h1 · Pact · h1) dZ . (146) H1 dR H1 Equation (145) can be re-written as   d H2 H2 (1 + tan2 φ) R Z (h · P · h ) dZ + φ tan φ sec2 φR Z (h · P · h ) dZ dR 1 act 1 1 act 1 H1 H1 H2 H2 3 = Z (h2 · Pact · h2) dZ − Rr tan φ sec φ (h1 · Pact · h1) dZ. (147) H1 H1 − Since 1 + tan2 φ = sec2 φ = (cos φ) 2, we can re-write this as   d H2 H2 R Z (h · P · h ) dZ = cos2 φ Z (h · P · h ) dZ dR 1 act 1 2 act 2 H1 H1   H2

− Rφ + Rr sec φ tan φ Z (h1 · Pact · h1) dZ. (148) H1

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 27

In summary, the balance laws for linear momentum and angular momentum for the plate are   d H2 H2 R sec2 φ (h · P · h ) dZ = (h · P · h ) dZ, (149) dR 1 act 1 2 act 2  H1  H1 d H2 H2 R Z (h · P · h ) dZ = cos2 φ Z (h · P · h ) dZ dR 1 act 1 2 act 2 H1 H1   H2

− Rφ + Rr sec φ tan φ Z (h1 · Pact · h1) dZ. (150) H1 We have already ensured (Equation (138)) that ! H2 H2 H2 H2 R R (k · Pact · h1) dZ + cos φ β dZ = k · (Pact · h1) dZ + d β dZ (151) H1 H1 H1 H1 is identically zero. One of the boundary conditions is that the analogous h1-projection is zero at the boundary of the middle surface: ! H2 H2 R h1 · (Pact · h1) dZ + d β dZ H1 H1 H2 H2 R = (h1 · Pact · h1) dZ − sin φ β dZ H1 H1 H2 H2 = (h1 · Pact · h1) dZ + sin φ sec φ (k · Pact · h1) dZ H1 H1   H2 2 = 1 + tan φ (h1 · Pact · h1) dZ H1 H2 2 = sec φ (h1 · Pact · h1) dZ. (152) H1 Hence, the boundary condition for stress is * H2 * 2 * sec φ (h1 · Pact · h1) dZ* = 0, (153) H 1 R∈{Rmin, Rmax} where we consider Rmax in every case and Rmin only in the case of an annulus. The boundary condition for moment is * H * 2 * Z (Pact · h1) dZ* = 0. (154) H 1 R∈{Rmin, Rmax}

We will find a scalar equation that ensures that the h1-andk-projections are both zero.

7. Applying the constitutive relation We consider an isotropic hyperelastic constitutive relation derived from a strain energy density of neo-Hookean type:   1   W (λ , λ , λ ) = μ λ2 + λ2 + λ3 − 3 − ln J + λ (J − 1 − ln J) , (155) 1 2 3 2 1 2 3 where λi, i = 1, 2, 3, are the principal stretches of the deformation, and J = λ1λ2λ3. The constants λ and μ correspond to the Lamé moduli of linear elasticity.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 28 Mathematics and Mechanics of Solids 000(00)

7.1. Without growth We first consider the case without growth and establish the form of Cauchy stress tensor as a function of the deformation gradient F. One of√ the polar decompositions of the deformation gradient is F = R · U,whereR is a proper rotation and U = FT · F is a symmetric positive-definite tensor called the right stretch tensor. Since the strain-energy + density W (λ1, λ2, λ3) is an isotropic function of the deformation gradient, it is equal to some function W(U). We compute the Biot stress or the Jaumann stress

∂W+ ,3 ∂W T(1) = = u(i) ⊗ u(i), (156) ∂U ∂λ i=1 i where the u(i) are the unit eigenvectors of U. In isotropic hyperelasticity without local constraints, the Cauchy stress and Biot/Jaumann stress are related by

