HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.

  • failed: epic

Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.

License: arXiv.org perpetual non-exclusive license
arXiv:2311.07500v2 [cond-mat.soft] 24 Mar 2024

Biaxial nematic order in fundamental measure theory

Anouar El Moumane Institut für Theoretische Physik II: Weiche Materie, Heinrich-Heine-Universität Düsseldorf, 40225 Düsseldorf, Germany    Michael te Vrugt DAMTP, Centre for Mathematical Sciences, University of Cambridge, Cambridge CB3 0WA, United Kingdom    Hartmut Löwen Institut für Theoretische Physik II: Weiche Materie, Heinrich-Heine-Universität Düsseldorf, 40225 Düsseldorf, Germany    René Wittmann [email protected] Institut für Theoretische Physik II: Weiche Materie, Heinrich-Heine-Universität Düsseldorf, 40225 Düsseldorf, Germany Institut für Sicherheit und Qualität bei Fleisch, Max Rubner-Institut, 95326 Kulmbach, Germany
(March 24, 2024)
Abstract

Liquid crystals consisting of biaxial particles can exhibit a much richer phase behavior than their uniaxial counterparts. Usually, one has to rely on simulation results to understand the phase diagram of these systems, since very few analytical results exist. In this work, we apply fundamental measure theory, which allows us to derive free energy functionals for hard particles from first principles and with high accuracy, to systems of hard cylinders, cones and spherotriangles. We provide a general recipe for incorporating biaxial liquid crystal order parameters into fundamental measure theory and use this framework to obtain the phase boundaries for the emergence of orientational order in the considered systems. Our results provide insights into the phase behavior of biaxial nematic liquid crystals and, in particular, into methods for their analytical investigation.

I Introduction

The rich collective behavior that results from anisotropic interactions between particles has fascinated liquid crystal researchers for decades de Gennes and Prost (1993). Even in spatially homogeneous systems, the particle’s orientations can be uncorrelated (isotropic phase) or correlated (ordered phases). A typical example for an orientationally ordered phase is the nematic phase, in which the main symmetry axis of the particles is aligned. Over the years, more and more phases with nontrivial ordering behavior have been identified Sebastián et al. (2022); Bruce (2004). Prominent examples that have attracted interest in recent years are ferroelectric nematics (global polar order in which additional top-down symmetry is broken) Chen et al. (2020); Lavrentovich (2020) and biaxial nematics (two main axes of preferred orientation) Freiser (1970); Madsen et al. (2004); Luckhurst and Sluckin (2015). Beyond these homogeneous ordered phases, many systems can also exhibit additional positional order, forming inhomogeneous liquid crystal phases such as columnars and smectics.

Liquid crystalline order has also been observed in colloidal systems Vroege and Lekkerkerker (1992); Smalyukh (2018), whose constituting particles can be synthesized nowadays with nearly any desired shape Sacanna et al. (2013); Hueckel et al. (2021). The entropic nature of the interactions governing these systems allows for a reliable theoretical modeling in terms of pure hard-core interactions Mederos et al. (2014). Therefore, the ordering behavior of various hard-body systems is well studied, in particular, for apolar uniaxial shapes such as rods or disks McGrother et al. (1996); Bolhuis and Frenkel (1997); Wensink and Vroege (2003); Pfleiderer and Schilling (2007); Marechal et al. (2011); Odriozola (2012). Moreover, while shape polarity is typically not sufficient to stabilize polar order in a homogeneous phase of purely hard particles, it can result in a rich behavior including polar domains at higher densities Kubala et al. (2023) or even periodic gyroidal structures for particular pear-like shapes Ellison et al. (2006); Schönhöfer et al. (2017). Arguably, biaxial shapes give rise to an even more complex phase behavior due to their lower symmetry. Even in two spatial dimensions, ordering phenomena related to two distinct particle axes reach from interlocked layers Monderkamp et al. (2023) to the emergence of nontrivial orientational symmetries Martínez-Ratón and Velasco (2021, 2022). In three spatial dimensions, the biaxial ordering of board-like particles (cuboids or spheroplatelets) is well studied Alben (1973); Taylor and Herzfeld (1991); Vanakaras et al. (2003); Martínez-Ratón et al. (2011); Belli et al. (2011, 2012); Peroukidis and Vanakaras (2013); Cuetos et al. (2017). These shapes are the common choice for investigating biaxiality due to their particular symmetry and their realization in colloidal experiments Van den Pol et al. (2009). However, quantitative theoretical results remain scarce.

The phase behavior of systems with complex particle shapes is typically investigated using computer simulations. For gaining a deeper understanding of the phase diagram of complex liquid crystals and for reducing the computational effort, it is advantageous to have available an analytic theory which also represents the actual geometric interactions between the particles. The best candidate for such a theoretical description is fundamental measure theory (FMT) Rosenfeld (1989); Roth (2010); Roth et al. (2012), which allows us to obtain quantitatively accurate free energy functionals for hard particle systems from first principles. Following its generalization from hard spheres to general convex hard particles Hansen-Goos and Mecke (2009, 2010), it has been successfully refined Wittmann et al. (2014); Wittmann et al. (2015a); Wittmann (2015); Wittmann et al. (2016); Marechal et al. (2017) and applied to study the bulk phase behavior and various inhomogeneous systems of (mostly) uniaxial rod-like particles Marechal and Löwen (2013); Wittmann and Mecke (2014); Wittmann et al. (2014); Wittmann et al. (2015a, b); Wittmann et al. (2016); Marechal et al. (2017); Schönhöfer et al. (2018). For hard spherocylinders in particular, FMT has been tested in detail and proven very reliable for a large range of problems Wittmann et al. (2016). A prominent example for polar particles investigated by FMT are those of pear-like shape Schönhöfer et al. (2018). Although triangular prisms have also been studied using FMT Marechal et al. (2017), they were implicitly assumed as uniaxial and only shapes were considered at which no biaxial order was found in simulations. In turn, a particular version of FMT for parallel hard parallelepipeds was used to treat biaxiality under the assumption of restricted orientations Martínez-Ratón et al. (2011).

Refer to caption
Figure 1: Parameterizations of the hard uniaxial and biaxial bodies considered in this work. Uniaxial cylinders (left) and cones (middle) are characterized by their (base) diameter D𝐷Ditalic_D and height H𝐻Hitalic_H, which determine their common aspect ratio l=H/D𝑙𝐻𝐷l=H/Ditalic_l = italic_H / italic_D. The mantle length L𝐿Litalic_L of a cylinder (L=H𝐿𝐻L=Hitalic_L = italic_H) differs from that of a cone (L=H2+D2/4𝐿superscript𝐻2superscript𝐷24L=\sqrt{H^{2}+D^{2}/4}italic_L = square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 end_ARG). Biaxial isosceles spherotriangles (right) of diameter D𝐷Ditalic_D are the parallel sets at distance D/2𝐷2D/2italic_D / 2 of flat isosceles triangles with base length A𝐴Aitalic_A and side lengths B𝐵Bitalic_B (i.e., particles comprised of all points with distance equal or less than D/2𝐷2D/2italic_D / 2 to a flat isosceles triangle). We define their aspect ratio l=(A+2B)/(2D)𝑙𝐴2𝐵2𝐷l=(A+2B)/(2D)italic_l = ( italic_A + 2 italic_B ) / ( 2 italic_D ), their shape ratio x=A/(2B)𝑥𝐴2𝐵x=A/(2B)italic_x = italic_A / ( 2 italic_B ) and their opening angle 2γ2𝛾2\gamma2 italic_γ, such that x=sinγ𝑥𝛾x=\sin\gammaitalic_x = roman_sin italic_γ.

The general strength of FMT is the possibility to gain analytic insight into the phase transitions between ordered phases, because it directly connects anisotropic shape information to the particle orientation. Analytic theories describing orientational ordering phenomena often rely on the use of orientational order parameters, which measure the degree of orientational order in a system. The most widely used one is the nematic order parameter measuring alignment of rod-like particles. In systems of particles with complex shapes, different types of orientational order are possible, such that it is useful to employ a set of several order parameters Rosso (2007); te Vrugt and Wittkowski (2020a) that measures orientational order with respect to different particle axes. While it is well understood by now how the free energy for a homogeneous fluid of uniaxial particles can be expressed as a function of the average number density and the nematic order parameter Hansen-Goos and Mecke (2010); Wittmann et al. (2014, 2017), the free energy and, consequently, the phase behavior of biaxial particles as a function of appropriate orientational order parameters remains to be investigated. In particular, it is not known how these order parameters can be incorporated into FMT.

In this work, we apply FMT to hard particles with uniaxial, polar and biaxial shapes, specifically to cylinders, cones and spherotriangles, as illustrated in Fig. 1, taking into account all orientational degrees of freedom. We demonstrate how shape polarity affects the isotropic–nematic phase boundary by comparing a polar hard cone to an apolar hard cylinder and provide detailed phase diagrams for the homogeneous phases of hard isosceles spherotriangles, also resolving the transitions between different uniaxial nematic phases and a biaxial phase. In doing so, we also provide in Sec. II a generally applicable recipe for how to incorporate a certain set of relevant order parameters into FMT. This general theory allows us to overcome the need to use different assumptions to describe different phase transitions, while still providing analytic insight. These results are discussed in Sec. III.

II Theory of orientational order

II.1 Order parameters

II.1.1 Uniaxial and biaxial particles

The orientation of a hard particle in three spatial dimensions can in general be specified using a set of three angles θ[0,π]𝜃0𝜋\theta\in[0,\pi]italic_θ ∈ [ 0 , italic_π ], ϕ[0,2π]italic-ϕ02𝜋\phi\in[0,2\pi]italic_ϕ ∈ [ 0 , 2 italic_π ] and ψ[0,2π]𝜓02𝜋\psi\in[0,2\pi]italic_ψ ∈ [ 0 , 2 italic_π ], known as the Euler angles Gray and Gubbins (1984). Here, the angles θ𝜃\thetaitalic_θ and ϕitalic-ϕ\phiitalic_ϕ are the angles of spherical polar coordinates determining the orientation of a particle’s main axis, the third angle ψ𝜓\psiitalic_ψ then specifies how the particle has to be rotated around this axis in order to get into a certain position. If the particle has an axis of continuous rotational symmetry - a typical example for this would be a rod - the axis whose orientation is specified by θ𝜃\thetaitalic_θ and ϕitalic-ϕ\phiitalic_ϕ can be conveniently chosen to be the symmetry axis. In this case, the angle ψ𝜓\psiitalic_ψ has no physical relevance and can be ignored, such that the particle’s orientation is fully specified by only two angles. Such a particle is referred to as an uniaxial particle, a particle without such a symmetry is a biaxial particle.

