Nonlinear Poisson effect in affine semiflexible polymer networks

Jordan L. Shivers Present affiliation: James Franck Institute and Department of Chemistry, University of Chicago, Chicago, Illinois 60637, USA Department of Chemical and Biomolecular Engineering, Rice University, Houston, TX 77005, USA Center for Theoretical Biological Physics, Rice University, Houston, TX 77005, USA James Franck Institute, University of Chicago, Chicago, IL 60637, USA Department of Chemistry, University of Chicago, Chicago, IL 60637, USA    Fred C. MacKintosh Department of Chemical and Biomolecular Engineering, Rice University, Houston, TX 77005, USA Center for Theoretical Biological Physics, Rice University, Houston, TX 77005, USA Department of Chemistry, Rice University, Houston, TX 77005, USA Department of Physics & Astronomy, Rice University, Houston, TX 77005, USA
Abstract

Stretching an elastic material along one axis typically induces contraction along the transverse axes, a phenomenon known as the Poisson effect. From these strains, one can compute the specific volume, which generally either increases or, in the incompressible limit, remains constant as the material is stretched. However, in networks of semiflexible or stiff polymers, which are typically highly compressible yet stiffen significantly when stretched, one instead sees a significant reduction in specific volume under finite strains. This volume reduction is accompanied by increasing alignment of filaments along the strain axis and a nonlinear elastic response, with stiffening of the apparent Young’s modulus. For semiflexible networks, in which entropic bending elasticity governs the linear elastic regime, the nonlinear Poisson effect is caused by the nonlinear force-extension relationship of the constituent filaments, which produces a highly asymmetric response of the constituent polymers to stretching and compression. The details of this relationship depend on the geometric and elastic properties of the underlying filaments, which can vary greatly in experimental systems. Here, we provide a comprehensive characterization of the nonlinear Poisson effect in an affine network model and explore the influence of filament properties on essential features of the macroscopic response, including strain-driven alignment and volume reduction.

Networks of semiflexible biopolymers provide elasticity to many biological materials, which they imbue with a variety of distinctive properties that set them apart from conventional elastic media [1, 2, 3, 4]. One of the more remarkable features of biopolymer networks is their highly asymmetric mechanical response: they tend to respond to increasing deformation with a strongly increasing stiffness, often by more than an order of magnitude [5, 6, 7], while providing a comparatively weak or even softening response to compression [8, 9, 10, 11, 12]. The stiffness of typical elastic materials, in contrast, usually depends weakly on the magnitude or nature of the applied load. A striking consequence of the asymmetric response of biopolymer networks is seen in the manner in which they exhibit the Poisson effect, which refers to the tendency of a material stretched along one axis to contract along the transverse axes. Unlike ordinary materials, for which this effect depends only mildly on strain, biopolymer networks exhibit a strongly nonlinear Poisson effect under finite extensional strain [13, 14, 15, 16].

Refer to caption
Figure 1: Extensile strain drives alignment and densification of networks of crosslinked semiflexible polymers. (a) A schematic of the nonlinear Poisson effect in biopolymer networks. Applying (vertical) extensile strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT to an initially isotropic semiflexible polymer network (left) and allowing the (horizontal) transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT to freely vary results in alignment of polymers along the vertical axis, along with local densification due to the transverse contraction of the network (right). The dashed box represents the same portion of the network in the initial and strained states. In the strained state, the stress along the (vertical) strain axis, σsubscript𝜎parallel-to\sigma_{\parallel}italic_σ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT is positive, whereas the transverse (horizontal) stress is relaxed, σ=0subscript𝜎perpendicular-to0\sigma_{\perp}=0italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0. (b) Sketch of a 3D cubic volume element of initial length L𝐿Litalic_L before and after the application of extensile strain ε>0subscript𝜀parallel-to0\varepsilon_{\parallel}>0italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT > 0 along the z𝑧zitalic_z-axis. If the Poisson’s ratio ν𝜈\nuitalic_ν is positive, requiring σ=0subscript𝜎perpendicular-to0\sigma_{\perp}=0italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0 results in transverse contraction quantified by strains ε<0subscript𝜀perpendicular-to0\varepsilon_{\perp}<0italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT < 0.

In the small strain limit, this effect is quantified by Poisson’s ratio ν=limε0+[ε/ε]𝜈subscriptsubscript𝜀parallel-tosuperscript0delimited-[]subscript𝜀parallel-tosubscript𝜀perpendicular-to\nu=\lim_{\varepsilon_{\parallel}\to 0^{+}}[-\varepsilon_{\parallel}/% \varepsilon_{\perp}]italic_ν = roman_lim start_POSTSUBSCRIPT italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT → 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT [ - italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ], in which εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT is the applied extensional strain and εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT is the resulting transverse strain in the absence of transverse stress [17, 18]. As with any stable, isotropic, three-dimensional material, the true (small strain) Poisson’s ratio of biopolymer networks is strictly constrained within the range ν[1,1/2]𝜈112\nu\in[-1,1/2]italic_ν ∈ [ - 1 , 1 / 2 ] [19]. However, under finite applied (extensional) strains, biopolymer networks in various contexts have been shown to exhibit an unusually large and highly strain-dependent apparent Poisson’s ratio [13, 14, 15, 16], corresponding to significant contraction along the transverse axes in response to small increases in extension. This behavior, sketched in Fig. 1, produces a contracted state with pronounced filament alignment along the strain axis [14] and a significant decrease in the local volume occupied by the network [13, 15] that is accompanied by an expulsion of solvent and a significant reduction in the average pore size [14]. In contrast, the apparent Poisson’s ratio changes very little under compression. Although the nonlinear Poisson effect observed in biopolymer networks is a fundamentally network-scale behavior, it nonetheless reflects the mechanical asymmetry of the underlying components: the individual filaments resist finite tension significantly more strongly than they resist finite compression.

The large configurational changes caused by the nonlinear Poisson effect could have meaningful biological consequences, e.g. in cell migration, mechanosensing, and intracellular transport. For example, the Poisson effect produces significant changes in network porosity, which in the extracellular matrix can strongly influence the motility of migrating cells [20, 21]. Likewise, the Poisson effect leads to the strain-driven formation of dense, highly oriented network regions between contractile cells [14, 16] which can influence mechanosensing [22, 23, 24] and induce directed cell motion [25, 26]. Within cells, changes in cytoskeletal network orientation and density caused by the Poisson effect could likewise influence the transport of intracellular cargo both by passive diffusion and by active (e.g., molecular motor-driven or polymerization-driven) transport processes [27]. The influence of the Poisson effect on filament orientation could also play an important role in other active mechanical processes involving the cytoskeleton, such as actomyosin-based cell motility [28, 29] and the organization of stress fibers in response to applied load [30, 31, 32, 33]. Develo** a more comprehensive understanding of how the nonlinear Poisson effect is controlled, e.g. by filament stiffness and network structure, is therefore of interest both for understanding existing biological systems and for designing reconstituted or synthetic materials that mimic these behaviors.

To date, studies addressing the Poisson effect in biopolymer gels have focused primarily on models relevant to networks of stiff athermal biopolymers, such as thick fibrin or collagen fibers, that exhibit enthalpic elasticity and tend to deform in a highly non-affine (inhomogeneous) manner [34, 15, 35, 16, 36]. While such models are appropriate for the stiff constituents of the extracellular matrix, many biopolymers, particularly on the subcellular scale, are semiflexible and thus exhibit entropic elasticity under small strains [37, 38]. For networks of semiflexible polymers, mechanical models neglecting non-affinity have been an effective starting point to capture the mechanics of reconstituted gels, even in highly nonlinear regimes [6, 7], but our understanding of the nonlinear Poisson effect in these systems remains limited. In this work, we aim to elucidate the mechanism underlying the Poisson effect in such systems. Specifically, we consider an affine model of a network of semiflexible filaments, in which the coordinates of the crosslinks (or entanglements) transform affinely with the macroscopic applied strain [39, 40, 41, 42, 43, 44, 45, 7, 46, 47, 48, 49]. We systematically characterize the strain-dependent mechanical and configurational behavior of initially isotropic, three-dimensional networks under varying applied uniaxial strain with a zero transverse stress condition. These conditions mimic a typical experiment used to measure a material’s Poisson’s ratio. Using this model, we explore the influence of various filament properties, including bending rigidity, stretch modulus, and contour length on the nonlinear Poisson effect and the associated strain dependence of the incremental Poisson’s ratio, nematic alignment, relative volume, and stress.

The manuscript is organized as follows. In Sec. I.1, we briefly describe the force-extension relationship of individual filaments, for which we adopt an existing model of stretchable wormlike chains. Sec. I.2 describes the affine network model, into which the force-extension relationship is fed as an input, and define the associated stress and nematic tensors. In Sec. I.3, we describe the modeled uniaxial strain protocol and the calculation of the strain-dependent stress and filament alignment. In Sec. II, we present and discuss our results.

I Model

I.1 Force-extension relations

We consider a simple mechanical model of extensible wormlike chains (WLCs) [38, 7], with which we can parametrically tune the filament force-extension relationship between the purely enthalpic Hookean spring limit (i.e., a linear force-extension relationship with spring constant μ/c𝜇subscript𝑐\mu/\ell_{c}italic_μ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) and the highly nonlinear limit of inextensible wormlike chains with entropic linear elasticity governed by a bending rigidity κ𝜅\kappaitalic_κ [37, 6]. Between these regimes, the model captures the more generic behavior of extensible wormlike chains, filaments that exhibit entropic linear elasticity governed by a bending rigidity κ𝜅\kappaitalic_κ but, under sufficient extension, cross over to an enthalpic stretching regime governed by the stretch modulus μ𝜇\muitalic_μ.

To obtain the force-extension relationship for extensible wormlike chains, we first need to derive the corresponding relationship for the inextensible limit. Consider an inextensible wormlike chain filament of contour length csubscript𝑐\ell_{c}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and persistence length psubscript𝑝\ell_{p}roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT under applied tension τ𝜏\tauitalic_τ at thermal equilibrium. Thermal fluctuations excite the filament’s bending modes, reducing the ensemble-averaged end-to-end length WLC(τ)subscriptWLC𝜏\ell_{\mathrm{WLC}}(\tau)roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ ) with respect to the contour length [50, 37]. Specifically,

WLC(τ)=cc2π2pn=11n2+τc2/(π2κ)=cκ2τp(cτ/κtanh(cτ/κ)1)subscriptWLC𝜏subscript𝑐superscriptsubscript𝑐2superscript𝜋2subscript𝑝subscriptsuperscript𝑛11superscript𝑛2𝜏superscriptsubscript𝑐2superscript𝜋2𝜅subscript𝑐𝜅2𝜏subscript𝑝subscript𝑐𝜏𝜅subscript𝑐𝜏𝜅1\displaystyle\begin{split}\ell_{\mathrm{WLC}}(\tau)&=\ell_{c}-\frac{\ell_{c}^{% 2}}{\pi^{2}\ell_{p}}\sum^{\infty}_{n=1}{\frac{1}{n^{2}+\tau\ell_{c}^{2}/(\pi^{% 2}\kappa)}}\\ &=\ell_{c}-\frac{\kappa}{2\tau\ell_{p}}\left(\frac{\ell_{c}\sqrt{\tau/\kappa}}% {\tanh{\left(\ell_{c}\sqrt{\tau/\kappa}\right)}}-1\right)\end{split}start_ROW start_CELL roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ ) end_CELL start_CELL = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - divide start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ∑ start_POSTSUPERSCRIPT ∞ end_POSTSUPERSCRIPT start_POSTSUBSCRIPT italic_n = 1 end_POSTSUBSCRIPT divide start_ARG 1 end_ARG start_ARG italic_n start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_τ roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_κ ) end_ARG end_CELL end_ROW start_ROW start_CELL end_CELL start_CELL = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - divide start_ARG italic_κ end_ARG start_ARG 2 italic_τ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( divide start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT square-root start_ARG italic_τ / italic_κ end_ARG end_ARG start_ARG roman_tanh ( roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT square-root start_ARG italic_τ / italic_κ end_ARG ) end_ARG - 1 ) end_CELL end_ROW (1)

in which κ=kBTp𝜅subscript𝑘𝐵𝑇subscript𝑝\kappa=k_{B}T\ell_{p}italic_κ = italic_k start_POSTSUBSCRIPT italic_B end_POSTSUBSCRIPT italic_T roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT is the filament’s bending rigidity and the sum is taken over all mode numbers. Under zero applied tension (τ=0𝜏0\tau=0italic_τ = 0), the ensemble-averaged end-to-end length is 0WLC(τ=0)=cc2/(6p)subscript0subscriptWLC𝜏0subscript𝑐superscriptsubscript𝑐26subscript𝑝\ell_{0}\equiv\ell_{\mathrm{WLC}}(\tau=0)=\ell_{c}-\ell_{c}^{2}/(6\ell_{p})roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≡ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ = 0 ) = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 6 roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ). We can thus write the extension-force relation δWLC(τ)=WLC0𝛿subscriptWLC𝜏subscriptWLCsubscript0\delta\ell_{\mathrm{WLC}}(\tau)=\ell_{\mathrm{WLC}}-\ell_{0}italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ ) = roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT - roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT for an inextensible wormlike chain as