JRT · T · R = U · T(1) ,3 ∂W   = · (i) ⊗ (i) ∂λ U u u i=1 i ,3 ∂W = λ (i) ⊗ (i) i ∂λ u u . (157) i=1 i

See Section 4.3 of Ogden [13]. The principal components of the active Cauchy stress (the constitutively defined portion of the Cauchy stress) are then λ ∂W 1     = i = μ λ2 − + λ − (Tact)i i 1 (J 1) . (158) J ∂λi J

7.2. With growth When growth is included via multiplicative decomposition of the deformation gradient F = A · G, the (active) Cauchy stress is a function of the elastic portion A alone. However, we can import the results found above. Equation (158) still holds, but the eigenvalues are eigenvalues of A (actually, square roots of eigenvalues of A · AT), and the principal directions are the eigenvectors of A · AT. By Equation (104), the full active Cauchy stress tensor is 1     T = μ λ2 a ⊗ a + λ2 h ⊗ h + λ2 d ⊗ d + (λJ − (λ + μ)) I , (159) act J 1 2 2 2 3 where the λi are the ‘pseudo-stretches’

r sec φ − Zφ r − Z sin φ λ1 = , λ2 = , λ3 = 1, (160) γ1 γ2R and J = det A = λ1λ2λ3. −T Equations (149) and (150) are written in terms of Pact = (det F) Tact · F , but we need only certain projections of this tensor. Note that   −T −1 F · h1 = r sec φ − Zφ a, (161)  − r − Z sin φ 1 F−T · h = h . (162) 2 R 2

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 29

We have

−T Pact · h1 = (det F) Tact · F · h1  − = Jγ γ T · r sec φ − Zφ 1 a 1 2 act  μ   λ   λ + μ = γ γ r sec φ − Zφ + r sec φ − Zφ − a, (163) 1 2 γ 2 γ γ φ − φ 1 1 2 r sec Z where

det F = (det A)(det G) = (J)(γ1γ2) . (164)

A similar calculation shows that       μ r − Z sin φ λ r − Z sin φ λ + μ P · h = γ γ + − h . (165) act 2 1 2 γ 2 γ γ − φ / 2 2 R 1 2 R (r Z sin ) R

The projections pertinent for the differential equations are     μγ   r − Z sin φ (λ + μ) γ γ h · P · h = cos φ 2 r sec φ − φ Z + λ − 1 2 , (166) 1 act 1 γ φ − φ  1  R r sec Z   μγ1 r − Z sin φ (λ + μ) γ1γ2 h2 · Pact · h2 = + λ r sec φ − Zφ − . (167) γ2 R (r − Z sin φ) /R

We also assume that the middle surface occupies the true middle of the plate, so that H1 =−H and H2 = H for some H > 0. The Z- then have the forms    ! H μγ λ + μ γ γ φ · · = φ 2 φ + λ r − 2 ( ) 1 2 H (h1 Pact h1) dZ cos 2H r sec arctanh (168) − γ R φ r sec φ H  1    H μγ1 r 2 (λ + μ) γ1γ2R H sin φ (h2 · Pact · h2) dZ = 2H + λr sec φ − arctanh (169) − γ R sin φ r H 2   H 2 μγ sin φ ( · · ) =− φ 3 2 φ + λ Z h1 Pact h1 dZ cos H γ −H 3 1 R   ! 2 (λ + μ) γ γ Hφ + 1 2 r sec φ arctanh − Hφ (170) (φ )2 r sec φ   H 2 μγ sin φ ( · · ) =− 3 1 + λφ Z h2 Pact h2 dZ H γ −H 3 2 R     2 (λ + μ) γ γ R H sin φ − 1 2 r arctanh − H sin φ . (171) sin2 φ r

