Helical close-packing of anisotropic tubes

Benjamin R. Greenvall & Gregory M. Grason Department of Polymer Science and Engineering, University of Massachusetts, Amherst, MA 01003 [email protected]
(May 3, 2024)
Abstract

Helically close-packed states of filaments are common in natural and engineered material systems, ranging from nanoscopic biomolecules to macroscopic structural components. While the simplest models of helical close-packing, described by the ideal rope model, neglect anisotropy perpendicular to the backbone, physical filaments are often quite far from circular in their cross-section. Here, we consider an anisotropic generalization of the ideal rope model and show that cross-section anisotropy has a strongly non-linear impact on the helical close-packing configurations of helical filaments. We show that the topology and composition of the close-packing landscape depends on the cross-sectional aspect ratio and is characterized by several distinct states of self-contact. We characterize the local density of these distinct states based on the notion of confinement within a ‘virtual’ cylindrical capillary, and show that states of optimal density vary strongly with the degree of anisotropy. While isotropic filaments are densest in a straight configuration, any measure of anisotropy leads to helicity of the maximal density state. We show the maximally dense states exhibit a sequence of transitions in helical geometry and cross-sectional tilt with increasing anisotropy, from spiral tape to spiral screw packings. Furthermore, we show that maximal capillary density saturates in a lower bound for volume fraction of π/4𝜋4\pi/4italic_π / 4 in the large-anisotropy, spiral-screw limit. While cross-sectional anisotropy is well-known to impact the mechanical properties of filaments, our study shows its strong effects to shape the configuration space and packing efficiency of this elementary material motif.

Keywords: filaments, helices, geometry of materials, packing

1 Introduction

Filamentous or fibrous structures are an essential building block of material systems, ranging from the molecular to human scale. At the microscopic end, natural biomolecules [1, 2, 3] and synthetic polymers [4, 5, 6] form the building blocks of life and of many engineered consumer products, respectively. At macroscopic scale, fibers, wires and ropes form structural materials [7, 8, 9] with diverse applications ranging from textiles [10, 11, 12] to tension bearing components in marine mooring lines [13, 14] and in architecture (e.g. cables in suspension bridges) [15, 16].

A filamentous material’s function, utility, or morphology is inherently intertwined with its packing, or space filling arrangement. Entanglements [17], knots [18], knits [19] and inter-filament orientation [20] have been shown to influence the mechanical properties of filament assemblies. Precisely folded proteins inherit both function and stability from their ordered arrangement [21]. Braiding string or tying our shoes requires specific spatio-temporal control of the strand. In each of the above, the packing of the material is essential.

Refer to caption
Figure 1: Dense helical coils across length scales. (A.) α𝛼\alphaitalic_α-helix protein shown using the “ribbon” representation; backbone and sidechain atoms rendered in yellow and gray, respectivley (PDB structure from  [22] rendered using VMD [23]). (B.) Self-assembled ferrocene-dipeptide “screw” nanostructure (scale bar = 1μm1𝜇𝑚1\mu m1 italic_μ italic_m). Adapted with permission from  [24]. Copyright 2015 American Chemical Society. (C.) Nested mesoscale polymer ribbon with tunable photocrease-mediated configuration (scale bar = 50μm50𝜇𝑚50\mu m50 italic_μ italic_m). Adapted from [25]. (D.) Twisted helical nanofiber yarn (scale bar = 500μm500𝜇𝑚500\mu m500 italic_μ italic_m). Adapted from [26]. (E.) Twisted Graphene Oxide/Alginate hydrogel fiber (scale bar = 500μm500𝜇𝑚500\mu m500 italic_μ italic_m). Adapted with permission from  [27]. Copyright 2020 American Chemical Society. (F.) ’Blue Bouquet’ Passion Flower Passiflora caerulea tendril self-coiling, photographed at University of Massachusetts Durfee Conservatory. (G.) Coiled telephone cable. (H.) Oriental Bittersweet (Celastrus orbiculatus) vine, photographed in Northampton, Massachusetts.

Indeed, the packing of material components has been a long-standing and broad interest in mathematics, science, and technology, dating at least as early as Kepler’s 17th century conjecture on sphere packing [28, 29], though likely beginning many centuries earlier, especially in the case of mosaics and tilings [30]. Historically studies on packing have largely focused on arrangements of discrete, rigid components (e.g. tiles, spheres, polyhedra). Recently, more attention has been given to the packing of continua, i.e. soft, deformable and extended objects [31], such as flexible sheets [32, 33, 34], membranes [35, 36, 37] and filaments [38, 39, 40, 41, 42]. However, despite the natural and technological relevance of filamentous materials, their packing has received considerably less attention, in part due to the complex and vast configuration space accessible to even a single deformable high aspect ratio structure. Due to their flexibility, the packing may be constrained by contact at any location in space and at either local or non-local positions along the filament length [43].

Key progress in understanding filament packing geometry is based on what is known as the Ideal Rope or Ideal Tube model of packing, which considers space filling configurations based on a filament centerline and a rigid, circular cross-section of uniform diameter d𝑑ditalic_d swept-out normal to the backbone curve. Early applications of this model have been applied to understand the geometry of closed curves in optimally “tight” knots and links [44, 45, 46, 47, 48, 49, 50, 51, 52, 53] This model has also been applied to model the complex close contact geometry in multi-filament clasps [54, 55], as well as plies and bundles  [56, 57, 58, 59, 60, 61], which have formed a basis for comparison to experimental systems [62, 63, 64, 65, 66, 67].

A particular focus of the ideal tube model has been on helical close-packing of single filaments, initially motivated as a generic, physical model of condensed helical motifs in biomacromolecules, e.g. α𝛼\alphaitalic_α-helices, nucleosomes and chromatids [68, 69, 70, 71, 72, 73, 74, 75, 76, 77]. In a seminal work, Przybyl and Pieranski determined the configuration space of helical packing in ideal tubes, solving for the conditions for self-contact that delineate allowed from self-overlap** states in terms of helical radius and pitch of the tubular centerline [78]. Notably, they showed that contact is defined by turn-to-turn stacking of consecutive pitches at large radius, while for small radius contact constraints are local, defined by a minimal curvature radius of the finite-thickness tube. These two contact conditions intersect in configuration space at a particular point, which they consider to be a state of ideal close-packing for helical tubes. This packing geometry is used to rationalize the stability of similar helical states in models of entropic or osmotically collapsed helical molecules [75, 79, 80].

While the ideal tube model is a natural starting point for consideration of filamentary close-packing, it is important to recognize that most physical realizations of helical packing present some measure of cross-sectional anisotropy, deviating by some measure from locally circular shapes. Fig. 1 shows several examples of anisotropic tightly-packed helical structures, across a range of size scales and materials systems. Some form of cross-sectional asymmetry is present in flexible, densely-coiled filamentous structures at the molecular and meso-engineered systems [81, 25], and in a variety of organism structures, including (but certainly not limited to) bacteria [82, 83], curly human hair [84], plant tendrils and seed pods [85, 86, 87, 88], climbing and swimming snakes [89, 90], elephant trunks [91], and squid and octopus tentacles[92, 93]. Naturally, such filament assemblies lead to the basic question, how does cross-section geometry influence the states of dense packing in helical filaments?

In this article, we study an extension of the (circular) ideal tube model to include filaments with elliptical cross-section, and consider the states of helical packing. In addition to the radius R𝑅Ritalic_R and pitch P𝑃Pitalic_P of the helical backbone, anisotropic tubes are described by two additional parameters: the elliptical aspect ratio ϵitalic-ϵ\epsilonitalic_ϵ; and relative orientation, or tilt angle α𝛼\alphaitalic_α, of the cross-section relative to local bending direction, as illustrated in Fig. 2. Like the circular ideal tube model, we model the packing configurations by assuming the filament to be perfectly flexible (i.e. it does not require any energy to bend the filament) but inextensible and unshearable (that is, the cross-section of the tube is everywhere elliptical of specified dimensions). Based on these assumptions, we fully determine the states of self-contact, which for a given anisotropy ϵitalic-ϵ\epsilonitalic_ϵ correspond to a two-dimensional manifold in helical configuration space defined by R𝑅Ritalic_R, P𝑃Pitalic_P and α𝛼\alphaitalic_α. We show that for any measure of anisotropy (ϵ1)italic-ϵ1(\epsilon\neq 1)( italic_ϵ ≠ 1 ) close contact configurations include a new class of local curvature singularities, corresponding to “outward folding” cusps on the surface of low-radius packings. Moreover, we find that the nature of the self-contact singularity that separates allowed from overlap** configurations is highly sensitive to the degree of anisotropy as well as the orientation of the cross-sectional axis.

Refer to caption
Figure 2: Anisotrpic helical tubes with helical radius R=0.6d0𝑅0.6subscript𝑑0R=0.6d_{0}italic_R = 0.6 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and helical pitch P=1.5d0𝑃1.5subscript𝑑0P=1.5d_{0}italic_P = 1.5 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (A.) Tube surface with helical pitch P𝑃Pitalic_P and helical radius R𝑅Ritalic_R; together, these parameters define the helical angle, θ=tan1(2πRP)𝜃superscript12𝜋𝑅𝑃\theta=\tan^{-1}{\big{(}}\frac{2\pi R}{P}{\big{)}}italic_θ = roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG 2 italic_π italic_R end_ARG start_ARG italic_P end_ARG ). Everywhere along the helical backbone, the cross-sectional elements are ellipses. (B.) Tilt angle (α𝛼\alphaitalic_α) of the cross-section, where α𝛼\alphaitalic_α measures the rotation of the material short axis away from 𝐫^^𝐫\hat{{\bf r}}over^ start_ARG bold_r end_ARG, see B. (C.) cross-section anisotropy. The (shorter) semi-minor elliptical axis has length a𝑎aitalic_a while the (longer) semi-major axis has length b𝑏bitalic_b. ϵitalic-ϵ\epsilonitalic_ϵ defines the aspect ratio, ϵ=a/bitalic-ϵ𝑎𝑏\epsilon=a/bitalic_ϵ = italic_a / italic_b.

To understand conditions for “optimal” helical close-packing, we compute a measure of local-density that we call capillary density, ϕitalic-ϕ\phiitalic_ϕ, which is simply the occupied volume fraction within a cylindrical volume enclosing the helical tube (of unlimited length) [94]. Previous studies of ideal tubes have employed a variety of metrics to define “maximally-dense” helical packing, including minimizing the local radius of gyration  [78, 69] and maximizing excluded volume in a bath of spherical particles [75]. Our use of capillary density is motivated by the fact that helical packing under cylindrical-like confinement is also prominent in a variety of natural and engineered systems, ranging from DNA coiled within viral phage capsids [95, 96], biopolymers during synthesis [97, 98], translocation  [99, 100], and incorporated within membranes  [101], as well as in variety of confined synthetic systems including colloids [102, 103, 104, 105, 106, 107, 108], block copolymers  [109, 110, 111], and liquid crystals  [112, 113, 114, 115]. Moreover, as we show, the states of optimal capillary density for isotropic versus anisotropic helical tubes are both qualitatively and quantitatively distinct, and capillary close-packing exhibits a strongly non-linear sensitivity to the degree of cross-sectional anisotropy.

While the densest configuration of an isotropic tube under capillary confinement is trivially a straight rod, we show that once cross-sectional isotropy is broken, the densest configurations within any size capillary are helical, with a geometry that depends on the level of anisotropy. When the tube is moderately anisotropic, the densest configuration occurs under high confinement (i.e. in a narrow capillary) while increased anisotropy prefers a larger capillary. The densest structures within these two regimes exhibit strikingly different packing motifs both in terms of the helical angle and cross-section orientation; light anisotropy prefers a “tape-like” motif reminiscent of Fig. 1E while high anistropy prefers a “screw-like” motif reminiscent of Fig. 1B. We determine that these packing regimes crossover when the cross-sectional aspect ratio is approximately 4:1:414:14 : 1. While the densest structure (for a given cross-section) never exhibits scrolled or nested packing (i.e. intermediate tilt, like Fig. 1C), we find such structures do optimize density at intermediate levels of confinement, and over a range of confinement that increases with asymmetry. This analysis reveals that the densest structures for any degree of anisotropy are bounded, exhibiting a packing fraction bounded between 1ϕπ/41italic-ϕ𝜋41\geq\phi\geq\pi/41 ≥ italic_ϕ ≥ italic_π / 4.

The remainder of this article is as follows. In Section 2, we describe the geometric model used to represent our helical tube, the method for finding the boundary between expanded and self-contacting helical states and the capillary packing fraction. In Section 3, we describe the self-contacting manifold and the packing motifs that constrain different regimes of the configuration space. In Section 4, we assess the capillary packing density of contacting structures, and the dependence on maximally dense structures on cross-section anisotropy. We discuss the potential implications of these key results in Section 5 for physical and mechanical properties of filamentary systems.

2 Parameterization and Modes of Contact

2.1 3D Helical Geometry

We model helical configurations of anisotropic tubes using a centerline (or backbone) described by a circular helix (i.e. one with constant radius, R𝑅Ritalic_R and pitch, P𝑃Pitalic_P):

𝐱(s)=R𝐫^(s)±Psz^𝐱𝑠plus-or-minus𝑅^𝐫𝑠𝑃𝑠^𝑧{\bf x}(s)=R\hat{\bf{r}}(s)\pm\frac{P}{\ell}s\hat{z}bold_x ( italic_s ) = italic_R over^ start_ARG bold_r end_ARG ( italic_s ) ± divide start_ARG italic_P end_ARG start_ARG roman_ℓ end_ARG italic_s over^ start_ARG italic_z end_ARG (1)

where 𝐫^(s)=cos(2πs/)x^+sin(2πs/)y^^𝐫𝑠2𝜋𝑠^𝑥2𝜋𝑠^𝑦\hat{\bf{r}}(s)=\cos(2\pi s/\ell)\hat{x}+\sin(2\pi s/\ell)\hat{y}over^ start_ARG bold_r end_ARG ( italic_s ) = roman_cos ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_x end_ARG + roman_sin ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_y end_ARG, s𝑠sitalic_s is an arclength coordinate along the backbone and =4π2R2+P24superscript𝜋2superscript𝑅2superscript𝑃2\ell=\sqrt{4\pi^{2}R^{2}+P^{2}}roman_ℓ = square-root start_ARG 4 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG is the arclength per helical turn 111Note that in this parameterization, the chirality is controlled by the sign of P𝑃Pitalic_P, where P>0𝑃0P>0italic_P > 0 corresponds to right-handed coils while P<0𝑃0P<0italic_P < 0 refers to left-handed coils. The contact geometry and packing density are, of course, invariant under changes in chirality. In the current manuscript, all schematics shown are left-handed; therefore, listed values of P𝑃Pitalic_P throughout the text can be take to be negative.. Here, we consider a filament backbone of unlimited length, or in effect, neglect contributions to packing from the ends of the structure, appropriate when the total length is much larger than the tube diameter. The geometry of the helix is independent of backbone position up to screw motions, so the overall configuration can be described by a single ratio, the helical angle, θ=tan1(2πRP)𝜃superscript12𝜋𝑅𝑃\theta=\tan^{-1}(\frac{2\pi R}{P})italic_θ = roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG 2 italic_π italic_R end_ARG start_ARG italic_P end_ARG ) bounded between θ=0𝜃0\theta=0italic_θ = 0 (a straight line) and θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2 (a closed circle).

