This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Mechanics of metric frustration in contorted filament bundles:
From local symmetry to columnar elasticity

Daria W. Atkinson Department of Physics and Astronomy, University of Pennsylvania atkinsod@sas.upenn.edu, she/her or they/them    Christian D. Santangelo Department of Physics, Syracuse University    Gregory M. Grason Department of Polymer Science and Engineering, University of Massachusetts Amherst
Abstract

Bundles of filaments are subject to geometric frustration: certain deformations (e.g. bending while twisted) require longitudinal variations in spacing between filaments. While bundles are common—from protein fibers to yarns—the mechanical consequences of longitudinal frustration are unknown. We derive a geometrically-nonlinear formalism for bundle mechanics, using a gauge-like symmetry under reptations along filament backbones. We relate force balance to orientational geometry and assess the elastic cost of frustration in twisted toroidal bundles.

Elastic zero modes are a ubiquitous feature of soft materials, from mechanical metamaterials [1, 2] to liquid crystal elastomers [3]. Such systems can undergo large deformations with minimal strain, as geometrically coupled rotations and translations preserve local spacing between microscopic constituents. The smectic and columnar liquid crystalline phases provide paradigmatic examples of zero modes in soft elastic systems, permitting relative “sliding” of 2D layers and 1D columns, respectively. The zero-cost sliding displacements of smectic and columnar phases are characteristic of a much broader class of laminated and filamentous structures, ranging from multi-layer graphene materials [4] and stacked paper [5] to biopolymer bundles [6, 7], nanotube yarns [8], wire ropes [9].

While there are well established frameworks which capture the geometric nonlinearities of smectic liquid crystals (i.e., the strain tensor accurately describes arbitrarily large and complex deformations) [10], no such framework exists for columnar and filamentous materials. The orientations of column backbones impose constraints on inter-filament spacing, generating rich modes of geometric frustration without counterpart in smectic liquid crystals. In the simplest non-trivial case of helical bundles, predictions from a minimally non-linear approximation of columnar elasticity [11] and tomographic analysis of elastic filament bundles [12] show that twist in straight bundles gives rise to non-uniform inter-filament stress and spacing in transverse sections (see Fig 1a). Except for the restrictive classes of straight, twisted bundles [13] and twist-free developable domains [14, 15], bundle textures also generate longitudinal frustration, requiring local spacings to vary along a bundle [16]. Although deformations that introduce longitudinal frustration are the rule rather than the exception—for example, wire ropes or toroidal biopolymer condensates are both twisted and bent (e.g. Fig. 1)—existing frameworks of columnar elasticity fail to capture this effect.

Refer to caption
Figure 1: In (a), an equilibrium twisted bundle, with ΩR=1\Omega R=1 and 2D Poisson ratio ν=0.8\nu=0.8, colored by the local pressure. Reptation of the orange filament by σ\sigma along its contour leaves local separation d𝚫d\boldsymbol{\Delta} unchanged. In (b), a twisted-toroidal bundle found by bending the same bundle to κ0=0.2/R\kappa_{0}=0.2/R and optimizing inter-filament elastic cost.

In this Letter, we develop a fully geometrically non-linear Lagrangian elasticity theory of columnar materials, which completely captures the interplay between orientation, and both lateral and longitudinal frustration of inter-filament spacing. We construct this theory by imposing a gauge-like local symmetry under reptations, deformations that slide filaments along their contours without changing the inter-filament spacing. The resultant equilibrium equations point to the role geometrical measures of non-equidistance play in bundles’ mechanics. Within this framework, we compute the energetic costs of longitudinal frustration in twisted, toroidal bundles, and give evidence that 1) optimal configurations generically incorporate splay and 2) the bending cost that derives from non-uniform compression depends non-monotonically on pretwist.

To construct the elastic theory, we divide space into points on curves (i.e. filament backbones) labeled by two coordinates: 𝐯\mathbf{v}, a 2D label of filaments; and ss, a length coordinate along filaments. Hence, the location of each point in the bundle is described by a function 𝐫(s,𝐯)\mathbf{r}(s,\mathbf{v}), with s𝐫(s,𝐯)\partial_{s}\mathbf{r}(s,\mathbf{v}) parallel to the tangent vector 𝐭=s𝐫/|s𝐫|\mathbf{t}={\partial_{s}\mathbf{r}}/{|\partial_{s}\mathbf{r}|}. Because they lack positional order along their backbone curves, filament bundles and columnar liquid crystals have a family of continuous zero modes, corresponding to reptations (Fig. 1),

𝐫(s,𝐯)=𝐫(s+σ(s,𝐯),𝐯).\mathbf{r}^{\prime}(s,\mathbf{v})=\mathbf{r}(s+\sigma(s,\mathbf{v}),\mathbf{v}). (1)

We assume that changes in local spacing can be described by a hyper-elastic energy density function 𝒲{\cal W}, that depends only on the deformation gradient [17].

To account for reptation symmetry, we demand that 𝒲{\cal W} depend on a modified deformation gradient, which transforms as a scalar under σ(s,𝐯)\sigma(s,\mathbf{v}), depends solely on the deformation, 𝐫\mathbf{r}, of the material itself, and recovers the well-established 2D elasticity of developable [14, 15] bundles. Specifically, we construct a covariant derivative DI𝐫=I𝐫𝐀ID_{I}\mathbf{r}=\nabla_{I}\mathbf{r}-\mathbf{A}_{I}, where I\nabla_{I} is the usual covariant derivative on tensors in the material space and 𝐫\mathbf{r} determines 𝐀I\mathbf{A}_{I}, such that if two configurations, 𝐫\mathbf{r} and 𝐫\mathbf{r}^{\prime}, are related by Eq. (1), then DI𝐫=DI𝐫D_{I}\mathbf{r}=D^{\prime}_{I}\mathbf{r}^{\prime}. In order for DI𝐫D_{I}\mathbf{r} to be reptation invariant, we must have that 𝐀I+𝐀I=Iσs𝐫-\mathbf{A}^{\prime}_{I}+\mathbf{A}_{I}=\nabla_{I}\sigma\partial_{s}\mathbf{r}. Therefore, 𝐀I+𝐀I=(𝐭I𝐫)𝐭+(𝐭I𝐫)𝐭-\mathbf{A}^{\prime}_{I}+\mathbf{A}_{I}=-(\mathbf{t}\cdot\nabla_{I}\mathbf{r}^{\prime})\mathbf{t}+(\mathbf{t}\cdot\nabla_{I}\mathbf{r})\mathbf{t}. To construct an elastic theory of columnar materials, we set 𝐀I=(𝐭I𝐫)𝐭\mathbf{A}_{I}=(\mathbf{t}\cdot\nabla_{I}\mathbf{r})\mathbf{t}, which is manifestly reptation invariant but also leads to a deformation gradient that only measures deformations transverse to the local backbones,

DI𝐫I𝐫(𝐭I𝐫)𝐭.D_{I}\mathbf{r}\equiv\nabla_{I}\mathbf{r}-(\mathbf{t}\cdot\nabla_{I}\mathbf{r})\mathbf{t}. (2)

Notably, for two nearby filaments at 𝐫(𝐯){\bf r}({\bf v}) and 𝐫(𝐯+d𝐯){\bf r}({\bf v}+d{\bf v}) in material coordinates, it is straightforward to show that the covariant derivative gives the local distance of closest approach d𝚫=dvIDI𝐫d{\bf\Delta}=dv^{I}~{}D_{I}\mathbf{r}, for which d𝚫𝐭=0d{\bf\Delta}\cdot{\bf t}=0 (see again inset of Fig. 1a). As shown explicitly in the Appendix, this covariant deformation gradient captures the standard 2D deformation gradients of developable domains (i.e. parallel arrays).