We get a further simplification if we divide by μ:    ! H γ + κ γ γ φ 1 · · = φ 2 φ + κ r − 2 (1 ) 1 2 H (h1 Pact h1) dZ cos 2H r sec arctanh (172) μ − γ R φ r sec φ H  1    H 1 γ1 r 2 (1 + κ) γ1γ2R H sin φ (h2 · Pact · h2) dZ = 2H + κr sec φ − arctanh (173) μ −H γ2 R sin φ r

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 30 Mathematics and Mechanics of Solids 000(00)   1 H 2 γ sin φ ( · · ) =− φ 3 2 φ + κ μ Z h1 Pact h1 dZ cos H γ −H 3 1 R   ! 2 (1 + κ) γ γ Hφ + 1 2 r sec φ arctanh − Hφ (174) (φ )2 r sec φ   1 H 2 γ sin φ ( · · ) =− 3 1 + κφ μ Z h2 Pact h2 dZ H γ −H 3 2 R     2 (1 + κ) γ γ R H sin φ − 1 2 rarctanh − H sin φ , (175) sin2 φ r where κ = λ/μ. The differential equations have the same form if each stress and moment is replaced by the same quantity divided by μ.

8. Stability of the flat plate 8.1. The flat plate

If the plate is un-buckled, then there is no moment, Pact ·h1 has no Z-dependence and is radial only, i.e. it has no vertical component. In the case of constant γ1 and γ2, the equations for the un-buckled plate can be converted into a pair of autonomous ordinary differential equations, as detailed in Antman and Negrón-Marrero [14]. We define new dependent variables

s−1 R = Rmaxe , (176) r τ = , (177) R γ + κ γ γ = 2 + κ r − (1 ) 1 2 N r . (178) γ1 R r

When τ and N are viewed as functions of the independent variable s = 1 + ln(R/Rmax), they satisfy   dτ γ = 1 N − κτ + (N − κτ)2 + 4(1 + κ)γ 2 − τ, (179) ds 2γ 2 2   dN γ κγ (1 + κ)γ γ = 1 τ + 1 − κτ + − κτ 2 + + κ γ 2 − 1 2 − N (N ) 4(1 ) 2 N. (180) ds γ2 2γ2 τ We consider an annulus that has zero radial stress at its inner and outer faces. We can find numerically an initial value for τ such that the boundary conditions n(s0) = 0andn(s1) = 0 are met. The functions n(s)and τ(s) correspond to functions N(R)andT(R)via       R R N(R) = n 1 + ln , T(R) = τ 1 + ln . (181) Rmax Rmax This correspondence also gives us a function r(R):    R r(R) = RT(R) = Rτ 1 + ln , (182) Rmax which also provides the derivative r (R).

8.2. Differential equations for bifurcation Finding numerical solutions of the boundary-value problem for the buckled plate has proved a daunting task. To demonstrate the existence of buckled solutions, we consider perturbation about the flat configuration.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 31

Equations (149) and (150) for the buckled plate, i.e. with stresses and moments given by Equations (168)– (171), can be expressed in the form ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ 10 0 0 r r ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ 01 0 0⎟ d ⎜ φ ⎟ ⎜ φ ⎟ ⎝ ⎠ ⎝ ⎠ = ⎝ ⎠ , (183) 00a33 a34 dR r f3 00a a φ f  43 44   4  A(R, r, φ, r , φ ) f(R, r, φ, r , φ ) with   2 γ2 sec φ γ1γ2(1 + κ) a33 = 2HR + , (184) γ 2 − 2 φ 2 2 φ 1 (r ) H ( ) cos  γ γ (1 + κ) sec φ Hφ cos φ a = 2R 1 2 arctanh 34 φ 2 ( ) ⎫ r ⎬ Hγ γ (1 + κ) − " 1 2 # (185) φ − H2(φ )2 cos2 φ ⎭ r 1 2 (r )   γ γ (1 + κ) Hφ cos φ a = 2R − 1 2 arctanh 43 φ 2 ( ) ⎫ r ⎬ Hγ γ (1 + κ)cosφ + "1 2 # (186) 2 φ 2 2 φ ⎭ r φ 1 − H ( ) cos (r )2