δWLC(τ)=c26p(13κc2τ(cτ/κtanh(cτ/κ)1))𝛿subscriptWLC𝜏superscriptsubscript𝑐26subscript𝑝13𝜅superscriptsubscript𝑐2𝜏subscript𝑐𝜏𝜅subscript𝑐𝜏𝜅1\delta\ell_{\mathrm{WLC}}(\tau)=\frac{\ell_{c}^{2}}{6\ell_{p}}\left(1-\frac{3% \kappa}{\ell_{c}^{2}\tau}\left(\frac{\ell_{c}\sqrt{\tau/\kappa}}{\tanh{\left(% \ell_{c}\sqrt{\tau/\kappa}\right)}}-1\right)\right)italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ ) = divide start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 6 roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG ( 1 - divide start_ARG 3 italic_κ end_ARG start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_τ end_ARG ( divide start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT square-root start_ARG italic_τ / italic_κ end_ARG end_ARG start_ARG roman_tanh ( roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT square-root start_ARG italic_τ / italic_κ end_ARG ) end_ARG - 1 ) ) (2)

The above relation assumes the filaments are inextensible, such that τ𝜏\tauitalic_τ diverges as the thermally contracted filament is stretched to its contour length, δWLCc2/(6p)𝛿subscriptWLCsuperscriptsubscript𝑐26subscript𝑝\delta\ell_{\mathrm{WLC}}\to\ell_{c}^{2}/(6\ell_{p})italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT → roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 6 roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ). Real filaments are of course not entirely inextensible, and with sufficient strain they instead transition from the entropic regime to an enthalpic regime governed by a stretching modulus μ𝜇\muitalic_μ 111For a homogeneous cylindrical elastic rod of radius r𝑟ritalic_r, the stretching modulus μ𝜇\muitalic_μ is related to the bending stiffness κ𝜅\kappaitalic_κ as μ=4κ/r2𝜇4𝜅superscript𝑟2\mu=4\kappa/r^{2}italic_μ = 4 italic_κ / italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [7].. To capture this behavior, we define a more general extension-force relation δ(τ)𝛿𝜏\delta\ell(\tau)italic_δ roman_ℓ ( italic_τ ) for extensible wormlike chains by renormalizing δWLC𝛿subscriptWLC\delta\ell_{\mathrm{WLC}}italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT as follows [38, 7, 3] :

δ(τ)=cτ/μ+δWLC(τ[1+τ/μ])𝛿𝜏subscript𝑐𝜏𝜇𝛿subscriptWLC𝜏delimited-[]1𝜏𝜇\delta\ell(\tau)=\ell_{c}\tau/\mu+\delta\ell_{\mathrm{WLC}}\left(\tau\left[1+% \tau/\mu\right]\right)italic_δ roman_ℓ ( italic_τ ) = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT italic_τ / italic_μ + italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ [ 1 + italic_τ / italic_μ ] ) (3)

Note that taking μ𝜇\mu\to\inftyitalic_μ → ∞ with finite κ𝜅\kappaitalic_κ recovers the inextensible wormlike chain limit, δWLC(τ)=limμδ(τ)𝛿subscriptWLC𝜏subscript𝜇𝛿𝜏\delta\ell_{\mathrm{WLC}}(\tau)=\lim_{\mu\to\infty}\delta\ell(\tau)italic_δ roman_ℓ start_POSTSUBSCRIPT roman_WLC end_POSTSUBSCRIPT ( italic_τ ) = roman_lim start_POSTSUBSCRIPT italic_μ → ∞ end_POSTSUBSCRIPT italic_δ roman_ℓ ( italic_τ ), while taking κ𝜅\kappa\to\inftyitalic_κ → ∞ with finite μ𝜇\muitalic_μ produces a simple linear Hookean relationship.

The force-extension relationship is essentially controlled by two dimensionless numbers: the dimensionless bending rigidity, κ~κ/(μc2)~𝜅𝜅𝜇superscriptsubscript𝑐2\tilde{\kappa}\equiv\kappa/(\mu\ell_{c}^{2})over~ start_ARG italic_κ end_ARG ≡ italic_κ / ( italic_μ roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ), which compares the relative strengths of the bending and stretching resistance of the filaments, and a dimensionless persistence length ~pp/csubscript~𝑝subscript𝑝subscript𝑐\tilde{\ell}_{p}\equiv\ell_{p}/\ell_{c}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≡ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, which compares the filament persistence length psubscript𝑝\ell_{p}roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT to the contour length csubscript𝑐\ell_{c}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the latter of which describes the backbone length of the filament between constraints (crosslinks or entanglements). The dimensionless bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG controls the degree to which the force-extension relation is nonlinear; the Hookean limit corresponds to κ~~𝜅\tilde{\kappa}\to\inftyover~ start_ARG italic_κ end_ARG → ∞, whereas the the inextensible limit corresponds to κ~0~𝜅0\tilde{\kappa}\to 0over~ start_ARG italic_κ end_ARG → 0. The dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT controls the degree to which thermal fluctuations reduce the rest length 0subscript0\ell_{0}roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT of the filament relative to the full contour length csubscript𝑐\ell_{c}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as 0=c(11/(6~p))subscript0subscript𝑐116subscript~𝑝\ell_{0}=\ell_{c}(1-1/(6\tilde{\ell}_{p}))roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ( 1 - 1 / ( 6 over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ). Equivalently, ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT governs the level of extensional strain (applied along the filament end-to-end vector) required to bring the initially contracted filament to its contour length,

εc=c00=16~p1.subscript𝜀𝑐subscript𝑐subscript0subscript016subscript~𝑝1\varepsilon_{c}=\frac{\ell_{c}-\ell_{0}}{\ell_{0}}=\frac{1}{6\tilde{\ell}_{p}-% 1}.italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = divide start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = divide start_ARG 1 end_ARG start_ARG 6 over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - 1 end_ARG . (4)

We will refer to εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as the “critical extension,” at which the resulting tension τ𝜏\tauitalic_τ diverges in the inextensible (κ~0~𝜅0\tilde{\kappa}\to 0over~ start_ARG italic_κ end_ARG → 0) limit or becomes stretching-dominated (μproportional-toabsent𝜇\propto\mu∝ italic_μ) for filaments with finite κ~>0~𝜅0\tilde{\kappa}>0over~ start_ARG italic_κ end_ARG > 0.

Refer to caption
Figure 2: A summary of the asymmetric force-extension relationship of semiflexible polymers: the wormlike chain model. (a) Force-extension relationships for inextensible (μ𝜇\mu\to\inftyitalic_μ → ∞) and extensible (μ=1𝜇1\mu=1italic_μ = 1) wormlike chains with bending rigidity κ=105𝜅superscript105\kappa=10^{-5}italic_κ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT and contour and persistence lengths c=psubscript𝑐subscript𝑝\ell_{c}=\ell_{p}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, under compression (dashed curves, δ<0𝛿0\delta\ell<0italic_δ roman_ℓ < 0) and extension (solid curves, δ>0𝛿0\delta\ell>0italic_δ roman_ℓ > 0), in which δ=0𝛿subscript0\delta\ell=\ell-\ell_{0}italic_δ roman_ℓ = roman_ℓ - roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the difference between the end-to-end distance \ellroman_ℓ and the rest length 0subscript0\ell_{0}roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, defined as the thermally contracted average length under zero tension, 0=cc2/(6p)subscript0subscript𝑐superscriptsubscript𝑐26subscript𝑝\ell_{0}=\ell_{c}-\ell_{c}^{2}/(6\ell_{p})roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 6 roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ). As an inextensible (μ𝜇\mu\to\inftyitalic_μ → ∞) wormlike chain is pulled to its full contour length csubscript𝑐\ell\to\ell_{c}roman_ℓ → roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, as indicated by the vertical dotted line, the tension τ𝜏\tauitalic_τ diverges. The thick solid grey line indicates the linear relationship τ=kδ/c𝜏𝑘𝛿subscript𝑐\tau=k\delta\ell/\ell_{c}italic_τ = italic_k italic_δ roman_ℓ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with k=1/(μ1+c3/(90κp))𝑘1superscript𝜇1superscriptsubscript𝑐390𝜅subscript𝑝k=1/(\mu^{-1}+\ell_{c}^{3}/(90\kappa\ell_{p}))italic_k = 1 / ( italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / ( 90 italic_κ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ). The horizontal dotted line indicates the Euler buckling force, τe=κπ2/c2subscript𝜏e𝜅superscript𝜋2superscriptsubscript𝑐2\tau_{\mathrm{e}}=-\kappa\pi^{2}/\ell_{c}^{2}italic_τ start_POSTSUBSCRIPT roman_e end_POSTSUBSCRIPT = - italic_κ italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. (b) Chain stiffness τ/𝜏\partial\tau/\partial\ell∂ italic_τ / ∂ roman_ℓ as a function of applied tension τ𝜏\tauitalic_τ. For small tensions, τ/k/c𝜏𝑘subscript𝑐\partial\tau/\partial\ell\to k/\ell_{c}∂ italic_τ / ∂ roman_ℓ → italic_k / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In the inextensible limit (μ𝜇\mu\to\inftyitalic_μ → ∞), the stiffness scales as τ/τ3/2proportional-to𝜏superscript𝜏32\partial\tau/\partial\ell\propto\tau^{3/2}∂ italic_τ / ∂ roman_ℓ ∝ italic_τ start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT for large τ𝜏\tauitalic_τ. For finite μ𝜇\muitalic_μ, one finds τ/μ/c𝜏𝜇subscript𝑐\partial\tau/\partial\ell\to\mu/\ell_{c}∂ italic_τ / ∂ roman_ℓ → italic_μ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT for large τ𝜏\tauitalic_τ.

Having defined the extension-force relationship δ(τ)𝛿𝜏\delta\ell(\tau)italic_δ roman_ℓ ( italic_τ ), we can obtain the corresponding force-extension relationship τ(δ)𝜏𝛿\tau(\delta\ell)italic_τ ( italic_δ roman_ℓ ) by numerical inversion [7]. In Fig. 2a, we plot the absolute value of the tension |τ|𝜏|\tau|| italic_τ | as a function of the normalized elongation |δ/c|𝛿subscript𝑐|\delta\ell/\ell_{c}|| italic_δ roman_ℓ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT | for extensible (μ=1𝜇1\mu=1italic_μ = 1) and inextensible (μ𝜇\mu\to\inftyitalic_μ → ∞) filaments under both extension and compression, with p=c=1subscript𝑝subscript𝑐1\ell_{p}=\ell_{c}=1roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1 (~p=1subscript~𝑝1\tilde{\ell}_{p}=1over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 1) and κ=105𝜅superscript105\kappa=10^{-5}italic_κ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT. For small δ/c𝛿subscript𝑐\delta\ell/\ell_{c}italic_δ roman_ℓ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, the force-extension relationship exhibits a simple linear dependence τ=kδ/c𝜏𝑘𝛿subscript𝑐\tau=k\delta\ell/\ell_{c}italic_τ = italic_k italic_δ roman_ℓ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, in which the effective stiffness k𝑘kitalic_k is given by k=1/(μ1+c3/(90κp))𝑘1superscript𝜇1superscriptsubscript𝑐390𝜅subscript𝑝k=1/(\mu^{-1}+\ell_{c}^{3}/(90\kappa\ell_{p}))italic_k = 1 / ( italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / ( 90 italic_κ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ). Note that for the Hookean limit, this yields k=μ𝑘𝜇k=\muitalic_k = italic_μ, and for the inextensible WLC limit, k=90κp/c3𝑘90𝜅subscript𝑝superscriptsubscript𝑐3k=90\kappa\ell_{p}/\ell_{c}^{3}italic_k = 90 italic_κ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT [37, 50]. As κ=105𝜅superscript105\kappa=10^{-5}italic_κ = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT for the plotted curves, the values of k𝑘kitalic_k for the inextensible and extensible cases are virtually identical. Under extension (δ>0𝛿0\delta\ell>0italic_δ roman_ℓ > 0), the tension τ𝜏\tauitalic_τ stiffens dramatically as the filament length \ellroman_ℓ approaches and, in the extensible case, exceeds the contour length csubscript𝑐\ell_{c}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. For μ𝜇\mu\to\inftyitalic_μ → ∞, τ𝜏\tauitalic_τ diverges as csubscript𝑐\ell\to\ell_{c}roman_ℓ → roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, whereas for μ=1𝜇1\mu=1italic_μ = 1, the filament crosses over to a regime in which the tension is dominated by the stretch modulus, with τμproportional-to𝜏𝜇\tau\propto\muitalic_τ ∝ italic_μ. In Fig. 2b, we plot the filament stiffness τ/𝜏\partial\tau/\partial\ell∂ italic_τ / ∂ roman_ℓ as a function of the tension τ𝜏\tauitalic_τ. In the limit of small τ𝜏\tauitalic_τ, the stiffness for both extensible and inextensible filaments is given by τ/=k/c𝜏𝑘subscript𝑐\partial\tau/\partial\ell=k/\ell_{c}∂ italic_τ / ∂ roman_ℓ = italic_k / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, whereas in the limit of large τ𝜏\tauitalic_τ, the stiffness of extensible filaments reaches an upper limit of τ/=μ/c𝜏𝜇subscript𝑐\partial\tau/\partial\ell=\mu/\ell_{c}∂ italic_τ / ∂ roman_ℓ = italic_μ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, while the stiffness of inextensible filaments exhibits a power law dependence on the tension with τ/τ3/2proportional-to𝜏superscript𝜏32\partial\tau/\partial\ell\propto\tau^{3/2}∂ italic_τ / ∂ roman_ℓ ∝ italic_τ start_POSTSUPERSCRIPT 3 / 2 end_POSTSUPERSCRIPT [52, 53]. Under compression (δ<0𝛿0\delta\ell<0italic_δ roman_ℓ < 0), τ𝜏\tauitalic_τ approaches a limiting value of τ=τe𝜏subscript𝜏𝑒\tau=\tau_{e}italic_τ = italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT (in the small κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG limit), in which τe=κπ2/c2subscript𝜏𝑒𝜅superscript𝜋2superscriptsubscript𝑐2\tau_{e}=-\kappa\pi^{2}/\ell_{c}^{2}italic_τ start_POSTSUBSCRIPT italic_e end_POSTSUBSCRIPT = - italic_κ italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is the critical compressive force for Euler buckling of a rod of length csubscript𝑐\ell_{c}roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT and bending rigidity κ𝜅\kappaitalic_κ [3].