From this deformation gradient, we construct an effective metric gIJeff=DI𝐫DJ𝐫g^{\text{eff}}_{IJ}=D_{I}\mathbf{r}\cdot D_{J}\mathbf{r}, which is naturally invariant under rotations of 𝐫\mathbf{r}, and which encodes the metric inherited by the local 2D section transverse to the filaments in the bundle 111Equidistant bundles where components of gIJeffg^{\text{eff}}_{IJ} are uniform along ss are known as Riemannian fibrations, see e.g.  [36, 37], and the transverse structure can be mapped on to single 2D base space. Non-equidistant bundles correspond to Sub-Riemannian generalizations [38] where the inherited metric measures distances in the 2-planes perpendicular to 𝐭{\bf t} at each point.. Because Ds𝐫=0D_{s}\mathbf{r}=0, the effective metric only has components for 2×22\times 2 block I,JsI,J\neq s, which we denote using index notation α\alpha, β{1,2}\beta\in\{1,2\}. With these definitions, we construct the Green-Saint-Venant strain tensor, [17, 19, 20]

ϵαβ=12[Dα𝐫Dβ𝐫gαβtar],\epsilon_{\alpha\beta}=\tfrac{1}{2}\big{[}D_{\alpha}\mathbf{r}\cdot D_{\beta}\mathbf{r}-g_{\alpha\beta}^{\text{tar}}\big{]}, (3)

where gαβtarg_{\alpha\beta}^{\text{tar}} is the target metric corresponding to strain-free state, which for this Letter, we take to be developable with uniform spacing, so gαβtar=δαβg_{\alpha\beta}^{\text{tar}}=\delta_{\alpha\beta}. For weak deflections from the uniform parallel state, Eq. (3) reduces to the small-tilt approximation to the non-linear columnar strain [21, 22, 11], which captures the lowest-order dependence of spacing on orientation (see Appendix).

Assuming that strains are small though deformations may be large, the Hookean elastic energy takes the usual form,

s=𝑑V𝒲(ϵ)=12𝑑VSαβϵαβ,{\cal E}_{s}=\int dV~{}{\cal W}(\epsilon)=\frac{1}{2}\int dVS^{\alpha\beta}\epsilon_{\alpha\beta}, (4)

where Sαβ=𝒲ϵαβ=CαβγδϵγδS^{\alpha\beta}=\tfrac{\partial{\cal W}}{\partial\epsilon_{\alpha\beta}}=C^{\alpha\beta\gamma\delta}\epsilon_{\gamma\delta} is the nominal stress tensor, and CαβγδC^{\alpha\beta\gamma\delta} is a tensor of elastic constants which depends on both the crystalline symmetries of the underlying columnar order and the target metric, gαβtarg_{\alpha\beta}^{\text{tar}} 222We raise and lower indices with gαβtarg^{\alpha\beta\text{tar}}, the inverse of gαβtarg_{\alpha\beta}^{\text{tar}}, rather than gαβeffg^{\alpha\beta\text{eff}}, as a matter of convenience, and note that the difference in the elastic energy is higher order in the strain tensor [20].. Energetics of columnar materials also include other gauge-invariant costs, including the Frank-Oseen orientational free energy and the cost of local density changes along columns 333The symmetry under local reptations in this theory is unlike a related treatment of both the nematic to smectic-A transition [39, 40] and its columnar analog [41, 42, 43] in which the gauge symmetry of a density-wave is explicitly broken by, e.g., the splay elasticity of the underlying nematic order.. Here, for clarity, we focus only on the energetics of columnar strain and detail the combined effects of other contributions elsewhere [25]. Given this gauge-invariant formulation of the columnar strain energy, we first illustrate the mechanical effects of orientational geometry on local forces. This follows from the bulk Euler-Lagrange equations of Eq. (21) (see Appendix for a complete derivation),

δsδ𝐫=α(SαβDβ𝐫)+s(SαβDβ𝐫𝐭α𝐫|s𝐫|).\frac{\delta\mathcal{E}_{s}}{\delta{\bf r}}=-\nabla_{\alpha}\big{(}S^{\alpha\beta}D_{\beta}\mathbf{r}\big{)}+\nabla_{s}\Big{(}S^{\alpha\beta}D_{\beta}\mathbf{r}\frac{\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r}}{|\nabla_{s}\mathbf{r}|}\Big{)}. (5)

The bulk terms represent body forces generated by the columnar strains, which must be balanced by other internal stresses or external forces. To cast them in a more geometrical light, we consider separately the components tangential and perpendicular to 𝐭{\bf t}, FF_{\parallel} and 𝐅{\bf F}_{\perp}, respectively.

Making use of the identity, 𝐭αDβ𝐫=α𝐭β𝐫{\bf t}\cdot\nabla_{\alpha}D_{\beta}{\bf r}=-\partial_{\alpha}{\bf t}\cdot\partial_{\beta}{\bf r}, the tangential forces can be recast simply as

F=Sαβhαβ,F_{\parallel}=S^{\alpha\beta}h_{\alpha\beta}, (6)

where

hαβ\displaystyle h_{\alpha\beta} =12[α𝐭β𝐫+β𝐭α𝐫\displaystyle=\tfrac{1}{2}\big{[}\partial_{\alpha}\mathbf{t}\cdot\partial_{\beta}\mathbf{r}+\partial_{\beta}\mathbf{t}\cdot\partial_{\alpha}\mathbf{r} (7)
𝐭α𝐫|s𝐫|s𝐭β𝐫𝐭β𝐫|s𝐫|s𝐭α𝐫],\displaystyle-\frac{\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\partial_{s}\mathbf{t}\cdot\partial_{\beta}\mathbf{r}-\frac{\mathbf{t}\cdot\partial_{\beta}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\partial_{s}\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}\big{]},

is the convective flow tensor that measures longitudinal variations in inter-filament spacing [16]. Just as the second fundamental form measures gradients of a surface’s normal vector [26], hαβh_{\alpha\beta} measure the symmetric gradients of 𝐭\mathbf{t} in its normal plane (i.e. its trace is the splay of filament tangents).