H3γ cos φ Hγ γ (1 + κ)cosφ a = 2R − 2 + 1 2 44 3γ φ 2 1 " ( ") # #

Hφ cos φ 2γ1γ2(1 + κ)cosφ arctanh r sec φ − Hφ + r φ 3 ⎫( ) ⎬ Hγ γ (1 + κ)cosφ − "1 2 # , (187) φ 2 − H2(φ )2 cos2 φ ⎭ ( ) 1 2 (r )   H sin φ 2Hκr sec φ f3 = 2Rγ1γ2(1 + κ)arctanh csc φ −  r  R γ r +2Hκr sec φ − 2H 1 + κr sec φ Rγ ⎛ 2 " #⎞ φ φ   γ γ + κ H cos κr γ r sec φ 2 1 2(1 )arctanh r + sec φ ⎝2H + 2 − ⎠ R γ1 φ γ φ 2 φ φ +2HR 2r sec tan γ 1 " # ⎛ ⎞   Hφ cos φ κ γ φ 2γ1γ2(1 + κ)arctanh + φ φ ⎝ r + 2r sec − r ⎠ φ R sec tan 2H R γ1 φ 2HRγ γ (1 + κ)φ tan φ + " 1 2 # (188) 2 φ 2 2 φ r 1 − H ( ) cos (r )2

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 32 Mathematics and Mechanics of Solids 000(00)

2H3κ cos φ sin φ f4 = 3R " # ⎛ ⎞   Hφ cos φ κ γ φ 2γ1γ2(1 + κ)arctanh + φ ⎝ r + 2r sec − r ⎠ Rr tan 2H R γ1 φ " #

Hφ cos φ 2Rγ1γ2(1 + κ)arctanh r tan φ − r φ 2 2HRγ γ (1 + κ)sinφ − H3κφ cos2 φ + 1 2 3 − H2(φ )2 cos2 φ 1 2 " (r") # # ⎧ ⎨ γ γ + κ Hφ cos φ φ − φ 2 1 2(1 ) arctanh r r sec H − cos φ ⎩ (φ )2  ! 2 κ sin φ γ φ + 3 + 2 H γ 3 R 1     φ 2 H sin − cos φ −2Rγ1γ2(1 + κ) csc φ rarctanh − H sin φ  ! r 2 γ sin φ − H3 1 + κφ . (189) 3 Rγ2

The matrix A is invertible symbolically, so Equation (183) can be expressed in quasilinear form: ⎛ ⎞ r ⎜ ⎟ d ⎜ φ ⎟ −1 ⎝ ⎠ = A f. (190) dR r φ

We consider expressions r = r0 + r1 and φ = φ1,wherer0 is the radius function in Equation (182), i.e. the solution of the boundary-value problem for the un-buckled plate. Inserting these into Equation (190), we have ⎛ ⎞ ⎛ ⎞ r r ⎜ 0 ⎟ ⎜ 1 ⎟ d ⎜ 0 ⎟ d ⎜ φ1 ⎟ ⎝ ⎠ +  ⎝ ⎠ dR r0 dR r1 0 φ 1 ⎛ ⎞ * r1 * −1 * ⎜ ⎟   − ∂A f φ 1 * * ⎜ 1 ⎟ 2 = A f r = r , r = r +  · + O  . (191) 0 0 * = = ⎝ ⎠ ∂ φ φ r r0, r r r φ = 0, φ = 0 (r, , r , ) 0 1 φ = 0, φ = 0 φ 1

 φ φ Equating terms with equal powers of , we are left with a system of linear equations for (r1, 1, r1, 1): ⎛ ⎞ ⎛ ⎞ r1 * r1 ⎜ ⎟ ∂ −1 * ⎜ ⎟ d ⎜ φ1 ⎟ A f * ⎜ φ1 ⎟ = · . (192) ⎝ ⎠ * = = ⎝ ⎠ r ∂ φ φ r r0, r r r dR 1 (r, , r , ) 0 1 φ φ = 0, φ = 0 φ 1 1