II.1.2 Orientational ordering tensors

Whether or not a system of particles is in an ordered phase can be measured using orientational order parameters. These can be systematically defined by expanding the orientational distribution function into Cartesian tensors Cremer et al. (2012); te Vrugt and Wittkowski (2020a, b). At second order and for uniaxial particles, this expansion gives rise to the nematic tensor Qijsubscript𝑄𝑖𝑗Q_{ij}italic_Q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT, which is a symmetric traceless tensor. The eigenvalue S𝑆Sitalic_S of Qijsubscript𝑄𝑖𝑗Q_{ij}italic_Q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT with the largest absolute value measures the degree of nematic order, the corresponding eigenvector is the nematic director n𝑛\vec{n}over→ start_ARG italic_n end_ARG Andrienko (2018). If the two other eigenvalues are equal to S/2𝑆2-S/2- italic_S / 2, the system is in an uniaxial (nematic) phase. On the other hand, if the system has three distinct eigenvalues, it is in a biaxial phase Luckhurst and Sluckin (2015).

The distinction between uniaxial and biaxial can thus be made both regarding the particle shapes and regarding the ordered phases. Biaxial particles can also form a uniaxial phase. In principle, uniaxial particles can also form a biaxial phase (if their axes are ordered in such a way that their nematic tensor has three distinct eigenvalues). However, in the absence of additional external influences, phase biaxiality in practice typically only arises from particle biaxiality.

Which order parameters are appropriate, and which ordered phases the system can exhibit, depends on the symmetries of the particles. For example, in a system of square cuboids where the edges have different lengths, one can distinguish a phase in which edges of the same length tend to be parallel from a phase in which edges of the same length are either parallel or orthogonal to each other. For cubes, where all edges have the same length, this distinction would not be meaningful since these phases would be physically indistinguishable.



Refer to caption
Figure 2: Relation between the body frame (m1,m2,m3)subscript𝑚1subscript𝑚2subscript𝑚3(\vec{m}_{1},\vec{m}_{2},\vec{m}_{3})( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) and the lab frame (l1,l2,l3)subscript𝑙1subscript𝑙2subscript𝑙3(\vec{l}_{1},\vec{l}_{2},\vec{l}_{3})( over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT , over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT , over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ). (a) Expected alignment of the uniaxial nematic director n𝑛\vec{n}over→ start_ARG italic_n end_ARG (specified in the lab frame such that l3||n\vec{l}_{3}\,||\,\vec{n}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | | over→ start_ARG italic_n end_ARG) for different shape ratios x𝑥xitalic_x of a spherotriangle, compare Fig. 1, as indicated by the bottom arrow. For small or large x𝑥xitalic_x, the body has a prolate shape and we expect the height or the base to align with n𝑛\vec{n}over→ start_ARG italic_n end_ARG, respectively. For oblate shapes at intermediate x𝑥xitalic_x, we expect n𝑛\vec{n}over→ start_ARG italic_n end_ARG to be perpendicular to the face of the spherotriangle. In each case, the blue coordinates indicate the body frame chosen such that the main particle axis is parallel to m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, while the red coordinates indicate the body frame chosen such that m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT always specifies the direction of the height. Both conventions result in a different interpretation of the order parameters, which can be identified according to substitutions in Eq. (8) and Eq. (10), as indicated by the double arrows with corresponding permutation operators τ1/2subscript𝜏12\tau_{1/2}italic_τ start_POSTSUBSCRIPT 1 / 2 end_POSTSUBSCRIPT annotated. Further details are given in the text and Ref. Rosso, 2007. In these transformations, the sign of misubscript𝑚𝑖\vec{m}_{i}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT is irrelevant for the set of order parameters considered here. (b) Illustration of the three Euler angles, transforming the body frame (blue) to the lab frame (black) via the rotation matrix ^^\hat{\mathcal{R}}over^ start_ARG caligraphic_R end_ARG from Eq. (11): (i) both coordinate frames are initially aligned, in particular l3=m3subscript𝑙3subscript𝑚3\vec{l}_{3}=\vec{m}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, which is assumed for a proper interpretation of the order parameters defined in Sec. II.1.3; (ii) a rotation by ψ𝜓\psiitalic_ψ around l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT specifies the orientation perpendicular to the body’s main axis (irrelevant for uniaxial particles); (iii) a rotation by θ𝜃\thetaitalic_θ around l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT specifies the polar orientation of the body’s main axis; (iv) a rotation by ϕitalic-ϕ\phiitalic_ϕ around l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT specifies the azimuthal orientation of the body’s main axis.

This implies that systems of particles whose shape is more complex than that of rods (the most widely studied particle type in liquid crystal physics) can exhibit a much richer phase behavior and require more sophisticated order parameters te Vrugt and Wittkowski (2020a); Rosso (2007). These can also be defined in terms of a systematic expansion of the distribution function in Wigner matrices or (equivalently) Cartesian tensors. Among experimentalists, the Saupe ordering matrix Allender and Lee (1984); Luckhurst and Sluckin (2015)

Sijαβ=123(milα)(mjlβ)δijδαβsuperscriptsubscript𝑆𝑖𝑗𝛼𝛽12expectation3subscript𝑚𝑖subscript𝑙𝛼subscript𝑚𝑗subscript𝑙𝛽subscript𝛿𝑖𝑗subscript𝛿𝛼𝛽S_{ij}^{\alpha\beta}=\frac{1}{2}\braket{3\,(\vec{m}_{i}\cdot\vec{l}_{\alpha})(% \vec{m}_{j}\cdot\vec{l}_{\beta})-\delta_{ij}\delta_{\alpha\beta}}italic_S start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ⟨ start_ARG 3 ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ) ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_β end_POSTSUBSCRIPT ) - italic_δ start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT italic_δ start_POSTSUBSCRIPT italic_α italic_β end_POSTSUBSCRIPT end_ARG ⟩ (1)

with the orientational average expectation\braket{\cdot}⟨ start_ARG ⋅ end_ARG ⟩ is particularly popular. Here, the orthonormal sets {mi}subscript𝑚𝑖\{\vec{m}_{i}\}{ over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT }, i=1,2,3𝑖123i=1,2,3italic_i = 1 , 2 , 3 and {lα}subscript𝑙𝛼\{\vec{l}_{\alpha}\}{ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT }, α=1,2,3𝛼123\alpha=1,2,3italic_α = 1 , 2 , 3 constitute a basis fixed in the lab frame and the molecular frame, respectively. The reason one needs a 3×3×3×333333\times 3\times 3\times 33 × 3 × 3 × 3 matrix (namely Sijαβsuperscriptsubscript𝑆𝑖𝑗𝛼𝛽S_{ij}^{\alpha\beta}italic_S start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT) in the case of biaxial particles rather than a 3×3333\times 33 × 3 matrix (namely Qijsubscript𝑄𝑖𝑗Q_{ij}italic_Q start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT) for uniaxial particle is that, in the biaxial case, the specification of the particle orientation requires two orthogonal orientation vectors, with the Euler angle ψ𝜓\psiitalic_ψ specifying the second one. The matrix Sijαβsuperscriptsubscript𝑆𝑖𝑗𝛼𝛽S_{ij}^{\alpha\beta}italic_S start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α italic_β end_POSTSUPERSCRIPT roughly corresponds to the second order in the Cartesian expansion of a distribution function depending on all three Euler angles, see Ref. Turzi, 2011 for more details.

II.1.3 Orientational order parameters

For a particle shape which belongs to the D2hsubscript𝐷2hD_{2\mathrm{h}}italic_D start_POSTSUBSCRIPT 2 roman_h end_POSTSUBSCRIPT symmetry group (i.e., with the same symmetry as a cuboid), one can show Rosso (2007) that the Saupe ordering matrix has only four independent elements. In this case, it is convenient to work not directly with elements of the Saupe ordering matrix, but with the order parameters

S𝑆\displaystyle Sitalic_S =32(m3l3)213,absent32expectationsuperscriptsubscript𝑚3subscript𝑙3213\displaystyle=\frac{3}{2}\braket{(\vec{m}_{3}\cdot\vec{l}_{3})^{2}-\frac{1}{3}% }\,,= divide start_ARG 3 end_ARG start_ARG 2 end_ARG ⟨ start_ARG ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - divide start_ARG 1 end_ARG start_ARG 3 end_ARG end_ARG ⟩ , (2)
U𝑈\displaystyle Uitalic_U =32(m1l3)2(m2l3)2,absent32expectationsuperscriptsubscript𝑚1subscript𝑙32superscriptsubscript𝑚2subscript𝑙32\displaystyle=\frac{\sqrt{3}}{2}\braket{(\vec{m}_{1}\cdot\vec{l}_{3})^{2}-(% \vec{m}_{2}\cdot\vec{l}_{3})^{2}}\,,= divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG ⟨ start_ARG ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ⟩ , (3)
P𝑃\displaystyle Pitalic_P =32(m3l1)2(m3l2)2,absent32expectationsuperscriptsubscript𝑚3subscript𝑙12superscriptsubscript𝑚3subscript𝑙22\displaystyle=\frac{\sqrt{3}}{2}\braket{(\vec{m}_{3}\cdot\vec{l}_{1})^{2}-(% \vec{m}_{3}\cdot\vec{l}_{2})^{2}}\,,= divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG ⟨ start_ARG ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ⟩ , (4)
F𝐹\displaystyle Fitalic_F =12(m1l1)2(m1l2)2(m2l1)2+(m2l2)2,absent12expectationsuperscriptsubscript𝑚1subscript𝑙12superscriptsubscript𝑚1subscript𝑙22superscriptsubscript𝑚2subscript𝑙12superscriptsubscript𝑚2subscript𝑙22\displaystyle=\frac{1}{2}\braket{(\vec{m}_{1}\cdot\vec{l}_{1})^{2}-(\vec{m}_{1% }\cdot\vec{l}_{2})^{2}-(\vec{m}_{2}\cdot\vec{l}_{1})^{2}+(\vec{m}_{2}\cdot\vec% {l}_{2})^{2}}\,,\ \ = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ⟨ start_ARG ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + ( over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ⟩ , (5)

which can be expressed as linear combinations of the four independent matrix elements. The physical meaning of these parameters is as follows:

  • S𝑆Sitalic_S, the uniaxial nematic order parameter, measures whether the axis m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT of the molecules is aligned with the lab axis l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. If m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is the symmetry axis of a rod and l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is chosen to align with the uniaxial nematic director n𝑛\vec{n}over→ start_ARG italic_n end_ARG, then S𝑆Sitalic_S is simply the standard nematic order parameter.

  • U𝑈Uitalic_U, the molecular biaxiality order parameter, measures whether there is molecular biaxiality in a uniaxial phase. If there is a physical difference between the molecular axes m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2\vec{m}_{2}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, then it is likely that, on average, there is a difference regarding their probability of being aligned with l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT.