Here, we see that tangent forces couple non-equidistance to the stress tensor much like the Young-Laplace law couples in-plane stresses to normal forces in curved membranes [19]. This analogy becomes exact for zero twist textures: when 𝐭(×𝐭)=0\mathbf{t}\cdot(\nabla\times\mathbf{t})=0, filaments can be described by a set of surfaces normal to 𝐭\mathbf{t}. In this case, it is possible to choose coordinates so that α𝐫𝐭=0\partial_{\alpha}{\bf r}\cdot{\bf t}=0, and tangential forces give the Young-Laplace force normal to each surface, with hαβh_{\alpha\beta} reducing to their second fundamental form. This illustrates the intuitive notion, shown schematically in Fig. 2, that columnar strain generates tangential body forces that push material points towards lower-stress locations in the array.

Refer to caption
Figure 2: A 2D section of filament bundle colored by the local pressure, with regions under compression in blue and those under extension in red. The forces parallel and tangent to 𝐭\mathbf{t} are shown at two points, and push material points to regions of vanishing stress.

The bulk components of Eq. (30) perpendicular to 𝐭\mathbf{t} give the transverse force

𝐅\displaystyle\mathbf{F}_{\perp} =Dα[SαβDβ𝐫]+Ds[Sαβ𝐭α𝐫|s𝐫|Dβ𝐫].\displaystyle=-D_{\alpha}\big{[}S^{\alpha\beta}D_{\beta}\mathbf{r}\big{]}+D_{s}\big{[}S^{\alpha\beta}\frac{\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}}{|\partial_{s}\mathbf{r}|}D_{\beta}\mathbf{r}\big{]}. (8)

This form captures the divergence of stress in the planes perpendicular to backbones. The second term accounts for the corrections arising from material derivatives that lie along the backbone, such as twisted textures, when α𝐫𝐭0\partial_{\alpha}{\bf r}\cdot{\bf t}\neq 0. Thus, the longitudinal derivatives in 𝐅\mathbf{F}_{\perp} are needed to capture transverse mechanics of even equidistant twisted bundles beyond the lowest order geometric non-linearity [11].

We now illustrate the energetics of longitudinal frustration by considering a prototypical non-equidistant geometry: twisted toroidal bundles. Motivated in large part by the morphologies of condensed biopolymers [6, 7], theoretical models of twisted toroids have focused on their orientational elasticity costs [27, 28, 29], ignoring the unavoidable frustration of inter-filament spacing in this geometry. While satisfying force balance in non-equidistant bundles requires physical ingredients beyond the columnar strain energy, which we consider elsewhere [25], for the purposes of this Letter we take advantage of the full geometric-nonlinearity of Eq. (30) to explore the specific costs of longitudinal gradients in spacing required by simultaneous twist and bend.

We construct twisted toroids from equilibrium twisted helical bundles of radius RR and constant pitch, 2π/Ω{2\pi}{/\Omega} by bending them such that their central curve 𝐫0{\bf r}_{0} is deformed from a straight line into a circle of radius κ01\kappa_{0}^{-1} (see Fig. 1b). We then define perturbative displacements 𝐫(s+δs,ρ+δρ,ϕ+δϕ)𝐫0+ρρ^+s𝐫(0)δs+δρρ^+δϕϕ^\mathbf{r}(s+\delta s,\rho+\delta\rho,\phi+\delta\phi)\simeq\mathbf{r}_{0}+\rho\hat{\rho}+\partial_{s}\mathbf{r}^{(0)}\delta s+\delta\rho\hat{\rho}+\delta\phi\hat{\phi} relative to the bent, pre-twisted bundles, where ρ\rho and ϕ\phi describe the (Eulerian) distance from the central curve and the angular position relative to its normal in the plane perpendicular to its tangent 𝐭0{\bf t}_{0}, and where s𝐫(0)=𝐭0+Ωρϕ^\partial_{s}\mathbf{r}^{(0)}={\bf t}_{0}+\Omega\rho\hat{\phi}. The small-ρ\rho limit of the force balance equations for the strain energy motivates the following displacements

(δsδρδϕ)=(asΩκ0ρ3sinϕaρΩ2κ0ρ4cosϕaϕΩ2κ0ρ3sinϕ)\begin{pmatrix}\delta s\\ \delta\rho\\ \delta\phi\end{pmatrix}=\begin{pmatrix}a_{s}\Omega\kappa_{0}\rho^{3}\sin{\phi}\\ a_{\rho}\Omega^{2}\kappa_{0}\rho^{4}\cos{\phi}\\ a_{\phi}\Omega^{2}\kappa_{0}\rho^{3}\sin{\phi}\end{pmatrix} (9)

where asa_{s}, aρa_{\rho}, and aϕa_{\phi} are variational parameters. Notably, to linear order in curvature, these parameterize the“almost equidistant” ansatzes considered previously, including splay-free (tr(h)=0\text{tr}(h)=0[27] configurations and det(h)=0\text{det}(h)=0 [16] ansatzes.

We expand the energy to quadratic order in κ0\kappa_{0}, holding the center of area at 𝐫0\mathbf{r}_{0}, which constrains as=aρaϕa_{s}=a_{\rho}-a_{\phi}, then minimize Eq. (21) with respect to the displacements for a given ΩR\Omega R. Examples of the distribution of pressure P=Sαα/2P=S^{\alpha}_{\alpha}/2 are shown in Fig. 1b. Relative to the axisymmetric pressure induced by helical twist in the straight bundle, bending into a twisted toroid requires bunching (spreading) of the filaments at the inner (outer) positions in the toroid, leading to a polarization of the pressure towards the normal.

Because bending and twisting of bundles introduces longitudinal strain variation, bending pre-twisted bundle introduces additional stresses whose elastic cost can be characterized by an effective bending stiffness BB, defined by Ebend=B2𝑑sκ02E_{\text{bend}}=\tfrac{B}{2}\int ds\kappa_{0}^{2}, which derives purely from columnar strain, rather than intra-filament deformations. Fig. 3 shows that longitudinal frustration leads to a bending cost that increases with small twist as B(ΩR)4B\sim(\Omega R)^{4}, but eventually gives way to remarkable non-monotonic behavior at large pre-twist. We note further that the bending cost grows with the 2D Poisson ratio ν\nu of the columnar array, highlighting the importance of local compressional deformations in optimal twisted toroids.

Refer to caption
Figure 3: The effective bending modulus BB which results from the frustration of constant spacing in twisted-toroidal bundles. The inset shows dimensionless measures of mean splay (Ψ\Psi) and biaxial splay (Γ\Gamma) defined in the text, for fixed twist ΩR=0.8\Omega R=0.8 for a range of 2D Poisson ratios, from ν=0.2\nu=0.2 to 1.0. Approaching incompressiblity (ν1\nu\to 1), Ψ\Psi goes to zero, maintaining uniform area-per-filament at the expense of expense of markedly increased strain energy.