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 33

8.3. Linear boundary conditions for bifurcation We employ first-order expansions of the scalar coefficient of the radial stress in Equation (168): * H H * 1 1 * (a · Pact · h1) dZ = (a · Pact · h1) dZ * = = μ μ r r0, r r −H −H 0 φ = 0, φ = 0 " # ⎛ ⎞  * r ∂ 1 H · · * 1 μ − (a Pact h1) dZ * ⎜ φ ⎟   + H * ·  ⎜ 1 ⎟ + O 2 ∂ φ φ * ⎝ r ⎠ (r, , r , ) * = = 1 r r0, r r 0 φ  φ =0, φ = 0 1 γ r (1 + κ)γ γ = 2H 2 r + κ 0 − 1 2 γ 0 R r  1  0   2Hκ 2Hγ 2H(1 + κ)γ γ   +  r + 2 +   1 2 r + O 2 . (193) 1 2 1 R γ1 r0

At R = R0 and R = R1, the radial stress of the flat annulus is zero, so for i = 0, 1,    * * * 1 H * 2Hκ 2Hγ 2H(1 + κ)γ γ *   · · * =  + 2 +   1 2 * + O 2 (a Pact h1) dZ* r1 2 r1 . (194) μ − R γ * H R=Ri 1 r 0 R=Ri

This provides one linear boundary condition for R = R0 and one for R = R1:   * * κ γ (1 + κ)γ γ * r + 2 +   1 2 r * = 0. (195) 1 γ 2 1* R 1 r 0 R=Ri

We perform the same expansion procedure for the moment in Equation (170): * H H * 1 1 * (a · Pact · h1) ZdZ= (a · Pact · h1) ZdZ * = = μ μ r r0, r r −H −H 0 φ = 0, φ = 0 " # ⎛ ⎞  * r ∂ 1 H · · * 1 μ − (a Pact h1) ZdZ * ⎜ φ ⎟   + H * ·  ⎜ 1 ⎟ + O 2 ∂ φ φ * ⎝ r ⎠ (r, , r , ) * = = 1 r r0, r r 0 φ   φ = 0, φ =0  1 2H3 κ γ (1 + κ)γ γ   = 0 + φ + 2 +   1 2 φ + O 2 . (196) 1 2 1 3 R γ1 r0

Recall that the un-buckled plate has identically zero moment. This provides linear boundary conditions at = = R R0 and R R1:   * * κ γ (1 + κ)γ γ * φ + 2 +   1 2 φ * = 0. (197) 1 γ 2 1* R 1 r 0 R=Ri

8.4. Evidence of bifurcation We now have a system of four linear ordinary differential equations and four linear boundary conditions. The linear boundary conditions at R = R0 have a two-dimensional vector space of solutions. Considering that the boundary conditions at R = R0 can be expressed as

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 34 Mathematics and Mechanics of Solids 000(00) ⎛ ⎞* ⎛ ⎞ * κ γ +κ γ γ r1 *   2 + (1 ) 1 2 γ 2 00 ⎜ ⎟* ⎝ R 1 (r ) ⎠ ⎜ r1 ⎟* 0 0 γ +κ γ γ = , (198) κ 2 (1 ) 1 2 ⎝ φ ⎠* 00 + 1 0 R γ (r )2 * 1 0 φ * 1 R=R0 we see that the two-dimensional solution space at R = R0 is spanned by, for example, ⎛ ⎞* ⎛ ⎞* ⎛ ⎞* ⎛ ⎞* * γ2 (1+κ)γ1γ2 * * * r + r 0 1 * γ1 (r )2 * 1 * * ⎜ ⎟* ⎜ κ 0 ⎟* ⎜ ⎟* ⎜ 0 ⎟* ⎜ r1 ⎟* ⎜ − ⎟* ⎜ r1 ⎟* ⎜ ⎟* = R and = γ2 (1+κ)γ1γ2 . (199) ⎝ φ ⎠* ⎝ ⎠* ⎝ φ ⎠* ⎝ + ⎠* 1 * 0 * 1 * γ1 (r )2 * φ * * φ * κ 0 * 1 R=R 0 1 R=R − 0 R=R0 0 R R=R0