  • P𝑃Pitalic_P, the phase biaxiality order parameter, measures whether the axis m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is aligned preferably with l1subscript𝑙1\vec{l}_{1}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT or l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. In a perfect uniaxial nematic phase, m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT would always be aligned with l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. If there are deviations from this alignment, they can either be random (P=0𝑃0P=0italic_P = 0) or have a preferred direction. In this scenario (biaxial nematic), there are two preferred axes rather than one. Note that P𝑃Pitalic_P can be nonzero even if the particles are uniaxial since it only depends on m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (whether that will happen in an actual physical system is another matter).

  • F𝐹Fitalic_F, the full biaxiality order parameter, measures whether there is orientational ordering with respect to the axes m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2\vec{m}_{2}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT. A system with nonzero F𝐹Fitalic_F is fully biaxial in the sense that both particle shape and phase are biaxial Rosso (2007). If, for example, m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT is perfectly aligned with l1subscript𝑙1\vec{l}_{1}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m2subscript𝑚2\vec{m}_{2}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT with l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (which, since l1subscript𝑙1\vec{l}_{1}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT can be chosen according to the preferred particle orientations, essentially just means hat the vectors m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT of all particles are aligned), then F𝐹Fitalic_F is maximal. On the other hand, if m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT has equal probability of being aligned with l1subscript𝑙1\vec{l}_{1}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT (uniform distribution of m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT), then F𝐹Fitalic_F vanishes.

Moreover, we will demonstrate in this work that the order parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F can be used not only for particles with D2hsubscript𝐷2hD_{2\mathrm{h}}italic_D start_POSTSUBSCRIPT 2 roman_h end_POSTSUBSCRIPT symmetry, but even for certain polar particles. Cone-like particles, for example, are polar (a mirror reflection at the base changes their physical state). However, while they may exhibit local polar order in spatially inhomogeneous configurations, they do not form spatially homogeneous polar phases Kubala et al. (2023). Since our focus lies on orientational order in spatially homogeneous phases here, we can, despite the fact that we consider particles with polar symmetries, restrict ourselves to the parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F.

II.1.4 Choice of coordinate systems

The description of a certain physical state of a system in terms of S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F is not unique but it depends on the alignment of both the lab and the body frame, as we will elaborate below and illustrate in Fig. 2. In choosing these coordinate frames, we aim to follow the convention that S𝑆Sitalic_S represents the standard nematic order parameter (the implication of alternative conventions are exemplified in Sec. II.3.3).

To fix the coordinate system of the lab frame specified by l1subscript𝑙1\vec{l}_{1}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, we must indicate the direction of the uniaxial director n𝑛\vec{n}over→ start_ARG italic_n end_ARG. For example, perfect uniaxial order only corresponds to the case S=1𝑆1S=1italic_S = 1 and P=0𝑃0P=0italic_P = 0 if we choose l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT to align with n𝑛\vec{n}over→ start_ARG italic_n end_ARG. Suppose that we instead choose l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT to be aligned with n𝑛\vec{n}over→ start_ARG italic_n end_ARG. Then we would — even though the system looks exactly the same — have S=1/2𝑆12S=-1/2italic_S = - 1 / 2 and P=3/2𝑃32P=\sqrt{3}/2italic_P = square-root start_ARG 3 end_ARG / 2 instead of S=1𝑆1S=1italic_S = 1 and P=0𝑃0P=0italic_P = 0. To avoid this problem and to ensure that the order parameters have a clear physical interpretation that coincides with the usual one in the uniaxial limit, we always choose the lab frame to be aligned with the director frame, i.e., l3||n\vec{l}_{3}\,||\,\vec{n}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT | | over→ start_ARG italic_n end_ARG. It will be shown that in this case the parameter P𝑃Pitalic_P is irrelevant for describing phase transitions in homogeneous bulk systems. The parameter U𝑈Uitalic_U, despite being relevant for characterizing the orientational distribution, will turn out not to play a major role for the phase behavior. This leaves us with the two central order parameters S𝑆Sitalic_S and F𝐹Fitalic_F, which we use to map out our phase diagrams, where S𝑆Sitalic_S measures nematic order in the standard way and a nonzero F𝐹Fitalic_F then shows that biaxial order is present.

The coordinate system needs to be fixed not only for the lab frame, but also for the body frame, specified by m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT, m2subscript𝑚2\vec{m}_{2}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. For uniaxial particles, the symmetry axis must be chosen to be parallel to m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT irrespective of their shape. Otherwise, one would treat the particle as if it was biaxial, since S𝑆Sitalic_S would not measure the alignment of the symmetry axis to the director and U𝑈Uitalic_U or F𝐹Fitalic_F may not vanish. This choice of the main axis is, however, not as obvious for biaxial particles. Take, as an example, an isosceles spherotriangle (made of a triangle with two sides of equal length). If the base of such a triangle (the side whose length is different) is much longer than the other two sides, the particles are expected to exhibit nematic order with the director pointing along the base of the triangles. If, on the other hand, the base is sufficiently short, the particles exhibit nematic order with the director pointing in a direction being orthogonal to the base.

Specifically, in this work, we will consider the three different uniaxial nematic phases, which we distinguish by their director orientations assumed for a certain typical particle shape, as depicted in Fig. 2 for three representative cases: (i) a prolate nematic, Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT, in which the triangles align to the director with their height (when it is much larger than the base length), (ii) an oblate nematic, Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT, in which the triangles align to the director perpendicular to their face (when height and base are of comparable length such that the particle has an oblate shape), and, (iii) a prolate nematic, Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT, in which the triangles align to the director with their base (when it is much larger than their height). We thereby use the words “prolate” and “oblate” for both phases and, as one usually does, for particle shapes. Note that, similar to the notion of “uniaxial” and “biaxial”, prolate/oblate particle shapes and the emergence of prolate/oblate nematic phases are closely related. This leaves us with two options for choosing the direction m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (which we identify with the z𝑧zitalic_z-axis) in the body frame. First, we can choose it parallel to the axis along which the particles presumably align in the case of uniaxial nematic order, which depends on the relative lengths of the different sides of the particle (black coordinate systems drawn in Fig. 2). This option allows us to follow the convention described above to disregard the order parameter P𝑃Pitalic_P for each presumed director orientation and is employed in this work. Second, we can choose to always align m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT with, say the height of the triangle (red coordinate systems drawn in Fig. 2). This option is more general as there is no need to input assumptions a priori, which may impose a bias towards certain director orientations, but it requires to take into account all four order parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F without a clear interpretation.

Finally, we note that there exists a well-defined relation between the order parameters obtained from different choices of unit vectors in the body frame, which we summarize below. Suppose we chose the second option and align the triangle height with the vector m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT of the body frame. If we then want to investigate uniaxial nematic order along the triangle height, Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT, the first option (aligning the particular axis of interest) is trivially equivalent, as indicated in the first panel of Fig. 2. For uniaxial nematic order along the triangle faces, Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT, choosing the first option would amount to exchanging the two vectors m2subscript𝑚2\vec{m}_{2}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT, which defines the action of the permutation operator τ1subscript𝜏1\tau_{1}italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT on the basis of the body frame, as indicated in the second panel of Fig. 2. As detailed in Ref. Rosso, 2007, this procedure allows us to obtain the appropriate order parameters by taking the result of the second option and making the substitutions

S𝑆\displaystyle Sitalic_S τ112S32U,Uτ132S+12U,formulae-sequencesuperscriptsubscript𝜏1absent12𝑆32𝑈superscriptsubscript𝜏1𝑈32𝑆12𝑈\displaystyle\stackrel{{\scriptstyle\tau_{1}}}{{\rightarrow}}-\frac{1}{2}S-% \frac{\sqrt{3}}{2}U\,,\ \ \ U\stackrel{{\scriptstyle\tau_{1}}}{{\rightarrow}}-% \frac{\sqrt{3}}{2}S+\frac{1}{2}U\,,start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_S - divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_U , italic_U start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_S + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_U , (7)
P𝑃\displaystyle Pitalic_P τ112P32F,Fτ132P+12F.formulae-sequencesuperscriptsubscript𝜏1absent12𝑃32𝐹superscriptsubscript𝜏1𝐹32𝑃12𝐹\displaystyle\stackrel{{\scriptstyle\tau_{1}}}{{\rightarrow}}-\frac{1}{2}P-% \frac{\sqrt{3}}{2}F\,,\ \ \ F\stackrel{{\scriptstyle\tau_{1}}}{{\rightarrow}}-% \frac{\sqrt{3}}{2}P+\frac{1}{2}F\,.start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_P - divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_F , italic_F start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_P + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_F . (8)

The same can be done for uniaxial nematic order along the triangle base, Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT, where m1subscript𝑚1\vec{m}_{1}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT are exchanged by τ2subscript𝜏2\tau_{2}italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, as indicated in the third panel of Fig. 2. The appropriate substitutions are then found to be

S𝑆\displaystyle Sitalic_S τ212S+32U,Uτ232S+12U,formulae-sequencesuperscriptsubscript𝜏2absent12𝑆32𝑈superscriptsubscript𝜏2𝑈32𝑆12𝑈\displaystyle\stackrel{{\scriptstyle\tau_{2}}}{{\rightarrow}}-\frac{1}{2}S+% \frac{\sqrt{3}}{2}U\,,\ \ \ U\stackrel{{\scriptstyle\tau_{2}}}{{\rightarrow}}% \frac{\sqrt{3}}{2}S+\frac{1}{2}U\,,start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_S + divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_U , italic_U start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_RELOP divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_S + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_U , (9)
P𝑃\displaystyle Pitalic_P τ212P+32F,Fτ232P+12F.formulae-sequencesuperscriptsubscript𝜏2absent12𝑃32𝐹superscriptsubscript𝜏2𝐹32𝑃12𝐹\displaystyle\stackrel{{\scriptstyle\tau_{2}}}{{\rightarrow}}-\frac{1}{2}P+% \frac{\sqrt{3}}{2}F\,,\ \ \ F\stackrel{{\scriptstyle\tau_{2}}}{{\rightarrow}}% \frac{\sqrt{3}}{2}P+\frac{1}{2}F\,.start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_RELOP - divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_P + divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_F , italic_F start_RELOP SUPERSCRIPTOP start_ARG → end_ARG start_ARG italic_τ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG end_RELOP divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG italic_P + divide start_ARG 1 end_ARG start_ARG 2 end_ARG italic_F . (10)

An alternative point of view on these substitutions is discussed in Sec. II.3.3 for a scenario with perfect uniaxial order.