To describe the material body anisotropically distributed around this backbone (Fig. 2A), we employ an orthonormal material frame, {𝐓(s),𝐞^1(s),𝐞^2(s)𝐓𝑠subscript^𝐞1𝑠subscript^𝐞2𝑠{\bf T}(s),\hat{{\bf e}}_{1}(s),\hat{{\bf e}}_{2}(s)bold_T ( italic_s ) , over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) , over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s )} where 𝐓(s)𝐓𝑠{\bf T}(s)bold_T ( italic_s ) is the backbone tangent vector, 𝐓(s)=𝐱(s)=sinθϕ^(s)+cosθz^𝐓𝑠superscript𝐱𝑠𝜃^italic-ϕ𝑠𝜃^𝑧{\bf T}(s)={\bf x}^{\prime}(s)=\sin\theta\hat{\bf{\phi}}(s)+\cos\theta\hat{z}bold_T ( italic_s ) = bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_s ) = roman_sin italic_θ over^ start_ARG italic_ϕ end_ARG ( italic_s ) + roman_cos italic_θ over^ start_ARG italic_z end_ARG and ϕ^(s)^italic-ϕ𝑠\hat{\bf{\phi}}(s)over^ start_ARG italic_ϕ end_ARG ( italic_s ) is the azimuthal unit vector. We constrain the (undeformable) cross-section to lie within the 𝐞^1(s)subscript^𝐞1𝑠\hat{{\bf e}}_{1}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ), 𝐞^2(s)subscript^𝐞2𝑠\hat{{\bf e}}_{2}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ) plane, defined in terms of the Frenet-Serret frame (where {𝐓(s),𝐍(s),𝐁(s)𝐓𝑠𝐍𝑠𝐁𝑠{\bf T}(s),{\bf N}(s),{\bf B}(s)bold_T ( italic_s ) , bold_N ( italic_s ) , bold_B ( italic_s )} form an orthonormal triad, see A and Ref [116])

𝐞^1(s)=cosα𝐍(s)+sinα𝐁(s),subscript^𝐞1𝑠𝛼𝐍𝑠𝛼𝐁𝑠\displaystyle\hat{{\bf e}}_{1}(s)=\cos\alpha{\bf N}(s)+\sin\alpha{\bf B}(s),over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) = roman_cos italic_α bold_N ( italic_s ) + roman_sin italic_α bold_B ( italic_s ) , (2a)
𝐞^2(s)=cosα𝐁(s)sinα𝐍(s)subscript^𝐞2𝑠𝛼𝐁𝑠𝛼𝐍𝑠\displaystyle\hat{{\bf e}}_{2}(s)=\cos\alpha{\bf B}(s)-\sin\alpha{\bf N}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ) = roman_cos italic_α bold_B ( italic_s ) - roman_sin italic_α bold_N ( italic_s ) (2b)

Thus, the backbone tangent is normal to the cross-section at all points along the curve. The parameter α𝛼\alphaitalic_α describes the “tilt” (angle) of the material frame, 𝐞^1(s)subscript^𝐞1𝑠\hat{{\bf e}}_{1}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) and 𝐞^2(s)subscript^𝐞2𝑠\hat{{\bf e}}_{2}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ) relative to the normal (𝐍(s{\bf N}(sbold_N ( italic_s)) and binormal (𝐁(s)𝐁𝑠{\bf B}(s)bold_B ( italic_s )) components of the Frenet-Serret frame. For helices, the normal points toward the pitch axis (i.e. 𝐍(s)=𝐫^𝐍𝑠^𝐫{\bf N}(s)=-\hat{\bf{r}}bold_N ( italic_s ) = - over^ start_ARG bold_r end_ARG), and hence, α𝛼\alphaitalic_α is easy to visualize as the local orientation of the material frame in the 𝐫^^𝐫\hat{\bf{r}}over^ start_ARG bold_r end_ARG, z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG plane, as shown in Fig. 2B.

We model the tubular cross-section as an ellipse with semi-minor axis of length a𝑎aitalic_a oriented in the direction of 𝐞^1(s)subscript^𝐞1𝑠\hat{{\bf e}}_{1}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) and semi-major axis of length b𝑏bitalic_b oriented in the direction of 𝐞^2(s)subscript^𝐞2𝑠\hat{{\bf e}}_{2}(s)over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ), see Fig. 2C. The 3D filament surface is then described by

𝐗(s,ψ)=𝐱(s)+acosψ𝐞^1(s)+bsinψ𝐞^2(s)𝐗𝑠𝜓𝐱𝑠𝑎𝜓subscript^𝐞1𝑠𝑏𝜓subscript^𝐞2𝑠{\bf X}(s,\psi)={\bf x}(s)+a\cos\psi\hat{{\bf e}}_{1}(s)+b\sin\psi\hat{{\bf e}% }_{2}(s)bold_X ( italic_s , italic_ψ ) = bold_x ( italic_s ) + italic_a roman_cos italic_ψ over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( italic_s ) + italic_b roman_sin italic_ψ over^ start_ARG bold_e end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_s ) (2c)

where ψ𝜓\psiitalic_ψ parameterizes the boundary of the elliptical cross-section at a backbone position s𝑠sitalic_s. The material body around the helical backbone is then defined by the aspect ratio of the cross-section,

ϵa/bitalic-ϵ𝑎𝑏\epsilon\equiv a/bitalic_ϵ ≡ italic_a / italic_b (2d)

and the tilt or rotation of the material frame, α𝛼\alphaitalic_α (Fig. 2B, C). Intuitively, when α=0𝛼0\alpha=0italic_α = 0, the short material axis is oriented in the direction of curvature (i.e. towards the core of the coil), while α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 corresponds to the long axis oriented in the curvature direction. Importantly, here we extend the “Ideal Tube” model, and therefore treat the filaments as inextensible (fixed length), unshearable (rigid cross-section), and perfectly flexible. In the following, we consider filaments with fixed cross-section area, Afilsubscript𝐴filA_{\rm fil}italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT. We define the effective diameter of filaments as,

d0=4Afilπ=2absubscript𝑑04subscript𝐴fil𝜋2𝑎𝑏d_{0}=\sqrt{\frac{4A_{\rm fil}}{\pi}}=2\sqrt{ab}italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = square-root start_ARG divide start_ARG 4 italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT end_ARG start_ARG italic_π end_ARG end_ARG = 2 square-root start_ARG italic_a italic_b end_ARG (2e)

that is independent of ϵitalic-ϵ\epsilonitalic_ϵ, and we consider all other length scales relative to this microscopic dimension.

2.2 Sectional geometry

Refer to caption
Figure 3: 3D and 2D helical configurations for ϵ=0.64italic-ϵ0.64\epsilon=0.64italic_ϵ = 0.64 all with R/d0=0.5𝑅subscript𝑑00.5R/d_{0}=0.5italic_R / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.5. (A.) Non-contacting structure with P>Pmin𝑃subscript𝑃minP>P_{\rm min}italic_P > italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT and α=5π/16𝛼5𝜋16\alpha=5\pi/16italic_α = 5 italic_π / 16. (B.) Locally contacting structures obtained from (A.) by increasing tilt (α=3π/8𝛼3𝜋8\alpha=3\pi/8italic_α = 3 italic_π / 8) and decreasing P𝑃Pitalic_P until contact, see also Fig. 13 (C.) Non-locally contacting structures obtained from (A.) by decreasing tilt (α=π/4𝛼𝜋4\alpha=\pi/4italic_α = italic_π / 4) and decreasing P𝑃Pitalic_P until contact.

Based on the parameterization of anisotropic tubes defined in eqn. 2c, we then determine helical configurations of self-contact by solving for two types of singularities in the tubular surface. The first correspond to over-curvature resulting in cusps or creases in the tubular surface, which we denote as local contact. The second correspond to overlap of the tube at two distinct surface points, which we denote as non-local contact

For the purposes of illustrating and analyzing these singular structures it is particular useful to consider the geometry of 2D sections perpendicular to the pitch axis z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG, as shown in Fig. 3. We refer to these perpendicular cuts as “croiss-sections” due to their crescent-like shape (see Fig. 3A). These are defined by solutions to fixed-height equation 𝐗(s,ψ)z^=const.𝐗𝑠𝜓^𝑧const{\bf X}(s,\psi)\cdot\hat{z}={\rm const.}bold_X ( italic_s , italic_ψ ) ⋅ over^ start_ARG italic_z end_ARG = roman_const ., which requires a parameteric relation s0(ψ)subscript𝑠0𝜓s_{0}(\psi)italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ψ ) between surface coordinates, from which define the 2D curves

𝐗(ψ)=𝐗(s0(ψ),ψ).subscript𝐗perpendicular-to𝜓𝐗subscript𝑠0𝜓𝜓{\bf X}_{\perp}(\psi)={\bf X}\big{(}s_{0}(\psi),\psi\big{)}.bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) = bold_X ( italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ψ ) , italic_ψ ) . (2f)

For a given helical tube, all croiss-sections are identical up to rigid rotations by screw symmetry. Fig. 3A shows an example for a non-contacting helical configuration. Fig. 3B shows an example of a local contact singularity for higher tilt and minimal pitch, which appears as an “outward” cusp in 𝐗(ψ)subscript𝐗perpendicular-to𝜓{\bf X}_{\perp}(\psi)bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) for this example. Fig. 3C shows an example of a non-local contact singularity for lower tilt and minimal pitch, which appears as two distinct locations on the curve 𝐗(ψ)subscript𝐗perpendicular-to𝜓{\bf X}_{\perp}(\psi)bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) meeting at single point in space.

2.3 Local Contact

Local contact occurs at a single ψ𝜓\psiitalic_ψ along the tube surface where the surface fails to be smooth; this discontinuity manifests as either an inward or outward crease in the 𝐗(s,ψ)𝐗𝑠𝜓{\bf X}(s,\psi)bold_X ( italic_s , italic_ψ ) surface (or a cusp of the same ”direction” in the 𝐗(ψ)subscript𝐗perpendicular-to𝜓{\bf X}_{\perp}(\psi)bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) slice, see Fig. 3B, Fig. 13). Operationally, these singularities can be selected by locations where the differential surface element of the tube vanishes. Using the surface metric, gg\rm groman_g, the magnitude of the surface normal can be expressed as

det(g)=|s𝐗(s,ψ)×ψ𝐗(s,ψ)|detgsubscript𝑠𝐗𝑠𝜓subscript𝜓𝐗𝑠𝜓\sqrt{\rm det(g)}=|\partial_{s}{\bf X}(s,\psi)\times\partial_{\psi}{\bf X}(s,% \psi)|square-root start_ARG roman_det ( roman_g ) end_ARG = | ∂ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT bold_X ( italic_s , italic_ψ ) × ∂ start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT bold_X ( italic_s , italic_ψ ) | (2g)

which yields the surface area of a differential tube element as dA=det(g)dsdψ𝑑𝐴detg𝑑𝑠𝑑𝜓dA=\sqrt{\rm det(g)}\,ds\,d\psiitalic_d italic_A = square-root start_ARG roman_det ( roman_g ) end_ARG italic_d italic_s italic_d italic_ψ.

The conditions for local contact are defined by the solutions to det(g)=0detg0\rm det(g)=0roman_det ( roman_g ) = 0 (see eqns. 2ab – 2ad in B). Notably, the Ideal Tube case (ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1) corresponds to a singular limit of these equations in which local-contact takes the form on an inward crease, at a minimum pitch derived by Przybyl and Pieranski [78]

Plocal=2πd02RR2,forϵ=1formulae-sequencesubscript𝑃local2𝜋subscript𝑑02𝑅superscript𝑅2foritalic-ϵ1P_{\rm local}=2\pi\sqrt{\frac{d_{0}}{2}R-{R}^{2}},\ \ \ \ {\rm for}\ \epsilon=1italic_P start_POSTSUBSCRIPT roman_local end_POSTSUBSCRIPT = 2 italic_π square-root start_ARG divide start_ARG italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 2 end_ARG italic_R - italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , roman_for italic_ϵ = 1 (2h)

More generally for ϵ1italic-ϵ1\epsilon\neq 1italic_ϵ ≠ 1 there are two solutions, corresponding to local singularities at distinct surface locations, given by

Plocal()=2πacosαRR2,forϵ1formulae-sequencesubscript𝑃local2𝜋𝑎𝛼𝑅superscript𝑅2foritalic-ϵ1\displaystyle P_{\rm local(-)}=2\pi\sqrt{a\cos\alpha R-{R}^{2}},\ \ {\rm for}% \ \ \epsilon\neq 1italic_P start_POSTSUBSCRIPT roman_local ( - ) end_POSTSUBSCRIPT = 2 italic_π square-root start_ARG italic_a roman_cos italic_α italic_R - italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , roman_for italic_ϵ ≠ 1 (2i)
Plocal(+)=2πbsinαRR2,forϵ1formulae-sequencesubscript𝑃local2𝜋𝑏𝛼𝑅superscript𝑅2foritalic-ϵ1\displaystyle P_{\rm local(+)}=2\pi\sqrt{b\sin\alpha R-{R}^{2}},\ \ {\rm for}% \ \ \epsilon\neq 1italic_P start_POSTSUBSCRIPT roman_local ( + ) end_POSTSUBSCRIPT = 2 italic_π square-root start_ARG italic_b roman_sin italic_α italic_R - italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG , roman_for italic_ϵ ≠ 1 (2j)

which also varies with orientation of the tubular cross-section .