Let u and v be solutions of Equation (192) with the initial (R = R0) conditions above. Then u and v have the forms ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ ⎛ ⎞ u1 r1 v1 0 ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ ⎟ ⎜ u2 ⎟ ⎜ r ⎟ ⎜ v2 ⎟ ⎜ 0 ⎟ ⎝ ⎠ = ⎝ 1 ⎠ and ⎝ ⎠ = ⎝ ⎠ . (200) u3 0 v3 φ1 φ u4 0 v4 1 Every linear combination αu + βv is a solution of the differential equations that satisfies the linear boundary conditions Equation (195) and (197) at R = R0. We seek a linear combination of this form that also satisfies the linear boundary conditions at R = R1.That is, we want ⎛ ⎞* ⎛ ⎞ α * γ +κ γ γ u1 *   κ 2 + (1 ) 1 2 ⎜ ⎟* R γ 2 00 α ⎝ 1 (r0) ⎠ ⎜ u2 ⎟* 0 κ γ +κ γ γ * = . (201) 2 + (1 ) 1 2 ⎝ βv ⎠ 0 00R γ 2 3 * 1 (r0) * βv4 R=R1 We can seek the existence of such a linear combination by re-writing this condition as ⎛ " # ⎞* κ γ2 (1+κ)γ1γ2 *     u + + u 0 * R 1 γ (r )2 2 α 0 ⎝ 1 0 " # ⎠* = γ +κ γ γ * . (202) κ 2 (1 ) 1 2 β 0 0 v1 + + v2 * R γ1 (r )2 0 R=R1

If the matrix has zero determinant, then there is a non-trivial solution (α, β)T of this pair of linear equations at R = R1. We know that if γ1 = γ2 = constant, then the growth tensor is a true deformation gradient, and there is no residual stress in the un-buckled annulus. We fix a value for γ2 and let γ1 vary until the determinant of the matrix in Equation (202) is zero. In truth, we seek the first γ1-value at which either of the non-zero entries in the matrix is zero.

8.5. Numerical results In all numerical results found so far, the first term to drop to zero is the (2, 2)-entry of the matrix in Equation (202). This means that we can choose a perturbation of the moment alone. Further, the (2, 2)-entry reaches zero at some γ1 >γ2, which indicates buckled solutions for the case in which incompatible growth corresponds to azimuthal contraction. In the following example, we set γ2 = 1.01, H = 0.01, R0 = 0.01, R1 = 1, and κ = 1. As seen in Figures 8 and 9, as functions of γ1, the stress perturbation (dependent on r1 alone) at R = R1 has no zeros, while the moment perturbation (dependent on φ1 alone) at R = R1 has multiple zeros, all greater than γ2. The first four zeros are found at γ1 = 1.01086, 1.01348, 1.01793, and 1.0242.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 35

κ r R γ 1 γ γ 1 1 2 1 2 r R γ 2 1 1 R1 1 r0 R1 420

415

410

405

γ 1.012 1.014 1.016 1.018 1.020 1.022 1.024 1

Figure 8. The stress perturbation (dependent on r1)atR = R1 has no zeros.

κφ R γ 1 γ γ 1 1 2 1 2 φ R γ 2 1 1 R1 1 r0 R1

150

100

50 γ 1.012 1.014 1.016 1.018 1.020 1.022 1.024 1 50

Figure 9. The moment perturbation (dependent on φ1)atR = R1 has multiple zeros.