II.1.5 Representation in terms of Euler angles

The definitions of the order parameters (LABEL:eq_Xall) are relatively general and require, for practical calculations, a specification of the average expectation\braket{\cdot}⟨ start_ARG ⋅ end_ARG ⟩ that is used. We assume that the system can be characterized by a distribution g(ϕ,θ,ψ)𝑔italic-ϕ𝜃𝜓g(\phi,\theta,\psi)italic_g ( italic_ϕ , italic_θ , italic_ψ ) that gives the probability of finding a particle with an orientation specified by the Euler angles ϕitalic-ϕ\phiitalic_ϕ, θ𝜃\thetaitalic_θ and ψ𝜓\psiitalic_ψ, which connect the body frame to the lab frame. Here, we follow the convention of Ref. Rosso, 2007 and use the y𝑦yitalic_y-notation. As illustrated in Fig. 2b, the body frame is first rotated by the angle ψ[0,2π]𝜓02𝜋\psi\in[0,2\pi]italic_ψ ∈ [ 0 , 2 italic_π ] around l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (which coincides by default with its z𝑧zitalic_z-axis, i.e., m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT). This is an identity map for uniaxial bodies. It is followed by a second rotation by the angle θ[0,π]𝜃0𝜋\theta\in[0,\pi]italic_θ ∈ [ 0 , italic_π ] around l2subscript𝑙2\vec{l}_{2}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and a third one by the angle ϕ[0,2π]italic-ϕ02𝜋\phi\in[0,2\pi]italic_ϕ ∈ [ 0 , 2 italic_π ] around l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. These latter two rotations map the z𝑧zitalic_z-axis of the body frame onto the unit sphere.

Any vector can be transferred from the body frame to the lab frame via the rotation matrix

^:=R^3(ϕ)R^2(θ)R^3(ψ)assign^subscript^𝑅3italic-ϕsubscript^𝑅2𝜃subscript^𝑅3𝜓\displaystyle\hat{\mathcal{R}}:=\hat{R}_{3}(\phi)\cdot\hat{R}_{2}(\theta)\cdot% \hat{R}_{3}(\psi)over^ start_ARG caligraphic_R end_ARG := over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_ϕ ) ⋅ over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( italic_θ ) ⋅ over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ( italic_ψ ) (11)

accounting for all three Euler angles. Following the convention of using extrinsic rotations, R^α(γ)subscript^𝑅𝛼𝛾\hat{R}_{\alpha}(\gamma)over^ start_ARG italic_R end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT ( italic_γ ) is the rotation matrix describing a clockwise rotation by the angle γ𝛾\gammaitalic_γ around the axis of the α𝛼\alphaitalic_αth unit vector lαsubscript𝑙𝛼\vec{l}_{\alpha}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT of the lab frame. Explicitly, the components of ^^\hat{\mathcal{R}}over^ start_ARG caligraphic_R end_ARG read

^11subscript^11\displaystyle\hat{\mathcal{R}}_{11}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =cos(ϕ)cos(θ)cos(ψ)sin(ϕ)sin(ψ),absentitalic-ϕ𝜃𝜓italic-ϕ𝜓\displaystyle=\cos(\phi)\cos(\theta)\cos(\psi)-\sin(\phi)\sin(\psi)\,,= roman_cos ( italic_ϕ ) roman_cos ( italic_θ ) roman_cos ( italic_ψ ) - roman_sin ( italic_ϕ ) roman_sin ( italic_ψ ) , (12)
^12subscript^12\displaystyle\hat{\mathcal{R}}_{12}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT =cos(ψ)sin(θ)cos(ϕ)cos(θ)sin(ψ),absent𝜓𝜃italic-ϕ𝜃𝜓\displaystyle=-\cos(\psi)\sin(\theta)-\cos(\phi)\cos(\theta)\sin(\psi)\,,= - roman_cos ( italic_ψ ) roman_sin ( italic_θ ) - roman_cos ( italic_ϕ ) roman_cos ( italic_θ ) roman_sin ( italic_ψ ) , (13)
^13subscript^13\displaystyle\hat{\mathcal{R}}_{13}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT =cos(ϕ)sin(θ),absentitalic-ϕ𝜃\displaystyle=\cos(\phi)\sin(\theta)\,,= roman_cos ( italic_ϕ ) roman_sin ( italic_θ ) , (14)
^21subscript^21\displaystyle\hat{\mathcal{R}}_{21}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT =cos(ϕ)sin(ψ)+cos(θ)cos(ψ)sin(ϕ),absentitalic-ϕ𝜓𝜃𝜓italic-ϕ\displaystyle=\cos(\phi)\sin(\psi)+\cos(\theta)\cos(\psi)\sin(\phi)\,,= roman_cos ( italic_ϕ ) roman_sin ( italic_ψ ) + roman_cos ( italic_θ ) roman_cos ( italic_ψ ) roman_sin ( italic_ϕ ) , (15)
^22subscript^22\displaystyle\hat{\mathcal{R}}_{22}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT =cos(ϕ)cos(ψ)cos(θ)sin(ϕ)sin(ψ),absentitalic-ϕ𝜓𝜃italic-ϕ𝜓\displaystyle=\cos(\phi)\cos(\psi)-\cos(\theta)\sin(\phi)\sin(\psi)\,,= roman_cos ( italic_ϕ ) roman_cos ( italic_ψ ) - roman_cos ( italic_θ ) roman_sin ( italic_ϕ ) roman_sin ( italic_ψ ) , (16)
^23subscript^23\displaystyle\hat{\mathcal{R}}_{23}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 23 end_POSTSUBSCRIPT =sin(ϕ)sin(θ),absentitalic-ϕ𝜃\displaystyle=\sin(\phi)\sin(\theta)\,,= roman_sin ( italic_ϕ ) roman_sin ( italic_θ ) , (17)
^31subscript^31\displaystyle\hat{\mathcal{R}}_{31}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT =cos(ψ)sin(θ),absent𝜓𝜃\displaystyle=-\cos(\psi)\sin(\theta)\,,= - roman_cos ( italic_ψ ) roman_sin ( italic_θ ) , (18)
^32subscript^32\displaystyle\hat{\mathcal{R}}_{32}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT =sin(θ)sin(ψ),absent𝜃𝜓\displaystyle=\sin(\theta)\sin(\psi)\,,= roman_sin ( italic_θ ) roman_sin ( italic_ψ ) , (19)
^33subscript^33\displaystyle\hat{\mathcal{R}}_{33}over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =cos(θ),absent𝜃\displaystyle=\cos(\theta)\,,= roman_cos ( italic_θ ) , (20)

and these can be conveniently expressed in terms of the basis vectors of the director and body frame as

^αi=milα,subscript^𝛼𝑖subscript𝑚𝑖subscript𝑙𝛼\displaystyle\hat{\mathcal{R}}_{\alpha i}=\vec{m}_{i}\cdot\vec{l}_{\alpha}\,,over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_α italic_i end_POSTSUBSCRIPT = over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT italic_α end_POSTSUBSCRIPT , (21)

which allows us to make contact to the order parameters, as defined in Eq. (LABEL:eq_Xall).

In terms of the Euler angles (ϕ,θ,ψ)italic-ϕ𝜃𝜓(\phi,\theta,\psi)( italic_ϕ , italic_θ , italic_ψ ) and an appropriate orientational distribution g(ϕ,θ,ψ)𝑔italic-ϕ𝜃𝜓g(\phi,\theta,\psi)italic_g ( italic_ϕ , italic_θ , italic_ψ ), we can rewrite the four order parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F as

X=d𝐎g(ϕ,θ,ψ)fX(ϕ,θ,ψ),X{S,U,P,F},formulae-sequence𝑋differential-d𝐎𝑔italic-ϕ𝜃𝜓subscript𝑓𝑋italic-ϕ𝜃𝜓𝑋𝑆𝑈𝑃𝐹X=\int\mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,f_{X}(\phi,\theta,\psi)\,,\ % \ \ X\in\{S,U,P,F\}\,,italic_X = ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) italic_f start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) , italic_X ∈ { italic_S , italic_U , italic_P , italic_F } , (22)

where the angular integral reads

d𝐎:=18π202πdϕ0πdθsin(θ)02πdψassigndifferential-d𝐎18superscript𝜋2superscriptsubscript02𝜋differential-ditalic-ϕsuperscriptsubscript0𝜋differential-d𝜃𝜃superscriptsubscript02𝜋differential-d𝜓\int\mathrm{d}\mathbf{O}:=\frac{1}{8\pi^{2}}\int_{0}^{2\pi}\mathrm{d}\phi\int_% {0}^{\pi}\mathrm{d}\theta\sin(\theta)\int_{0}^{2\pi}\mathrm{d}\psi∫ roman_d bold_O := divide start_ARG 1 end_ARG start_ARG 8 italic_π start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT roman_d italic_ϕ ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_π end_POSTSUPERSCRIPT roman_d italic_θ roman_sin ( italic_θ ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 italic_π end_POSTSUPERSCRIPT roman_d italic_ψ (23)

and the functions

fS(ϕ,θ,ψ)subscript𝑓𝑆italic-ϕ𝜃𝜓\displaystyle f_{S}(\phi,\theta,\psi)italic_f start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =12(3cos2(θ)1),absent123superscript2𝜃1\displaystyle=\frac{1}{2}\left(3\cos^{2}(\theta)-1\right)\,,= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 3 roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ ) - 1 ) , (24)
fU(ϕ,θ,ψ)subscript𝑓𝑈italic-ϕ𝜃𝜓\displaystyle f_{U}(\phi,\theta,\psi)italic_f start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =32sin2(θ)cos(2ψ),absent32superscript2𝜃2𝜓\displaystyle=\frac{\sqrt{3}}{2}\sin^{2}(\theta)\cos(2\psi)\,,= divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ ) roman_cos ( 2 italic_ψ ) , (25)
fP(ϕ,θ,ψ)subscript𝑓𝑃italic-ϕ𝜃𝜓\displaystyle f_{P}(\phi,\theta,\psi)italic_f start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =32sin2(θ)cos(2ϕ),absent32superscript2𝜃2italic-ϕ\displaystyle=\frac{\sqrt{3}}{2}\sin^{2}(\theta)\cos(2\phi)\,,= divide start_ARG square-root start_ARG 3 end_ARG end_ARG start_ARG 2 end_ARG roman_sin start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ ) roman_cos ( 2 italic_ϕ ) , (26)
fF(ϕ,θ,ψ)subscript𝑓𝐹italic-ϕ𝜃𝜓\displaystyle f_{F}(\phi,\theta,\psi)italic_f start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =12(1+cos2(θ))cos(2ϕ)cos(2ψ)absent121superscript2𝜃2italic-ϕ2𝜓\displaystyle=\frac{1}{2}\left(1+\cos^{2}(\theta)\right)\cos(2\phi)\cos(2\psi)= divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( 1 + roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_θ ) ) roman_cos ( 2 italic_ϕ ) roman_cos ( 2 italic_ψ ) (27)
cos(θ)sin(2ϕ)sin(2ψ),𝜃2italic-ϕ2𝜓\displaystyle\ \ \ \;-\cos(\theta)\sin(2\phi)\sin(2\psi)\,,- roman_cos ( italic_θ ) roman_sin ( 2 italic_ϕ ) roman_sin ( 2 italic_ψ ) , (28)

are specified by inserting the relation from Eq. (21) and the explicit expressions from Eq. (20) into Eq. (LABEL:eq_Xall). In this explicit representation of the order parameters, the choice of coordinate systems discussed in Sec. II.1.4 becomes more intuitive. The only relevant angle for the uniaxial nematic order parameter (24) is θ𝜃\thetaitalic_θ. For θ=0𝜃0\theta=0italic_θ = 0, i.e., in the case of perfect uniaxial order, m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT equals l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT (as illustrated in Fig. 2). Hence the particle axis chosen to align with m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT in the body frame sets the uniaxial director, i.e., the maximum of g𝑔gitalic_g at θ=0𝜃0\theta=0italic_θ = 0.