While these local-contact solutions share similarities with the isotropic (Ideal Tube) case, there are notable differences in which elliptical axis sets the curvature constraint. The first solution (eqn. 2i) is constrained by the cross-section semi-minor axis (a𝑎aitalic_a) and is maximized when this length is oriented in the radial (𝐫^^𝐫\hat{{\bf r}}over^ start_ARG bold_r end_ARG) direction (i.e. α=0𝛼0\alpha=0italic_α = 0). On the other head, the second solution (eqn. 2j) is constrained by the cross-section semi-major axis (b𝑏bitalic_b) and is maximized when the cross-section is rotated by α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2, such that this length is oriented in the 𝐫^^𝐫\hat{{\bf r}}over^ start_ARG bold_r end_ARG direction. These two regimes have distinct geometrical structure and consequences of constraining close-packed geometries, discussed in Section 3.

2.4 Non-local Contact

The tube surface makes contact non-locally when two distinct elements of the croiss-section occupy the same point in space (Fig. 3C), which like the local-contact case, implies contact along a 1D helical curve where two parts of the same tube meet along its length.

In searching for non-local contact, we compute the distance of closest approach between these distinct points along the tube surface. When the cross-section is isotropic, it is sufficient to calculate the distance between points on the backbone and require that they be separated by a distance equal to the tube diameter d0subscript𝑑0d_{0}italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [78, 48]. Because of the asymmetry in the cross-section, we must instead compute distances between surface elements. Our approach employs several coupled criteria which determine points of distance of closest approach along the tube surface (e.g. that the separation is normal to the tube surface at both points ) then vary the configuration parameters (i.e. R𝑅Ritalic_R, P𝑃Pitalic_P and α𝛼\alphaitalic_α) until this closest distance is zero (i.e. the points are in contact, see B). In practice, we fixed combinations of R𝑅Ritalic_R and α𝛼\alphaitalic_α, and vary P𝑃Pitalic_P, numerically solving for Pnonlocal(R,α)subscript𝑃nonlocal𝑅𝛼P_{\rm non-local}(R,\alpha)italic_P start_POSTSUBSCRIPT roman_non - roman_local end_POSTSUBSCRIPT ( italic_R , italic_α ).

While it is straightforward to see that for sufficiently large P𝑃Pitalic_P any combination of R𝑅Ritalic_R and α𝛼\alphaitalic_α will be embeddable (i.e. non-overlap**), in general, it is not a priori known what type of contact first limits the reduction of pitch at a given R𝑅Ritalic_R and α𝛼\alphaitalic_α (i.e. local or non-local contact). We determine the pitch at which the tube first makes contact at a given given R𝑅Ritalic_R and α𝛼\alphaitalic_α, i.e. minimum non-overlap** pitch, as the largest pitch at which the tube either makes local or non-local contact, Pmin(R,α)=Max[Plocal(±)(R,α),Pnonlocal(R,α)]subscript𝑃min𝑅𝛼Maxsubscript𝑃localplus-or-minus𝑅𝛼subscript𝑃nonlocal𝑅𝛼P_{\rm min}(R,\alpha)={\rm Max}\big{[}P_{\rm local(\pm)}(R,\alpha),P_{\rm non-% local}(R,\alpha)\big{]}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R , italic_α ) = roman_Max [ italic_P start_POSTSUBSCRIPT roman_local ( ± ) end_POSTSUBSCRIPT ( italic_R , italic_α ) , italic_P start_POSTSUBSCRIPT roman_non - roman_local end_POSTSUBSCRIPT ( italic_R , italic_α ) ].

2.5 Capillary Confinement

For illustrative purposes, consider a expanded, loosely coiled filament. If this filament is compressed until self-contact is made, then a confining capillary is constricted around assembly until the capillary is in contact with the filament surface. In a sense, the capillary contact constrains the structure from dilating radially, while the filament self-contact prevents further radial compression; this conceptual paradigm is presented in Fig 4A.

Operationally, we numerically determine the smallest cylinder (with radius RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT) which could confine the structure, shown in Fig 4B,C. Because each planar section of the packing is identical up to a rigid rotation about the pitch axis, the occupied volume fraction in the capillary is identical to the occupied area fraction of the croiss-section within the circular boundary of the cylinder. Because the filament cross-sections are unshearable, and by Pappus’s theorem the enclosed volume per unit backbone length is constant [117], it follows that AcroissP=Afilsubscript𝐴croiss𝑃subscript𝐴filA_{{\rm croiss}}P=A_{{\rm fil}}\ellitalic_A start_POSTSUBSCRIPT roman_croiss end_POSTSUBSCRIPT italic_P = italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT roman_ℓ, where Acroisssubscript𝐴croissA_{{\rm croiss}}italic_A start_POSTSUBSCRIPT roman_croiss end_POSTSUBSCRIPT is the area enclosed by the cross-section normal to the pitch axis and Afilsubscript𝐴filA_{{\rm fil}}italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT is the area of (elliptical) cross-section normal to the backbone. Using the fact that Acroiss/Afil=secθsubscript𝐴croisssubscript𝐴fil𝜃A_{{\rm croiss}}/A_{{\rm fil}}=\sec\thetaitalic_A start_POSTSUBSCRIPT roman_croiss end_POSTSUBSCRIPT / italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT = roman_sec italic_θ we compute the capillary packing fraction as

ϕ=AcroissπRC2=d02secθ4RC2italic-ϕsubscript𝐴croiss𝜋superscriptsubscript𝑅𝐶2superscriptsubscript𝑑02𝜃4superscriptsubscript𝑅𝐶2\phi=\frac{A_{{\rm croiss}}}{\pi{R_{C}}^{2}}=\frac{d_{0}^{2}\sec\theta}{4R_{C}% ^{2}}italic_ϕ = divide start_ARG italic_A start_POSTSUBSCRIPT roman_croiss end_POSTSUBSCRIPT end_ARG start_ARG italic_π italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = divide start_ARG italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sec italic_θ end_ARG start_ARG 4 italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (2k)

where RC=maxψ|𝐗(ψ)|subscript𝑅𝐶subscriptmax𝜓subscript𝐗perpendicular-to𝜓R_{C}={\rm max}_{\psi}|{\bf X}_{\perp}(\psi)|italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = roman_max start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) |. Rationale for the use of the minimal size capillary is discussed in C.

Refer to caption
Figure 4: Dense capillary packing. (A.) Hypothetical filament constrained by capillary- and self-contact. Schematic of uniform filament confined in a (B.) 3D and (C.) 2D cylindrical pore of radius RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT

3 Close-packed configuration landscapes

In this section, we analyze the manifold self-contacting helical geometries, specifically, analyzing the minimal non-overlap** pitch Pmin(R,α)subscript𝑃min𝑅𝛼P_{\rm min}(R,\alpha)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R , italic_α )for for a given helical backbone radius and cross-sectional tilt α𝛼\alphaitalic_α for a sequence of cross-sectional anisotropy ϵitalic-ϵ\epsilonitalic_ϵ.

3.1 Isotropic tubes

When the cross-section is isotropic (e.g. circular, a=b𝑎𝑏a=bitalic_a = italic_b), the packing landscape is determined only by helical radius (e.g. it is invariant to tilt). Przybyl and Pieranski observed that contact is delineated into two regimes [78]; when R𝑅Ritalic_R is small, the structure is curvature-limited and makes self-contact locally, or along an inward crease at the center of helix (analogous to overbending a garden hose) as illustrated in Fig. 5A i.-iii. Packing at large R𝑅Ritalic_R is limited by (non-local) contact between successive turns of the tube (analogous to coil stacking of a garden hose), illustrated in Fig. 5A v.-vi. Here, the single boundary between local and non-local contact occurs at R/d00.431similar-to-or-equals𝑅subscript𝑑00.431R/d_{0}\simeq 0.431italic_R / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 0.431 and P/d01.083similar-to-or-equals𝑃subscript𝑑01.083P/d_{0}\simeq 1.083italic_P / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 1.083 (Fig. 5A iv.); a single local maximum in Pmin/d0=π/2subscript𝑃minsubscript𝑑0𝜋2P_{\rm min}/d_{0}=\pi/2italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_π / 2 occurs at R/d0=0.25𝑅subscript𝑑00.25R/d_{0}=0.25italic_R / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.25. Together, these contact criteria generate the minimum pitch bounding hull, reproduced in the style of [78] in Fig. 5B, where structures below Pmin(R)subscript𝑃min𝑅P_{\rm min}(R)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R ) are forbidden (due to self-overlap) and structures above Pmin(R)subscript𝑃min𝑅P_{\rm min}(R)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R ) are said to be “expanded”.

Refer to caption
Figure 5: Packing landscape for isotropic tube, ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1, originally calculated in ref. [78]. (A.) Maximal pitch structures for increasing values of R/d0𝑅subscript𝑑0R/d_{0}italic_R / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (B.) Minimal pitch contact hull indicating the limiting mode of contact.

3.2 Slightly anisotropic tubes

Refer to caption
Figure 6: Packing landscape for ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95. (A.) Contact type phase diagram and (B.) minimal pitch contact hull. (C.) Maximal pitch structures for increasing values of α𝛼\alphaitalic_α.

With apparently any perturbation to the cross-section shape (or at least with as small as 0.5%percent0.50.5\%0.5 % deviation, ϵ=0.995italic-ϵ0.995\epsilon=0.995italic_ϵ = 0.995), we find that the modes of contact change significantly; notably, now with a vital dependence on cross-section tilt. These large changes for weak anisotropy illustrate that the isotropic tube case (ϵ1italic-ϵ1\epsilon\to 1italic_ϵ → 1) is in fact a singular limit of the more general variable anisotropy behavior. As shown in Fig. 6B, the large R𝑅Ritalic_R behavior is largely unchanged (albeit with a slight tilt-bias which intuitively grants large α𝛼\alphaitalic_α structures the ability to pack at lower pitch). However, when R𝑅Ritalic_R is small in the regime corresponding to local contact of the isotropic filament, we observe three distinct contact regimes, shown in Fig. 6A. Two of these regimes are determined by local contact corresponding to solutions of vanishing area element of the metric tensor(eqn. 2i and 2j), however these represent distinct types of contact, in particular driven by the long material axis (b𝑏bitalic_b, at high α𝛼\alphaitalic_α) and the short material axis (a𝑎aitalic_a, at low α𝛼\alphaitalic_α). Surprisingly, the low R𝑅Ritalic_R - intermediate α𝛼\alphaitalic_α behavior exhibits non-local contact; this seemingly stems from variation in tube-surface curvature (resulting from cross-section anisotropy) thereby smoothing what was otherwise an inward local kink. This packing paradigm allows for a slight softening of the previous global maximum in Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT; in other words, non-local contact at intermediate tilt allows for a lower pitch configuration. While these changes are challenging to recognize visually at the 3D scale (structures along the Pmin(R,α)subscript𝑃min𝑅𝛼P_{\rm min}(R,\alpha)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R , italic_α ) ridge are shown in Fig. 6C), their modes of contact are distinct when magnified; we note that the low-α𝛼\alphaitalic_α local contact creases “inward” in the structure, manifesting as an intruding groove (and hence labeled “Local (-)”) while the high-α𝛼\alphaitalic_α local contact creases “outward”, manifesting as a protruding seam (and hence labeled “Local (+)”). It is also important to note that despite the minuscule differences in contact geometry and fairly modest shifts in the values of Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT at a given R𝑅Ritalic_R, the boundaries between distinct contact topologies exhibit a large shift in then R𝑅Ritalic_R-α𝛼\alphaitalic_α plane, even when the cross-section anisotropy is small. These shifts in the contact manifold become even more pronounced as the tubular cross-section becomes more anisotropic, as we discuss next.

3.3 Highly anisotropic tubes

When the cross-section is symmetric (with respect to tilt), the packing landscapes are also symmetric (see Fig. 5B). Introducing asymmetry in the cross-section results in asymmetry in the landscape; the Local (+) regime occupies slightly more area than the Local (-) regime in Fig. 6A and the corresponding maxima in Fig. 6B is more pronounced.

Refer to caption
Figure 7: Packing landscape for ϵ=0.8175italic-ϵ0.8175\epsilon=0.8175italic_ϵ = 0.8175. (A.) Contact type phase diagram and (B.) minimal pitch contact hull. (C.) Maximal pitch structures for increasing values of α𝛼\alphaitalic_α.

As the filament becomes more anisotropic (see Fig. 7A,B for ϵ=0.8175italic-ϵ0.8175\epsilon=0.8175italic_ϵ = 0.8175), the asymmetry between + and -- local contact grows. With decreasing ϵitalic-ϵ\epsilonitalic_ϵ, the Local (+) and Non-local regions become more prominent while the Local (-) shrinks, and the differences in Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT between large- and small-α𝛼\alphaitalic_α become significant. The minimal pitch also exhibits a stronger tilt-dependence; while two maxima separated by a saddle still dominated the low-R𝑅Ritalic_R behavior, the magnitude of the maxima (corresponding to structures Fig. 7C.i and C.vii now differ by approximately 20 percent (with Pmin1.42d0similar-to-or-equalssubscript𝑃min1.42subscript𝑑0P_{\rm min}\simeq 1.42d_{0}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≃ 1.42 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and 1.73d01.73subscript𝑑01.73d_{0}1.73 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, respectively). At this scale, the 3D structures exhibit some perceptible asymmetry; structures with intermediate tilt (Fig. 7C.ii-vi) display uneven protrusions in the radial direction, while structures constrained by Local (+) curvature (Fig. 7C.v-vii) display a noticeable seam.

When the cross-section is increasingly asymmetric (below ϵ4/5similar-to-or-equalsitalic-ϵ45\epsilon\simeq 4/5italic_ϵ ≃ 4 / 5), the Local (-) branch ceases to constrain the packing manifold, at which point, the entire low-tilt regime is limited by non-local turn-to-turn contact exclusively. When the cross-section has at least this degree of anisotropy, it is not possible to construct a viable configuration that is more curved than the short material length scale would allow because when the structure is in the low-tilt regime, the curvature of the filament is primarily in the direction of the short material axis. There simply is not enough material in this direction to curve such that a kink arises, while in the pitch-direction, the cross-section is wide which obstructs larger regions of turn-to-turn configurations.