I.2 Affine network model

We model the network as a collection of independent filament segments with random (isotropically distributed) initial orientations. The positions of the filament endpoints are assumed to transform affinely according to the macroscopically applied strain, and the resulting tension is assumed to act along the transformed end-to-end vector. The network-level stress is then obtained by averaging over all filament orientations in the deformed configuration. This model dates back to early work on rubber [54, 39, 40, 41, 55, 42, 43, 44] and has been successfully used to describe the mechanics of semiflexible networks in a variety of contexts [37, 56, 6, 7, 57, 58, 59, 46, 60, 61, 62, 47, 48].

In this model, each filament segment is treated as a central-force elastic element of initial end-to-end distance 0subscript0\ell_{0}roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT with initial orientation 𝐧^0subscript^𝐧0\hat{\mathbf{n}}_{0}over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT that obeys a force-extension relation τ(δ)𝜏𝛿\tau(\delta\ell)italic_τ ( italic_δ roman_ℓ ), with δ=0𝛿subscript0\delta\ell=\ell-\ell_{0}italic_δ roman_ℓ = roman_ℓ - roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT denoting the change in the deformed filament end-to-end distance \ellroman_ℓ with respect to the initial distance 0subscript0\ell_{0}roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In the unstrained initial configuration, δ=0𝛿0\delta\ell=0italic_δ roman_ℓ = 0 and τ=0𝜏0\tau=0italic_τ = 0. The deformation gradient tensor 𝚲𝚲\bm{\mathrm{\Lambda}}bold_Λ transforms the initial filament end-to-end vector 𝐫0=0𝐧^0subscript𝐫0subscript0subscript^𝐧0\mathbf{r}_{0}=\ell_{0}\hat{\mathbf{n}}_{0}bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT into the deformed end-to-end vector 𝐫=𝚲𝐫0=𝐧^𝐫𝚲subscript𝐫0^𝐧\mathbf{r}=\bm{\Lambda}\mathbf{r}_{0}=\ell\hat{\mathbf{n}}bold_r = bold_Λ bold_r start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℓ over^ start_ARG bold_n end_ARG with a new orientation 𝐧^=𝚲𝐧^0/|𝚲𝐧^0|^𝐧𝚲subscript^𝐧0𝚲subscript^𝐧0\hat{\mathbf{n}}=\bm{\Lambda}\hat{\mathbf{n}}_{0}/|\bm{\Lambda}\hat{\mathbf{n}% }_{0}|over^ start_ARG bold_n end_ARG = bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | and end-to-end length =0|𝚲𝐧^0|subscript0𝚲subscript^𝐧0\ell=\ell_{0}|\bm{\Lambda}\hat{\mathbf{n}}_{0}|roman_ℓ = roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT |, such that the change in length is δ=0(|𝚲𝐧^0|1)𝛿subscript0𝚲subscript^𝐧01\delta\ell=\ell_{0}(|\mathbf{\Lambda}\hat{\mathbf{n}}_{0}|-1)italic_δ roman_ℓ = roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | - 1 ) with corresponding tension 𝝉(δ)=τ(δ)𝐧^𝝉𝛿𝜏𝛿^𝐧\bm{\tau}(\delta\ell)=\tau(\delta\ell)\hat{\mathbf{n}}bold_italic_τ ( italic_δ roman_ℓ ) = italic_τ ( italic_δ roman_ℓ ) over^ start_ARG bold_n end_ARG. We assume that the initial orientations 𝐧^0subscript^𝐧0\hat{\mathbf{n}}_{0}over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT are isotropically distributed and define the initial length density ρ0=0/V0subscript𝜌0subscript0subscript𝑉0\rho_{0}=\ell_{0}/V_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, in which V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial volume per filament segment. The length density in the deformed configuration becomes ρ/V=ρ0|𝚲𝐧^0|/det𝚲𝜌delimited-⟨⟩𝑉subscript𝜌0delimited-⟨⟩𝚲subscript^𝐧0𝚲\rho\equiv\langle\ell\rangle/V=\rho_{0}\langle|\bm{\Lambda}\hat{\mathbf{n}}_{0% }|\rangle/\det\bm{\Lambda}italic_ρ ≡ ⟨ roman_ℓ ⟩ / italic_V = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ⟨ | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | ⟩ / roman_det bold_Λ. Averaging over all filaments, we compute the Cauchy stress tensor 𝝈=V1𝝉(δ)𝐫=ρ𝝉(δ)𝐧^𝝈superscript𝑉1delimited-⟨⟩tensor-product𝝉𝛿𝐫𝜌delimited-⟨⟩tensor-product𝝉𝛿^𝐧\bm{\sigma}=V^{-1}\langle\bm{\tau}(\delta\ell)\otimes\mathbf{r}\rangle=\rho% \langle\bm{\tau}(\delta\ell)\otimes\hat{\mathbf{n}}\ranglebold_italic_σ = italic_V start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ⟨ bold_italic_τ ( italic_δ roman_ℓ ) ⊗ bold_r ⟩ = italic_ρ ⟨ bold_italic_τ ( italic_δ roman_ℓ ) ⊗ over^ start_ARG bold_n end_ARG ⟩ [63], or equivalently [64, 65, 7, 3]

𝝈=ρ0det𝚲𝝉(δ)(𝚲𝐧^0),𝝈subscript𝜌0det𝚲delimited-⟨⟩tensor-product𝝉𝛿𝚲subscript^𝐧0\bm{\sigma}=\frac{\rho_{0}}{\rm{det}\mathbf{\Lambda}}\big{\langle}\bm{\tau}(% \delta\ell)\otimes\left(\bm{\Lambda}\hat{\mathbf{n}}_{0}\right)\big{\rangle},bold_italic_σ = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_det bold_Λ end_ARG ⟨ bold_italic_τ ( italic_δ roman_ℓ ) ⊗ ( bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) ⟩ , (5)

and the first Piola-Kirchhoff stress tensor 𝐏𝐏\mathbf{P}bold_P as

𝐏=det𝚲𝝈T(𝚲T)1,𝐏𝚲superscript𝝈Tsuperscriptsuperscript𝚲T1\mathbf{P}=\det\mathbf{\Lambda}\bm{\sigma}^{\mathrm{T}}(\mathbf{\Lambda}^{% \mathrm{T}})^{-1},bold_P = roman_det bold_Λ bold_italic_σ start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT ( bold_Λ start_POSTSUPERSCRIPT roman_T end_POSTSUPERSCRIPT ) start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT , (6)

the latter of which can be interpreted as the force in the deformed configuration per unit area in the undeformed configuration [46]. Note that 𝐏=𝝈𝐏𝝈\mathbf{P}=\bm{\sigma}bold_P = bold_italic_σ in the small strain limit.

To quantify the alignment of filaments as a function of strain, we compute the symmetric and traceless nematic tensor 𝐐𝐐\bm{\mathrm{Q}}bold_Q [66, 67] as

𝐐𝐧^𝐧^1d𝐈𝐐delimited-⟨⟩tensor-product^𝐧^𝐧1𝑑𝐈\bm{\mathrm{Q}}\equiv\bigg{\langle}\hat{\mathbf{n}}\otimes\hat{\mathbf{n}}-% \frac{1}{d}\bm{\mathrm{I}}\bigg{\rangle}bold_Q ≡ ⟨ over^ start_ARG bold_n end_ARG ⊗ over^ start_ARG bold_n end_ARG - divide start_ARG 1 end_ARG start_ARG italic_d end_ARG bold_I ⟩ (7)

in which d=3𝑑3d=3italic_d = 3 is the dimensionality and 𝐈𝐈\bm{\mathrm{I}}bold_I is the identity tensor. From the dominant eigenvalue λ𝜆\lambdaitalic_λ of 𝐐𝐐\bm{\mathrm{Q}}bold_Q, we obtain the nematic order parameter S=(d/(d1))λ𝑆𝑑𝑑1𝜆S=(d/(d-1))\lambdaitalic_S = ( italic_d / ( italic_d - 1 ) ) italic_λ, which quantifies the filament alignment [66].

In the linear (small strain) regime, the nematic tensor 𝐐𝐐\mathbf{Q}bold_Q and Cauchy stress tensor 𝝈𝝈\bm{\sigma}bold_italic_σ are related by the stress-optical law [68, 69],

𝐐=C(𝝈p𝐈)𝐐𝐶𝝈𝑝𝐈\mathbf{Q}=C\left(\bm{\sigma}-p\mathbf{I}\right)bold_Q = italic_C ( bold_italic_σ - italic_p bold_I ) (8)

in which C𝐶Citalic_C is a material-dependent proportionality coefficient and p=𝝈:𝐈:𝑝𝝈𝐈p=\bm{\sigma}:\mathbf{I}italic_p = bold_italic_σ : bold_I.

Refer to caption
Figure 3: Response of initially isotropic networks of extensible wormlike chains with varying dimensionless bending rigidity κ~=κ/(μc2)~𝜅𝜅𝜇superscriptsubscript𝑐2\tilde{\kappa}=\kappa/(\mu\ell_{c}^{2})over~ start_ARG italic_κ end_ARG = italic_κ / ( italic_μ roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) to applied uniaxial strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT under the condition of zero transverse stress (σ=0subscript𝜎perpendicular-to0\sigma_{\perp}=0italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0). Curves are also plotted for filaments in the Hookean limit (κ~~𝜅\tilde{\kappa}\to\inftyover~ start_ARG italic_κ end_ARG → ∞ with finite μ𝜇\muitalic_μ) and the inextensible WLC limit ( μ𝜇\mu\to\inftyitalic_μ → ∞ with finite κ𝜅\kappaitalic_κ, such that κ~0~𝜅0\tilde{\kappa}\to 0over~ start_ARG italic_κ end_ARG → 0). All curves correspond to c=p=1subscript𝑐subscript𝑝1\ell_{c}=\ell_{p}=1roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = 1. (a) Transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT vs. applied longitudinal strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. In the limit of large κ𝜅\kappaitalic_κ, extensible WLC networks behave as Hookean spring networks, satisfying ε=νεsubscript𝜀perpendicular-to𝜈subscript𝜀parallel-to\varepsilon_{\perp}=-\nu\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - italic_ν italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT up to relatively large strains. With decreasing κ𝜅\kappaitalic_κ, extensible wormlike chain networks exhibit increasingly rapid transverse contraction in the vicinity of the critical extensile strain εc1/(6~p1)subscript𝜀𝑐16subscript~𝑝1\varepsilon_{c}\equiv 1/(6\tilde{\ell}_{p}-1)italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ≡ 1 / ( 6 over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT - 1 ). In the inextensible WLC limit (κ0𝜅0\kappa\to 0italic_κ → 0), the transverse collapse is complete (ε1)\varepsilon_{\perp}\to-1)italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT → - 1 ) at ε=εcsubscript𝜀parallel-tosubscript𝜀𝑐\varepsilon_{\parallel}=\varepsilon_{c}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. (b) The differential Poisson’s ratio ν~=ε/ε~𝜈subscript𝜀perpendicular-tosubscript𝜀parallel-to\tilde{\nu}=-\partial\varepsilon_{\perp}/\partial\varepsilon_{\parallel}over~ start_ARG italic_ν end_ARG = - ∂ italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / ∂ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT exhibits a peak that diverges in the limit of κ0𝜅0\kappa\to 0italic_κ → 0 (or μ𝜇\mu\to\inftyitalic_μ → ∞) at the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. For small strains, ν~=ν=1/4~𝜈𝜈14\tilde{\nu}=\nu=1/4over~ start_ARG italic_ν end_ARG = italic_ν = 1 / 4 for all models considered, as is expected for a Cauchy solid [18] (c) The relative volume V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT initially increases with applied extension, before rapidly vanishing at the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in the κ0𝜅0\kappa\to 0italic_κ → 0 (μ)\mu\to\infty)italic_μ → ∞ ) limit. (d) The nematic alignment S𝑆Sitalic_S increases with applied extension before rapidly approaching 1111 (perfect alignment) at the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in the κ0𝜅0\kappa\to 0italic_κ → 0 (μ𝜇\mu\to\inftyitalic_μ → ∞) limit.