II.2 Fundamental measure theory (FMT)

II.2.1 Density functional theory (DFT)

Based on the celebrated Hohenberg-Kohn theorem Hohenberg and Kohn (1964) originally developed for quantum-mechanical systems, classical density functional theory (DFT) Evans (1979) states that there exists a unique functional Ω[ρ]Ωdelimited-[]𝜌\Omega[\rho]roman_Ω [ italic_ρ ] of the classical number density ρ𝜌\rhoitalic_ρ, which gets minimal in equilibrium, i.e., when inserting the equilibrium density ρ=ρ0𝜌subscript𝜌0\rho=\rho_{0}italic_ρ = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT. Then, the value of the functional equals the grand potential of the system, Ω=Ω[ρ0]ΩΩdelimited-[]subscript𝜌0\Omega=\Omega[\rho_{0}]roman_Ω = roman_Ω [ italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ]. For systems of anisotropic particles as considered here, the density ρ(𝐫,𝐎)𝜌𝐫𝐎\rho({\mathbf{r}},{\mathbf{O}})italic_ρ ( bold_r , bold_O ) depends in general on both positions 𝐫𝐫{\mathbf{r}}bold_r and orientations 𝐎𝐎{\mathbf{O}}bold_O. Thus, to determine the equilibrium configuration of such a system from a given functional Ω[ρ]Ωdelimited-[]𝜌\Omega[\rho]roman_Ω [ italic_ρ ], one has to perform a basic minimization by solving the Euler-Lagrange equation

δΩ[ρ]δρ(𝐫,𝐎)|ρ=ρ0=0.evaluated-at𝛿Ωdelimited-[]𝜌𝛿𝜌𝐫𝐎𝜌subscript𝜌00\frac{\delta\Omega[\rho]}{\delta\rho({\mathbf{r}},{\mathbf{O}})}\bigg{|}_{\rho% =\rho_{0}}=0\,.divide start_ARG italic_δ roman_Ω [ italic_ρ ] end_ARG start_ARG italic_δ italic_ρ ( bold_r , bold_O ) end_ARG | start_POSTSUBSCRIPT italic_ρ = italic_ρ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT end_POSTSUBSCRIPT = 0 . (29)

The general form of the density functional is

βΩ[ρ]=d𝐫𝛽Ωdelimited-[]𝜌differential-d𝐫\displaystyle\beta\Omega[\rho]=\int\mathrm{d}{\mathbf{r}}italic_β roman_Ω [ italic_ρ ] = ∫ roman_d bold_r (Φid(𝐫)+Φex(𝐫))subscriptΦid𝐫subscriptΦex𝐫\displaystyle\left(\Phi_{\text{id}}({\mathbf{r}})+\Phi_{\text{ex}}({\mathbf{r}% })\right)( roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT ( bold_r ) + roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r ) ) (30)
+d𝐫differential-d𝐫\displaystyle+\int\mathrm{d}{\mathbf{r}}+ ∫ roman_d bold_r d𝐎ρ(𝐫,𝐎)(βVext(𝐫,𝐎)βμ),differential-d𝐎𝜌𝐫𝐎𝛽subscript𝑉ext𝐫𝐎𝛽𝜇\displaystyle\int\mathrm{d}\mathbf{O}\,\rho({\mathbf{r}},{\mathbf{O}})(\beta V% _{\text{ext}}({\mathbf{r}},{\mathbf{O}})-\beta\mu)\,,∫ roman_d bold_O italic_ρ ( bold_r , bold_O ) ( italic_β italic_V start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , bold_O ) - italic_β italic_μ ) , (31)

where Vext(𝐫,𝐎)subscript𝑉ext𝐫𝐎V_{\text{ext}}({\mathbf{r}},{\mathbf{O}})italic_V start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , bold_O ) is an external potential acting on the particles, μ𝜇\muitalic_μ is their chemical potential and β𝛽\betaitalic_β is the inverse temperature. The free energy density of the system is split into the exactly known contribution

Φid(𝐫)=d𝐎ρ(𝐫,𝐎)(ln(λ3ρ(𝐫,𝐎))1)subscriptΦid𝐫differential-d𝐎𝜌𝐫𝐎superscript𝜆3𝜌𝐫𝐎1\Phi_{\text{id}}({\mathbf{r}})=\int\mathrm{d}\mathbf{O}\,\rho({\mathbf{r}},{% \mathbf{O}})(\ln(\lambda^{3}\rho({\mathbf{r}},{\mathbf{O}}))-1)roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT ( bold_r ) = ∫ roman_d bold_O italic_ρ ( bold_r , bold_O ) ( roman_ln ( italic_λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_ρ ( bold_r , bold_O ) ) - 1 ) (32)

of an anisotropic ideal gas with the (irrelevant) thermal volume λ3superscript𝜆3\lambda^{3}italic_λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT (comprising trivial integrals over momenta and angular momenta) and the excess free energy density Φex(𝐫)subscriptΦex𝐫\Phi_{\text{ex}}({\mathbf{r}})roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r ), which describes the interactions in the system and is not known in general.

II.2.2 Excess free energy for hard bodies

For convex hard bodies, a sophisticated expression for Φex(𝐫)subscriptΦex𝐫\Phi_{\text{ex}}({\mathbf{r}})roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( bold_r ) exists that is based on the seminal work of Rosenfeld, who found an elegant way to decompose the interaction of hard spheres in 1989 Rosenfeld (1989). His FMT functional has been further refined in later work Tarazona (2000); Hansen-Goos and Roth (2006); Roth (2010); Roth et al. (2012) and eventually generalized to more general shapes which can assemble in orientationally ordered phases Hansen-Goos and Mecke (2009); Wittmann et al. (2015a); Wittmann (2015); Wittmann et al. (2016).

In the framework of FMT, the hard-body interaction can be conveniently expressed in terms of a set (labeled by the index ν𝜈\nuitalic_ν) of weight functions ω(ν)(𝐫,𝐎)superscript𝜔𝜈𝐫𝐎\omega^{(\nu)}({\mathbf{r}},{\mathbf{O}})italic_ω start_POSTSUPERSCRIPT ( italic_ν ) end_POSTSUPERSCRIPT ( bold_r , bold_O ). These are local geometrical measures of a hard body \mathcal{B}caligraphic_B with surface \partial\mathcal{B}∂ caligraphic_B. These functions represent an interacting particle and thus both depend on its position in space and its orientation, as we will specify in Sec. II.2.3. This allows us to shift the functional dependence on the density from the excess free energy to a set of weighted densities

nν(𝐫)=d𝐫d𝐎ρ(𝐫,𝐎)ω(ν)(𝐫𝐫,𝐎),subscript𝑛𝜈𝐫differential-d𝐫differential-d𝐎𝜌superscript𝐫𝐎superscript𝜔𝜈𝐫superscript𝐫𝐎n_{\nu}({\mathbf{r}})=\int\mathrm{d}{\mathbf{r}}\int\mathrm{d}{\mathbf{O}}\,% \rho({\mathbf{r}}^{\prime},{\mathbf{O}})\,\omega^{(\nu)}({\mathbf{r}}-{\mathbf% {r}}^{\prime},{\mathbf{O}})\,,italic_n start_POSTSUBSCRIPT italic_ν end_POSTSUBSCRIPT ( bold_r ) = ∫ roman_d bold_r ∫ roman_d bold_O italic_ρ ( bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_O ) italic_ω start_POSTSUPERSCRIPT ( italic_ν ) end_POSTSUPERSCRIPT ( bold_r - bold_r start_POSTSUPERSCRIPT ′ end_POSTSUPERSCRIPT , bold_O ) , (33)

which are defined as an orientationally averaged convolution of the density and the weight functions.

The key feature of FMT is that we can write the excess free energy density in the general form

Φex=n0ln(1n3)+ϕ2(1n3)+ϕ3(1n3)2subscriptΦexsubscript𝑛01subscript𝑛3subscriptitalic-ϕ21subscript𝑛3subscriptitalic-ϕ3superscript1subscript𝑛32\Phi_{\text{ex}}=-n_{0}\ln(1-n_{3})+\frac{\phi_{2}}{(1-n_{3})}+\frac{\phi_{3}}% {(1-n_{3})^{2}}roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT = - italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT roman_ln ( 1 - italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) + divide start_ARG italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT end_ARG start_ARG ( 1 - italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) end_ARG + divide start_ARG italic_ϕ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT end_ARG start_ARG ( 1 - italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG (34)

solely as a function of weighted densities (33). Here, we use the truncated expansion

ϕ2=n1n2n1n2ζTr[n1n2]subscriptitalic-ϕ2subscript𝑛1subscript𝑛2subscript𝑛1subscript𝑛2𝜁Trdelimited-[]subscript𝑛1subscript𝑛2\phi_{2}=n_{1}n_{2}-\overrightarrow{n}_{1}\cdot\overrightarrow{n}_{2}-\zeta\,% \mbox{Tr}\!\left[\overleftrightarrow{n}_{1}\overleftrightarrow{n}_{2}\right]italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ⋅ over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT - italic_ζ Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] (35)

involving vectorial and tensorial weighted densities of rank two, the correction factor ζ=5/4𝜁54\zeta=5/4italic_ζ = 5 / 4 and the Tarazona-Rosenfeld Tarazona and Rosenfeld (1997) version of the term

ϕ3=316π(n233n2Tr[n22]+2Tr[n23]).subscriptitalic-ϕ3316𝜋superscriptsubscript𝑛233subscript𝑛2Trdelimited-[]superscriptsubscript𝑛222Trdelimited-[]superscriptsubscript𝑛23\phi_{3}=\frac{3}{16\pi}\left(n_{2}^{3}-3n_{2}\mbox{Tr}\!\left[% \overleftrightarrow{n}_{2}^{2}\right]+2\mbox{Tr}\!\left[\overleftrightarrow{n}% _{2}^{3}\right]\right)\,.italic_ϕ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT = divide start_ARG 3 end_ARG start_ARG 16 italic_π end_ARG ( italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT - 3 italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ] + 2 Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ] ) . (36)

Note that a proper choice of the terms ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and ϕ3subscriptitalic-ϕ3\phi_{3}italic_ϕ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT is particularly important for anisotropic particles, since the standard version of ϕ3subscriptitalic-ϕ3\phi_{3}italic_ϕ start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT for hard spheres, as proposed by Tarazona Tarazona (2000), leads to qualitatively and quantitatively poor results in the anisotropic case Wittmann et al. (2014) (while all versions of Eq. (34) reduce to an appropriate functional for hard spheres if the according weighted densities are used). On the other hand, more sophisticated choices of both terms are available in terms of more complicated geometrical two- and three-body measures and their appropriate expansions into rotational invariants (in the case of uniaxial particles) Schönhöfer et al. (2018). We choose to work here with the functional based on Eqs. (35) and (36) for the following reasons: It has been demonstrated in detail for hard spherocylinders that the chosen approximation of ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in terms of rank-two tensors does not result in a major offset when locating the isotropic–nematic transition (while it slightly underestimates the difference of coexistence densities). Most notably, we will demonstrate below that only using rank-two tensors allows us to make direct contact to the general order parameters introduced in Sec. II.1, which is one of the main goals of the present work. In doing so, we will gain direct analytic insight for spatially homogeneous systems instead of having to perform a tedious numerical integration, which would be required for other versions of ΦexsubscriptΦex\Phi_{\text{ex}}roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT.