An example of the “two-phase” contact topology for highly-anisotropic tubes (ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25) is shown in Fig. 8. As the filament becomes more anisotriopic, the global maxima of Pmin(R,α)subscript𝑃min𝑅𝛼P_{\rm min}(R,\alpha)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R , italic_α ) at α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 become more pronounced (Fig. 8B). The secondary maxima at low-R𝑅Ritalic_R for α=0𝛼0\alpha=0italic_α = 0 persists, despite the change between modes of contact (from Local (-) to Non-local), though the feature continues to softens relative to the global maximum (Fig. 8C.i and C.v have Pmin2.25d0similar-to-or-equalssubscript𝑃min2.25subscript𝑑0P_{\rm min}\simeq 2.25d_{0}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ≃ 2.25 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT and 3.15d0similar-to-or-equalsabsent3.15subscript𝑑0\simeq 3.15d_{0}≃ 3.15 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT at α=0𝛼0\alpha=0italic_α = 0 and α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2, respectively). In this asymmetry regime, these extrema remain separated by a saddle which spans intermediate tilt; however, the span is not smooth (i.e. Pmin/αsubscript𝑃min𝛼\partial P_{\rm min}/\partial\alpha∂ italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT / ∂ italic_α is discontinuous), owing to the boundary between modes of contact. In these significantly asymmetric cases, the corresponding structures exhibit exotic morphologies; a selection that spans the Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT saddle are shown in Fig. 8C. In particular, the global maxima of Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT Fig. 8C.v., corresponding to Local (+) contact at small-R𝑅Ritalic_R, α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 presents a relatively large Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT characterized by elaborate outward cusps at the high curvature edge of the cross-section. This behavior is in stark contrast to the Non-local large-R𝑅Ritalic_R, α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 packing, in which Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT is minimal (i.e. Pmin(R,α=π/2)2asubscript𝑃minformulae-sequence𝑅𝛼𝜋22𝑎P_{\rm min}(R\to\infty,\alpha=\pi/2)\rightarrow 2aitalic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R → ∞ , italic_α = italic_π / 2 ) → 2 italic_a, or “small-axis stacking”). Intuitively, as the radius of curvature is compressed beyond the width of cross-section in radial direction, the packing must expand correspondingly into the z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG dimension (resulting in larger Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT).

Refer to caption
Figure 8: Packing landscape for ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25. (A.) Contact type phase diagram and (B.) minimal pitch contact hull. (C.) Maximal pitch structures for increasing values of α𝛼\alphaitalic_α.

Przybyl and Pieranki considered the transition point between local and non-local contact branches to correspond to a state of optimally dense helical packing. In the next section, we consider the density of these much richer close-packed landscapes, and in particular, the behavior in proximity to packing motif coexistence branches. Additionally, a more complete analysis of contacts along the coexistence branch are provided in C.

4 Capillary packing

Refer to caption
Figure 9: Multiplicity of states within a given capillary of RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT shown for ϵ=0.25.italic-ϵ0.25\epsilon=0.25.italic_ϵ = 0.25 .. Traces of (A.) density and (B.) helical angle for fixed tilt angle as a function of confining radius. (C.) Dense and loose structures within the same capillary RC/d0=0.79subscript𝑅𝐶subscript𝑑00.79R_{C}/d_{0}=0.79italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = 0.79 (i.), 0.840.840.840.84 (ii, iii), 0.890.890.890.89 (iv, v), 0.940.940.940.94 (vi, vii), 0.990.990.990.99 (viii, ix). (D.) RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT evaluated along the contact hull; the region at low-R𝑅Ritalic_R low-α𝛼\alphaitalic_α corresponds to bifurcated regions, (E.) Landscape of maximum with variable RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and α𝛼\alphaitalic_α. A plot of the maximal density at fixed RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is overlaid.

Having found that helical geometry of filament self-contact is strongly sensitive to the anisotropy of the cross-section, we now consider its effects on capillary packing fraction, ϕitalic-ϕ\phiitalic_ϕ.

In this section, we first briefly summarize the approach to computing optimally dense configurations. We then describe features of the packing density manifold and the underlying structural motifs as a function of the size of the confining capillary, RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. Finally, we analyze and compare the overall densest configurations as a function of tube aspect ratio, ϵitalic-ϵ\epsilonitalic_ϵ and classify the distinct packing motifs selected for variable filament asymmetry and confinement.

4.1 Measuring local density by capillary confinement

In the prior sections, it was most convenient to parameterize packing in terms of the purely geometric parameters that describe the configuration, in particular, the radius R𝑅Ritalic_R and pitch P𝑃Pitalic_P of the tubular “centerline”. This description is sufficient for capillary confinement of isotropic tubes (ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1) because the tightest capillary radius for a given helical radius is simply RC=R+d0/2subscript𝑅𝐶𝑅subscript𝑑02R_{C}=R+d_{0}/2italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = italic_R + italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2. For anisotropic tubes (ϵ1italic-ϵ1\epsilon\neq 1italic_ϵ ≠ 1), the map** between helical centerline radius R𝑅Ritalic_R and capillary radius, is a more complex and non-linear function RC(R,P,α)subscript𝑅𝐶𝑅𝑃𝛼R_{C}(R,P,\alpha)italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_R , italic_P , italic_α ). In practice, this is computed from the maximal radial distance of the tubular surface from the pitch axis, i.e. RC=maxψ|𝐗(ψ)|subscript𝑅𝐶subscriptmax𝜓subscript𝐗perpendicular-to𝜓R_{C}={\rm max}_{\psi}|{\bf X}_{\perp}(\psi)|italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = roman_max start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) | for a given R,P𝑅𝑃R,Pitalic_R , italic_P and α𝛼\alphaitalic_α. That is, the capillary radius is the smallest possible to enclose the helical filament, or other words, the minimal possible value of RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is chosen so as to maximize the packing fraction for that R,P𝑅𝑃R,Pitalic_R , italic_P and α𝛼\alphaitalic_α.

For all close-packed (i.e. self-contacting) structures for a given ϵitalic-ϵ\epsilonitalic_ϵ, that is, all combinations of R𝑅Ritalic_R, α𝛼\alphaitalic_α, and the resulting minimal pitch Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, we calculate ϕ(R,α)italic-ϕ𝑅𝛼\phi(R,\alpha)italic_ϕ ( italic_R , italic_α ) and RC(R,α)subscript𝑅𝐶𝑅𝛼R_{C}(R,\alpha)italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_R , italic_α )222Note that by constraining our optimization to close-packed structures, ϕ(R,α)italic-ϕ𝑅𝛼\phi(R,\alpha)italic_ϕ ( italic_R , italic_α ) and RC(R,α)subscript𝑅𝐶𝑅𝛼R_{C}(R,\alpha)italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_R , italic_α ) have an implicit dependence on Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT, where Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT = Pmin(R,α)subscript𝑃min𝑅𝛼P_{\rm min}(R,\alpha)italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT ( italic_R , italic_α ) . By inverting the map** RC(R,α)subscript𝑅𝐶𝑅𝛼R_{C}(R,\alpha)italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ( italic_R , italic_α ) (to yield R(RC,α)𝑅subscript𝑅𝐶𝛼R(R_{C},\alpha)italic_R ( italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT , italic_α )), we analyze the relationship between confinement size for a close-packed structure and its packing fraction. As an example, we show several traces of ϕ(RC,α)italic-ϕsubscript𝑅𝐶𝛼\phi(R_{C},\alpha)italic_ϕ ( italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT , italic_α ) for fixed α𝛼\alphaitalic_α and ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25 in Fig. 9A-B. Importantly, when both α𝛼\alphaitalic_α and RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT are small, ϕ(RC)italic-ϕsubscript𝑅𝐶\phi(R_{C})italic_ϕ ( italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) is multivalued since there are multiple R𝑅Ritalic_R values with the same α𝛼\alphaitalic_α that perfectly fit within a specific capillary (see schematic examples in Fig. 9C.).

As we are presently interested in the densest packing, we take only the higher ϕCsubscriptitalic-ϕ𝐶\phi_{C}italic_ϕ start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT configuration in this multi-valued region in our analysis, neglecting the features of the “loose” (larger pitch) branch of capillary confinement.

For ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25 the map between helical radius and capillary radius is shown in Fig.  9D, where the “loose” structures occupy a region at low-R𝑅Ritalic_R low-α𝛼\alphaitalic_α. The boundary of this region (white dashed line in Fig.  9) corresponds to the minimal capillary radius for a given α𝛼\alphaitalic_α in this regime. For all confined packings along this boundary between single- and multi-valued helical radius, we observe helical geometry (θ>0𝜃0\theta>0italic_θ > 0, marked with black dots in Fig. 9B), which is not the case for larger tilt and lower asymmetry, where the tightest capillary radius is achieved for R0𝑅0R\to 0italic_R → 0. For example, at values of tilt larger than the bifurcation region for ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25, α0.6greater-than-or-equivalent-to𝛼0.6\alpha\gtrsim 0.6italic_α ≳ 0.6, the minimal capillary size is independent of α𝛼\alphaitalic_α; all structures exhibit a minimal RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT governed by θ=0𝜃0\theta=0italic_θ = 0 “rod-like” packing, in which RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is exactly the semi-major cross-sectional axis, b𝑏bitalic_b. Regardless of aspect ratio (and tilt), these rod-like packings exhibit a density ϕ=Afil/πRC2=πab/πb2=ϵitalic-ϕsubscript𝐴fil𝜋superscriptsubscript𝑅𝐶2𝜋𝑎𝑏𝜋superscript𝑏2italic-ϵ\phi=A_{{\rm fil}}/\pi{R_{C}}^{2}=\pi ab/\pi b^{2}=\epsilonitalic_ϕ = italic_A start_POSTSUBSCRIPT roman_fil end_POSTSUBSCRIPT / italic_π italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_π italic_a italic_b / italic_π italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = italic_ϵ.

Taken together, these effects lead to complex landscapes of capillary packing density as a function of cross-section tilt and capillary radius, as shown for example for ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25 in Fig. 9E. Here, ϕitalic-ϕ\phiitalic_ϕ spans 0.8ϕ0.1greater-than-or-equivalent-to0.8italic-ϕgreater-than-or-equivalent-to0.10.8\gtrsim\phi\gtrsim 0.10.8 ≳ italic_ϕ ≳ 0.1. Within the parameter range shown, ϕitalic-ϕ\phiitalic_ϕ exhibits a minimum when RC/d01.5similar-to-or-equalssubscript𝑅𝐶subscript𝑑01.5R_{C}/d_{0}\simeq 1.5italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≃ 1.5, α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2; structures in this neighborhood correspond to high-Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT Local (+) structures, like Fig. 8C.v. Of course, these configurations only correspond to a local minima; ϕ0italic-ϕ0\phi\rightarrow 0italic_ϕ → 0 as RC/d0subscript𝑅𝐶subscript𝑑0R_{C}/d_{0}\rightarrow\inftyitalic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT → ∞, resulting from the center of the coil opening.

In the following, we next consider α𝛼\alphaitalic_α as a configurational degree of freedom (like R𝑅Ritalic_R and P𝑃Pitalic_P) and determine the optimal tilt and helical geometry for dense packing of a tube with a given anisotropy ϵitalic-ϵ\epsilonitalic_ϵ and in a given capillary radius RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. For the example of ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25 shown in Fig. 9E, the optimal α𝛼\alphaitalic_α value for each given RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is shown as blue curve, which transitions from α=0𝛼0\alpha=0italic_α = 0 for small RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT to α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 for large RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, a generic feature for all ϵitalic-ϵ\epsilonitalic_ϵ as we discuss below.

4.2 Variable-Radius Capillary Packing

Fig. 10A shows plots of the optimal capillary packing density and configurations as function of capillary radius (RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT) and for a sequence of increasing cross-sectional asymmetry. For all aspect ratios of the filament, packing fraction ϕitalic-ϕ\phiitalic_ϕ exhibits a non-monotonic dependence on the confining radius. When the tube is isotropic (ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1), the packing exhibits four extrema (Fig. 10A,top):

  1. 1.

    when the capillary is the same size as the tube, the tube perfectly packs within the capillary (ϕ=1italic-ϕ1\phi=1italic_ϕ = 1). We consider this to be a “rod-like” packing motif.

  2. 2.

    when the capillary size is intermediate to the tube radius and tube diameter, the structure is helical with lower packing fraction, corresponding to high-Pminsubscript𝑃minP_{\rm min}italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT local contact structures.

  3. 3.

    when the capillary size is comparable to the tube diameter, the packing fraction reaches a local maxima of ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4; based on the density, we consider this to be a “toroid-like” packing motif 333Toroid-like packing exhibits a packing density ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4, equivalent to that of a horn-torus within a cylinder of radius and height equal to the tube diameter; interestingly, the helical capillary exhibits the same ratio of size (a cylinder of radius equal to the tube diameter).

  4. 4.

    when the RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is larger than the tube diameter, the core of the helical structure begins to open, leading to a monotonic decrease in density with increasing capillary size (ϕ0italic-ϕ0\phi\rightarrow 0italic_ϕ → 0 with RCsubscript𝑅𝐶R_{C}\rightarrow\inftyitalic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT → ∞).

The helical angle (θ=tan1(2πRP))𝜃superscript12𝜋𝑅𝑃(\theta=\tan^{-1}(\frac{2\pi R}{P}))( italic_θ = roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG 2 italic_π italic_R end_ARG start_ARG italic_P end_ARG ) ) exhibits a monotonic dependence on RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT. When the filament is isotropic and the cylinder is in the small limit (RC=d0/2subscript𝑅𝐶subscript𝑑02R_{C}=d_{0}/2italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT = italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / 2), the tube is straight (i.e, R=0𝑅0R=0italic_R = 0) thus θ=0𝜃0\theta=0italic_θ = 0. In the large limit (RCsubscript𝑅𝐶R_{C}\rightarrow\inftyitalic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT → ∞), the tube approaches a toriodal geometry (i.e, a helix with P0𝑃0P\rightarrow 0italic_P → 0) thus θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2.

Refer to caption
Figure 10: Densest configurations within variable sized capillaries. (A.) Maximal density (ϕitalic-ϕ\phiitalic_ϕ, black, left axis), tilt angle (α𝛼\alphaitalic_α, red, right axis), helical angle (θ𝜃\thetaitalic_θ, blue, right axis) within capillary RC/d0subscript𝑅𝐶subscript𝑑0R_{C}/d_{0}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for selected values of ϵitalic-ϵ\epsilonitalic_ϵ. Global maxima of ϕitalic-ϕ\phiitalic_ϕ marked with solid dots; local maxima marked with open dots. (B.) Structures for ϵ=0.316italic-ϵ0.316\epsilon=0.316italic_ϵ = 0.316 within increasing RC/d0subscript𝑅𝐶subscript𝑑0R_{C}/d_{0}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, displaying increasing α𝛼\alphaitalic_α and θ𝜃\thetaitalic_θ.