I.3 Uniaxial strain protocol

Experimentally, Poisson’s ratio can be obtained by stretching a material along one axis and measuring the resulting transverse strain [18]. To characterize the Poisson effect using the affine model, we apply a uniaxial strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT along the z𝑧zitalic_z axis and solve for the transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT that satisfies a condition of zero stress along the transverse axes. The corresponding deformation gradient tensor 𝚲(ε,ε)𝚲subscript𝜀parallel-tosubscript𝜀perpendicular-to\bm{\Lambda}(\varepsilon_{\parallel},\varepsilon_{\perp})bold_Λ ( italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) is

𝚲=(1+ε0001+ε0001+ε).𝚲matrix1subscript𝜀perpendicular-to0001subscript𝜀perpendicular-to0001subscript𝜀parallel-to\bm{\Lambda}=\begin{pmatrix}1+\varepsilon_{\perp}&0&0\\ 0&1+\varepsilon_{\perp}&0\\ 0&0&1+\varepsilon_{\parallel}\end{pmatrix}.bold_Λ = ( start_ARG start_ROW start_CELL 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_CELL start_CELL 0 end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) . (9)

As the cross-sectional area varies considerably as a function of strain for these systems, the strain dependence of the first Piola-Kirchhoff stress tensor 𝐏𝐏\mathbf{P}bold_P (also called the engineering stress), given by Eq. 6, is more informative for our purposes than the Cauchy stress tensor 𝝈𝝈\bm{\sigma}bold_italic_σ (also called the true stress). This is because the Cauchy stress diverges as the cross-sectional area vanishes, even if the total tension remains finite. Note that Pij=0subscript𝑃𝑖𝑗0P_{ij}=0italic_P start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 0 if σij=0subscript𝜎𝑖𝑗0\sigma_{ij}=0italic_σ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = 0. According to Eqs. 6 and 9, the relevant components of 𝐏𝐏\mathbf{P}bold_P are given by

P=(1+ε)2σsubscript𝑃parallel-tosuperscript1subscript𝜀perpendicular-to2subscript𝜎parallel-toP_{\parallel}=(1+\varepsilon_{\perp})^{2}\sigma_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_σ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT (10)

and

P=(1+ε)(1+ε)σsubscript𝑃perpendicular-to1subscript𝜀perpendicular-to1subscript𝜀parallel-tosubscript𝜎perpendicular-toP_{\perp}=(1+\varepsilon_{\perp})(1+\varepsilon_{\parallel})\sigma_{\perp}italic_P start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT (11)

in which the components of 𝝈𝝈\bm{\sigma}bold_italic_σ are calculated using Eq. 5. Full details of the stress calculation are given in Appendix A. For a given applied εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, we numerically solve Eq. 18 for ε(ε)subscript𝜀perpendicular-tosubscript𝜀parallel-to\varepsilon_{\perp}(\varepsilon_{\parallel})italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) with P(ε,ε)=0subscript𝑃perpendicular-tosubscript𝜀parallel-tosubscript𝜀perpendicular-to0P_{\perp}(\varepsilon_{\parallel},\varepsilon_{\perp})=0italic_P start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ( italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) = 0. For small strains, PE0εsubscript𝑃parallel-tosubscript𝐸0subscript𝜀parallel-toP_{\parallel}\approx E_{0}\varepsilon_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, in which we have defined the linear Young’s modulus E0=16ρ0k0/csubscript𝐸016subscript𝜌0𝑘subscript0subscript𝑐E_{0}=\frac{1}{6}\rho_{0}k\ell_{0}/\ell_{c}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 6 end_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_k roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (see Appendix A).

We can then calculate the differential Poisson’s ratio ν~~𝜈\tilde{\nu}over~ start_ARG italic_ν end_ARG, defined as

ν~=εε~𝜈subscript𝜀perpendicular-tosubscript𝜀parallel-to\tilde{\nu}=\frac{\partial\varepsilon_{\perp}}{\partial\varepsilon_{\parallel}}over~ start_ARG italic_ν end_ARG = divide start_ARG ∂ italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT end_ARG (12)

and the relative volume,

VV0=ρ0ρ=det𝚲.𝑉subscript𝑉0subscript𝜌0𝜌𝚲\frac{V}{V_{0}}=\frac{\rho_{0}}{\rho}=\det\bm{\Lambda}.divide start_ARG italic_V end_ARG start_ARG italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG italic_ρ end_ARG = roman_det bold_Λ . (13)

In the limit of small strains, limε0+ν~=ν=1/4subscript𝜀superscript0~𝜈𝜈14\lim_{\varepsilon\to 0^{+}}\tilde{\nu}=\nu=1/4roman_lim start_POSTSUBSCRIPT italic_ε → 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT over~ start_ARG italic_ν end_ARG = italic_ν = 1 / 4 for this model and VV0(1+ε/2)𝑉subscript𝑉01subscript𝜀parallel-to2V\approx V_{0}(1+\varepsilon_{\parallel}/2)italic_V ≈ italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2 ).

We calculate the strain-dependent nematic alignment S𝑆Sitalic_S as [66]

S=32cos2θ13𝑆32delimited-⟨⟩superscript2𝜃13S=\frac{3}{2}\bigg{\langle}\cos^{2}\theta-\frac{1}{3}\bigg{\rangle}italic_S = divide start_ARG 3 end_ARG start_ARG 2 end_ARG ⟨ roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ - divide start_ARG 1 end_ARG start_ARG 3 end_ARG ⟩ (14)

in which θ𝜃\thetaitalic_θ is the angle between the unit orientation vector of a given filament 𝐧^^𝐧\hat{\mathbf{n}}over^ start_ARG bold_n end_ARG and the axis of applied strain (𝐳^^𝐳\hat{\mathbf{z}}over^ start_ARG bold_z end_ARG) and the average is taken over all filaments. For small strains, Sε/2𝑆subscript𝜀parallel-to2S\approx\varepsilon_{\parallel}/2italic_S ≈ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2. See Appendix B for additional details.

Since PE0εsubscript𝑃parallel-tosubscript𝐸0subscript𝜀parallel-toP_{\parallel}\approx E_{0}\varepsilon_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT and Sε/2𝑆subscript𝜀parallel-to2S\approx\varepsilon_{\parallel}/2italic_S ≈ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2 for small strains, the stress in this regime can be written as a function of the alignment, P2E0Ssubscript𝑃parallel-to2subscript𝐸0𝑆P_{\parallel}\approx 2E_{0}Sitalic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_S. Consequently, the proportionality coefficient C𝐶Citalic_C in the stress-optical law (Eq. 8) is given by C=1/(2E0)𝐶12subscript𝐸0C=1/(2E_{0})italic_C = 1 / ( 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

Refer to caption
Figure 4: Response of initially isotropic networks with varying dimensionless persistence length ~pp/csubscript~𝑝subscript𝑝subscript𝑐\tilde{\ell}_{p}\equiv\ell_{p}/\ell_{c}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ≡ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT to applied uniaxial strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT under the condition of zero transverse stress (σ=0subscript𝜎perpendicular-to0\sigma_{\perp}=0italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0). For the plotted curves, the dimensionless bending rigidity is κ~=105~𝜅superscript105\tilde{\kappa}=10^{-5}over~ start_ARG italic_κ end_ARG = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT except in the Hookean case(κ~~𝜅\tilde{\kappa}\to\inftyover~ start_ARG italic_κ end_ARG → ∞ with μ=1𝜇1\mu=1italic_μ = 1). Increasing the dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT leads to a decrease in the critical extensile strain εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, i.e. alignment and transverse contraction at lower values of applied uniaxial extensile strain. (a) Transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT vs. applied uniaxial strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. (b) Differential Poisson’s ratio ν~~𝜈\tilde{\nu}over~ start_ARG italic_ν end_ARG. (c) Relative volume V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. (d) Nematic alignment S𝑆Sitalic_S.

II Results and discussion

Using the affine model, we now explore the effects of varying filament properties on the response of a network to uniaxial strain. Specifically, we will consider the effects of independently varying the dimensionless parameters defined in Sec. I.1: the dimensionless bending rigidity κ~κ/(μc2)~𝜅𝜅𝜇superscriptsubscript𝑐2\tilde{\kappa}\equiv\kappa/(\mu\ell_{c}^{2})over~ start_ARG italic_κ end_ARG ≡ italic_κ / ( italic_μ roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) and dimensionless persistence length p~=p/c~subscript𝑝subscript𝑝subscript𝑐\tilde{\ell_{p}}=\ell_{p}/\ell_{c}over~ start_ARG roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG = roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. Note that independently varying these quantities in experiments may be challenging, as both may depend on one or more of the same experimental variables, e.g. the polymer concentration [37]. Our aim here is simply to provide an intuitive feel for how each of these dimensionless quantities influences the network-level response.

First, we consider the effects of varying the dimensionless bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG with a fixed value of the dimensionless persistence length p~=1~subscript𝑝1\tilde{\ell_{p}}=1over~ start_ARG roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG = 1. In Fig. 3a, we plot the transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT as a function of applied longitudinal strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT for networks of extensible filaments over a wide range of κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG values, in addition to the Hookean spring limit (κ𝜅\kappa\to\inftyitalic_κ → ∞ with finite μ𝜇\muitalic_μ) and the inextensible WLC limit (μ𝜇\mu\to\inftyitalic_μ → ∞ with finite κ𝜅\kappaitalic_κ). For small strains, all exhibit a linear regime in which ε=νεsubscript𝜀perpendicular-to𝜈subscript𝜀parallel-to\varepsilon_{\perp}=-\nu\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - italic_ν italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, where the linear Poisson’s ratio ν=1/4𝜈14\nu=1/4italic_ν = 1 / 4 is the expected value for a Cauchy solid [18, 70]. We find that as κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG increases, the behavior of the finite-κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG networks approaches that of simple Hookean spring networks, exhibiting an approximately linear dependence of transverse strain εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT on applied strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ( ενεsubscript𝜀perpendicular-to𝜈subscript𝜀parallel-to\varepsilon_{\perp}\approx-\nu\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ≈ - italic_ν italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT) over a large range of εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. In contrast, as κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG is reduced, the behavior of the finite-κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG networks approaches that of inextensible WLC networks, exhibiting increasingly dramatic transverse contraction under finite values of applied strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. Furthermore, we find that the value of applied strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT corresponding to the inflection point in the εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT vs. εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT curve decreases with decreasing κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG. In the low-κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG limit, this inflection point approaches the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (Eq. 4) corresponding to the applied extensional strain at which the end-to-end length of the most highly stretched filaments in the network (those oriented along the applied strain axis) approaches the filament contour length, δc2/(6p)𝛿superscriptsubscript𝑐26subscript𝑝\delta\ell\to\ell_{c}^{2}/(6\ell_{p})italic_δ roman_ℓ → roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / ( 6 roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ). For inextensible WLC networks, complete transverse collapse occurs at εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

This behavior is also reflected in the differential Poisson’s ratio ν~=ε/ε~𝜈subscript𝜀perpendicular-tosubscript𝜀parallel-to\tilde{\nu}=-\partial\varepsilon_{\perp}/\partial\varepsilon_{\parallel}over~ start_ARG italic_ν end_ARG = - ∂ italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT / ∂ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT (Eq. 12), as we show in Fig. 3b. The location of the inflection point in the εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT vs. εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT curve corresponds to a peak in the differential Poisson’s ratio ν~~𝜈\tilde{\nu}over~ start_ARG italic_ν end_ARG. This peak increases in magnitude and shifts to lower values of εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT with decreasing κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG, approaching εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT from above in the κ~0~𝜅0\tilde{\kappa}\to 0over~ start_ARG italic_κ end_ARG → 0 limit. The rapid transverse collapse that occurs at εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT in the low-κ𝜅\kappaitalic_κ limit corresponds to a diverging differential Poisson’s ratio ν~~𝜈\tilde{\nu}over~ start_ARG italic_ν end_ARG. We can understand this behavior with a simple qualitative argument. Under finite applied strain, maintaining the zero transverse stress condition σ=0subscript𝜎perpendicular-to0\sigma_{\perp}=0italic_σ start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = 0 requires a balance, in the transverse plane, between outward forces generated by compressed filaments and inward forces generated by stretched filaments. For networks of extensible wormlike chains (finite κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG), increasing the applied extensional strain beyond the critical strain (ε>εcsubscript𝜀parallel-tosubscript𝜀𝑐\varepsilon_{\parallel}>\varepsilon_{c}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT > italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT) causes an increasing fraction of the stretched filaments to enter the enthalpic stretching regime (τμsimilar-to𝜏𝜇\tau\sim\muitalic_τ ∼ italic_μ), such that the inward transverse component of the resulting tension becomes proportional to μ𝜇\muitalic_μ. However, the outward forces generated by compressed filaments remain proportional to κ𝜅\kappaitalic_κ even in the large compression limit. For κμmuch-less-than𝜅𝜇\kappa\ll\muitalic_κ ≪ italic_μ, achieving force balance thus requires significant transverse contraction as the elongated filaments enter the stretching-dominated regime. In the inextensible WLC limit (μ𝜇\mu\to\inftyitalic_μ → ∞) such a balance is impossible, as τ𝜏\tau\to\inftyitalic_τ → ∞ as each filament approaches its contour length, so the network collapses completely (ε1subscript𝜀perpendicular-to1\varepsilon_{\perp}\to-1italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT → - 1) precisely at the critical extensional strain εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT.