II.2.3 Homogeneous weighted densities

In what follows, we are particularly interested in the bulk phase behavior. Thus, we set Vext(𝐫,𝐎)=0subscript𝑉ext𝐫𝐎0V_{\text{ext}}({\mathbf{r}},{\mathbf{O}})=0italic_V start_POSTSUBSCRIPT ext end_POSTSUBSCRIPT ( bold_r , bold_O ) = 0 in Eq. (31) and focus on spatially homogeneous (but orientationally ordered) phases. Hence, we express the orientation-dependent density ρ(𝐎)=ρg(ϕ,θ,ψ)𝜌𝐎𝜌𝑔italic-ϕ𝜃𝜓\rho({\mathbf{O}})=\rho\,g(\phi,\theta,\psi)italic_ρ ( bold_O ) = italic_ρ italic_g ( italic_ϕ , italic_θ , italic_ψ ) in terms of the homogeneous bulk density ρ𝜌\rhoitalic_ρ and the normalized orientational distribution function g(ϕ,θ,ψ)𝑔italic-ϕ𝜃𝜓g(\phi,\theta,\psi)italic_g ( italic_ϕ , italic_θ , italic_ψ ) entering in Eq. (22). With this definition, the free energy density of the ideal gas (32) simplifies to

Φid=ρ(ln(λ3ρ)1+d𝐎g(ϕ,θ,ψ)lng(ϕ,θ,ψ))subscriptΦid𝜌superscript𝜆3𝜌1differential-d𝐎𝑔italic-ϕ𝜃𝜓𝑔italic-ϕ𝜃𝜓\Phi_{\text{id}}=\rho\left(\ln(\lambda^{3}\rho)-1+\int\mathrm{d}\mathbf{O}\,g(% \phi,\theta,\psi)\ln g(\phi,\theta,\psi)\right)roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT = italic_ρ ( roman_ln ( italic_λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT italic_ρ ) - 1 + ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) roman_ln italic_g ( italic_ϕ , italic_θ , italic_ψ ) ) (37)

and the convolution in Eq. (33) turns into an ordinary integral over space. It is thus possible to write all required weighted densities in an instructive form, which directly incorporates the appropriate orientation dependence of the corresponding weight functions.

The four scalar weighted densities

n3subscript𝑛3\displaystyle n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =ρd𝐫=ρv=η,absent𝜌subscriptdifferential-d𝐫𝜌𝑣𝜂\displaystyle=\rho\int_{\mathcal{B}}\mathrm{d}{\mathbf{r}}=\rho\,v=\eta\,,= italic_ρ ∫ start_POSTSUBSCRIPT caligraphic_B end_POSTSUBSCRIPT roman_d bold_r = italic_ρ italic_v = italic_η , (38)
n2subscript𝑛2\displaystyle n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρd𝐫,absent𝜌subscriptdifferential-d𝐫\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\,,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r , (39)
n1subscript𝑛1\displaystyle n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρd𝐫(𝐫)4π,absent𝜌subscriptdifferential-d𝐫𝐫4𝜋\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\frac{% \mathcal{H}({\mathbf{r}})}{4\pi}\,,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r divide start_ARG caligraphic_H ( bold_r ) end_ARG start_ARG 4 italic_π end_ARG , (40)
n0subscript𝑛0\displaystyle n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =ρd𝐫𝒦(𝐫)4π=ρabsent𝜌subscriptdifferential-d𝐫𝒦𝐫4𝜋𝜌\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\frac{% \mathcal{K}({\mathbf{r}})}{4\pi}=\rho\,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r divide start_ARG caligraphic_K ( bold_r ) end_ARG start_ARG 4 italic_π end_ARG = italic_ρ (41)

do not depend on the orientational distribution and represent the particle volume v𝑣vitalic_v, its surface area, the surface average of the local mean curvature (𝐫)𝐫\mathcal{H}({\mathbf{r}})caligraphic_H ( bold_r ) and the surface average of the local Gaussian curvature 𝒦(𝐫)𝒦𝐫\mathcal{K}({\mathbf{r}})caligraphic_K ( bold_r ), respectively. The latter is equal to one for any (simply connected) hard body. Moreover, we have defined the packing fraction η𝜂\etaitalic_η obtained for n3subscript𝑛3n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT as the product of the density and the particle volume. The two vector densities

n2subscript𝑛2\displaystyle\overrightarrow{n}_{2}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρd𝐫d𝐎g(ϕ,θ,ψ)𝐧,absent𝜌subscriptdifferential-d𝐫differential-d𝐎𝑔italic-ϕ𝜃𝜓𝐧\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\,\int% \mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,{\mathbf{n}}\,,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) bold_n , (42)
n1subscript𝑛1\displaystyle\overrightarrow{n}_{1}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρd𝐫d𝐎g(ϕ,θ,ψ)(𝐫)4π𝐧absent𝜌subscriptdifferential-d𝐫differential-d𝐎𝑔italic-ϕ𝜃𝜓𝐫4𝜋𝐧\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\,\int% \mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,\frac{\mathcal{H}({\mathbf{r}})}{4% \pi}{\mathbf{n}}\,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) divide start_ARG caligraphic_H ( bold_r ) end_ARG start_ARG 4 italic_π end_ARG bold_n (43)

represent averages of the surface unit normal vector 𝐧(𝐫,𝐎)𝐧𝐫𝐎{\mathbf{n}}({\mathbf{r}},{\mathbf{O}})bold_n ( bold_r , bold_O ). Finally, the tensor densities

n2subscript𝑛2\displaystyle\overleftrightarrow{n}_{2}over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρd𝐫d𝐎g(ϕ,θ,ψ)𝐧𝐧T,absent𝜌subscriptdifferential-d𝐫differential-d𝐎𝑔italic-ϕ𝜃𝜓superscript𝐧𝐧𝑇\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\,\int% \mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,{\mathbf{n}}{\mathbf{n}}^{T}\,,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) bold_nn start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT , (44)
n1subscript𝑛1\displaystyle\overleftrightarrow{n}_{1}over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρd𝐫d𝐎g(ϕ,θ,ψ)absent𝜌subscriptdifferential-d𝐫differential-d𝐎𝑔italic-ϕ𝜃𝜓\displaystyle=\rho\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}\,\int% \mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,= italic_ρ ∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) (45)
×κ1(𝐫)κ2(𝐫)8π(𝐯1𝐯1T𝐯2𝐯2T)absentsubscript𝜅1𝐫subscript𝜅2𝐫8𝜋subscript𝐯1superscriptsubscript𝐯1𝑇subscript𝐯2superscriptsubscript𝐯2𝑇\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \times\frac{\kappa_{1}({\mathbf{r}}% )-\kappa_{2}({\mathbf{r}})}{8\pi}\left({\mathbf{v}}_{1}{\mathbf{v}}_{1}^{T}-{% \mathbf{v}}_{2}{\mathbf{v}}_{2}^{T}\right)\,× divide start_ARG italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ( bold_r ) - italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ( bold_r ) end_ARG start_ARG 8 italic_π end_ARG ( bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT - bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT ) (46)

are quadratic in the vectors characterizing the particle surface (𝐧𝐧Tsuperscript𝐧𝐧𝑇{\mathbf{n}}{\mathbf{n}}^{T}bold_nn start_POSTSUPERSCRIPT italic_T end_POSTSUPERSCRIPT denotes a tensor product). Here, the κi(𝐫)subscript𝜅𝑖𝐫\kappa_{i}({\mathbf{r}})italic_κ start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r ) denote the local principal curvatures of the surface in the direction of the unit vectors 𝐯i(𝐫,𝐎)subscript𝐯𝑖𝐫𝐎{\mathbf{v}}_{i}({\mathbf{r}},{\mathbf{O}})bold_v start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r , bold_O ), i=1,2𝑖12i=1,2italic_i = 1 , 2.

Note that the vectorial and tensorial weighted densities represent orientational averages and are therefore crucial for describing orientationally ordered phases. While the former will always vanish for apolar particle shapes and average to zero for the nematic phases of interest here, the latter will allow us to identify appropriate order parameters for uniaxial and biaxial order. To show this, we provide below a detailed recipe for how to calculate the weighted densities.

II.2.4 Explicit calculation of weighted densities: orientational order in FMT

To explicitly calculate the weighted densities in Eqs. (41)-(46), the position dependence in the integrands is to be understood in the sense that 𝐫𝐫{\mathbf{r}}bold_r is a coordinate in the body frame and the orientation dependence of vectors and tensors stems from the relation between the body and the lab frames. While the volume measure n3subscript𝑛3n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT always equals the packing fraction, the surface measures (involving positional integrals over \partial\mathcal{B}∂ caligraphic_B) can be calculated by a six-step procedure explained in the following. Additional information is provided in appendix A). To calculate the scalar measures, it is sufficient to follow steps two, three and five, as they do not require an explicit orientational averaging.