Helical packing of asymmetric tubes (ϵ1italic-ϵ1\epsilon\neq 1italic_ϵ ≠ 1) leads to new features of the capillary dependent packing, shown in Fig. 10A. With any degree of anisotropy, we observe that the minimal capillary radius no longer corresponds to a local maxima in density; this maxima shifts to larger capillaries and decreases in density with increasing anisotropy, first very slightly for low anistropy (e.g. ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95) and then more obviously with larger anisotropy (ϵ=0.49)italic-ϵ0.49(\epsilon=0.49)( italic_ϵ = 0.49 ) The minimum helical angle (which occurs within the smallest possible capillary radius) increases with anisotropy, never limiting to the isotropic value of θ=0𝜃0\theta=0italic_θ = 0; in other words, the straight, non-helical tube is never a density maximizer (for an aniostropic tube), regardless of size of the capillary. For any degree of cross-sectional anisotropy, capillary density favor helical over straight configurations.

Relative to small-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT, the large-capillary behavior is qualitatively insensitive to anisotropy; the density exhibits a local maxima when the capillary radius is equal to twice the semi-major tube axis. Naturally, this value of RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT grows as with increasingly tubes anisotropy.

Given a tube anisotropy, optimal capillary density requires strongly non-linear variation in the tilt angle (α𝛼\alphaitalic_α) of the cross-section. Near maxima of ϕ(RC)italic-ϕsubscript𝑅𝐶\phi(R_{C})italic_ϕ ( italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ), we observe extremes in the tilt angle. When RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is small, the cross-section is oriented with the short axis pointing (inward) in the curvature direction (α=0𝛼0\alpha=0italic_α = 0). This general motif, observed at the low RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT density maxima, is reminiscent of a coiled tape, on for example, bicycle handlebars, and shown for ϵ=0.316italic-ϵ0.316\epsilon=0.316italic_ϵ = 0.316 in Fig. 10B.i. When RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT is larger, the dense structures exhibit a cross-section with the long axis oriented (inward) in the curvature direction (α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2; this motif is reminiscent of the threads on a screw, see Fig. 10B.v; the onset of this regime universally exhibits the toroidal density, ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4.

For large anisotropies (ϵ4/5less-than-or-similar-toitalic-ϵ45\epsilon\lesssim 4/5italic_ϵ ≲ 4 / 5, these states are connected at intermediate RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT via a monotonic increase in the tilt angle, visualized in Fig. 10B.ii-iv. The width of the transition broadens with anisotropy, revealing an increasing number of dense, intermediate-tilt states that exhibit a nested or “scroll”-like packing. Interestingly, these states with large but sub-maximal tilt exhibit a packing density nearly degenerate to the ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4 toroidal packing. When the tube is moderately anistropic, we also note the existence of a set of low-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT nested states; these structures correspond to a slight preference for non-local contact over Local (-) contact, and persists until the Local (-) regime no longer constrains the close-packing landscape (ϵ0.81similar-to-or-equalsitalic-ϵ0.81\epsilon\simeq 0.81italic_ϵ ≃ 0.81). Given the slight amount of anisotropy, these configurations are structurally similar to the untilted structures; the packing benefit and resulting structural changes are slight in this case.

We mark the locations of low-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT (tape-like) and high-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT (screw-like) density maxima with black markers in Fig. 10A. For each value of ϵitalic-ϵ\epsilonitalic_ϵ, the global maximum is marked with a solid dot, while the secondary maximum is marked with an open dot. When ϵitalic-ϵ\epsilonitalic_ϵ is high (i.e. nearly isotropic), the low-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT maxima is most dense, and when ϵitalic-ϵ\epsilonitalic_ϵ is small (i.e. large anisotropy), the high-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT maxima is most dense. This transition in the geometry of maximally dense packing is examined in more detail in the next section.

4.3 Optimal Capillary Packing

Refer to caption
Figure 11: Competing dense packing motifs. (A.) Densest configurations for selected values of ϵitalic-ϵ\epsilonitalic_ϵ. (B. top) Density (ϕitalic-ϕ\phiitalic_ϕ) global (“dense”) and local (“loose”) maxima. Global maxima are shown as solid dots while submaximal packings are shown as open dots. (B. bottom) Helical angle (θ𝜃\thetaitalic_θ) for densest configurations

As shown in Fig. 10, ϕ(RC)italic-ϕsubscript𝑅𝐶\phi(R_{C})italic_ϕ ( italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ) is characterized by two local maxima, one at low-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT with small θ𝜃\thetaitalic_θ and a second at large-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and large θ𝜃\thetaitalic_θ, which are respectively closer to nearly-straight versus “gardenhose-like” coiling. In Fig. 11, we track each of these two localy dense states (i.e the ϕitalic-ϕ\phiitalic_ϕ maxima in Fig. 10). Again, global maxima (“dense”) are marked with a solid point while the secondary maxima (“loose”) are marked with an open point. When the tube is either isotropic or only moderately elliptical, the density of the small-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT (α=0𝛼0\alpha=0italic_α = 0, low θ𝜃\thetaitalic_θ) state is higher than that of the large-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT (α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2, high θ𝜃\thetaitalic_θ) state. However, the small-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT density decreases with anisotropy, while the large-RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT density remains constant; at a critical aspect ratio, ϵcr0.316similar-to-or-equalssubscriptitalic-ϵcr0.316\epsilon_{\rm cr}\simeq 0.316italic_ϵ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ≃ 0.316, the densest state transitions between solution branches. This dependence is intuitive to understand by simply considering the capillary fraction of a straight tube, ϕ(θ=0)=ϵitalic-ϕ𝜃0italic-ϵ\phi(\theta=0)=\epsilonitalic_ϕ ( italic_θ = 0 ) = italic_ϵ, which obviously decreases due to the inefficiently of filling a circular capillary with an elliptic cross-section. Hence, secondary “gardenhose” maxima eventually overtakes the nearly-straight state in density. This transition between packing motifs shows a characteristic tilt- and helical angle dependence; the densest high-ϵitalic-ϵ\epsilonitalic_ϵ structures pack with α=0𝛼0\alpha=0italic_α = 0 and low θ𝜃\thetaitalic_θ (or a steep incline) while the low-ϵitalic-ϵ\epsilonitalic_ϵ structures with α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 orientation and high θ𝜃\thetaitalic_θ (or a gradual incline); the globally optimal configurations for assorted values of ϵitalic-ϵ\epsilonitalic_ϵ are shown in Fig. 11A. These competing motifs are shown at approximately the transition point, ϵcr0.316similar-to-or-equalssubscriptitalic-ϵcr0.316\epsilon_{\rm cr}\simeq 0.316italic_ϵ start_POSTSUBSCRIPT roman_cr end_POSTSUBSCRIPT ≃ 0.316, in Fig. 10B.i and B.v. Curiously, this transition occurs when the aspect ratio is π1similar-to-or-equalsabsentsuperscript𝜋1\simeq\pi^{-1}≃ italic_π start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT, at which point the density of both the tape-like and screw-like structure exhibit ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4.

Refer to caption
Figure 12: Anisotropy-confinement landscapes. (A.) Density of packed filaments with aspect ratio ϵitalic-ϵ\epsilonitalic_ϵ constrained to capillary with radius RC/d0subscript𝑅𝐶subscript𝑑0R_{C}/d_{0}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT / italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (B.) Configurational phase diagram consists of spiral tapes (α=0𝛼0\alpha=0italic_α = 0), spiral scrolls ((π/2>α>0𝜋2𝛼0\pi/2>\alpha>0italic_π / 2 > italic_α > 0)), and spiral screws (α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2).

5 Discussion and concluding remarks

In this article, we have inventoried the dense-packing parameter space of helical tubes with cross-sections of various aspect ratio (or flatness). We find that once cross-section symmetry is broken, the tube configuration space is constrained by novel and non-trivial variants of the local (curvature-limited) and non-local (turn-to-turn) contact for isotropic filaments reported in [78]. This complex spectrum of closed-packing geometries, and their non-linear dependence on anisotropy has important implications for the “local density” of helical close-packing, measured in terms of capillary packing density. While the densest structure is straight for isotropic tubes, once cross-section symmetry is broken, the densest configuration is either tape-like (α=0𝛼0\alpha=0italic_α = 0) when the tube is mildly anisotropic or screw-like (α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2) if the tube is sufficiently anisotropic. Importantly, the density of the weakly-anisotropic regime decreases with ϵitalic-ϵ\epsilonitalic_ϵ while the latter is independent of ϵitalic-ϵ\epsilonitalic_ϵ (and displays a packing fraction of ϕ=π/4italic-ϕ𝜋4\phi=\pi/4italic_ϕ = italic_π / 4, equivalent to that of a horn torus confined to a tight cylinder). Therefore, the densest configuration for any filament is bounded 1ϕπ/41italic-ϕ𝜋41\geq\phi\geq\pi/41 ≥ italic_ϕ ≥ italic_π / 4, assuming an optimal capillary. Regardless of the morphology, the “locally dense” packing motif required an increasingly larger capillary as the tube became more anisotropic. These trends of optimal density are summarized in Fig. 12A. Based on our highest aspect ratio studied here (ϵ=0.09italic-ϵ0.09\epsilon=0.09italic_ϵ = 0.09), it seems reasonable that the low ϵitalic-ϵ\epsilonitalic_ϵ preference for α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 packing should hold in the limit of thin ribbon- or sheet-like helices (i.e. ϵ0italic-ϵ0\epsilon\rightarrow 0italic_ϵ → 0); perhaps, not surprisingly, these packings seem to be abundant in biological membranes, including in the endoplasmic reticulum [35] and plant (photosynthetic) thylakoid membranes [36].

The strong influence of cross-sectional anisotropy on optimal close-packing, raises additional question about how (density and morphology) evolves under sub-optimal confinement. Though the optimal structures always required extremal (tape-like with α=0𝛼0\alpha=0italic_α = 0 or screw-like with α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2) tilt, intermediate levels of confinement often preferred intermediately tilted, nested scroll-like configurations, as shown in the morphological phase diagram in Fig. 12B (see B for a discussion of the small tape/scroll region at low RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and high ϵitalic-ϵ\epsilonitalic_ϵ). Many of these nested configurations (especially at large α𝛼\alphaitalic_α) exhibit a density nearly degenerate to the screw-like structure (see the soft gradients in Fig. 12A at low ϵitalic-ϵ\epsilonitalic_ϵ, intermediate RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT). Further, the nested states occupy a larger portion of confinement-parameter space as the cross-section becomes more asymmetric. Given a driving force for densification (like some applied compressive force), these findings hint at the potential to control the assembly morphology via the degree of confinement.

In considering the anisotropic filament packing geometry, we account for more complex material shape, however we have thus far restricted ourselves to very simple mechanical descriptions, namely, neglecting considerations of bend and twist elasticity of filaments. A more complete mechanical model should take into elastic costs of helical deformations, and specifically account for elastic anisotropies that result from the geometric anisotropies [31, 118]. Based on simple Kirchoff rod theory, one should expect (for a filament composed of an isotropic elastic medium) bending is stiffer in the wide axis relative to the narrow axis by a factor of ϵ2.superscriptitalic-ϵ2\epsilon^{-2}.italic_ϵ start_POSTSUPERSCRIPT - 2 end_POSTSUPERSCRIPT . On these grounds, and assuming that filaments are straight and untwised in their rest configuration, it is reasonable to expect that the α=π/2𝛼𝜋2\alpha=\pi/2italic_α = italic_π / 2 configurations would impose a higher elastic cost (assuming a straight rest configuration) than α=0𝛼0\alpha=0italic_α = 0 packings. This suggests a basic antagonism between geometries that favor optimal dense packing and those that require low elastic energy. Further, we can imagine a class of situations for which the rest state of filaments is helical [119, 66, 120, 121, 27], but the geometry of close-packing may indeed frustrate the minimal elastic energy. One example of this scenario is likely to be elastically programmed microfilaments that target inaccessible values of helical pitch and radius [25]. Notably, accounting for both density and elasticity in this situation is likely to direct the structures towards the tilted, nested structures shown in Fig. 1C.

Of course, the current question of dense packings underlies a range of other interesting phenomena, including the mechanical response of helical structures under an applied load, the impact of incompatible coiling (e.g. geometric frustration arising from an inaccessible elastic ground state) on the resulting structure, multi-component (and size-disperse) interacting systems including bundles [38, 122] and arrays [123]. In particular, it remains to be explored how the map** between packing of filaments in helically-twisted structures and packing on non-Euclidean surfaces [56, 38] sheds useful light on optimal states of packing anisotropic filaments. For example, for the present model of helical filament configurations with elliptical cross-sectional asymmetry, it can be expected that local and global geometric constraints can be understood by packing of geodesic ellipses on positive curvature surfaces [124].

The authors are grateful to K. T. Sullivan for helpful comments on this manuscript. Numerical calculations were carried out on the Unity Cluster, a collaborative, multi-institutional high-performance computing cluster managed by the UMass Amherst Research Computing and Data team. This research was supported by a National Science Foundation Graduate Research Fellowship awarded to B.R.G.

Conflict of Interest

The authors declare no conflict of interest.

Data Availability

The analysis code that was used for this study is openly available in UMass Amherst ScholarWorks at https://doi.org/10.7275/gr0k-fv13.