An important consequence of the nonlinear Poisson effect is a reduction in the volume occupied by the network as it stiffens. Having determined εsubscript𝜀perpendicular-to\varepsilon_{\perp}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT as a function of εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, we compute the relative volume V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT given by Eq. 13, i.e. the ratio of the volume V𝑉Vitalic_V occupied by a portion of the strained network to its initial volume V0subscript𝑉0V_{0}italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. In Fig. 3c, we plot V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT as a function of applied extension εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT for the same systems as in the upper two panels. In the linear regime, V/V01+ε/2𝑉subscript𝑉01subscript𝜀parallel-to2V/V_{0}\approx 1+\varepsilon_{\parallel}/2italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≈ 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2. For a network of Hookean springs (κ~~𝜅\tilde{\kappa}\to\inftyover~ start_ARG italic_κ end_ARG → ∞ with finite μ𝜇\muitalic_μ), the relative volume V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT decreases monotonically with compression and increases monotonically with extension over a large range of εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. For networks with finite or zero κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG under applied compression, V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT likewise decreases and exhibits only a weak dependence on κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG. Under applied extension, however, the relative volume initially increases in the linear regime before decreasing dramatically in the vicinity of the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT as ε1subscript𝜀perpendicular-to1\varepsilon_{\perp}\to-1italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT → - 1. Note that, because our model does not account for steric interaction between filaments, the relative volume vanishes completely in the case of highly nonlinear filaments under large extension. In reality, if we assume the filaments are negligibly compressible, then V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT should not decrease below ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, the initial volume fraction of filaments.

The nonlinear Poisson effect is also characterized by the onset of filament alignment along the strain axis [14], quantified by the nematic alignment S𝑆Sitalic_S given by Eq. 14. In the linear regime, S𝑆Sitalic_S increases with applied extensional strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT as Sε/2𝑆subscript𝜀parallel-to2S\approx\varepsilon_{\parallel}/2italic_S ≈ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2. In the Hookean spring network limit (κ~~𝜅\tilde{\kappa}\to\inftyover~ start_ARG italic_κ end_ARG → ∞ with finite μ𝜇\muitalic_μ), the linear regime extends over a large range of extensile and compressive strain. For wormlike chain filaments with small or zero κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG, identical behavior is seen in the linear regime, whereas we see rapidly increasing alignment in the vicinity of the critical extensional strain εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. In the κ~0~𝜅0\tilde{\kappa}\to 0over~ start_ARG italic_κ end_ARG → 0 limit (inextensible WLCs), the network aligns completely (S1𝑆1S\to 1italic_S → 1) as εεcsubscript𝜀parallel-tosubscript𝜀𝑐\varepsilon_{\parallel}\to\varepsilon_{c}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT → italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT. This enhanced alignment is a direct consequence of the enhanced transverse contraction that occurs near εεcsimilar-tosubscript𝜀parallel-tosubscript𝜀𝑐\varepsilon_{\parallel}\sim\varepsilon_{c}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ∼ italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT due to the asymmetric, nonlinear force-extension relationship of the individual filaments. Notably, significant alignment that coincides with stiffening and the nonlinear Poisson effect has been observed experimentally in uniaxially stretched fibrin [14] and collagen gels [16]. Interestingly, under compression, we find that S𝑆Sitalic_S depends very little on κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG or, as we will soon see, ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT.

We now consider the effects of varying the dimensionless persistence length ~p=p/csubscript~𝑝subscript𝑝subscript𝑐\tilde{\ell}_{p}=\ell_{p}/\ell_{c}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT = roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT while the dimensionless bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG remains fixed. Recall, the critical extensional strain εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT (Eq. 4) decreases as the dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT increases. Thus, increasing p~~subscript𝑝\tilde{\ell_{p}}over~ start_ARG roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT end_ARG should cause the rapid transverse contraction, peak in the Poisson’s ratio, and alignment to occur at lower values of applied extensional strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. This is precisely what is shown in Fig. 4, in which we plot the same quantities as in Fig. 3 for networks with fixed dimensionless bending rigidity (κ~=105~𝜅superscript105\tilde{\kappa}=10^{-5}over~ start_ARG italic_κ end_ARG = 10 start_POSTSUPERSCRIPT - 5 end_POSTSUPERSCRIPT) and varying ~p[0.5,2]subscript~𝑝0.52\tilde{\ell}_{p}\in[0.5,2]over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ∈ [ 0.5 , 2 ]. Aside from the shift in εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT with varying ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, the qualitative features (e.g. dramatic alignment and a large peak in the differential Poisson’s ratio) remain unchanged.

We also examine the utility of the nematic alignment S𝑆Sitalic_S as an indicator of the stress Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. In the linear regime, the two quantities are directly proportional, with P2E0Ssubscript𝑃parallel-to2subscript𝐸0𝑆P_{\parallel}\approx 2E_{0}Sitalic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_S as we outline in Appendix B. Thus, if the linear Young’s modulus E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (Eq. 20) is known and the applied strain is sufficiently small, one can in principle use optical measurements of the nematic alignment S𝑆Sitalic_S as an indirect measure of the stress Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT. This is simply another way of stating the “stress optical law” [68, 69] (Eq. 8), which relates the nematic and stress tensors for simple polymeric systems under small strains. For our systems, the proportionality coefficient in Eq. 8 is given by C=1/(2E0)𝐶12subscript𝐸0C=1/(2E_{0})italic_C = 1 / ( 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ). As is shown in Fig. 5, beyond the linear regime, this relationship no longer remains valid, and the deviation becomes more significant as κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG decreases. Nevertheless, if E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG, and ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT are known, one could in principle use a plot like Fig. 5 to estimate local values of the Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT from S𝑆Sitalic_S even beyond the linear regime.

Refer to caption
Figure 5: Filament alignment S𝑆Sitalic_S acts as a sensitive measure of the stress Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, and the stress-alignment relationship becomes increasingly nonlinear with decreasing bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG. Here, we plot the longitudinal component of the first Piola-Kirchhoff stress, Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT normalized by the linear Young’s modulus E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, as a function of the nematic alignment S𝑆Sitalic_S. The dashed line corresponds to the linear relationship valid for small strains, P=2E0Ssubscript𝑃parallel-to2subscript𝐸0𝑆P_{\parallel}=2E_{0}Sitalic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_S.

III Conclusions

Here, we have provided a detailed characterization of the nonlinear Poisson effect in a simple affine model of semiflexible polymer networks over a broad range of parameters. The mechanical behavior of the model network is highly sensitive to the asymmetric force-extension relationship of the constituent filaments, which is governed by two key dimensionless parameters: the dimensionless bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG and the dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, which together control the force-extension relationship of individual chains. Microscopically, the dimensionless bending rigidity κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG essentially controls the asymmetry of the filament force-extension relationship, whereas the dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT controls the amount of initial contraction and thus the elongation required to straighten the filament. We find that, on the level of the full network, decreasing κ~~𝜅\tilde{\kappa}over~ start_ARG italic_κ end_ARG intensifies the nonlinear Poisson effect and, in turn, intensifies the simultaneous alignment, densification, and stiffening of the network. The dimensionless persistence length ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT, in contrast, controls the amount of applied strain εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT required to induce this effect: increasing ~psubscript~𝑝\tilde{\ell}_{p}over~ start_ARG roman_ℓ end_ARG start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT reduces the critical extension εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, shifting the peak in the differential Poisson’s ratio ν~~𝜈\tilde{\nu}over~ start_ARG italic_ν end_ARG to lower values of εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT.

In future work, several extensions can be made to the model, both at the filament and network level, to improve its quantitative accuracy. For example, alternative models for the force-extension relationship of semiflexible filaments that properly incorporate buckling [71, 72, 73, 74, 3, 75] and more accurately account for filament extensibility [76, 77, 62] could be adopted in place of the relationship used here. We also note a recent model introducing a simplified force-extension relation based on pre-bent filaments, which permits simpler analytic results for some network properties [78]. On the network level, contributions to the total stress from the steric repulsion acting between filaments could also be included; this would prevent the complete transverse collapse of the network, with the initial filament volume fraction ϕ0subscriptitalic-ϕ0\phi_{0}italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT acting as a lower bound for the relative volume V/V0𝑉subscript𝑉0V/V_{0}italic_V / italic_V start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, assuming incompressibility of the filaments. For hydrogels of collagen, fibrin or other biopolymers, the network volume reduction seen under extensional strain requires the expulsion of water, which will tend to be suppressed on short time scales [79, 80, 50]. Thus, our results here are limited to the long-time elastic regime of network response. In future work, it would be interesting to consider effects of both network and solvent [81, 82] and poroelasticity [80, 50, 83], which would both improve accuracy of the model and allow for consideration of the dynamics associated with the nonlinear Poisson effect.

We note that, for the affine network model, the critical extensional strain εcsubscript𝜀𝑐\varepsilon_{c}italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT corresponding to the nonlinear Poisson effect can be related to the critical shear strain γcsubscript𝛾𝑐\gamma_{c}italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT corresponding to strain stiffening under applied simple shear. The maximum extension λmaxsubscript𝜆max\lambda_{\mathrm{max}}italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT for a given deformation 𝚲𝚲\mathbf{\Lambda}bold_Λ can be determined from the maximum eigenvalue λmax2superscriptsubscript𝜆max2\lambda_{\mathrm{max}}^{2}italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT of the right Cauchy-Green tensor 𝐂=𝚲T𝚲𝐂superscript𝚲𝑇𝚲\mathbf{C}=\mathbf{\Lambda}^{T}\mathbf{\Lambda}bold_C = bold_Λ start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT bold_Λ. For 𝚲(γ)𝚲𝛾\mathbf{\Lambda}(\gamma)bold_Λ ( italic_γ ) corresponding to simple shear γ𝛾\gammaitalic_γ, λmax2=1+(γ2+γγ2+4)/2superscriptsubscript𝜆max21superscript𝛾2𝛾superscript𝛾242\lambda_{\mathrm{max}}^{2}=1+(\gamma^{2}+\gamma\sqrt{\gamma^{2}+4})/2italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 1 + ( italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_γ square-root start_ARG italic_γ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 end_ARG ) / 2, and equivalently γ=λmax1/λmax𝛾subscript𝜆max1subscript𝜆max\gamma=\lambda_{\mathrm{max}}-1/\lambda_{\mathrm{max}}italic_γ = italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT - 1 / italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT. As strain-stiffening occurs when λmax=1+εcsubscript𝜆max1subscript𝜀𝑐\lambda_{\mathrm{max}}=1+\varepsilon_{c}italic_λ start_POSTSUBSCRIPT roman_max end_POSTSUBSCRIPT = 1 + italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, we can therefore relate the critical shear strain to the critical extension corresponding to the nonlinear Poisson’s ratio as γc=1+εc1/(1+εc)subscript𝛾𝑐1subscript𝜀𝑐11subscript𝜀𝑐\gamma_{c}=1+\varepsilon_{c}-1/(1+\varepsilon_{c})italic_γ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT = 1 + italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT - 1 / ( 1 + italic_ε start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT ).

While an affine model such as the one we propose here can be a useful, and in some cases, even predictive approach to modeling networks, we note that non-affine deformations are known to be important in some cases [56, 84, 85, 86, 3]. When considering non-affine effects, it is important to include the effects of mechanical integrity and bending rigidity of semiflexible filaments across network nodes, which leads to possible long-range mechanical effects [87]. This makes the modeling of such networks substantially more challenging. Nevertheless, it may be possible to extend a recent method for modeling such effects on linear elasticity to the nonlinear regime [88]. We speculate that non-affine effects can result in a possible delayed onset of network nonlinearity [6] and a possible reduction of the the peak in the Poisson ratio we see above.