The first step is to identify (in the case of uniaxial particles) or choose (in the case of biaxial particles, where this selection may be ambiguous, see Sec. II.1.4) the primary uniaxial nematic axis for a given particle geometry and ensure that it points in the z𝑧zitalic_z-direction of the body frame (i.e., the basis vector m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT). The second step is to parameterize the surface \partial\mathcal{B}∂ caligraphic_B in the body frame using two appropriate parameters. Due to the additivity of the weight functions, different parts of the surface can be treated independently. Specifically, for the bodies of interest depicted in Fig. 1, we need to consider the following two-dimensional manifolds: a cone mantle, disks, portions of a torus (in the limit of circular rings), portions of a sphere, (parts of) cylinder mantles and flat triangles. The corresponding parameterizations are summarized in appendix A.1. The third step is to calculate all geometrical measures on the surface of each body part using the chosen parameterization in the body frame. This is exemplified for the cone mantle in appendix A.2. The fourth step is to transfer the surface vectors to the lab frame via the rotation matrix ^^\hat{\mathcal{R}}over^ start_ARG caligraphic_R end_ARG, defined in Eq. (11). To do so, we calculate xi(𝐫,𝐎)=^ijx¯j(𝐫)subscript𝑥𝑖𝐫𝐎subscript^𝑖𝑗subscript¯𝑥𝑗𝐫x_{i}({\mathbf{r}},{\mathbf{O}})=\hat{\mathcal{R}}_{ij}\bar{x}_{j}({\mathbf{r}})italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT ( bold_r , bold_O ) = over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over¯ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ( bold_r ), where xisubscript𝑥𝑖x_{i}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT represents the desired i𝑖iitalic_ith component of 𝐧𝐧{\mathbf{n}}bold_n, 𝐯1subscript𝐯1{\mathbf{v}}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT or 𝐯2subscript𝐯2{\mathbf{v}}_{2}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT with full orientation dependence in the lab frame, x¯jsubscript¯𝑥𝑗\bar{x}_{j}over¯ start_ARG italic_x end_ARG start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT is the j𝑗jitalic_jth component of these vectors in the body frame and we have implied summation over the repeated index j𝑗jitalic_j.

The fifth step is to integrate over the surface of all body parts using the parameterization in appendix A.1. The sixth and final step is to perform the orientational average and thereby specify how we can explicitly identify the appropriate order parameters in our density functional. This can be conveniently achieved by recognizing the relation (21) between the components of ^^\hat{\mathcal{R}}over^ start_ARG caligraphic_R end_ARG and the basis vectors of the two coordinate frames. In general, the calculation of a tensorial weighted density of rank k𝑘kitalic_k (with k=0𝑘0k=0italic_k = 0 for scalars and k=1𝑘1k=1italic_k = 1 for vectors) will involve k𝑘kitalic_k factors of these components. For our purpose, we find after some algebra that the average of the squared components can be directly related to the order parameters defined in Eq. (LABEL:eq_Xall), such that we can set

(^11)2delimited-⟨⟩superscriptsubscript^112\displaystyle\left\langle(\hat{\mathcal{R}}_{11})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =SPU+3F6+13,absent𝑆𝑃𝑈3𝐹613\displaystyle=\,\frac{S-\sqrt{P}-\sqrt{U}+3F}{6}+\frac{1}{3}\,,= divide start_ARG italic_S - square-root start_ARG italic_P end_ARG - square-root start_ARG italic_U end_ARG + 3 italic_F end_ARG start_ARG 6 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (47)
(^12)2delimited-⟨⟩superscriptsubscript^122\displaystyle\left\langle(\hat{\mathcal{R}}_{12})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =SP+U+3F6+13,absent𝑆𝑃𝑈3𝐹613\displaystyle=\,\frac{S-\sqrt{P}+\sqrt{U}+3F}{6}+\frac{1}{3}\,,= divide start_ARG italic_S - square-root start_ARG italic_P end_ARG + square-root start_ARG italic_U end_ARG + 3 italic_F end_ARG start_ARG 6 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (48)
(^13)2delimited-⟨⟩superscriptsubscript^132\displaystyle\left\langle(\hat{\mathcal{R}}_{13})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 13 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =S+P3+13,absent𝑆𝑃313\displaystyle=\frac{-S+\sqrt{P}}{3}+\frac{1}{3}\,,= divide start_ARG - italic_S + square-root start_ARG italic_P end_ARG end_ARG start_ARG 3 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (49)
(^21)2delimited-⟨⟩superscriptsubscript^212\displaystyle\left\langle(\hat{\mathcal{R}}_{21})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =S+P+U3F6+13,absent𝑆𝑃𝑈3𝐹613\displaystyle=\,\frac{S+\sqrt{P}+\sqrt{U}-3F}{6}+\frac{1}{3}\,,= divide start_ARG italic_S + square-root start_ARG italic_P end_ARG + square-root start_ARG italic_U end_ARG - 3 italic_F end_ARG start_ARG 6 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (50)
(^22)2delimited-⟨⟩superscriptsubscript^222\displaystyle\left\langle(\hat{\mathcal{R}}_{22})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =S+P+U+3F6+13,absent𝑆𝑃𝑈3𝐹613\displaystyle=\,\frac{S+\sqrt{P}+\sqrt{U}+3F}{6}+\frac{1}{3}\,,= divide start_ARG italic_S + square-root start_ARG italic_P end_ARG + square-root start_ARG italic_U end_ARG + 3 italic_F end_ARG start_ARG 6 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (51)
(^23)2delimited-⟨⟩superscriptsubscript^232\displaystyle\left\langle(\hat{\mathcal{R}}_{23})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 23 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =SP3+13,absent𝑆𝑃313\displaystyle=\frac{-S-\sqrt{P}}{3}+\frac{1}{3}\,,= divide start_ARG - italic_S - square-root start_ARG italic_P end_ARG end_ARG start_ARG 3 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (52)
(^31)2delimited-⟨⟩superscriptsubscript^312\displaystyle\left\langle(\hat{\mathcal{R}}_{31})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 31 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =S+U3+13,absent𝑆𝑈313\displaystyle=\frac{-S+\sqrt{U}}{3}+\frac{1}{3}\,,= divide start_ARG - italic_S + square-root start_ARG italic_U end_ARG end_ARG start_ARG 3 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (53)
(^32)2delimited-⟨⟩superscriptsubscript^322\displaystyle\left\langle(\hat{\mathcal{R}}_{32})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 32 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =S+U3+13,absent𝑆𝑈313\displaystyle=\frac{S+\sqrt{U}}{3}+\frac{1}{3}\,,= divide start_ARG italic_S + square-root start_ARG italic_U end_ARG end_ARG start_ARG 3 end_ARG + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (54)
(^33)2delimited-⟨⟩superscriptsubscript^332\displaystyle\left\langle(\hat{\mathcal{R}}_{33})^{2}\right\rangle⟨ ( over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ⟩ =23S+13,absent23𝑆13\displaystyle=\frac{2}{3}S+\frac{1}{3}\,,= divide start_ARG 2 end_ARG start_ARG 3 end_ARG italic_S + divide start_ARG 1 end_ARG start_ARG 3 end_ARG , (55)

while all linear and mixed terms must vanish if no other order parameters do contribute.

The weighted densities calculated from the procedure outlined above are collected in appendix B for the bodies depicted in Fig. 1. As expected, the tensorial weighted densities of hard cones and hard cylinders (see appendices B.1 and B.2) do not depend on U𝑈Uitalic_U and F𝐹Fitalic_F, since their shape is uniaxial, while these order parameters become important for hard isosceles spherotriangles (see appendix B.3), whose shape is biaxial. A comprehensive account on the order parameters for uniaxial particles is given in appendix B.4, where generalized expressions for the weighted densities of hard spherocylinders are found by taking different limits of the results for spherotriangles. In short, we can obtain analytic results for the weighted densities, which can be inserted into Eq. (34). Hence, as a final result, we find the general form

Φex[g]=Φex(ρ,S,U,P,F)subscriptΦexdelimited-[]𝑔subscriptΦex𝜌𝑆𝑈𝑃𝐹\Phi_{\text{ex}}[g]=\Phi_{\text{ex}}(\rho,S,U,P,F)roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT [ italic_g ] = roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( italic_ρ , italic_S , italic_U , italic_P , italic_F ) (56)

of the excess free energy density, which functionally depends on the orientational distribution, as a function of the (homogeneous) density and the four order parameters, which are themselves functionals of the orientational distribution, as they are calculated in FMT according to Eq. (22).

Recall from Sec. II.1.3 that the number of order parameters may increase when drop** the restriction to particles with a D2hsubscript𝐷2hD_{2\mathrm{h}}italic_D start_POSTSUBSCRIPT 2 roman_h end_POSTSUBSCRIPT symmetry. However, we argue that isosceles spherotriangles can still be fully described by S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F as the only deviation from D2hsubscript𝐷2hD_{2\mathrm{h}}italic_D start_POSTSUBSCRIPT 2 roman_h end_POSTSUBSCRIPT symmetry is due to their polar shape, which is not reflected by tensors of even rank. In turn, vectorial weighted densities do, in principle, provide additional order parameters measuring polarity, but their contribution is found to be negligibly small (see appendix B.1 for a more detailed discussion). This is in line with our expectation that shape polarity is not sufficient stabilize phases with global polar order for hard interactions. For lower particle symmetries, such as for arbitrary spherotriangles, we describe in appendix B.5 that we get additional terms proportional to ^ij^kldelimited-⟨⟩subscript^𝑖𝑗subscript^𝑘𝑙\langle\hat{\mathcal{R}}_{ij}\hat{\mathcal{R}}_{kl}\rangle⟨ over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT ⟩ with ik𝑖𝑘i\neq kitalic_i ≠ italic_k and/or jl𝑗𝑙j\neq litalic_j ≠ italic_l in the tensorial weighted densities of rank two, which represent nondiagonal elements of the Saupe matrix (1) and can thus not be expressed in terms of our four order parameters alone. More generally, extending the functional to tensors of higher rank in Eq. (35) would also give rise to a plethora of additional order parameters. However, their inclusion would blur the compact analytical picture we wish to provide in the following.