We analyze the optimal modes of deformation via the convective flow tensor, in particular, the trace tr(h){\rm tr}(h) (splay) and deviatoric components hdev=hαβtr(h)δαβ/2h^{\text{dev}}=h_{\alpha\beta}-{\rm tr}(h)\delta_{\alpha\beta}/2 (biaxial splay) [30], which characterize longitudinal gradients of dilatory and shear stress in the columnar array. In contrast to a heuristic view that optimal packings should favor the uniform area per filament of splay-free textures, the inset of Fig. 3 instead shows that optimal toroids incorporate a mixture of both splay and biaxial splay where we define the respective measures of average splay and biaxial splay, Ψ1κ02V𝑑Vtr(h)2\Psi\equiv\tfrac{1}{\kappa_{0}^{2}V}\int dV\text{tr}(h)^{2} and Γ1κ02V𝑑Vtr(hdev2)\Gamma\equiv\tfrac{1}{\kappa_{0}^{2}V}\int dV\text{tr}(h_{\text{dev}}^{2}). Only in the incompressible limit, as ν1\nu\to 1, does the splay vanish, and only at the expense of additional biaxial splay and energetic cost, implying counterintuitively that splayed textures are in fact energetically favorable in longitudinally frustrated twisted toroids. Indeed, the energetic preference for splay in non-equidistant bundles, can be traced to force balance conditions in this geometry [25].

In summary, we have shown that gauge-theoretic principles underlie the geometrically-nonlinear theory of columnar elasticity, providing a means to quantify the cost of longitudinal frustration in the mechanics of bundles. Unlike phase field models of nonlinear elasticity (such as [31, 32]), this description depends neither on the presence of a planar reference state, nor presupposes uniform crystalline order, allowing us to both accommodate the effective curvature of bundles of constant pitch helices [33], and providing a natural generalization to arbitrary target metrics [19]. Finally, because this approach to elasticity relies only on the existence of local, continuous zero modes, we note that it can be generalized to other liquid crystals, like smectics, and anticipate that it may have applications beyond liquid crystals, including mechanical metamaterials.

Acknowledgements.
The authors gratefully acknowledge useful discussions with R. Kusner, S. Zhou, and B. Davidovitch. M. Dimitriyev, D. Hall, and R. Kamien provided feedback on an early draft of this manuscript. This work was supported by the NSF under grants DMR-1608862 and NSF DMR-1822638; D.A. received additional funding from the Penn Provost’s postdoctoral fellowship and NSF MRSEC DMR-1720530.

References

Appendix A The form of the deformation gradient

To construct the covariant derivative in Eq. (2), we demand, in addition to reptation symmetry and dependence only on gradients of the deformation, that DI𝐫=𝐫𝐀ID_{I}\mathbf{r}=\nabla\mathbf{r}-\mathbf{A}_{I} reduce to the usual 2D deformation gradient for the well studied developable bundles [34, 35], where the cross-sectional geometry is Euclidean [14, 15]. This constrains the value of 𝐀I=(𝐭I𝐫)𝐭\mathbf{A}_{I}=(\mathbf{t}\cdot\partial_{I}\mathbf{r})\mathbf{t} in Eq. (2). Here, we illustrate that this reduces to the expected 2D planar elasticity of developable bundles.

In a developable bundle, all curves are everywhere to a common set of planes and therefore, share a common tangent in those planes. Choosing a curve in the bundle 𝐫0(s)\mathbf{r}_{0}(s) with tangent vector T^0\hat{T}_{0}, this condition requires that 𝐭(s,v)=T^0(s){\bf t}(s,\vec{v})=\hat{T}_{0}(s). Introducing a “twist-free”, right-handed, orthonormal frame {e^1(s),e^2(s),T^0(s)}\{\hat{e}_{1}(s),\hat{e}_{2}(s),\hat{T}_{0}(s)\},

sT^0\displaystyle\partial_{s}\hat{T}_{0} =κ1e^1+κ2e^2\displaystyle=\kappa_{1}\hat{e}_{1}+\kappa_{2}\hat{e}_{2}
se^1\displaystyle\partial_{s}\hat{e}_{1} =κ1T^0\displaystyle=-\kappa_{1}\hat{T}_{0} (10)
se^2\displaystyle\partial_{s}\hat{e}_{2} =κ2T^0.\displaystyle=-\kappa_{2}\hat{T}_{0}.

We can now see that a deformation 𝐫\mathbf{r} yields a developable bundle (up to reptations) when it can be written as

𝐫(s,v)=𝐫0(s)+w1(v)e^1(s)+w2(v)e^2(s),\mathbf{r}(s,\vec{v})=\mathbf{r}_{0}(s)+w_{1}(\vec{v})\hat{e}_{1}(s)+w_{2}(\vec{v})\hat{e}_{2}(s), (11)

where w1w_{1} and w2w_{2} are deformation fields depending only on the orthogonal (v\vec{v}) position, as the tangent field 𝐭=s𝐫/|s𝐫|\mathbf{t}=\partial_{s}\mathbf{r}/|\partial_{s}\mathbf{r}| is independent of v1v_{1} and v2v_{2}.

The most generic form of the covariant deformation gradient invariant under reptations is

D~I𝐫=I𝐫𝐀~I[𝐫](𝐭I𝐫)𝐭,\tilde{D}_{I}\mathbf{r}=\partial_{I}\mathbf{r}-\tilde{\mathbf{A}}_{I}\big{[}\nabla\mathbf{r}\big{]}-(\mathbf{t}\cdot\partial_{I}\mathbf{r})\mathbf{t}, (12)

where 𝐀~I[𝐫]\tilde{\mathbf{A}}_{I}\big{[}\nabla\mathbf{r}\big{]} is an unknown function of 𝐫\nabla\mathbf{r} (i.e. a potential non-zero value of 𝐀I(𝐭I𝐫)𝐭\mathbf{A}_{I}-(\mathbf{t}\cdot\partial_{I}\mathbf{r})\mathbf{t}). From eq. (11) we find that

D~s𝐫\displaystyle\tilde{D}_{s}\mathbf{r} =𝐀~s\displaystyle=-\tilde{\mathbf{A}}_{s} (13)
D~v1𝐫\displaystyle\tilde{D}_{v_{1}}\mathbf{r} =v1w1e^1+v1w2e^2𝐀~v1\displaystyle=\partial_{v_{1}}w_{1}~{}\hat{e}_{1}+\partial_{v_{1}}w_{2}~{}\hat{e}_{2}-\tilde{\mathbf{A}}_{v_{1}} (14)
D~v2𝐫\displaystyle\tilde{D}_{v_{2}}\mathbf{r} =v2w1e^1+v2w2e^2𝐀~v2.\displaystyle=\partial_{v_{2}}w_{1}~{}\hat{e}_{1}+\partial_{v_{2}}w_{2}~{}\hat{e}_{2}-\tilde{\mathbf{A}}_{v_{2}}. (15)

Namely, 𝐀~s-\tilde{\mathbf{A}}_{s} is the only term remaining in the ss component, while in the v1v_{1} and v2v_{2} components, 𝐀~vα\tilde{\mathbf{A}}_{v_{\alpha}} is subtracted from the standard 2D deformation gradient in the planes normal to T^0\hat{T}_{0}. Hence, in order that strains recover the elastic distortions of transverse distances in the columnar structure, for developable structure we must have 𝐀~v1=𝐀~v2=0\tilde{\mathbf{A}}_{v_{1}}=\tilde{\mathbf{A}}_{v_{2}}=0.