γ1 first critical value

1.10

1.08

1.06

1.04

1.02 H 0.02 0.04 0.06 0.08 0.10

Figure 10. The first bifurcation value of γ1, as a function of half-thickness H

γ1 critical value 1.7 1.6 1.5 1.4 1.3 1.2 1.1 H 0.02 0.04 0.06 0.08 0.10

Figure 11. The first three critical values of γ1, as functions of H

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 36 Mathematics and Mechanics of Solids 000(00)

φ

2.5 10 6

2. 10 6

1.5 10 6

1. 10 6

0.2 0.4 0.6 0.8 1.0 R

Figure 12. The buckling angle φ as a function of R,withγ2 = 1.01 and γ1 ≈ 1.01086.

ζ

2. 10 6

1.5 10 6

1. 10 6

5. 10 7

0.2 0.4 0.6 0.8 1.0 R

Figure 13. The height ζ of the middle surface as a function of R.

integrated stress μ 0.00002

0.2 0.4 0.6 0.8 1.0 R 0.00002

0.00004

0.00006

0.00008

Figure 14. Radial (rapid dashed) and azimuthal (slower dashes) stresses as function of R. The results of perturbation and of solving the full non-linear system of equations are overlaid.

Figures (10) and (11) show that an increase in the half-thickness H of the plate causes a delay in the onset of buckling. Figures 12–15 show the values of the buckling angle φ, the middle surface height ζ , and the thickness- integrated radial and azimuthal stresses and radial and azimuthal moments for γ1 = 56026090/55424431 ≈ 1.01086, which is very near a bifurcation value. In these figures, quantities computed with solutions of the perturbation equations are plotted with dashed lines, and those computed with solutions of the full non-linear equations are plotted with solid lines. Since γ1 is so close to the bifurcation value, the buckling is very small.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 37

integrated moments μ

2. 10 11

1.5 10 11

1. 10 11

5. 10 12

0.2 0.4 0.6 0.8 1.0 R 5. 10 12

1. 10 11

Figure 15. Radial (rapid dashes) and azimuthal (slower dashes) moments as function of R. The results of perturbation and of solving the full non-linear system of equations are overlaid. φ

7.0

6.8

6.6

6.4

6.2

6.0

0.3 0.4 0.5 0.6 0.7 R

Figure 16. The buckling angle φ as a function of R. See Figure 20. ζ

0.06

0.05

0.04

0.03

0.02

0.01

0.3 0.4 0.5 0.6 0.7 R

Figure 17. The height ζ of the middle surface as a function of R. See Figure 20.

9. Large buckling due to immersibility-precluding growth We consider a growth field of the kind in Equation (51): γ R 2R Z G = γ e ⊗ E +  eθ ⊗ E + e ⊗ E 1 r R γ z 0 1(s) ds γ R h = γ h ⊗ h +  2 rh ⊗ 2 + k ⊗ k 1 1 1 R γ 2 R 0 1(s) ds = γ1 h1 ⊗ h1 + γ2 h2 ⊗ h2 + k ⊗ k, (203)

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 38 Mathematics and Mechanics of Solids 000(00)

integrated stress μ 0.020

0.015

0.010

0.005

0.3 0.4 0.5 0.6 0.7 R 0.005

0.010

Figure 18. Radial (solid) and azimuthal (dashed) stresses as function of R

integrated moment μ

0.00004

0.00002

0.3 0.4 0.5 0.6 0.7 R

–0.00002

–0.00004

Figure 19. Radial (solid) and azimuthal (dashed) moments as function of R

where γ1 and γ2 are not constant but are continuously differentiable positive functions of R,and

R γ (R)R = γ θ =  2 = r(R) 1(s) ds, (R, ) R , z Z. (204) Rmin γ1(s) ds Rmin This growth tensor is incorporated into the constitutive relation in the same fashion, and the differential equations are Equations (149) and (150), and the boundary conditions are Equations (153) and (154). With parameters Rmin = 0.2, Rmax = 0.709 H = 0.01, and κ = 1, and growth factors

2 γ1(R) ≡ 1.9, γ2(R) = 1.9 − 0.9R , (205) we found a solution exhibiting significant buckling. Figures 16–19 show plots of the buckling angle, height of the buckled middle surface, stresses, and moments in the final configuration. The largest value of ζ is 0.064, around 9% of the radius of the reference configuration. Buckling of this size is visible in Figure 20, which features a top view of the buckled middle surface.