Acknowledgements.
We thank Michael Rubinstein and Sergey Panyukov for stimulating discussions. This work was supported in part by the National Science Foundation Division of Materials Research (Grant No. DMR-2224030) and the National Science Foundation Center for Theoretical Biological Physics (Grant No. PHY-2019745). J.L.S. is grateful for the support of the Eric and Wendy Schmidt AI in Science Postdoctoral Fellowship at the University of Chicago. We would also like to thank the Isaac Newton Institute for Mathematical Sciences, Cambridge, for support and hospitality during the programme “New statistical physics in living matter: non equilibrium states under adaptive control,” where work on this paper was undertaken. This work was supported by EPSRC grant no. EP/R014604/1.

Appendix A Calculation of the stress components

According to the affine network model (Eq. 5), the ij𝑖𝑗ijitalic_i italic_j component of the Cauchy stress tensor 𝝈𝝈\bm{\sigma}bold_italic_σ for an initially isotropic network can be written as

σij=ρ04πdet𝚲𝑑θ0𝑑φ0sinθ0[τ(δ)(𝚲𝐧^0)i(𝚲𝐧^0)j|𝚲𝐧^0|].subscript𝜎𝑖𝑗subscript𝜌04𝜋det𝚲double-integraldifferential-dsubscript𝜃0differential-dsubscript𝜑0subscript𝜃0delimited-[]𝜏𝛿subscript𝚲subscript^𝐧0𝑖subscript𝚲subscript^𝐧0𝑗𝚲subscript^𝐧0\sigma_{ij}=\frac{\rho_{0}}{4\pi\mathrm{det}\bm{\Lambda}}\iint d\theta_{0}d% \varphi_{0}\sin\theta_{0}\left[\tau(\delta\ell)\frac{(\bm{\Lambda}\hat{\mathbf% {n}}_{0})_{i}(\bm{\Lambda}\hat{\mathbf{n}}_{0})_{j}}{\left|\bm{\Lambda}\hat{% \mathbf{n}}_{0}\right|}\right].italic_σ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG 4 italic_π roman_det bold_Λ end_ARG ∬ italic_d italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ italic_τ ( italic_δ roman_ℓ ) divide start_ARG ( bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG start_ARG | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | end_ARG ] . (15)

in which ρ0subscript𝜌0\rho_{0}italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT is the initial length density, 𝚲𝚲\bm{\Lambda}bold_Λ is the deformation gradient tensor, τ(δ)𝜏𝛿\tau(\delta\ell)italic_τ ( italic_δ roman_ℓ ) is the force-extension relationship, 𝐧^0(θ0,ϕ0)subscript^𝐧0subscript𝜃0subscriptitalic-ϕ0\hat{\mathbf{n}}_{0}(\theta_{0},\phi_{0})over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_ϕ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) is the initial filament orientation defined as

𝐧^0(θ0,φ0)=(sinθ0cosφ0sinθ0sinφ0cosθ0),subscript^𝐧0subscript𝜃0subscript𝜑0matrixsubscript𝜃0subscript𝜑0subscript𝜃0subscript𝜑0subscript𝜃0\hat{\mathbf{n}}_{0}(\theta_{0},\varphi_{0})=\begin{pmatrix}\sin\theta_{0}\cos% \varphi_{0}\\ \sin\theta_{0}\sin\varphi_{0}\\ \cos\theta_{0}\end{pmatrix},over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT , italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ) = ( start_ARG start_ROW start_CELL roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_cos italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW start_ROW start_CELL roman_cos italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_CELL end_ROW end_ARG ) , (16)

and the integrals are taken over the ranges 0θ0π0subscript𝜃0𝜋0\leq\theta_{0}\leq\pi0 ≤ italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ italic_π and 0φ02π0subscript𝜑02𝜋0\leq\varphi_{0}\leq 2\pi0 ≤ italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ≤ 2 italic_π.

For the uniaxial strain scenario considered in this work, the appropriate deformation gradient tensor 𝚲(ε,ε)𝚲subscript𝜀parallel-tosubscript𝜀perpendicular-to\bm{\Lambda}(\varepsilon_{\parallel},\varepsilon_{\perp})bold_Λ ( italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT , italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) is given by Eq. 9, for which det𝚲=(1+ε)2(1+ε)det𝚲superscript1subscript𝜀perpendicular-to21subscript𝜀parallel-to\mathrm{det}\bm{\Lambda}=(1+\varepsilon_{\perp})^{2}(1+\varepsilon_{\parallel})roman_det bold_Λ = ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ). Taking advantage of the axial symmetry of the system, we can write the normal components of the first Piola-Kirchhoff stress tensor as

P=ρ0(1+ε)2𝑑θ0sinθ0[τ(δ)cos2θ0|𝚲𝐧^0|]subscript𝑃parallel-tosubscript𝜌01subscript𝜀parallel-to2differential-dsubscript𝜃0subscript𝜃0delimited-[]𝜏𝛿superscript2subscript𝜃0𝚲subscript^𝐧0\displaystyle P_{\parallel}=\frac{\rho_{0}(1+\varepsilon_{\parallel})}{2}\int d% \theta_{0}\sin\theta_{0}\left[\tau(\delta\ell)\frac{\cos^{2}\theta_{0}}{|\bm{% \Lambda}\hat{\mathbf{n}}_{0}|}\right]italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) end_ARG start_ARG 2 end_ARG ∫ italic_d italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ italic_τ ( italic_δ roman_ℓ ) divide start_ARG roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | end_ARG ] (17)
P=ρ0(1+ε)4𝑑θ0sinθ0[τ(δ)sin2θ0|𝚲𝐧^0|]subscript𝑃perpendicular-tosubscript𝜌01subscript𝜀perpendicular-to4differential-dsubscript𝜃0subscript𝜃0delimited-[]𝜏𝛿superscript2subscript𝜃0𝚲subscript^𝐧0\displaystyle P_{\perp}=\frac{\rho_{0}(1+\varepsilon_{\perp})}{4}\int d\theta_% {0}\sin\theta_{0}\left[\tau(\delta\ell)\frac{\sin^{2}\theta_{0}}{|\bm{\Lambda}% \hat{\mathbf{n}}_{0}|}\right]italic_P start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = divide start_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) end_ARG start_ARG 4 end_ARG ∫ italic_d italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ italic_τ ( italic_δ roman_ℓ ) divide start_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | end_ARG ] (18)

in which |𝚲𝐧^0|=(1+ε)2sin2θ0+(1+ε)2cos2θ0𝚲subscript^𝐧0superscript1subscript𝜀perpendicular-to2superscript2subscript𝜃0superscript1subscript𝜀parallel-to2superscript2subscript𝜃0|\bm{\Lambda}\hat{\mathbf{n}}_{0}|=\sqrt{(1+\varepsilon_{\perp})^{2}\sin^{2}% \theta_{0}+(1+\varepsilon_{\parallel})^{2}\cos^{2}\theta_{0}}| bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | = square-root start_ARG ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT + ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG. As described in the main text, the change in length is δ=0(|𝚲𝐧^0|1)𝛿subscript0𝚲subscript^𝐧01\delta\ell=\ell_{0}\left(|\bm{\Lambda}\hat{\mathbf{n}}_{0}|-1\right)italic_δ roman_ℓ = roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | - 1 ) and the tension τ(δ)𝜏𝛿\tau(\delta\ell)italic_τ ( italic_δ roman_ℓ ) is determined by inversion of Eq. 3.

In the small strain regime, we can solve Eqs. 17 and 18 using the linear force-extension relationship τ=kδ/c𝜏𝑘𝛿subscript𝑐\tau=k\delta\ell/\ell_{c}italic_τ = italic_k italic_δ roman_ℓ / roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT, with k=1/(μ1+c3/(90κp))𝑘1superscript𝜇1superscriptsubscript𝑐390𝜅subscript𝑝k=1/(\mu^{-1}+\ell_{c}^{3}/(90\kappa\ell_{p}))italic_k = 1 / ( italic_μ start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT + roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT / ( 90 italic_κ roman_ℓ start_POSTSUBSCRIPT italic_p end_POSTSUBSCRIPT ) ), the linear strain dependence ε=νεsubscript𝜀perpendicular-to𝜈subscript𝜀parallel-to\varepsilon_{\perp}=-\nu\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - italic_ν italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT with ν=1/4𝜈14\nu=1/4italic_ν = 1 / 4. To linear order in εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, the parallel stress component Psubscript𝑃parallel-toP_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT behaves as

Pσ16ρ0k0cεsubscript𝑃parallel-tosubscript𝜎parallel-to16subscript𝜌0𝑘subscript0subscript𝑐subscript𝜀parallel-toP_{\parallel}\approx\sigma_{\parallel}\approx\frac{1}{6}\rho_{0}k\frac{\ell_{0% }}{\ell_{c}}\varepsilon_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ italic_σ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ divide start_ARG 1 end_ARG start_ARG 6 end_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_k divide start_ARG roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT (19)

from which we see that the linear Young’s modulus E0=limε0+σ/εsubscript𝐸0subscript𝜀superscript0subscript𝜎parallel-tosubscript𝜀parallel-toE_{0}=\lim_{\varepsilon\to 0^{+}}\sigma_{\parallel}/\varepsilon_{\parallel}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = roman_lim start_POSTSUBSCRIPT italic_ε → 0 start_POSTSUPERSCRIPT + end_POSTSUPERSCRIPT end_POSTSUBSCRIPT italic_σ start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT is

E0=16ρ0k0c.subscript𝐸016subscript𝜌0𝑘subscript0subscript𝑐E_{0}=\frac{1}{6}\rho_{0}k\frac{\ell_{0}}{\ell_{c}}.italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 6 end_ARG italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_k divide start_ARG roman_ℓ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG roman_ℓ start_POSTSUBSCRIPT italic_c end_POSTSUBSCRIPT end_ARG . (20)

Appendix B Calculation of the nematic alignment

As described in the main text, we consider an initially isotropic network with orientations 𝐧^0subscript^𝐧0\hat{\mathbf{n}}_{0}over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT given by Eq. 16 under deformation gradient tensor 𝚲𝚲\bm{\Lambda}bold_Λ, such that the filament orientations in the deformed configuration are 𝐧^=𝚲𝐧^0/|𝚲𝐧^0|^𝐧𝚲subscript^𝐧0𝚲subscript^𝐧0\hat{\mathbf{n}}=\bm{\Lambda}\hat{\mathbf{n}}_{0}/|\bm{\Lambda}\hat{\mathbf{n}% }_{0}|over^ start_ARG bold_n end_ARG = bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT / | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT |. In the deformed configuration, the traceless and symmetric nematic tensor 𝐐𝐐\mathbf{Q}bold_Q is given by Eq. 7 and can be written in terms of its eigenvalues λisubscript𝜆𝑖\lambda_{i}italic_λ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT and orthonormal eigenvectors 𝐯^isubscript^𝐯𝑖\hat{\mathbf{v}}_{i}over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT as

𝐐=λ1𝐯^1𝐯^1+λ2𝐯^2𝐯^2+λ3𝐯^3𝐯^3.𝐐tensor-productsubscript𝜆1subscript^𝐯1subscript^𝐯1tensor-productsubscript𝜆2subscript^𝐯2subscript^𝐯2tensor-productsubscript𝜆3subscript^𝐯3subscript^𝐯3\mathbf{Q}=\lambda_{1}\hat{\mathbf{v}}_{1}\otimes\hat{\mathbf{v}}_{1}+\lambda_% {2}\hat{\mathbf{v}}_{2}\otimes\hat{\mathbf{v}}_{2}+\lambda_{3}\hat{\mathbf{v}}% _{3}\otimes\hat{\mathbf{v}}_{3}.bold_Q = italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⊗ over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⊗ over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⊗ over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT . (21)

with λ1+λ2+λ3=0subscript𝜆1subscript𝜆2subscript𝜆30\lambda_{1}+\lambda_{2}+\lambda_{3}=0italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT + italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 0. For uniaxial deformation applied along the z𝑧zitalic_z-axis, symmetry requires that two of the eigenvalues of 𝐐𝐐\mathbf{Q}bold_Q are equal, so we can write λ1=λ2=S/3subscript𝜆1subscript𝜆2𝑆3\lambda_{1}=\lambda_{2}=-S/3italic_λ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = italic_λ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = - italic_S / 3 and λ3=2S/3subscript𝜆32𝑆3\lambda_{3}=2S/3italic_λ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = 2 italic_S / 3 [89]. Noting that 𝐯^3=𝐳^subscript^𝐯3^𝐳\hat{\mathbf{v}}_{3}=\hat{\mathbf{z}}over^ start_ARG bold_v end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = over^ start_ARG bold_z end_ARG in this case, we can write the tensor 𝐐𝐐\mathbf{Q}bold_Q as