A similar argument constrains 𝐀~s\tilde{\mathbf{A}}_{s}. Again, because we expect that well established descriptions of 2D elasticity hold for developable bundles, we have that 𝐀~s\tilde{\mathbf{A}}_{s} should be independent of ss independent displacements 𝐰(v1,v2)\mathbf{w}(v_{1},v_{2}), as well as reptations, and that 𝐀~sDvα𝐫=0\tilde{\mathbf{A}}_{s}\cdot D_{v_{\alpha}}\mathbf{r}=0. This leaves 𝐀~s=c𝐭\tilde{\mathbf{A}}_{s}=c\mathbf{t}, where cc is independent of the deformation. Since, at the level of metric, gsIeff=c2δsIg^{\text{eff}}_{sI}=c^{2}\delta_{sI} is independent of the deformation, no matter what cc is, it will not appear in the strain tensor, which has by definition of eq. (3) components only for IsI\neq s. For simplicity, we take 𝐀~s=0\tilde{\mathbf{A}}_{s}=0.

Appendix B Small tilt limit

We can recover the small-tilt approximation of the strain tensor from [21, 22, 11] starting from Eq. (3), where the fully geometrically-nonlinear strain, ϵαβ\epsilon_{\alpha\beta}, is

ϵαβ=12[Dα𝐫Dβ𝐫gαβtar].\epsilon_{\alpha\beta}=\tfrac{1}{2}\big{[}D_{\alpha}\mathbf{r}\cdot D_{\beta}\mathbf{r}-g_{\alpha\beta}^{\text{tar}}\big{]}. (16)

We break the deformation 𝐫\mathbf{r} up into the unstrained, cartesian coordinates 𝐱=xx^+yy^+zz^\mathbf{x}=x~{}\hat{x}+y~{}\hat{y}+z~{}\hat{z} and a displacement field 𝐮\mathbf{u} in the xyx-y plane, taking the arc-coordinate s=zs=z. Then,

Dα𝐫Dβ𝐫=δαβ+αuβ+βuα+α𝐮β𝐮(𝐭α𝐫)(𝐭β𝐫),D_{\alpha}\mathbf{r}\cdot D_{\beta}\mathbf{r}=\delta_{\alpha\beta}+\partial_{\alpha}u_{\beta}+\partial_{\beta}u_{\alpha}+\partial_{\alpha}\mathbf{u}\cdot\partial_{\beta}\mathbf{u}-(\mathbf{t}\cdot\partial_{\alpha}\mathbf{r})(\mathbf{t}\cdot\partial_{\beta}\mathbf{r}), (17)

where 𝐭=(z^+s𝐮)/1+(s𝐮)2\mathbf{t}=(\hat{z}+\partial_{s}\mathbf{u})/\sqrt{1+(\partial_{s}\mathbf{u})^{2}} and uβ=𝐮(β𝐱)u_{\beta}=\mathbf{u}\cdot(\partial_{\beta}\mathbf{x}), and α,β=x,y\alpha,\beta=x,y. Subtracting off the Euclidean target metric, gαβtar=δαβg_{\alpha\beta}^{\text{tar}}=\delta_{\alpha\beta}, we have

ϵαβ=12[αuβ+βuα+α𝐮β𝐮(𝐭α𝐫)(𝐭β𝐫)].\epsilon_{\alpha\beta}=\tfrac{1}{2}\big{[}\partial_{\alpha}u_{\beta}+\partial_{\beta}u_{\alpha}+\partial_{\alpha}\mathbf{u}\cdot\partial_{\beta}\mathbf{u}-(\mathbf{t}\cdot\partial_{\alpha}\mathbf{r})(\mathbf{t}\cdot\partial_{\beta}\mathbf{r})\big{]}. (18)

Now substituting for 𝐭\mathbf{t} and 𝐫\mathbf{r} in the last term, and grouping terms by power of the displacement field 𝐮\mathbf{u}, we have:

ϵαβ\displaystyle\epsilon_{\alpha\beta} =12[αuβ+βuα+α𝐮β𝐮suαsuβ1+|s𝐮|2\displaystyle=\tfrac{1}{2}\big{[}\partial_{\alpha}u_{\beta}+\partial_{\beta}u_{\alpha}+\partial_{\alpha}\mathbf{u}\cdot\partial_{\beta}\mathbf{u}-\frac{\partial_{s}u_{\alpha}\partial_{s}u_{\beta}}{1+|\partial_{s}\mathbf{u}|^{2}}
s𝐮α𝐮suβ1+|s𝐮|2(s𝐮β𝐮)suα1+|s𝐮|2(s𝐮α𝐮)(s𝐮β𝐮)1+|s𝐮|2.\displaystyle-\frac{\partial_{s}\mathbf{u}\cdot\partial_{\alpha}\mathbf{u}\partial_{s}u_{\beta}}{1+|\partial_{s}\mathbf{u}|^{2}}-\frac{(\partial_{s}\mathbf{u}\cdot\partial_{\beta}\mathbf{u})\partial_{s}u_{\alpha}}{1+|\partial_{s}\mathbf{u}|^{2}}-\frac{(\partial_{s}\mathbf{u}\cdot\partial_{\alpha}\mathbf{u})(\partial_{s}\mathbf{u}\cdot\partial_{\beta}\mathbf{u})}{1+|\partial_{s}\mathbf{u}|^{2}}. (19)

Expanding the denominator for small displacement fields, and keeping only terms which are quadratic in 𝐮\mathbf{u} recovers the rotationally invariant strain tensor,

ϵαβ12[αuβ+βuα+α𝐮β𝐮suαsuβ].\epsilon_{\alpha\beta}\simeq\tfrac{1}{2}\big{[}\partial_{\alpha}u_{\beta}+\partial_{\beta}u_{\alpha}+\partial_{\alpha}\mathbf{u}\cdot\partial_{\beta}\mathbf{u}-\partial_{s}u_{\alpha}\partial_{s}u_{\beta}\big{]}. (20)

Appendix C Derivation of the force-balance equations

We are looking for local extrema of the elastic energy

s=𝑑V𝒲(ϵ)=12𝑑VSαβϵαβ,{\cal E}_{s}=\int dV~{}{\cal W}(\epsilon)=\frac{1}{2}\int dVS^{\alpha\beta}\epsilon_{\alpha\beta}, (21)

with respect to the deformation, 𝐫\mathbf{r}, where dV=dsdv1dv2det(gtar)dV=dsdv^{1}dv^{2}\sqrt{\text{det}(g^{\text{tar}})}, gtarg^{\text{tar}} is the target metric, as in Eq. (3), ϵαβ\epsilon_{\alpha\beta} is the strain tensor, as in Eq. (3), and SαβS^{\alpha\beta} is the nominal stress tensor, as defined following Eq. (5). As such, we consider arbitrary variations δ𝐫\delta\mathbf{r} of the energy around these local extrema, so that the restoring force on a small material volume is given by:

δsδ𝐫=𝑑VSαβδϵαβδ𝐫.\frac{\delta{\cal E}_{s}}{\delta\mathbf{r}}=\int dVS^{\alpha\beta}\frac{\delta\epsilon_{\alpha\beta}}{\delta\mathbf{r}}. (22)

What remains then is to work out δϵαβδ𝐫\frac{\delta\epsilon_{\alpha\beta}}{\delta\mathbf{r}}, and apply the divergence theorem to derive the conditions of force balance. First, we note that ϵαβ=12[Dα𝐫Dβ𝐫gαβtar],\epsilon_{\alpha\beta}=\tfrac{1}{2}\big{[}D_{\alpha}\mathbf{r}\cdot D_{\beta}\mathbf{r}-g_{\alpha\beta}^{\text{t}ar}], and that the covariant derivative \nabla is just the usual partial derivative on scalars in the material space, so I𝐫=I𝐫\nabla_{I}\mathbf{r}=\partial_{I}\mathbf{r}. We then have

δϵαβ=Dβ𝐫δ[Dα𝐫].\delta\epsilon_{\alpha\beta}=D_{\beta}\mathbf{r}\cdot\delta\big{[}D_{\alpha}\mathbf{r}\big{]}. (23)

Since Dβ𝐫𝐭=0D_{\beta}\mathbf{r}\cdot\mathbf{t}=0, only two terms in δ[Dα𝐫]\delta\big{[}D_{\alpha}\mathbf{r}\big{]} contribute to δϵαβ\delta\epsilon_{\alpha\beta}:

δ(α𝐫)=δ(α𝐫)=αδ𝐫,\delta(\nabla_{\alpha}\mathbf{r})=\delta(\partial_{\alpha}\mathbf{r})=\partial_{\alpha}\delta\mathbf{r}, (24)

and

(𝐭α𝐫)δ𝐭=(𝐭α𝐫)1|s𝐫|[sδ𝐫𝐭(𝐭sδ𝐫)].(\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r})\delta\mathbf{t}=(\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r})\frac{1}{|\partial_{s}\mathbf{r}|}\big{[}\partial_{s}\delta\mathbf{r}-\mathbf{t}(\mathbf{t}\cdot\partial_{s}\delta\mathbf{r})\big{]}. (25)

All together, we have that

δϵαβ=DJ𝐫[αδ𝐫𝐭α𝐫sδ𝐫|s𝐫|],\delta\epsilon_{\alpha\beta}=D_{J}\mathbf{r}\Big{[}\partial_{\alpha}\delta\mathbf{r}-\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}\frac{\partial_{s}\delta\mathbf{r}}{|\partial_{s}\mathbf{r}|}\Big{]}, (26)

where again we use that Dβ𝐫𝐭=0D_{\beta}\mathbf{r}\cdot\mathbf{t}=0. Substituting this back into our integral, we find that

δs=𝑑VSαβDβ𝐫[αδ𝐫𝐭α𝐫sδ𝐫|s𝐫|].\delta{\cal E}_{s}=\int dVS^{\alpha\beta}D_{\beta}\mathbf{r}\cdot\Big{[}\partial_{\alpha}\delta\mathbf{r}-\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r}\frac{\partial_{s}\delta\mathbf{r}}{|\partial_{s}\mathbf{r}|}\Big{]}. (27)

Now we apply the divergence theorem, finding that

0\displaystyle 0 =𝑑s𝑑v1𝑑v2α[det(gtar)SαβDβ𝐫]δ𝐫+𝑑s𝑑v1𝑑v2𝐬[det(gtar)SαβDβ𝐫𝐭α𝐫|s𝐫|]δ𝐫\displaystyle=-\int dsdv^{1}dv^{2}\partial_{\alpha}\Big{[}\sqrt{\text{det}(g^{\text{tar}})}S^{\alpha\beta}D_{\beta}\mathbf{r}\Big{]}\cdot\delta\mathbf{r}+\int dsdv^{1}dv^{2}\partial_{\mathbf{s}}\Big{[}\sqrt{\text{det}(g^{\text{tar}})}S^{\alpha\beta}D_{\beta}\mathbf{r}\frac{\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\Big{]}\cdot\delta\mathbf{r} (28)
+𝑑An^α[det(gtar)SαβDβ𝐫]δ𝐫𝑑An^s[det(gtar)SαβDβ𝐫𝐭α𝐫|s𝐫|]δ𝐫,\displaystyle+\int dA~{}\hat{n}_{\alpha}\big{[}\sqrt{\text{det}(g^{\text{tar}})}S^{\alpha\beta}D_{\beta}\mathbf{r}\big{]}\cdot\delta\mathbf{r}-\int dA~{}\hat{n}_{s}\big{[}\sqrt{\text{det}(g^{\text{tar}})}S^{\alpha\beta}D_{\beta}\mathbf{r}\frac{\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\big{]}\cdot\delta\mathbf{r}, (29)

where n^\hat{n} is the vector normal to the boundary of the material. Eq. (29), gives the boundary forces associated with residual stresses at the boundary, either directly from the stress tensor (first term), or from a non-zero flux of filament ends (second term). Now noting that for a vector fIf^{I} in the material space, I(det(gtar)fI)=det(gtar)IfI\partial_{I}\big{(}\sqrt{\text{det}(g^{\text{tar}})}f^{I})=\sqrt{\text{det}(g^{\text{tar}})}\nabla_{I}f^{I}, and that, since δ𝐫\delta\mathbf{r} is an arbitrary variation, everything dotted into it must vanish, Eq. (28) reduces to Eq (5):

δsδ𝐫=α(SαβDβ𝐫)+s(SαβDβ𝐫𝐭α𝐫|s𝐫|).\frac{\delta\mathcal{E}_{s}}{\delta{\bf r}}=-\nabla_{\alpha}\big{(}S^{\alpha\beta}D_{\beta}\mathbf{r}\big{)}+\nabla_{s}\Big{(}S^{\alpha\beta}D_{\beta}\mathbf{r}\frac{\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r}}{|\nabla_{s}\mathbf{r}|}\Big{)}. (30)

We obtain Eq (6)–(8) by projecting Eq. (30) along and perpendicular to 𝐭\mathbf{t}. The 𝐭\mathbf{t} component is