10. Conclusion We have demonstrated the relation of the standard kinematics of finite elasticity to the geometry of differentiable manifolds, and we have seen that kinematics can be adapted slightly to include incompatible growth. We have shown that incompatible growth is equivalent to a change of metric tensor on a Riemannian manifold such that the manifold can no longer be isometrically embedded in Euclidean three-dimensional space. We have also shown that the elastic response found in the multiplicative decomposition F = A · G of the deformation gradient, amounts to a second change of metric, so that the manifold can again be isometrically embedded in E3. We have imported incompatible growth into a geometrically exact model of nonlinearly elastic plates and have solved models showing that incompatible growth alone can induce buckling, without the assistance of applied tractions or moments.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016 McMahon et al. 39

Figure 20. A top view of the buckled middle surface reveals buckling on the order of one-tenth of the radius of the plate.

Funding

This work was supported by the National Science Foundation under grant DMS-0907773. AG also gratefully acknowledges partial support from Award No. KUK-C1-013-04, made by the King Abdullah University of Science and Technology (KAUST).

Notes 1. Infinite differentiability is required in Riemannian geometry, but several continuous derivatives are sufficient in much of solid mechanics.

References [1] McMahon, J, Goriely, A, and Tabor, M. Growing nonlinearly elastic plates I: the genesis of residual stress. Preprint, 2010. [2] Goriely, A, and Ben Amar, M. Differential growth and instability in elastic shells. Phys Rev Lett. 2005; 94: 198103. [3] Ben Amar, M, and Goriely, A. Growth and instability in elastic tissues. J Mech Phys Solids 2005; 53: 2284–2319. [4] Efrati, E, Sharon, E, and Kupferman, R. Elastic theory of unconstrained non-Euclidean plates. J Mech Phys Solids 2009; 57: 762–775. [5] Dervaux, J, Ciarletta, P, and Ben Amar, M. Morphogenesis of thin hyperelastic plates: A constitutive theory of biological growth in the Föppl–von Kármán limit. J Mech Phys Solids 2009; 57: 458–471. [6] Antman, SS, and Marlow, RS. Material constraints, Lagrange multipliers, and compatibility. Arch Rational Mech Anal 1991; 116: 257–299. [7] Marlow, RS. Global material constraints and their associated reactive stresses. J Elasticity 1993; 33: 203–212. [8] Ciarlet, PG. An Introduction to Differential Geometry with Applications to Elasticity. Dordrecht: Springer, 2005. [9] McMahon, J. Geometry and Mechanics of Growing, Nonlinearly Elastic Plates and Membranes. PhD thesis, The University of Arizona, Tucson, Arizona, 2009. [10] Rodriguez, EK, Hoger, A, and McCulloch, AD. Stress-dependent finite growth law in soft elastic tissue. J Biomechanics 1994; 27: 455–467. [11] Chen, Y-C, and Hoger, A. Constitutive functions of elastic materials in finite growth and deformation. J Elasticity 2000; 59: 173–193. [12] Antman, SS. Nonlinear Problems of Elasticity. Berlin: Springer, 1995. [13] Ogden, RW. Non-linear Elastic Deformations. New York: Dover, 1984. [14] Antman, SS, and Negrón-Marrero, PV. The remarkable nature of radially symmetric equilibrium states of nonlinearly elastic bodies. J Elasticity 1987; 18: 131–164.

Downloaded from mms.sagepub.com at Oxford University Libraries on June 9, 2016