𝐐=S(𝐳^𝐳^13𝐈)𝐐𝑆tensor-product^𝐳^𝐳13𝐈\mathbf{Q}=S\left(\hat{\mathbf{z}}\otimes\hat{\mathbf{z}}-\frac{1}{3}\mathbf{I% }\right)bold_Q = italic_S ( over^ start_ARG bold_z end_ARG ⊗ over^ start_ARG bold_z end_ARG - divide start_ARG 1 end_ARG start_ARG 3 end_ARG bold_I ) (22)

where the nematic alignment S𝑆Sitalic_S is given by Eq. 14. Integrating over the initial orientations 𝐧^0subscript^𝐧0\hat{\mathbf{n}}_{0}over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT, we find

S𝑆\displaystyle Sitalic_S =14π𝑑φ0𝑑θ0sinθ0[32((1+ε)2cos2θ0|𝚲𝐧^0|213)]absent14𝜋double-integraldifferential-dsubscript𝜑0differential-dsubscript𝜃0subscript𝜃0delimited-[]32superscript1subscript𝜀parallel-to2superscript2subscript𝜃0superscript𝚲subscript^𝐧0213\displaystyle=\frac{1}{4\pi}\iint d\varphi_{0}d\theta_{0}\sin\theta_{0}\left[% \frac{3}{2}\left(\frac{(1+\varepsilon_{\parallel})^{2}\cos^{2}\theta_{0}}{|\bm% {\Lambda}\hat{\mathbf{n}}_{0}|^{2}}-\frac{1}{3}\right)\right]= divide start_ARG 1 end_ARG start_ARG 4 italic_π end_ARG ∬ italic_d italic_φ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_d italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ divide start_ARG 3 end_ARG start_ARG 2 end_ARG ( divide start_ARG ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 3 end_ARG ) ]
=12𝑑θ0sinθ0[32((1+ε)2cos2θ0|𝚲𝐧^0|213)]absent12differential-dsubscript𝜃0subscript𝜃0delimited-[]32superscript1subscript𝜀parallel-to2superscript2subscript𝜃0superscript𝚲subscript^𝐧0213\displaystyle=\frac{1}{2}\int d\theta_{0}\sin\theta_{0}\left[\frac{3}{2}\left(% \frac{(1+\varepsilon_{\parallel})^{2}\cos^{2}\theta_{0}}{|\bm{\Lambda}\hat{% \mathbf{n}}_{0}|^{2}}-\frac{1}{3}\right)\right]= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ∫ italic_d italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_sin italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT [ divide start_ARG 3 end_ARG start_ARG 2 end_ARG ( divide start_ARG ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_ARG start_ARG | bold_Λ over^ start_ARG bold_n end_ARG start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT | start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 3 end_ARG ) ]
=32(1+ε)2a2(1(1+εa)tan1(a1+ε))12absent32superscript1subscript𝜀parallel-to2superscript𝑎211subscript𝜀perpendicular-to𝑎superscript1𝑎1subscript𝜀perpendicular-to12\displaystyle=\frac{3}{2}\frac{(1+\varepsilon_{\parallel})^{2}}{a^{2}}\left(1-% \left(\frac{1+\varepsilon_{\perp}}{a}\right)\tan^{-1}\left(\frac{a}{1+% \varepsilon_{\perp}}\right)\right)-\frac{1}{2}= divide start_ARG 3 end_ARG start_ARG 2 end_ARG divide start_ARG ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_a start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 1 - ( divide start_ARG 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG start_ARG italic_a end_ARG ) roman_tan start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ( divide start_ARG italic_a end_ARG start_ARG 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT end_ARG ) ) - divide start_ARG 1 end_ARG start_ARG 2 end_ARG (23)

in which a=(1+ε)2(1+ε)2𝑎superscript1subscript𝜀parallel-to2superscript1subscript𝜀perpendicular-to2a=\sqrt{(1+\varepsilon_{\parallel})^{2}-(1+\varepsilon_{\perp})^{2}}italic_a = square-root start_ARG ( 1 + italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( 1 + italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG. For small εsubscript𝜀parallel-to\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT, this yields Sε/2𝑆subscript𝜀parallel-to2S\approx\varepsilon_{\parallel}/2italic_S ≈ italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT / 2. Note that the alignment S𝑆Sitalic_S can vary in the range S[1/2,1]𝑆121S\in[-1/2,1]italic_S ∈ [ - 1 / 2 , 1 ], in which S=1/2𝑆12S=-1/2italic_S = - 1 / 2 corresponds to a configuration in which all filaments are oriented perpendicular to 𝐳^^𝐳\hat{\mathbf{z}}over^ start_ARG bold_z end_ARG and S=1𝑆1S=1italic_S = 1 corresponds to a configuration in which all filaments are parallel to 𝐳^^𝐳\hat{\mathbf{z}}over^ start_ARG bold_z end_ARG.

In the small strain regime, in which ε=νεsubscript𝜀perpendicular-to𝜈subscript𝜀parallel-to\varepsilon_{\perp}=-\nu\varepsilon_{\parallel}italic_ε start_POSTSUBSCRIPT ⟂ end_POSTSUBSCRIPT = - italic_ν italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT with ν=1/4𝜈14\nu=1/4italic_ν = 1 / 4, we have

S12ε𝑆12subscript𝜀parallel-toS\approx\frac{1}{2}\varepsilon_{\parallel}italic_S ≈ divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT (24)

Since PE0εsubscript𝑃parallel-tosubscript𝐸0subscript𝜀parallel-toP_{\parallel}\approx E_{0}\varepsilon_{\parallel}italic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_ε start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT in this regime, we find that the relationship between stress and alignment in the linear regime is

P2E0Ssubscript𝑃parallel-to2subscript𝐸0𝑆P_{\parallel}\approx 2E_{0}Sitalic_P start_POSTSUBSCRIPT ∥ end_POSTSUBSCRIPT ≈ 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT italic_S (25)

with E0subscript𝐸0E_{0}italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT given by Eq. 20. This relationship is plotted in Fig. 5. As we noted in the main text, Eq. 25 implies that the proportionality coefficient C𝐶Citalic_C in Eq. 8 is given by C=1/(2E0)𝐶12subscript𝐸0C=1/(2E_{0})italic_C = 1 / ( 2 italic_E start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ).