Appendix A Surface and Sectional Geometry

A.1 Frenet-Serret Formulas

To describe the material body anisotropically inflated around the helical backbone (Fig. 2A), we employ the orthonormal (geometric) Frenet-Serret frame, composed of the tangent, normal, and binormal triad, {𝐓,𝐍,𝐁𝐓𝐍𝐁{\bf T},{\bf N},{\bf B}bold_T , bold_N , bold_B}:

𝐓(s)=𝐱(s)=sinθϕ^(s)+cosθz^(s)𝐓𝑠superscript𝐱𝑠𝜃^italic-ϕ𝑠𝜃^𝑧𝑠{\bf T}(s)={\bf x}^{\prime}(s)=\sin\theta\hat{\bf{\phi}}(s)+\cos\theta\hat{z}(s)bold_T ( italic_s ) = bold_x start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_s ) = roman_sin italic_θ over^ start_ARG italic_ϕ end_ARG ( italic_s ) + roman_cos italic_θ over^ start_ARG italic_z end_ARG ( italic_s ) (2l)
𝐍(s)=𝐓(s)=𝐫^(s)𝐍𝑠superscript𝐓𝑠^𝐫𝑠{\bf N}(s)={\bf T}^{\prime}(s)=-\hat{\bf{r}}(s)bold_N ( italic_s ) = bold_T start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_s ) = - over^ start_ARG bold_r end_ARG ( italic_s ) (2m)
𝐁(s)=𝐓(s)×𝐍(s)=cosθϕ^(s)+sinθz^(s)𝐁𝑠𝐓𝑠𝐍𝑠𝜃^italic-ϕ𝑠𝜃^𝑧𝑠{\bf B}(s)={\bf T}(s)\times{\bf N}(s)=-\cos\theta\hat{\bf{\phi}}(s)+\sin\theta% \hat{z}(s)bold_B ( italic_s ) = bold_T ( italic_s ) × bold_N ( italic_s ) = - roman_cos italic_θ over^ start_ARG italic_ϕ end_ARG ( italic_s ) + roman_sin italic_θ over^ start_ARG italic_z end_ARG ( italic_s ) (2n)

where {𝐫^,ϕ^,z^}\hat{\bf{r}},\hat{\bf{\phi}},\hat{z}\}over^ start_ARG bold_r end_ARG , over^ start_ARG italic_ϕ end_ARG , over^ start_ARG italic_z end_ARG } are the canonical cylindrical coordinate basis vectors and tanθ=2πR/P𝜃2𝜋𝑅𝑃\tan\theta=2\pi R/Proman_tan italic_θ = 2 italic_π italic_R / italic_P:

𝐫^=cos(2πs/)x^+sin(2πs/)y^^𝐫2𝜋𝑠^𝑥2𝜋𝑠^𝑦\hat{\bf{r}}=\cos(2\pi s/\ell)\hat{x}+\sin(2\pi s/\ell)\hat{y}over^ start_ARG bold_r end_ARG = roman_cos ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_x end_ARG + roman_sin ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_y end_ARG (2o)
ϕ^=sin(2πs/)x^+cos(2πs/)y^^italic-ϕ2𝜋𝑠^𝑥2𝜋𝑠^𝑦\hat{\bf{\phi}}=-\sin(2\pi s/\ell)\hat{x}+\cos(2\pi s/\ell)\hat{y}over^ start_ARG italic_ϕ end_ARG = - roman_sin ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_x end_ARG + roman_cos ( 2 italic_π italic_s / roman_ℓ ) over^ start_ARG italic_y end_ARG (2p)

A.2 2D “Croiss-section” Derivation

We select the contents of the tube that lie in the x^^𝑥\hat{x}over^ start_ARG italic_x end_ARG, y^^𝑦\hat{y}over^ start_ARG italic_y end_ARG plane at fixed height, z0subscript𝑧0z_{0}italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT

𝐗(s,ψ)z^=z0𝐗𝑠𝜓^𝑧subscript𝑧0{\bf X}(s,\psi)\cdot\hat{z}=z_{0}bold_X ( italic_s , italic_ψ ) ⋅ over^ start_ARG italic_z end_ARG = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (2q)
P2πls+acosψsinαsinθ+bsinψcosαsinθ=z0𝑃2𝜋𝑙𝑠𝑎𝜓𝛼𝜃𝑏𝜓𝛼𝜃subscript𝑧0\frac{P}{2\pi l}s+a\cos\psi\sin\alpha\sin\theta+b\sin\psi\cos\alpha\sin\theta=% z_{0}divide start_ARG italic_P end_ARG start_ARG 2 italic_π italic_l end_ARG italic_s + italic_a roman_cos italic_ψ roman_sin italic_α roman_sin italic_θ + italic_b roman_sin italic_ψ roman_cos italic_α roman_sin italic_θ = italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (2r)

This constraint allows us to eliminate one of the two parametric variable by yielding the relationship between backbone (s𝑠sitalic_s) and surface (ψ𝜓\psiitalic_ψ) coordinate (that satisfies the planar constraint).

s0(ψ)=2πsinθP(acosψsinα+bsinψcosα+z0)subscript𝑠0𝜓2𝜋𝜃𝑃𝑎𝜓𝛼𝑏𝜓𝛼subscript𝑧0s_{0}(\psi)=-\frac{2\pi\ell\sin\theta}{P}(a\cos\psi\sin\alpha+b\sin\psi\cos% \alpha+z_{0})italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ψ ) = - divide start_ARG 2 italic_π roman_ℓ roman_sin italic_θ end_ARG start_ARG italic_P end_ARG ( italic_a roman_cos italic_ψ roman_sin italic_α + italic_b roman_sin italic_ψ roman_cos italic_α + italic_z start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) (2s)

By substitution of this constraint, the full 3D model reduces to a 2D one-coordinate function (see Fig. 3), where the (in-plane) boundary of the tube is given by

𝐗(ψ)=𝐗(s0(ψ),ψ)subscript𝐗perpendicular-to𝜓𝐗subscript𝑠0𝜓𝜓{\bf X}_{\perp}(\psi)={\bf X}\big{(}s_{0}(\psi),\psi\big{)}bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) = bold_X ( italic_s start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ψ ) , italic_ψ ) (2t)

And the 2D tangent of the surface is given by

𝐓(ψ)=𝐗(ψ)|𝐗(ψ)|subscript𝐓perpendicular-to𝜓superscriptsubscript𝐗perpendicular-to𝜓superscriptsubscript𝐗perpendicular-to𝜓{\bf T}_{\perp}(\psi)=\frac{{\bf X}_{\perp}^{\prime}(\psi)}{|{\bf X}_{\perp}^{% \prime}(\psi)|}bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) = divide start_ARG bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ψ ) end_ARG start_ARG | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ψ ) | end_ARG (2u)

Appendix B Determination of Self-Contact

B.1 Non-local Contact

To determine the distance between any two in-plane points, we define the separation vector,

𝚫(ψ1,ψ2)=𝐗(ψ1)𝐗(ψ2)subscript𝚫perpendicular-tosubscript𝜓1subscript𝜓2subscript𝐗perpendicular-tosubscript𝜓1subscript𝐗perpendicular-tosubscript𝜓2{\bf{\Delta_{\perp}}}(\psi_{1},\psi_{2})={\bf X}_{\perp}(\psi_{1})-{\bf X}_{% \perp}(\psi_{2})bold_Δ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) - bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (2v)

The distance of closest approach is defined between surface coordinates (ψ1(\psi_{1}( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ψ2)\psi_{2})italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) when 𝚫(ψ1,ψ2)subscript𝚫perpendicular-tosubscript𝜓1subscript𝜓2{\bf{\Delta_{\perp}}}(\psi_{1},\psi_{2})bold_Δ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) is orthogonal to both surface tangents, 𝐓(ψ1)subscript𝐓perpendicular-tosubscript𝜓1{\bf T}_{\perp}(\psi_{1})bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) and 𝐓(ψ2)subscript𝐓perpendicular-tosubscript𝜓2{\bf T}_{\perp}(\psi_{2})bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) (or, equivalently, 𝐓(ψ1)𝐓(ψ2)conditionalsubscript𝐓perpendicular-tosubscript𝜓1subscript𝐓perpendicular-tosubscript𝜓2{\bf T}_{\perp}(\psi_{1})\parallel{\bf T}_{\perp}(\psi_{2})bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ∥ bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ). In determining instances of non-local contact, we seek to simultaneously satisfy 𝐓(ψ1)𝐓(ψ2)conditionalsubscript𝐓perpendicular-tosubscript𝜓1subscript𝐓perpendicular-tosubscript𝜓2{\bf T}_{\perp}(\psi_{1})\parallel{\bf T}_{\perp}(\psi_{2})bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ∥ bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ), 𝚫(ψ1,ψ2)0subscript𝚫perpendicular-tosubscript𝜓1subscript𝜓20{\bf{\Delta_{\perp}}}(\psi_{1},\psi_{2})\rightarrow 0bold_Δ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) → 0, and ψ1ψ2subscript𝜓1subscript𝜓2\psi_{1}\neq\psi_{2}italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT.

Recalling that the 𝐗(ψ)subscript𝐗perpendicular-to𝜓{\bf X}_{\perp}(\psi)bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) croiss-section plane has z^^𝑧\hat{z}over^ start_ARG italic_z end_ARG as a normal, we define three inequalities to capture the distance of closest approach

|𝐓(ψ1)𝚫(ψ1,ψ2)|20superscriptsubscript𝐓perpendicular-tosubscript𝜓1subscript𝚫perpendicular-tosubscript𝜓1subscript𝜓220{\big{|}}{\bf T}_{\perp}(\psi_{1})\cdot{\bf{\Delta_{\perp}}}(\psi_{1},\psi_{2}% ){\big{|}}^{2}\geq 0| bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) ⋅ bold_Δ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≥ 0 (2w)
|[𝐓(ψ1)×𝐓(ψ2)]z^|20superscriptdelimited-[]subscript𝐓perpendicular-tosubscript𝜓1subscript𝐓perpendicular-tosubscript𝜓2^𝑧20{\Big{|}}{\big{[}}{\bf T}_{\perp}(\psi_{1})\times{\bf T}_{\perp}(\psi_{2}){% \big{]}}\cdot\hat{z}{\Big{|}}^{2}\geq 0| [ bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) × bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] ⋅ over^ start_ARG italic_z end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≥ 0 (2x)
|[𝐓(ψ2)×Δ(ψ1,ψ2)]z^|20superscriptdelimited-[]subscript𝐓perpendicular-tosubscript𝜓2subscriptΔperpendicular-tosubscript𝜓1subscript𝜓2^𝑧20{\Big{|}}{\big{[}}{\bf T}_{\perp}(\psi_{2})\times{\Delta_{\perp}}(\psi_{1},% \psi_{2}){\big{]}}\cdot\hat{z}{\Big{|}}^{2}\geq 0| [ bold_T start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) × roman_Δ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) ] ⋅ over^ start_ARG italic_z end_ARG | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ≥ 0 (2y)

When the first two quantities saturate 00, ψ1subscript𝜓1\psi_{1}italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ψ2subscript𝜓2\psi_{2}italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT correspond to a location of distance of closest approach; these points are in the neighborhood of non-local contact (given a sufficiently small P𝑃Pitalic_P). The third quantity extracts the separation distance between the points; when this quantity also equals 00, non-local contact occurs. We numerically solve this coupled system for saturation of all three constraints (in terms of the contact pitch, Pnonlocalsubscript𝑃nonlocalP_{\rm non-local}italic_P start_POSTSUBSCRIPT roman_non - roman_local end_POSTSUBSCRIPT and the surface contact coordinates, ψ1subscript𝜓1\psi_{1}italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ψ2subscript𝜓2\psi_{2}italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT) for every combination of R𝑅Ritalic_R, α𝛼\alphaitalic_α, and ϵitalic-ϵ\epsilonitalic_ϵ. In this procedure, we impose a constraint that ψ1ψ2subscript𝜓1subscript𝜓2\psi_{1}\neq\psi_{2}italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ≠ italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT; the angular constraints baked into our inequalities then maintain that ψ1subscript𝜓1\psi_{1}italic_ψ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and ψ2subscript𝜓2\psi_{2}italic_ψ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT are in fact non-local along the tube surface.

B.2 Local Contact

Local contact occurs at a single location along the tube surface where the surface fails to be smooth; this discontinuity manifests as a crease in the surface or cusp singularity in the in-plane representation. Operationally, these singularities can be selected by locations where the differential surface area of the tube vanishes, corresponding to the surface being multivalued.

Using the surface metric, gg\rm groman_g, the magnitude of the surface normal can be expressed as

det(g)=|s𝐗(s,ψ)×ψ𝐗(s,ψ)|detgsubscript𝑠𝐗𝑠𝜓subscript𝜓𝐗𝑠𝜓\sqrt{\rm det(g)}=|\partial_{s}{\bf X}(s,\psi)\times\partial_{\psi}{\bf X}(s,% \psi)|square-root start_ARG roman_det ( roman_g ) end_ARG = | ∂ start_POSTSUBSCRIPT italic_s end_POSTSUBSCRIPT bold_X ( italic_s , italic_ψ ) × ∂ start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT bold_X ( italic_s , italic_ψ ) | (2z)

which yields the surface area of a differential element as

dA=det(g)dsdψ𝑑𝐴detg𝑑𝑠𝑑𝜓dA=\sqrt{\rm det(g)}\,ds\,d\psiitalic_d italic_A = square-root start_ARG roman_det ( roman_g ) end_ARG italic_d italic_s italic_d italic_ψ (2aa)

We select vanishing area elements via solutions to det(g)=0detg0\rm det(g)=0roman_det ( roman_g ) = 0, which yields a sum of three squares (corresponding to component for each of the orthonormal basis components), c12(ψ)+c22(ψ)+c32(ψ)=0superscriptsubscript𝑐12𝜓superscriptsubscript𝑐22𝜓superscriptsubscript𝑐32𝜓0{c_{1}}^{2}(\psi)+{c_{2}}^{2}(\psi)+{c_{3}}^{2}(\psi)=0italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) + italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = 0.

c12(ψ)=b2cosψsinψa2cosψsinψsuperscriptsubscript𝑐12𝜓superscript𝑏2𝜓𝜓superscript𝑎2𝜓𝜓{c_{1}}^{2}(\psi)=b^{2}\cos\psi\sin\psi-a^{2}\cos\psi\sin\psi\\ italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos italic_ψ roman_sin italic_ψ - italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos italic_ψ roman_sin italic_ψ (2ab)
c22(ψ)=abcos2ψcosαsinθb2cosψsinψsinαsinθbcosψsuperscriptsubscript𝑐22𝜓𝑎𝑏superscript2𝜓𝛼𝜃superscript𝑏2𝜓𝜓𝛼𝜃𝑏𝜓{c_{2}}^{2}(\psi)=ab\cos^{2}\psi\cos\alpha\sin\theta-b^{2}\cos\psi\sin\psi\sin% \alpha\sin\theta-\ell b\cos\psi\\ italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = italic_a italic_b roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ roman_cos italic_α roman_sin italic_θ - italic_b start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos italic_ψ roman_sin italic_ψ roman_sin italic_α roman_sin italic_θ - roman_ℓ italic_b roman_cos italic_ψ (2ac)
c32(ψ)=absin2ψsinαsinθ+a2cosψsinψcosαsinθasinψsuperscriptsubscript𝑐32𝜓𝑎𝑏superscript2𝜓𝛼𝜃superscript𝑎2𝜓𝜓𝛼𝜃𝑎𝜓{c_{3}}^{2}(\psi)=-ab\sin^{2}\psi\sin\alpha\sin\theta+a^{2}\cos\psi\sin\psi% \cos\alpha\sin\theta-\ell a\sin\psiitalic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = - italic_a italic_b roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_ψ roman_sin italic_α roman_sin italic_θ + italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos italic_ψ roman_sin italic_ψ roman_cos italic_α roman_sin italic_θ - roman_ℓ italic_a roman_sin italic_ψ (2ad)

Of course, satisfying c12(ψ)+c22(ψ)+c32(ψ)=0superscriptsubscript𝑐12𝜓superscriptsubscript𝑐22𝜓superscriptsubscript𝑐32𝜓0{c_{1}}^{2}(\psi)+{c_{2}}^{2}(\psi)+{c_{3}}^{2}(\psi)=0italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) + italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) + italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = 0 requires that each term equal zero.