Sαβ(𝐭[αDβ𝐫])+Sαβ(𝐭[sDβ𝐫])𝐭α𝐫|s𝐫|.-S^{\alpha\beta}\Big{(}\mathbf{t}\cdot\big{[}\nabla_{\alpha}D_{\beta}\mathbf{r}\big{]}\big{)}+S^{\alpha\beta}\Big{(}\mathbf{t}\cdot\big{[}\nabla_{s}D_{\beta}\mathbf{r}\big{]}\Big{)}\frac{\mathbf{t}\cdot\nabla_{\alpha}\mathbf{r}}{|\nabla_{s}\mathbf{r}|}. (31)

Because 𝐭αDβ𝐫=α𝐭Dβ𝐫\mathbf{t}\cdot\nabla_{\alpha}D_{\beta}\mathbf{r}=-\partial_{\alpha}\mathbf{t}\cdot D_{\beta}\mathbf{r}, 𝐭sDβ𝐫=s𝐭Dβ𝐫\mathbf{t}\cdot\nabla_{s}D_{\beta}\mathbf{r}=-\partial_{s}\mathbf{t}\cdot D_{\beta}\mathbf{r}, and because the stress tensor is symmetric, we can rewrite this in terms of the convective flow tensor of the bundle,

hαβ=12[α𝐭β𝐫+β𝐭α𝐫𝐭α𝐫|s𝐫|s𝐭β𝐫𝐭β𝐫|s𝐫|s𝐭α𝐫],h_{\alpha\beta}=\tfrac{1}{2}\Big{[}\partial_{\alpha}\mathbf{t}\cdot\partial_{\beta}\mathbf{r}+\partial_{\beta}\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}-\frac{\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\partial_{s}\mathbf{t}\cdot\partial_{\beta}\mathbf{r}-\frac{\mathbf{t}\cdot\partial_{\beta}\mathbf{r}}{|\partial_{s}\mathbf{r}|}\partial_{s}\mathbf{t}\cdot\partial_{\alpha}\mathbf{r}\Big{]}, (32)

as

F=Sαβhαβ,F_{\parallel}=S^{\alpha\beta}h_{\alpha\beta}, (33)

recovering Eq. (6).

The orthogonal forces follow straightforwardly by projecting out the tangent component and the definition of the gauge covariant derivative, as Dα𝐅α=α𝐅α(𝐭α𝐅α)𝐭D_{\alpha}\mathbf{F}^{\alpha}=\nabla_{\alpha}\mathbf{F}^{\alpha}-(\mathbf{t}\cdot\nabla_{\alpha}\mathbf{F}^{\alpha})\mathbf{t}, from which we recover Eq. (8).

Appendix D The energetics of twisted-toroidal bundles with small curvatures

From Eqs. (6) and (8), we can solve for the stable configurations of bundles of helices with constant pitch 2π/Ω2\pi/\Omega (i.e., the twist axis is unbent). For a hexagonal columnar phase, the tensor of elastic constants in the material frame is

Cαβγδ=Y1+ν[ν(1ν)gtarαβgtarγδ+12(gtarαγgtarβδ+gtarαδgtarβγ)],C^{\alpha\beta\gamma\delta}=\frac{Y}{1+\nu}\Big{[}\frac{\nu}{(1-\nu)}g_{\text{tar}}^{\alpha\beta}g_{\text{tar}}^{\gamma\delta}+\tfrac{1}{2}\big{(}g_{\text{tar}}^{\alpha\gamma}g_{\text{tar}}^{\beta\delta}+g_{\text{tar}}^{\alpha\delta}g_{\text{tar}}^{\beta\gamma}\big{)}\Big{]}, (34)

where YY and ν\nu are the 2D Young’s modulus and Poisson ratio, respectively. The deformation field for a bundle of helices with uniform pitch is

𝐫(0)=𝐫0(s)+ρ(r)ρ^(ϕ),\mathbf{r}^{(0)}=\mathbf{r}_{0}(s)+\rho(r)\hat{\rho}(\phi), (35)

where here 𝐫0(s)\mathbf{r}_{0}(s) is a central reference curve which we take to be a straight line with s2𝐫0=0\partial_{s}^{2}\mathbf{r}_{0}=0, ρ^\hat{\rho} is the usual radial unit vector in cylindrical coordinates in the target space, ϕ=φ+Ωs\phi=\varphi+\Omega s, and rr, φ\varphi, and ss are the radial, polar, and cylyndrical coordinates in the material space. The tangent field then lies along s𝐫(0)=𝐭0+Ωρϕ^\partial_{s}\mathbf{r}^{(0)}={\bf t}_{0}+\Omega\rho\hat{\phi}, so 𝐭(0)=(𝐭0+Ωρϕ^)/1+Ω2ρ2\mathbf{t}^{(0)}=\big{(}{\bf t}_{0}+\Omega\rho\hat{\phi}\big{)}/\sqrt{1+\Omega^{2}\rho^{2}}. In the absence of external forces, Eqs. (6) and (8) now reduce to a nonlinear boundary value problem (BVP) for ρ(r)\rho(r):

0\displaystyle 0 =r{rρ(r)[Y22ν2(ρ(r)21)+νY22ν2(ρ(r)2r2(1+Ω2ρ(r)2)1)]}\displaystyle=-\partial_{r}\Big{\{}r\rho^{\prime}(r)\big{[}\frac{Y}{2-2\nu^{2}}(\rho^{\prime}(r)^{2}-1)+\frac{\nu Y}{2-2\nu^{2}}(\frac{\rho(r)^{2}}{r^{2}(1+\Omega^{2}\rho(r)^{2})}-1)\big{]}\Big{\}}
ρ(r)r(1+Ω2ρ(r)2)2[νY22ν2(ρ(r)21)+Y22ν2(ρ(r)2r2(1+Ω2ρ(r)2)1)]\displaystyle-\frac{\rho(r)}{r(1+\Omega^{2}\rho(r)^{2})^{2}}\big{[}\frac{\nu Y}{2-2\nu^{2}}(\rho^{\prime}(r)^{2}-1)+\frac{Y}{2-2\nu^{2}}(\frac{\rho(r)^{2}}{r^{2}(1+\Omega^{2}\rho(r)^{2})}-1)\big{]} (36)
0\displaystyle 0 =ρ(0)\displaystyle=\rho(0) (37)
0\displaystyle 0 =Y22ν2(ρ(R)21)+νY22ν2(ρ(R)2R2(1+Ω2ρ(r)2)1).\displaystyle=\frac{Y}{2-2\nu^{2}}(\rho^{\prime}(R)^{2}-1)+\frac{\nu Y}{2-2\nu^{2}}\Big{(}\frac{\rho(R)^{2}}{R^{2}(1+\Omega^{2}\rho(r)^{2})}-1\Big{)}. (38)

Solving this BVP numerically gives the equilibrium deformation field for the straight helical bundle, as shown in Fig. 1.