References

  • Kasza et al. [2007] K. E. Kasza, A. C. Rowat, J. Liu, T. E. Angelini, C. P. Brangwynne, G. H. Koenderink, and D. A. Weitz, The cell as a material, Current Opinion in Cell Biology 19, 101 (2007).
  • Pritchard et al. [2014] R. H. Pritchard, Y. Y. Shery Huang, and E. M. Terentjev, Mechanics of biological networks: From the cell cytoskeleton to connective tissue, Soft Matter 10, 1864 (2014).
  • Broedersz and MacKintosh [2014] C. P. Broedersz and F. C. MacKintosh, Modeling semiflexible polymer networks, Reviews of Modern Physics 86, 995 (2014).
  • Burla et al. [2019] F. Burla, Y. Mulla, B. E. Vos, A. Aufderhorst-Roberts, and G. H. Koenderink, From mechanical resilience to active material properties in biopolymer networks, Nature Reviews Physics 1, 249 (2019).
  • Fung [1967] Y. C. Fung, Elasticity of soft tissues in simple elongation, American Journal of Physiology 213, 1532 (1967).
  • Gardel et al. [2004] M. L. Gardel, J. H. Shin, F. C. MacKintosh, L. Mahadevan, P. Matsudaira, and D. A. Weitz, Elastic Behavior of Cross-Linked and Bundled Actin Networks, Science 304, 1301 (2004).
  • Storm et al. [2005] C. Storm, J. J. Pastore, F. C. MacKintosh, T. C. Lubensky, and P. A. Janmey, Nonlinear elasticity in biological gels, Nature 435, 191 (2005).
  • Van Oosten et al. [2016] A. S. G. Van Oosten, M. Vahabi, A. J. Licup, A. Sharma, P. A. Galie, F. C. MacKintosh, and P. A. Janmey, Uncoupling shear and uniaxial elastic moduli of semiflexible biopolymer networks: Compression-softening and stretch-stiffening, Scientific Reports 6, 19270 (2016).
  • Xu and Safran [2017] X. Xu and S. A. Safran, Compressive elasticity of polydisperse biopolymer gels, Physical Review E 95, 052415 (2017).
  • Vahabi et al. [2016] M. Vahabi, A. Sharma, A. J. Licup, A. S. G. Van Oosten, P. A. Galie, P. A. Janmey, and F. C. MacKintosh, Elasticity of fibrous networks under uniaxial prestress, Soft Matter 12, 5050 (2016).
  • Van Oosten et al. [2019] A. S. G. Van Oosten, X. Chen, L. Chin, K. Cruz, A. E. Patteson, K. Pogoda, V. B. Shenoy, and P. A. Janmey, Emergence of tissue-like mechanics from fibrous networks confined by close-packed cells, Nature 573, 96 (2019).
  • Ed-Daoui and Snabre [2021] A. Ed-Daoui and P. Snabre, Poroviscoelasticity and compression-softening of agarose hydrogels, Rheologica Acta 60, 327 (2021).
  • Brown et al. [2009] A. E. X. Brown, R. I. Litvinov, D. E. Discher, P. K. Purohit, and J. W. Weisel, Multiscale mechanics of fibrin polymer: Gel stretching with protein unfolding and loss of water, Science 325, 741 (2009).
  • Vader et al. [2009] D. Vader, A. Kabla, D. Weitz, and L. Mahadevan, Strain-Induced Alignment in Collagen Gels, PLoS ONE 4, e5902 (2009).
  • Steinwachs et al. [2016] J. Steinwachs, C. Metzner, K. Skodzek, N. Lang, I. Thievessen, C. Mark, S. Münster, K. E. Aifantis, and B. Fabry, Three-dimensional force microscopy of cells in biopolymer networks, Nature Methods 13, 171 (2016).
  • Ban et al. [2019] E. Ban, H. Wang, J. M. Franklin, J. T. Liphardt, P. A. Janmey, and V. B. Shenoy, Strong triaxial coupling and anomalous Poisson effect in collagen networks, Proceedings of the National Academy of Sciences 116, 6790 (2019).
  • Poisson [1827] S. D. Poisson, Note sur l’extension des fils et des plaques élastiques, Annales de Chimie et de Physique 36, 384 (1827).
  • Love [1906] A. E. H. Love, A Treatise on the Mathematical Theory of Elasticity, 2nd ed. (Cambridge University Press, 1906).
  • Greaves et al. [2011] G. N. Greaves, A. L. Greer, R. S. Lakes, and T. Rouxel, Poisson’s ratio and modern materials, Nature Materials 10, 823 (2011).
  • Wolf et al. [2013] K. Wolf, M. te Lindert, M. Krause, S. Alexander, J. te Riet, A. L. Willis, R. M. Hoffman, C. G. Figdor, S. J. Weiss, and P. Friedl, Physical limits of cell migration: Control by ECM space and nuclear deformation and tuning by proteolysis and traction force, The Journal of Cell Biology 201, 1069 (2013).
  • Paul et al. [2017] C. D. Paul, P. Mistriotis, and K. Konstantopoulos, Cancer cell motility: Lessons from migration in confined spaces, Nature Reviews Cancer 17, 131 (2017).
  • Abhilash et al. [2014] A. S. Abhilash, B. M. Baker, B. Trappmann, C. S. Chen, and V. B. Shenoy, Remodeling of fibrous extracellular matrices by contractile cells: Predictions from discrete fiber network simulations, Biophysical Journal 107, 1829 (2014).
  • Wang et al. [2014] H. Wang, A. S. Abhilash, C. S. Chen, R. G. Wells, and V. B. Shenoy, Long-range force transmission in fibrous matrices enabled by tension-driven alignment of fibers, Biophysical Journal 107, 2592 (2014).
  • Hall et al. [2016] M. S. Hall, F. Alisafaei, E. Ban, X. Feng, C.-Y. Hui, V. B. Shenoy, and M. Wu, Fibrous nonlinear elasticity enables positive mechanical feedback between cells and ECMs, Proceedings of the National Academy of Sciences 113, 14043 (2016).
  • Notbohm et al. [2015] J. Notbohm, A. Lesman, P. Rosakis, D. A. Tirrell, and G. Ravichandran, Microbuckling of fibrin provides a mechanism for cell mechanosensing, Journal of the Royal Society, Interface 12, 20150320 (2015).
  • SenGupta et al. [2021] S. SenGupta, C. A. Parent, and J. E. Bear, The principles of directed cell migration, Nature Reviews Molecular Cell Biology 22, 529 (2021).
  • Ahmed and Saif [2014] W. W. Ahmed and T. A. Saif, Active transport of vesicles in neurons is modulated by mechanical tension, Scientific Reports 4, 4481 (2014).
  • Svitkina et al. [1997] T. M. Svitkina, A. B. Verkhovsky, K. M. McQuade, and G. G. Borisy, Analysis of the Actin–Myosin II System in Fish Epidermal Keratocytes: Mechanism of Cell Body Translocation, Journal of Cell Biology 139, 397 (1997).
  • Keren and Theriot [2008] K. Keren and J. A. Theriot, Biophysical aspects of actin-based cell motility in fish epithelial keratocytes, in Cell Motility, Biological and Medical Physics, Biomedical Engineering, edited by P. Lenz (Springer, New York, NY, 2008) pp. 31–58.
  • Lee et al. [2010] C.-F. Lee, C. Haase, S. Deguchi, and R. Kaunas, Cyclic stretch-induced stress fiber dynamics – Dependence on strain rate, Rho-kinase and MLCK, Biochemical and Biophysical Research Communications 401, 344 (2010).
  • Nagayama et al. [2012] K. Nagayama, Y. Kimura, N. Makino, and T. Matsumoto, Strain waveform dependence of stress fiber reorientation in cyclically stretched osteoblastic cells: Effects of viscoelastic compression of stress fibers, American Journal of Physiology-Cell Physiology 302, C1469 (2012).
  • Kaunas [2015] R. Kaunas, Dynamic stress fiber reorganization on stretched matrices, in Cell and Matrix Mechanics, edited by R. Kaunas and A. Zemel (CRC Press, Taylor & Francis Group, Boca Raton, 2015).
  • Kumar et al. [2019] A. Kumar, M. S. Shutova, K. Tanaka, D. V. Iwamoto, D. A. Calderwood, T. M. Svitkina, and M. A. Schwartz, Filamin A mediates isotropic distribution of applied force across the actin network, Journal of Cell Biology 218, 2481 (2019).
  • Kabla and Mahadevan [2007] A. Kabla and L. Mahadevan, Nonlinear mechanics of soft fibrous networks, Journal of The Royal Society Interface 4, 99 (2007).
  • Picu et al. [2018] R. C. Picu, S. Deogekar, and M. R. Islam, Poisson’s Contraction and Fiber Kinematics in Tissue: Insight From Collagen Network Simulations, Journal of Biomechanical Engineering 140, 021002 (2018).
  • Shivers et al. [2020] J. L. Shivers, S. Arzash, and F. C. MacKintosh, Nonlinear Poisson Effect Governed by a Mechanical Critical Transition, Physical Review Letters 124, 038002 (2020).
  • MacKintosh et al. [1995] F. C. MacKintosh, J. Käs, and P. A. Janmey, Elasticity of Semiflexible Biopolymer Networks, Physical Review Letters 75, 4425 (1995).
  • Odijk [1995] T. Odijk, Stiff Chains and Filaments under Tension, Macromolecules 28, 7016 (1995).
  • Treloar [1954] L. R. G. Treloar, The photoelastic properties of short-chain molecular networks, Transactions of the Faraday Society 50, 881 (1954).
  • Treloar [1975] L. R. G. Treloar, The Physics of Rubber Elasticity, 3rd ed. (Oxford University Press, Oxford, 1975).
  • Treloar et al. [1979] L. R. G. Treloar, G. Riding, and G. Gee, A non-Gaussian theory for rubber in biaxial strain. I. Mechanical properties, Proceedings of the Royal Society of London. A 369, 261 (1979).
  • Wu and van der Giessen [1993] P. D. Wu and E. van der Giessen, On improved network models for rubber elasticity and their applications to orientation hardening in glassy polymers, Journal of the Mechanics and Physics of Solids 41, 427 (1993).
  • Wu and van der Giessen [1995] P. D. Wu and E. van der Giessen, On network descriptions of mechanical and optical properties of rubbers, Philosophical Magazine A 71, 1191 (1995).
  • Boyce and Arruda [2000] M. C. Boyce and E. M. Arruda, Constitutive Models of Rubber Elasticity: A Review, Rubber Chemistry and Technology 73, 504 (2000).
  • Palmer and Boyce [2008] J. S. Palmer and M. C. Boyce, Constitutive modeling of the stress–strain behavior of F-actin filament networks, Acta Biomaterialia 4, 597 (2008).
  • Cioroianu and Storm [2013] A. R. Cioroianu and C. Storm, Normal stresses in elastic networks, Physical Review E 88, 052601 (2013).
  • Meng and Terentjev [2017] F. Meng and E. Terentjev, Theory of Semiflexible Filaments and Networks, Polymers 9, 52 (2017).
  • Song et al. [2022] D. Song, A. A. Oberai, and P. A. Janmey, Hyperelastic continuum models for isotropic athermal fibrous networks, Interface Focus 12, 20220043 (2022).
  • Rubinstein and Colby [2006] M. Rubinstein and R. H. Colby, Polymer Physics (Oxford University Press, Oxford, 2006).
  • Vahabi et al. [2018] M. Vahabi, B. E. Vos, H. C. G. De Cagny, D. Bonn, G. H. Koenderink, and F. C. MacKintosh, Normal stresses in semiflexible polymer hydrogels, Physical Review E 97, 032418 (2018).
  • Note [1] For a homogeneous cylindrical elastic rod of radius r𝑟ritalic_r, the stretching modulus μ𝜇\muitalic_μ is related to the bending stiffness κ𝜅\kappaitalic_κ as μ=4κ/r2𝜇4𝜅superscript𝑟2\mu=4\kappa/r^{2}italic_μ = 4 italic_κ / italic_r start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT [7].
  • Bustamante et al. [1994] C. Bustamante, J. F. Marko, E. D. Siggia, and S. Smith, Entropic Elasticity of λ𝜆\lambdaitalic_λ-Phage DNA, Science 265, 1599 (1994).
  • Fixman and Kovac [1973] M. Fixman and J. Kovac, Polymer conformational statistics. III. Modified Gaussian models of stiff chains, The Journal of Chemical Physics 58, 1564 (1973).
  • James and Guth [1943] H. M. James and E. Guth, Theory of the Elastic Properties of Rubber, The Journal of Chemical Physics 11, 455 (1943).
  • Wu and van der Giessen [1992] P. D. Wu and E. van der Giessen, On improved 3-D non-Gaussian network models for rubber elasticity, Mechanics Research Communications 19, 427 (1992).
  • Kroy and Frey [1996] K. Kroy and E. Frey, Force-Extension Relation and Plateau Modulus for Wormlike Chains, Physical Review Letters 77, 306 (1996).
  • Broedersz et al. [2009] C. P. Broedersz, C. Storm, and F. C. MacKintosh, Effective-medium approach for stiff polymer networks with flexible cross-links, Physical Review E 79, 061914 (2009).
  • Lin et al. [2010] Y.-C. Lin, N. Y. Yao, C. P. Broedersz, H. Herrmann, F. C. MacKintosh, and D. A. Weitz, Origins of Elasticity in Intermediate Filament Networks, Physical Review Letters 104, 058101 (2010).
  • Yao et al. [2010] N. Y. Yao, C. P. Broedersz, Y.-C. Lin, K. E. Kasza, F. C. MacKintosh, and D. A. Weitz, Elasticity in Ionically Cross-Linked Neurofilament Networks, Biophysical Journal 98, 2147 (2010).
  • Van Oosterwyck et al. [2013] H. Van Oosterwyck, J. F. Rodríguez, M. Doblaré, and J. M. García Aznar, An affine micro-sphere-based constitutive model, accounting for junctional sliding, can capture F-actin network mechanics, Computer Methods in Biomechanics and Biomedical Engineering 16, 1002 (2013).
  • Holzapfel et al. [2014] G. A. Holzapfel, M. J. Unterberger, and R. W. Ogden, An affine continuum mechanical model for cross-linked F-actin networks with compliant linker proteins, Journal of the Mechanical Behavior of Biomedical Materials 38, 78 (2014).
  • Unterberger and Holzapfel [2014] M. J. Unterberger and G. A. Holzapfel, Advances in the mechanical modeling of filamentous actin and its cross-linked networks on multiple scales, Biomechanics and Modeling in Mechanobiology 13, 1155 (2014).
  • Larson [1999] R. G. Larson, The Structure and Rheology of Complex Fluids (Oxford University Press, 1999).
  • Morse [1998] D. C. Morse, Viscoelasticity of Concentrated Isotropic Solutions of Semiflexible Polymers. 1. Model and Stress Tensor, Macromolecules 31, 7030 (1998).
  • Gittes and MacKintosh [1998] F. Gittes and F. C. MacKintosh, Dynamic shear modulus of a semiflexible polymer network, Physical Review E 58, R1241 (1998).
  • de Gennes and Prost [2013] P.-G. de Gennes and J. Prost, The Physics of Liquid Crystals, 83 (Clarendon Press, Oxford, 2013).
  • Feng et al. [2015] J. Feng, H. Levine, X. Mao, and L. M. Sander, Alignment and nonlinear elasticity in biopolymer gels, Physical Review E 91, 042710 (2015).
  • Doi [2004] M. Doi, Introduction to Polymer Physics (Clarendon Press, Oxford, 2004).
  • Macosko [1994] C. W. Macosko, Rheology: Principles, Measurements, and Applications, Advances in Interfacial Engineering Series (VCH, New York, NY, 1994).
  • Das and MacKintosh [2010] M. Das and F. C. MacKintosh, Poisson’s Ratio in Composite Elastic Media with Rigid Rods, Physical Review Letters 105, 138102 (2010).
  • Odijk [1998] T. Odijk, Microfibrillar buckling within fibers under compression, The Journal of Chemical Physics 108, 6923 (1998).
  • Emanuel et al. [2007] M. Emanuel, H. Mohrbach, M. Sayar, H. Schiessel, and I. M. Kulić, Buckling of stiff polymers: Influence of thermal fluctuations, Physical Review E 76, 061907 (2007).
  • Baczynski et al. [2008] K. Baczynski, R. Lipowsky, and J. Kierfeld, Stretching of buckled filaments by thermal fluctuations, Physical Review E 76, 1 (2008).
  • Blundell and Terentjev [2009] J. R. Blundell and E. M. Terentjev, Buckling of semiflexible filaments under compression, Soft Matter 5, 4015 (2009).
  • Kurzthaler [2018] C. Kurzthaler, Elastic behavior of a semiflexible polymer in 3D subject to compression and stretching forces, Soft Matter 14, 7634 (2018).
  • Unterberger et al. [2013] M. J. Unterberger, K. M. Schmoller, A. R. Bausch, and G. A. Holzapfel, A new approach to model cross-linked actin networks: Multi-scale continuum formulation and computational analysis, Journal of the Mechanical Behavior of Biomedical Materials 22, 95 (2013).
  • Holzapfel and Ogden [2013] G. A. Holzapfel and R. W. Ogden, Elasticity of biopolymer filaments, Acta Biomaterialia 9, 7320 (2013).
  • Li et al. [2022] Y. Li, Y. Li, E. Prince, J. I. Weitz, S. Panyukov, A. Ramachandran, M. Rubinstein, and E. Kumacheva, Fibrous hydrogels under biaxial confinement, Nature Communications 13, 3264 (2022).
  • Janmey et al. [2007] P. A. Janmey, M. E. McCormick, S. Rammensee, J. L. Leight, P. C. Georges, and F. C. MacKintosh, Negative normal stress in semiflexible biopolymer gels, Nature Materials 6, 48 (2007).
  • De Cagny et al. [2016] H. C. G. De Cagny, B. E. Vos, M. Vahabi, N. A. Kurniawan, M. Doi, G. H. Koenderink, F. C. MacKintosh, and D. Bonn, Porosity Governs Normal Stresses in Polymer Gels, Physical Review Letters 117, 217802 (2016).
  • Doi [2013] M. Doi, Soft Matter Physics (Oxford University Press, 2013).
  • Yamamoto et al. [2017] T. Yamamoto, Y. Masubuchi, and M. Doi, Large Network Swelling and Solvent Redistribution Are Necessary for Polymer Gels to Show Negative Normal Stress, ACS Macro Letters 6, 512 (2017).
  • Punter et al. [2020] M. T. J. J. M. Punter, B. E. Vos, B. M. Mulder, and G. H. Koenderink, Poroelasticity of (bio)polymer networks during compression: Theory and experiment, Soft Matter 16, 1298 (2020).
  • Head et al. [2003] D. A. Head, A. J. Levine, and F. C. MacKintosh, Deformation of Cross-Linked Semiflexible Polymer Networks, Physical Review Letters 91, 108102 (2003).
  • Wilhelm and Frey [2003] J. Wilhelm and E. Frey, Elasticity of Stiff Polymer Networks, Physical Review Letters 91, 108103 (2003).
  • Heussinger and Frey [2006] C. Heussinger and E. Frey, Floppy Modes and Nonaffine Deformations in Random Fiber Networks, Physical Review Letters 97, 105501 (2006).
  • Broedersz et al. [2012] C. P. Broedersz, M. Sheinman, and F. C. MacKintosh, Filament-length-controlled elasticity in 3D fiber networks, Physical Review Letters 108, 078102 (2012).
  • Chen et al. [2023] S. Chen, T. Markovich, and F. C. MacKintosh, Nonaffine Deformation of Semiflexible Polymer and Fiber Networks, Physical Review Letters 130, 088101 (2023).
  • Ball [2012] J. Ball, Mathematics of Liquid Crystals, Cambridge CCA Course (2012).