Importantly, the first condition, c12superscriptsubscript𝑐12{c_{1}}^{2}italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, exhibits a singular dependence on the aspect ratio, ϵitalic-ϵ\epsilonitalic_ϵ; when the cross-section is circular (ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1), the condition is automatically satisfied for any value of ψ𝜓\psiitalic_ψ and the minimal pitch (Plocalsubscript𝑃localP_{\rm local}italic_P start_POSTSUBSCRIPT roman_local end_POSTSUBSCRIPT takes the form of a helix “limited by it’s curvature” presented by Przybyl and Pieranski [78], see eqn. 2h.

When ϵ<1italic-ϵ1\epsilon<1italic_ϵ < 1, satisfying the initial constraint (c12(ψ)=0superscriptsubscript𝑐12𝜓0{c_{1}}^{2}(\psi)=0italic_c start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) = 0) requires that the product cosψsinψ=0𝜓𝜓0\cos\psi\sin\psi=0roman_cos italic_ψ roman_sin italic_ψ = 0; naturally, this requires that either cosψ𝜓\cos\psiroman_cos italic_ψ or sinψ=0𝜓0\sin\psi=0roman_sin italic_ψ = 0 (i.e. ψ=0,π/2,π,3π/2,2π𝜓0𝜋2𝜋3𝜋22𝜋\psi={0,\pi/2,\pi,3\pi/2,2\pi}italic_ψ = 0 , italic_π / 2 , italic_π , 3 italic_π / 2 , 2 italic_π). Substitution of these constraints into c22(ψ)superscriptsubscript𝑐22𝜓{c_{2}}^{2}(\psi)italic_c start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) and c32(ψ)superscriptsubscript𝑐32𝜓{c_{3}}^{2}(\psi)italic_c start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_ψ ) yields two physical solutions for Plocalsubscript𝑃localP_{\rm local}italic_P start_POSTSUBSCRIPT roman_local end_POSTSUBSCRIPT, see eqn. 2i and 2j:

These solutions present as ellipses in the (R,P)𝑅𝑃(R,P)( italic_R , italic_P ) plane, making it trivial to identify their center and vertices. The ellipses are described by

P2(aπcosα)2+(R12acosα)2(12acosα)2=1superscript𝑃2superscript𝑎𝜋𝛼2superscript𝑅12𝑎𝛼2superscript12𝑎𝛼21\frac{{P}^{2}}{(a\pi\cos\alpha)^{2}}+\frac{(R-\frac{1}{2}a\cos\alpha)^{2}}{(% \frac{1}{2}a\cos\alpha)^{2}}=1divide start_ARG italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_a italic_π roman_cos italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG ( italic_R - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_a roman_cos italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_a roman_cos italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 1 (2ae)

and

P2(bπsinα)2+(R12bsinα)2(12bsinα)2=1superscript𝑃2superscript𝑏𝜋𝛼2superscript𝑅12𝑏𝛼2superscript12𝑏𝛼21\frac{{P}^{2}}{(b\pi\sin\alpha)^{2}}+\frac{(R-\frac{1}{2}b\sin\alpha)^{2}}{(% \frac{1}{2}b\sin\alpha)^{2}}=1divide start_ARG italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_b italic_π roman_sin italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG ( italic_R - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_b roman_sin italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_b roman_sin italic_α ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 1 (2af)

Perhaps more intuitively (or at least more graphically), this contact feature can also be detected numerically by a diverging in-plane curvature; the geometric significance of both operations are the same. In Fig 13, the in-plane radius of curvature, κ1=|𝐗(ψ)|3/2/|𝐗(ψ)×𝐗′′(ψ)|superscriptsubscript𝜅perpendicular-to1superscriptsuperscriptsubscript𝐗perpendicular-to𝜓32superscriptsubscript𝐗perpendicular-to𝜓superscriptsubscript𝐗perpendicular-to′′𝜓\kappa_{\perp}^{-1}=|{\bf X}_{\perp}^{\prime}(\psi)|^{3/2}/|{\bf X}_{\perp}^{% \prime}(\psi)\times{\bf X}_{\perp}^{\prime\prime}(\psi)|italic_κ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ψ ) | start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT / | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT ( italic_ψ ) × bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT ′ ′ end_POSTSUPERSCRIPT ( italic_ψ ) | is shown for the locally self-contacting structure presented in Fig. 3B.

Refer to caption
Figure 13: Radius of curvature (κ1superscriptsubscript𝜅perpendicular-to1\kappa_{\perp}^{-1}italic_κ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT) along 𝐗(ψ)subscript𝐗perpendicular-to𝜓{\bf X}_{\perp}(\psi)bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) corresponding to Fig. 3B(top); In this structure, κ1superscriptsubscript𝜅perpendicular-to1\kappa_{\perp}^{-1}italic_κ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT spans 0 (cusp) to \infty (inflection); for visualization the color bar scales logarithmically and was truncated to saturate at κ1=105superscriptsubscript𝜅perpendicular-to1superscript105\kappa_{\perp}^{-1}=10^{-5}italic_κ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT (dark blue) and κ1=105superscriptsubscript𝜅perpendicular-to1superscript105\kappa_{\perp}^{-1}=10^{5}italic_κ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = 10 start_POSTSUPERSCRIPT 5 end_POSTSUPERSCRIPT (dark red).

B.3 The Local - Non-local Boundary

As reported in [78], the isotropic filament contact regimes meet a single point, R0.431d0similar-to-or-equals𝑅0.431subscript𝑑0R\simeq 0.431d_{0}italic_R ≃ 0.431 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, P1.083d0similar-to-or-equals𝑃1.083subscript𝑑0P\simeq 1.083d_{0}italic_P ≃ 1.083 italic_d start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. This structure, shown in Fig 5A.iv simultaneously exhibits local and non-local. Perhaps intuitively, these two modes occur at distinct locations along the tube surface (i.e distinct ψ𝜓\psiitalic_ψ coordinates); the local contact (or curvature constraint) occurs inside the helical core (i.e in the direction of curvature) while the non-local contact occurs between successive turns (i.e roughly in the vertical direction). Several anisotropic filaments with local/non-local coexistence were presented in the main text (Fig. 6C.ii, Fig. 6C.iv, Fig. 7C.ii, Fig. 7C.v, and Fig. 8C.iii), however all of these structures appear to exhibit simultaneous contact at a single point in the croiss-section; the two sites of contact are indistinct. This curious difference occurs due to the specific pathway shown in Fig. 6C - 8C, in particular that the coexistent structure is shown at a relatively low value for R𝑅Ritalic_R. In Fig. 14, we show additional structures along the local/non-local coexistence line. For both ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95 and ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25, we find that the boundary exhibits regions of distinct and indistinct coexistence, at large and small R𝑅Ritalic_R, respectively; transitions between these motifs can be determined by inflections in the coexistence boundaries shown in Fig. 6B and 8B. It seems then that this indistinct coexistence is then a consequence (or in fact, a cause) of the non-local contact regime’s ability to persist to smaller helical radii; this structural motif is an inherent feature of cross-sectional anisotropy.

Refer to caption
Figure 14: Coexistence between Local (L) and Non-local (NL) contact. (A.) Single point of isotropic (ϵ=1italic-ϵ1\epsilon=1italic_ϵ = 1) L/NL coexistence; here, the local contact points “inward” (hence classified as local(-)) and the contacts occur at distinct locations along the tube surface. (B.) States along the L/NL coexistence boundary for slightly anisotropic tubes (ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95); the boundary pans both local(-) and local(+), as discussed in Section 3.2. Most states (i.e. intermediate α𝛼\alphaitalic_α)coexistent contact at a indistinct (i.e. the same) location along the tube surface exhibit; extreme values of α𝛼\alphaitalic_α exhibit distinct contact. (C.) States along the L/NL coexistence boundary for highly anisotropic tubes (ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25). States at smaller α𝛼\alphaitalic_α exhibit indistinct contact while states at larger α𝛼\alphaitalic_α exhibit distinct contact.

B.4 Nested packings at low RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT and high ϵitalic-ϵ\epsilonitalic_ϵ?

Highly anistropic structures follow a straightforward progression, tape \rightarrow scroll \rightarrow screw as RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT increases, as shown in e.g. Fig. 10A,bottom for ϵ=0.25italic-ϵ0.25\epsilon=0.25italic_ϵ = 0.25. Curiously, when ϵitalic-ϵ\epsilonitalic_ϵ is large, we observe a second state of scroll-like packing, which is apparently more dense than the corresponding tape-like packing. This phenomena is observed under the entire ϵitalic-ϵ\epsilonitalic_ϵ scale where the Local(-) branch constrains the contact manifold, 1>ϵ0.81751italic-ϵgreater-than-or-equivalent-to0.81751>\epsilon\gtrsim 0.81751 > italic_ϵ ≳ 0.8175, and corresponds for a preference for non-local contact over local(-) contact; we therefore dub this transition as the “Cusp Escape” regime.

Refer to caption
Figure 15: Optimal packing at high ϵitalic-ϵ\epsilonitalic_ϵ yields two nested regimes. (A.) ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95 contact mode phase diagram (black, see Fig. 6B), optimal packing (red, corresponding to Fig. 10A), and the “Ideal” low ϵitalic-ϵ\epsilonitalic_ϵ tape \rightarrow scroll \rightarrow screw progression (gray, dashed). The preferred packing exhibits intermediate tilt to “escape” the local(-) regime. (B. top) ϕitalic-ϕ\phiitalic_ϕ (black) and θ𝜃\thetaitalic_θ (blue) along the optimal branch, as well as along the “Ideal” branch (gray, dashed). (B. bottom) Normalized deviation of the quantities (q𝑞qitalic_q), ϕitalic-ϕ\phiitalic_ϕ and θ𝜃\thetaitalic_θ.

Based on Fig. 15B, it is obvious that this preference for intermediate tilt at low RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT does not significantly influence the packing; for ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95, the helical angle (θ𝜃\thetaitalic_θ) and the packing density (ϕitalic-ϕ\phiitalic_ϕ) only deviate by at most 2%similar-to-or-equalsabsentpercent2\simeq 2\%≃ 2 %; this largely is a consequence of the weak anisotropy. While this secondary tilt motif persists down to ϵ0.8175similar-to-or-equalsitalic-ϵ0.8175\epsilon\simeq 0.8175italic_ϵ ≃ 0.8175 (where tilt should have a larger consequence), the local(-) regime decrease in size with decreasing ϵitalic-ϵ\epsilonitalic_ϵ (see Fig. 7); as a result, the magnitude of α𝛼\alphaitalic_α (and therefore deviation from “Ideal” packing) needed for “Cusp Escape” also decreases (compare, e.g. ϵ=0.95italic-ϵ0.95\epsilon=0.95italic_ϵ = 0.95 and ϵ=0.85italic-ϵ0.85\epsilon=0.85italic_ϵ = 0.85 in Fig. 10A. As such, we consider these low RCsubscript𝑅𝐶R_{C}italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT/high ϵitalic-ϵ\epsilonitalic_ϵ nested states to be a relatively inconsequential deviation from the more intuitive tape \rightarrow scroll \rightarrow screw progression with increasing capillary size.

Appendix C Capillary Contact Maximizes Density

In the present work, we specifically report capillary density of dense (i.e. self-contacting) helices within their tightest capillary. However, our approach is in fact far more general; the capillary density, given in expression Eqn. 2k can be applied to helices of arbitrary configuration (in particular, “expanded helices” with P>Pmin𝑃subscript𝑃minP>P_{\rm min}italic_P > italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT) within arbitrary capillaries (in particular, with RCmaxψ|𝐗(ψ)|subscript𝑅𝐶subscriptmax𝜓subscript𝐗perpendicular-to𝜓R_{C}\geq{\rm max}_{\psi}|{\bf X}_{\perp}(\psi)|italic_R start_POSTSUBSCRIPT italic_C end_POSTSUBSCRIPT ≥ roman_max start_POSTSUBSCRIPT italic_ψ end_POSTSUBSCRIPT | bold_X start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ψ ) |).

While this larger ensemble of states may certainly be of interest in some physical phenomena or systems, our current focus is on states that maximize density.

It can be shown analytically that

ϕP0italic-ϕ𝑃0\frac{\partial\phi}{\partial P}\leq 0divide start_ARG ∂ italic_ϕ end_ARG start_ARG ∂ italic_P end_ARG ≤ 0 (2ag)

for PPmin𝑃subscript𝑃minP\geq P_{\rm min}italic_P ≥ italic_P start_POSTSUBSCRIPT roman_min end_POSTSUBSCRIPT (independent of the values of R𝑅Ritalic_R, α𝛼\alphaitalic_α, ϵitalic-ϵ\epsilonitalic_ϵ), so in surveying the densest configurations, it is appropriate to only consider packings which simultaneously exhibit self- and capillary-contact.