To find the low energy configurations of weakly curved twisted-toroidal bundles, we now introduce a perturbation at 𝒪(κ0)\mathcal{O}(\kappa_{0}) to Eq. (35) by taking s2𝐫0=κ0N^\partial_{s}^{2}\mathbf{r}_{0}=\kappa_{0}\hat{N} and 𝐫=𝐫(0)+𝐫(1)+𝒪(κ02)\mathbf{r}=\mathbf{r}^{(0)}+\mathbf{r}^{(1)}+\mathcal{O}(\kappa_{0}^{2}), with

𝐫(1)=δss𝐫(0)+δρρ^+δϕϕ^,\mathbf{r}^{(1)}=\delta s\partial_{s}\mathbf{r}^{(0)}+\delta\rho\hat{\rho}+\delta\phi\hat{\phi}, (39)

and expand the elastic energy in Eq. (4) to quadratic order in κ0\kappa_{0}. The linear correction to the elastic energy vanishes because the constant pitch helical bundles are in force balance, and so the resulting elastic energy takes the form

s=helical+12κ02correction[δs,δρ,δϕ].\mathcal{E}_{s}=\mathcal{E}_{\text{helical}}+\tfrac{1}{2}\kappa_{0}^{2}\mathcal{E}_{\text{correction}}[\delta s,\delta\rho,\delta\phi]. (40)

Scaling analysis of the force balance equations shows that, at small ρ\rho, the components of 𝐫(1)\mathbf{r}^{(1)} are

δs\displaystyle\delta s =asΩκ0ρ3sinϕ\displaystyle=a_{s}\Omega\kappa_{0}\rho^{3}\sin{\phi}
δρ\displaystyle\delta\rho =aρΩ2κ0ρ4cosϕ\displaystyle=a_{\rho}\Omega^{2}\kappa_{0}\rho^{4}\cos{\phi} (41)
δϕ\displaystyle\delta\phi =aϕΩ2κ0ρ3sinϕ,\displaystyle=a_{\phi}\Omega^{2}\kappa_{0}\rho^{3}\sin{\phi},

as in Eq. (9). To stop the bundle from unbending and effectively decreasing its curvature, we fix the average position along the normal vector, N^\hat{N}, at 𝐫0\mathbf{r}_{0}, so that

1A𝑑AN^𝐫(1)=0.\frac{1}{A}\int dA\hat{N}\cdot\mathbf{r}^{(1)}=0. (42)

This constrains as=aρaϕa_{s}=a_{\rho}-a_{\phi}, since

1A𝑑AN^𝐫(1)=(asaρ+aϕ)Ω2κ0R20R𝑑rrρ(r)4.\frac{1}{A}\int dA\hat{N}\cdot\mathbf{r}^{(1)}=-(a_{s}-a_{\rho}+a_{\phi})\frac{\Omega^{2}\kappa_{0}}{R^{2}}\int_{0}^{R}drr\rho(r)^{4}. (43)

Having found ρ(r)\rho(r) from Eq. (36), we can subtitute the ansatz from Eq. (9) into the elastic energy in Eq. (4) and integrate over the volume for a given Young’s modulus, YY, 2D Poisson’s ratio, ν\nu, and reciprocal pitch, ΩR\Omega R, to obtain an elastic energy

s=helical+12κ02correction(aρ,aϕ).\mathcal{E}_{s}=\mathcal{E}_{\text{helical}}+\tfrac{1}{2}\kappa_{0}^{2}\mathcal{E}_{\text{correction}}(a_{\rho},a_{\phi}). (44)

Eq (44) is quadratic in aρa_{\rho} and aϕa_{\phi} and has a minima at aρa_{\rho}, aϕ0a_{\phi}\neq 0 whenever ΩR\Omega R and κR\kappa R are nonzero. These energy minimizing displacements, and the resultant pressure in the cross-section, are shown in Fig. 1 for a given κ0R\kappa_{0}R. The resultant elastic energy per unit length takes the form of an effective bending modulus, as shown Fig. 3.

Twisted toroidal bundles are non-equidistant, and we can calculate the components of the convective flow tensor from the perturbative displacements in Eq. (41). Since the uniform pitch helices are equidistant, the leading contribution to the convective flow tensor is linear in κ0\kappa_{0}:

hρρ(1)\displaystyle h^{(1)}_{\rho\rho} =ρ(r)srδρ1+Ω2ρ2\displaystyle=\frac{\rho^{\prime}(r)\partial_{s}\partial_{r}\delta\rho}{\sqrt{1+\Omega^{2}\rho^{2}}}
hρϕ(1)\displaystyle h^{(1)}_{\rho\phi} =ρ2srδϕ+ρ(r)(1+Ω2ρ2)sφδρΩρ2ρ(r)s2δρ2(1+Ω2ρ2)3/2\displaystyle=\frac{\rho^{2}\partial_{s}\partial_{r}\delta\phi+\rho^{\prime}(r)(1+\Omega^{2}\rho^{2})\partial_{s}\partial_{\varphi}\delta\rho-\Omega\rho^{2}\rho^{\prime}(r)\partial_{s}^{2}\delta\rho}{2(1+\Omega^{2}\rho^{2})^{3/2}} (45)
hϕϕ(1)\displaystyle h^{(1)}_{\phi\phi} =Ω3κ0ρ5+ρsδρ+ρ2(1+Ω2ρ2)sφδϕΩρ3s2δϕ(1+Ω2ρ2)5/2.\displaystyle=\frac{\Omega^{3}\kappa_{0}\rho^{5}+\rho\partial_{s}\delta\rho+\rho^{2}(1+\Omega^{2}\rho^{2})\partial_{s}\partial_{\varphi}\delta\phi-\Omega\rho^{3}\partial_{s}^{2}\delta\phi}{(1+\Omega^{2}\rho^{2})^{5/2}}.

By substituting Eq (41) into the above expression for hαβh_{\alpha\beta}, and using the lowest energy values of aρa_{\rho} and aϕa_{\phi}, we can calculate the elasticity-mediated response of twisted-toroidal bundles to the geometric constraints on constant spacing. The contributions to non-equidistance can be broken up into the splay, tr(h)\text{tr}(h), and biaxial splay, (hαβ12tr(h)δαβ)=hdev\big{(}h_{\alpha\beta}-\tfrac{1}{2}\text{tr}(h)\delta_{\alpha\beta}\big{)}=h_{\text{dev}}, of the tangent field, 𝐭\mathbf{t}. To measure the contributions to these two modes from the linear displacements driven by elastic interactions of twisted-toroidal bundles, we compare their integrated, dedimensionalized contributions to the Frank free energy,

Γ\displaystyle\Gamma =1κ02V𝑑Vtr(hdev2)\displaystyle=\frac{1}{\kappa_{0}^{2}V}\int dV\text{tr}(h_{\text{dev}}^{2}) (46)
Ψ\displaystyle\Psi =1κ02V𝑑Vtr(h)2,\displaystyle=\frac{1}{\kappa_{0}^{2}V}\int dV\text{tr}(h)^{2}, (47)

where dV=dsdv1dv2det(gtar)dV=dsdv^{1}dv^{2}\sqrt{\text{det}(g^{\text{tar}})}, as shown in the inset of Fig. 3.