References

References

  • [1] Neville A 1993 Biology of Fibrous Composites: Development Beyond the Cell Membrane (Cambridge University Press) ISBN 9780521410519
  • [2] Marth J D 2008 Nature Cell Biology 10 1015–1015
  • [3] Fratzl P 2003 Current Opinion in Colloid & Interface Science 8 32–39 ISSN 1359-0294
  • [4] Zhang M, Biesold G M, Choi W, Yu J, Deng Y, Silvestre C and Lin Z 2022 Materials Today 53 134–161 ISSN 1369-7021
  • [5] Paderes M, Ahirwal D and Prieto S F 2017 Physical Sciences Reviews 2 20170021
  • [6] Saiter J M, Shit S C and Shah P M 2014 Journal of Polymers 2014 427259
  • [7] Hearle J, Grosberg P and Backer S 1969 Structural Mechanics of Fibers, Yarns, and Fabrics (Structural Mechanics of Fibers, Yarns, and Fabrics no v. 1) (Wiley-Interscience) ISBN 9780471366690
  • [8] Pan N and Brookstein D 2002 Journal of Applied Polymer Science 83 610–630
  • [9] Pan N 2014 Applied Physics Reviews 1 021302 ISSN 1931-9401
  • [10] Briggs-Goode A and Townsend K 2011 Textile design: principles, advances and applications
  • [11] Schneider J 1987 Annual Review of Anthropology 16 409–448 ISSN 00846570
  • [12] Patti A and Acierno D 2023 Macromol 3 665–680 ISSN 2673-6209
  • [13] Natarajan R and Ganapathy C 1995 Marine Structures 8 481–499 ISSN 0951-8339
  • [14] Chaplin C 1999 Engineering Failure Analysis 6 67–82 ISSN 1350-6307
  • [15] Costello G 2012 Theory of Wire Rope Mechanical Engineering Series (Springer New York) ISBN 9781468403503
  • [16] Krishna P 2001 Journal of Constructional Steel Research 57 1123–1140 ISSN 0143-974X
  • [17] Rodney D, Fivel M and Dendievel R 2005 Phys. Rev. Lett. 95(10) 108004
  • [18] Jawed M K, Dieleman P, Audoly B and Reis P M 2015 Phys. Rev. Lett. 115(11) 118302
  • [19] Singal K, Dimitriyev M S, Gonzalez S E, Cachine A P, Quinn S and Matsumoto E A 2024 Nature Communications 15 1–9
  • [20] Gusev A A, Hine P J and Ward I M 2000 Composites Science and Technology 60 535–541 ISSN 0266-3538
  • [21] Chaffey N 2003 Annals of Botany 91 401–401
  • [22] Montserret R, McLeish M J, Böckmann A, Geourjon C and Penin F 2000 Biochemistry 39 8362–8373
  • [23] Humphrey W, Dalke A and Schulten K 1996 Journal of Molecular Graphics 14 33–38
  • [24] Wang Y, Qi W, Huang R, Yang X, Wang M, Su R and He Z 2015 Journal of the American Chemical Society 137 7869–7880 pMID: 26018930
  • [25] Barber D M, Emrick T, Grason G M and Crosby A J 2023 Nature Communications 14 625
  • [26] Li Y, Guo F, Hao Y, Gupta S K, Hu J, Wang Y, Wang N, Zhao Y and Guo M 2019 Proceedings of the National Academy of Sciences 116 9245–9250
  • [27] An Y, Gao L and Wang T 2020 ACS Applied Nano Materials 3 5079–5087
  • [28] Kepler J 1611 Strena seu de nive sexangula (Tambach, Gottfried) ISBN 9780198506256
  • [29] Aste T and Weaire D 2000 The pursuit of perfect packing
  • [30] Pickover C 2009 The Math Book: From Pythagoras to the 57th Dimension, 250 Milestones in the History of Mathematics Sterling Milestones Series (Sterling) ISBN 9781402757969
  • [31] Audoly B and Pomeau Y 2010 Elasticity and Geometry: From Hair Curls to the Non-linear Response of Shells (OUP Oxford) ISBN 9780198506256
  • [32] Cambou A D and Menon N 2011 Proceedings of the National Academy of Sciences 108 14741–14745
  • [33] Boué L, Adda-Bedia M, Boudaoud A, Cassani D, Couder Y, Eddi A and Trejo M 2006 Phys. Rev. Lett. 97(16) 166104
  • [34] Vliegenthart G A and Gompper G 2006 Nature Materials 5 216–221
  • [35] Terasaki M, Shemesh T, Kasthuri N, Klemm R W, Schalek R, Hayworth K J, Hand A R, Yankova M, Huber G, Lichtman J W, Rapoport T A and Kozlov M M 2013 Cell 154 285–296 ISSN 0092-8674
  • [36] Bussi Y, Shimoni E, Weiner A, Kapon R, Charuvi D, Nevo R, Efrati E and Reich Z 2019 Proceedings of the National Academy of Sciences 116 22366–22375
  • [37] Tallinen T, Chung J Y, Rousseau F, Girard N, Lefèvre J and Mahadevan L 2016 Nature Physics 12 588–593
  • [38] Grason G M 2015 Rev. Mod. Phys. 87(2) 401–419
  • [39] Vetter R, Wittel F K and Herrmann H J 2015 Europhysics Letters 112 44003
  • [40] Weiner N, Bhosale Y, Gazzola M and King H 2020 Journal of Applied Physics 127 050902 ISSN 0021-8979
  • [41] Guerra A, Slim A C, Holmes D P and Kodio O 2023 Phys. Rev. Lett. 130(14) 148201
  • [42] Grason G M 2009 Phys. Rev. E 79(4) 041919
  • [43] Atkinson D W, Santangelo C D and Grason G M 2019 New Journal of Physics 21 062001
  • [44] Katritch V, Bednar J, Michoud D, Scharein R G, Dubochet J and Stasiak A 1996 Nature 384 142–145
  • [45] Moffatt H K 1996 Nature 384 114–114
  • [46] Stasiak A, Katritch V and Kauffman L H 1998 Ideal Knots (WORLD SCIENTIFIC)
  • [47] Cantarella J, Kusner R B and Sullivan J M 1998 Nature 392 237–238
  • [48] Gonzalez O and Maddocks J H 1999 Proceedings of the National Academy of Sciences of the United States of America 96 9 4769–73
  • [49] Pieranski P 1998 In search of ideal knots pp 20–41
  • [50] Vargas-Lara F, Hassan A M, Mansfield M L and Douglas J F 2017 Scientific Reports 7 13374
  • [51] Johanns P, Grandgeorge P, Baek C, Sano T G, Maddocks J H and Reis P M 2021 Extreme Mechanics Letters 43 101172 ISSN 2352-4316
  • [52] Klotz A R and Maldonado M 2021 Journal of Physics A: Mathematical and Theoretical 54 445201
  • [53] Klotz A R 2022 Journal of Knot Theory and Its Ramifications 31 2250102
  • [54] Starostin E 2003 PAMM 3 479–480
  • [55] Grandgeorge P, Baek C, Singh H, Johanns P, Sano T G, Flynn A, Maddocks J H and Reis P M 2021 Proceedings of the National Academy of Sciences 118 e2021684118
  • [56] Bruss I R and Grason G M 2012 Proceedings of the National Academy of Sciences 109 10781–10786
  • [57] Starostin E L 2006 Journal of Physics: Condensed Matter 18 S187
  • [58] Bohr J and Olsen K 2011 Europhysics Letters 93 60004
  • [59] Neukirch S and van der Heijden G 2002 Journal of Elasticity 69 ISSN 03743535
  • [60] Cajamarca L and Grason G M 2014 The Journal of Chemical Physics 141 174904
  • [61] Seguin A and Crassous J 2022 Phys. Rev. Lett. 128(7) 078002
  • [62] Ghatak A and Mahadevan L 2005 Phys. Rev. Lett. 95(5) 057801
  • [63] Panaitescu A, Grason G M and Kudrolli A 2017 Phys. Rev. E 95(5) 052503
  • [64] Panaitescu A, Grason G M and Kudrolli A 2018 Phys. Rev. Lett. 120(24) 248002
  • [65] Chopin J, Biswas A and Kudrolli A 2024 Phys. Rev. E 109(2) 025003
  • [66] Miller J T, Lazarus A, Audoly B and Reis P M 2014 Phys. Rev. Lett. 112(6) 068103
  • [67] Jawed M K, Da F, Joo J, Grinspun E and Reis P M 2014 Proceedings of the National Academy of Sciences 111 14663–14668
  • [68] Stasiak A and Maddocks J H 2000 Nature 406 251–252 ISSN 1476-4687
  • [69] Maritan A, Micheletti C, Trovato A and Banavar J R 2000 Nature 406 287–290 ISSN 1476-4687
  • [70] Poletto C, Giacometti A, Trovato A, Banavar J R and Maritan A 2008 Physical review. E, Statistical, nonlinear, and soft matter physics 77 6 Pt 1 061804
  • [71] Bohr J and Olsen K W 2011 The size of the nucleosome
  • [72] Olsen K W and Bohr J 2012 New Journal of Physics 15
  • [73] Banavar J R, Giacometti A, Hoang T X, Maritan A and Škrbić T Proteins: Structure, Function, and Bioinformatics
  • [74] Kubalová I, Câmara A S, Cápal P, Beseda T, Rouillard J M, Krause G M, Holušová K, Toegelová H, Himmelbach A, Stein N, Houben A, Doležel J, Mascher M, Šimková H and Schubert V 2023 Nucleic Acids Research 51 2641 – 2654
  • [75] Snir Y and Kamien R D 2006 Physical review. E, Statistical, nonlinear, and soft matter physics 75 5 Pt 1 051114
  • [76] Banavar J R and Maritan A 2003 Rev. Mod. Phys. 75(1) 23–34
  • [77] Chouaieb N, Goriely A and Maddocks J H 2006 Proceedings of the National Academy of Sciences 103 9398–9403
  • [78] Przybyl S and Pieranski P 2001 Eur. Phys. J. E 4 445–449
  • [79] Hansen-Goos H, Roth R, Mecke K and Dietrich S 2007 Phys. Rev. Lett. 99(12) 128101
  • [80] Poletto C, Giacometti A, Trovato A, Banavar J R and Maritan A 2008 Phys. Rev. E 77(6) 061804
  • [81] Yashima E, Ousaka N, Taura D, Shimomura K, Ikai T and Maeda K 2016 Chemical Reviews 116 13752–13990
  • [82] Guyard C, Raffel S, Schrumpf M, Dahlstrom E, Sturdevant D, Ricklefs S, Martens C, Hayes S, Fischer E, Hansen B, Porcella S and Schwan T 2013 PloS one 8 e72550
  • [83] Johnson R C, Hyde F W and Rumpel C M 2008 Taxonomy of the lyme disease spirochetes
  • [84] Gaines M K, Page I Y, Miller N A, Greenvall B R, Medina J J, Irschick D J, Southard A, Ribbe A E, Grason G M and Crosby A J 2023 Accounts of Chemical Research 56 1330–1339
  • [85] Smyth D R 2016 Development 143 3272 – 3282
  • [86] Sousa‐Baena M S, Hernandes-Lopes J and Van Sluys M A 2021 Current Opinion in Plant Biology 59 101982 ISSN 1369-5266
  • [87] Isnard S and Silk W K 2009 American Journal of Botany 96 1205–1221
  • [88] Abraham Y, Tamburu C, Klein E, Dunlop J, Fratzl P, Raviv U and Elbaum R 2011 Journal of the Royal Society, Interface / the Royal Society 9 640–7
  • [89] Jayne B C, Newman S J, Zentkovich M M and Berns H M 2015 Journal of Experimental Biology 218 3978–3986 ISSN 0022-0949
  • [90] Brischoux F and Shine R 2011 Journal of Morphology 272 566–572
  • [91] Schulz A, Boyle M, Boyle C, Sordilla S, Rincon C, Hooper S, Aubuchon C, Reidenberg J, Higgins C and Hu D 2022 Proceedings of the National Academy of Sciences 119
  • [92] Kier W M and Stella M P 2007 Journal of Morphology 268 831–843
  • [93] Kier W M 2012 Journal of Experimental Biology 215 1247–1257 ISSN 0022-0949
  • [94] Olsen K W and Bohr J 2010 Theoretical Chemistry Accounts 125 207–215
  • [95] Cerritelli M E, Cheng N, Rosenberg A H, McPherson C E, Booy F P and Steven A C 1997 Cell 91 271–280
  • [96] Purohit P K, Inamdar M M, Grayson P, Squires T M, Kondev J and Phillips R 2004 Biophysical journal 88 2 851–66
  • [97] Hartl F and Hayer-Hartl M 2002 Science (New York, N.Y.) 295 1852–8
  • [98] Ziv G, Haran G and Thirumalai D 2005 Proceedings of the National Academy of Sciences 102 18956–18961
  • [99] Kumar H, Lansac Y, Glaser M and Maiti P 2011 Soft Matter 7 5898–5907
  • [100] Feld G K, Brown M J and Krantz B A 2012 Protein Science 21 606–624
  • [101] Ren Z, Ren P X, Balusu R and Yang X 2016 Scientific Reports 6 34129
  • [102] Wood D A, Santangelo C D and Dinsmore A D 2013 Soft Matter 9(42) 10016–10024
  • [103] ** W, Chan H K and Zhong Z 2020 Physical Review Letters 124
  • [104] Jiménez-Millán S, García-Alcántara C, Ramírez-Hernández A, Sambriski E and Hernández S 2021 Journal of Molecular Liquids 335 116219 ISSN 0167-7322
  • [105] Li F, Badel X, Linnros J and Wiley J B 2005 Journal of the American Chemical Society 127 3268–3269
  • [106] Mughal A, Chan H K, Weaire D and Hutzler S 2012 Physical Review E 85
  • [107] Tymczenko M, Marsal L F, Trifonov T, Rodriguez I, Ramiro-Manzano F, Pallares J, Rodriguez A, Alcubilla R and Meseguer F 2008 Advanced Materials 20 2315–2318
  • [108] Jiang L, de Folter J W J, Huang J, Philipse A P, Kegel W K and Petukhov A V 2013 Angewandte Chemie International Edition 52 3364–3368
  • [109] Schneider L, Lichtenberg G, Vega D and Müller M 2020 ACS Applied Materials & Interfaces 12 50077–50095
  • [110] Lin Y, Böker A, He J, Sill K, Xiang H, Abetz C, Li X, Wang J, Emrick T, Long S, Wang Q, Balazs A and Russell T P 2005 Nature 434 55–59
  • [111] Cheng M H, Hsu Y C, Chang C W, Ko H W, Chung P Y and Chen J T 2017 ACS Applied Materials & Interfaces 9 21010–21016
  • [112] Kim H, Lee S, Shin T J, Korblova E, Walba D M, Clark N A, Lee S B and Yoon D K 2014 Proceedings of the National Academy of Sciences 111 14342–14347
  • [113] Kim H, Ryu S H, Tuchband M, Shin T J, Korblova E, Walba D M, Clark N A and Yoon D K 2017 Science Advances 3 e1602102
  • [114] Palacio-Betancur V, Armas-Pérez J C, Hernández-Ortiz J P and de Pablo J J 2023 Soft Matter 19(32) 6066–6073
  • [115] Chen H Q, Wang X Y, Bisoyi H K, Chen L J and Li Q 2021 Langmuir 37 3789–3807
  • [116] Kamien R D 2002 Rev. Mod. Phys. 74(4) 953–971
  • [117] Goodman A W and Goodman G 1969 The American Mathematical Monthly 76 355–366
  • [118] Palmer B and Pampano A 2020 Annals of Global Analysis and Geometry
  • [119] Lazarus A, Miller J T, Metlitz M M and Reis P M 2013 Soft Matter 9(34) 8274–8281
  • [120] Chouaieb N and Maddocks J H 2004 Journal of Elasticity 77 221–247
  • [121] Domokos G and Healey T J 2005 International Journal of Bifurcation and Chaos 15 871–890
  • [122] Hall D M, Bruss I R, Barone J R and Grason G M 2016 Nature Materials 15 727–732
  • [123] Moed D E, Dimitriyev M S, Greenvall B R, Grason G M and Crosby A J 2024
  • [124] Gnidovec A, Božič A and Čopar S 2022 Soft Matter 18(39) 7670–7678