II.3 Identification of phase transitions

II.3.1 Free minimization with all order parameters

Given the homogeneous ideal gas free energy (37) and an excess free energy (56) that is a function of the four order parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F, which can be written as in Eq. (22), the Euler-Lagrange equation  (29) becomes

ρ(ln(g(ϕ,θ,ψ))+1)+δΦexδg(ϕ,θ,ψ)=0.𝜌𝑔italic-ϕ𝜃𝜓1𝛿subscriptΦex𝛿𝑔italic-ϕ𝜃𝜓0\rho\left(\ln(g(\phi,\theta,\psi))+1\right)+\frac{\delta\Phi_{\text{ex}}}{% \delta g(\phi,\theta,\psi)}=0\,.italic_ρ ( roman_ln ( italic_g ( italic_ϕ , italic_θ , italic_ψ ) ) + 1 ) + divide start_ARG italic_δ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG italic_δ italic_g ( italic_ϕ , italic_θ , italic_ψ ) end_ARG = 0 . (57)

Using the chain rule for the functional derivative, we obtain

g(ϕ,θ,ψ)=𝒩1XeαX2fX(ϕ,θ,ψ),𝑔italic-ϕ𝜃𝜓superscript𝒩1subscriptproduct𝑋superscript𝑒superscriptsubscript𝛼𝑋2subscript𝑓𝑋italic-ϕ𝜃𝜓g(\phi,\theta,\psi)=\mathcal{N}^{-1}\prod_{X}\,e^{\alpha_{X}^{2}\,f_{X}(\phi,% \theta,\psi)}\,,italic_g ( italic_ϕ , italic_θ , italic_ψ ) = caligraphic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ∏ start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT italic_e start_POSTSUPERSCRIPT italic_α start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) end_POSTSUPERSCRIPT , (58)

where 𝒩𝒩\mathcal{N}caligraphic_N is a normalization constant that can be determined by the condition

1=d𝐎g(ϕ,θ,ψ)1differential-d𝐎𝑔italic-ϕ𝜃𝜓1=\int\mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,1 = ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) (59)

and where we have defined the intrinsic order parameters

αX2:=1ρΦexX,X{S,U,P,F}.formulae-sequenceassignsuperscriptsubscript𝛼𝑋21𝜌subscriptΦex𝑋𝑋𝑆𝑈𝑃𝐹\alpha_{X}^{2}:=-\frac{1}{\rho}\frac{\partial\Phi_{\text{ex}}}{\partial X}\,,% \ \ \ X\in\{S,U,P,F\}\,.italic_α start_POSTSUBSCRIPT italic_X end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT := - divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_X end_ARG , italic_X ∈ { italic_S , italic_U , italic_P , italic_F } . (60)

Inserting the obtained orientational distribution, Eq. (58), into the definition (22) of the order parameters and solving the five coupled equations (59) and (60) for the five unknowns αSsubscript𝛼𝑆\alpha_{S}italic_α start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT, αUsubscript𝛼𝑈\alpha_{U}italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT, αPsubscript𝛼𝑃\alpha_{P}italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT, αFsubscript𝛼𝐹\alpha_{F}italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT and 𝒩𝒩\mathcal{N}caligraphic_N yields the orientational distribution function of the stable homogeneous phase with full information on biaxiality. Unfortunately, it is in general not possible to find an explicit expression for 𝒩𝒩\mathcal{N}caligraphic_N and thus determine this solution in a closed form.

One option to determine the phase boundaries of a model fluid would be a numerical solution of Eqs. (59) and (60), where the global minimum could be identified from comparing the corresponding free energy Φ=Φid+ΦexΦsubscriptΦidsubscriptΦex\Phi=\Phi_{\text{id}}+\Phi_{\text{ex}}roman_Φ = roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT (evaluated for the obtained orientational distributions) in the case of multiple solutions. Moreover, as discussed in Sec. II.1.4, a minimization with respect to all four order parameters would result in a multitude of physically equivalent solutions due to spontaneous symmetry breaking. Instead, our goal is to get analytic insight and, in particular, also understand which are the most relevant order parameters for characterizing the different phases. Thus we proceed step by step and first investigate in Sec. II.3.2 the scenario of uniaxial nematic order, solely characterized by S𝑆Sitalic_S, before considering in Secs. II.3.3 and II.3.4 the role of the other order parameters Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F } with respect to this reference state.

II.3.2 Phase transitions involving uniaxial order

For uniaxial particles and an appropriate choice of the coordinate systems (in which both the director and the symmetry axis point in the z𝑧zitalic_z-direction of the laboratory and molecular frame, respectively), the only order parameter relevant for the homogeneous bulk phase behavior is S𝑆Sitalic_S. Thus, we may assume Φex[g]=Φex(ρ,S)subscriptΦexdelimited-[]𝑔subscriptΦex𝜌𝑆\Phi_{\text{ex}}[g]=\Phi_{\text{ex}}(\rho,S)roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT [ italic_g ] = roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ( italic_ρ , italic_S ) instead of the general form (56). We are left with only two equations (Eq. (59) and Eq. (60) for X=S𝑋𝑆X=Sitalic_X = italic_S), which can be solved explicitly Hansen-Goos and Mecke (2010); Wittmann et al. (2014), as we briefly recapitulate below.

Let us first define α2:=3αS2/2assignsuperscript𝛼23superscriptsubscript𝛼𝑆22\alpha^{2}:=3\alpha_{S}^{2}/2italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT := 3 italic_α start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 2 for later notational convenience, such that Eq. (60) becomes

α2:=23ρΦexS.assignsuperscript𝛼223𝜌subscriptΦex𝑆\alpha^{2}:=-\frac{2}{3\rho}\frac{\partial\Phi_{\text{ex}}}{\partial S}\,.italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT := - divide start_ARG 2 end_ARG start_ARG 3 italic_ρ end_ARG divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_S end_ARG . (61)

Normalization of the orientational distribution (58) with αU=αP=αF=0subscript𝛼𝑈subscript𝛼𝑃subscript𝛼𝐹0\alpha_{U}=\alpha_{P}=\alpha_{F}=0italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0 yields the explicit expression

g=g(α,cosθ)=α𝒟(α)exp(α2(1cos2θ))𝑔𝑔𝛼𝜃𝛼𝒟𝛼superscript𝛼21superscript2𝜃g=g(\alpha,\cos\theta)=\frac{\alpha}{\mathcal{D}(\alpha)}\exp\!\left(-\alpha^{% 2}\left(1-\cos^{2}\theta\right)\right)italic_g = italic_g ( italic_α , roman_cos italic_θ ) = divide start_ARG italic_α end_ARG start_ARG caligraphic_D ( italic_α ) end_ARG roman_exp ( - italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - roman_cos start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_θ ) ) (62)

with Dawson’s integral

𝒟(α)=exp(α2)0αduexp(u2).𝒟𝛼superscript𝛼2superscriptsubscript0𝛼differential-d𝑢superscript𝑢2\mathcal{D}(\alpha)=\exp(-\alpha^{2})\int_{0}^{\alpha}\mathrm{d}u\exp(u^{2})\,.caligraphic_D ( italic_α ) = roman_exp ( - italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ∫ start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT start_POSTSUPERSCRIPT italic_α end_POSTSUPERSCRIPT roman_d italic_u roman_exp ( italic_u start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (63)

Inserting Eq. (62) into Eq. (22), the nematic order parameter

S(α)=34α𝒟(α)34α212𝑆𝛼34𝛼𝒟𝛼34superscript𝛼212\displaystyle S(\alpha)=\frac{3}{4\alpha\mathcal{D}(\alpha)}-\frac{3}{4\alpha^% {2}}-\frac{1}{2}italic_S ( italic_α ) = divide start_ARG 3 end_ARG start_ARG 4 italic_α caligraphic_D ( italic_α ) end_ARG - divide start_ARG 3 end_ARG start_ARG 4 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - divide start_ARG 1 end_ARG start_ARG 2 end_ARG (64)

is obtained as a function of α𝛼\alphaitalic_α. With this result, we are left with the task to solve a single equation, Eq. (61), in a self-consistent way to obtain the solution α𝛼\alphaitalic_α that minimizes the free energy at a given density ρ𝜌\rhoitalic_ρ. Vice versa, the solution for the density as a function of α𝛼\alphaitalic_α can even be found in a closed analytic form.

To determine the densities of two coexisting phases we need to impose equilibrium conditions. As the temperature T𝑇Titalic_T only enters as a scaling factor that has the same value everywhere in the system, thermal equilibrium is always ensured.However, we need to demand that the two phases are in chemical and mechanical equilibrium. Hence, the chemical potential

βμ=(Φid+Φex)ρ𝛽𝜇subscriptΦidsubscriptΦex𝜌\displaystyle\beta\mu=\frac{\partial(\Phi_{\text{id}}+\Phi_{\text{ex}})}{% \partial\rho}italic_β italic_μ = divide start_ARG ∂ ( roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ) end_ARG start_ARG ∂ italic_ρ end_ARG (65)

and the pressure

βp=βμρ(Φid+Φex)𝛽𝑝𝛽𝜇𝜌subscriptΦidsubscriptΦex\displaystyle\beta p=\beta\mu\rho-(\Phi_{\text{id}}+\Phi_{\text{ex}})italic_β italic_p = italic_β italic_μ italic_ρ - ( roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT + roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT ) (66)

must be equal. The phases that can be compared in this way are the isotropic phase, characterized by the absence of any orientational order (α=0𝛼0\alpha=0italic_α = 0) and (uniaxial) nematic phases found as the nontrivial solution α𝛼\alphaitalic_α of Eq. (61) for a given functional.

For particles with biaxial shape, let us recall that, by choosing different particle axes which may align with the director, we can also examine coexistence between different uniaxial nematic phases in this way (assuming that no other order parameters are relevant).

II.3.3 Biaxiality for perfect uniaxial order

A simple way to investigate the uniaxial–biaxial transition is to assume perfect uniaxial order by fixing the main particle axis in space. This approximation reduces the complexity of the problem to that of locating the isotropic–nematic transition in two spatial dimensions. To make this apparent, we align the desired axis in the m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT-direction of the molecular frame (first option discussed in Sec. II.1.4 and blue coordinate frames in Fig. 2) and then set θ=ψ=0𝜃𝜓0\theta=\psi=0italic_θ = italic_ψ = 0 in Eqs. (24)-(28), such that the orientational order (perpendicular to the main axis) is characterized solely by the remaining polar angle ϕitalic-ϕ\phiitalic_ϕ. Doing so, we directly find that

S=1,U=P=0,F=S2d,formulae-sequenceformulae-sequence𝑆1𝑈𝑃0𝐹subscript𝑆2𝑑S=1\,,\ U=P=0\,,\ F=S_{2d}\,,italic_S = 1 , italic_U = italic_P = 0 , italic_F = italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT , (67)

where

S2d=12πdϕg2d(ϕ)cos(2ϕ)subscript𝑆2𝑑12𝜋differential-ditalic-ϕsubscript𝑔2𝑑italic-ϕ2italic-ϕS_{2d}=\frac{1}{2\pi}\int\mathrm{d}\phi\,g_{2d}(\phi)\,\cos(2\phi)\,italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT = divide start_ARG 1 end_ARG start_ARG 2 italic_π end_ARG ∫ roman_d italic_ϕ italic_g start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT ( italic_ϕ ) roman_cos ( 2 italic_ϕ ) (68)

is a two-dimensional nematic order parameter and g2d(ϕ)subscript𝑔2𝑑italic-ϕg_{2d}(\phi)italic_g start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT ( italic_ϕ ) the according orientational distribution. The remaining analysis is analogous to that in Sec. II.3.2 and has been performed in Ref. Wittmann et al., 2017 for two-dimensional rods, where it was demonstrated that the resulting self-consistency equation only has one stable solution. Hence, the uniaxial–biaxial transition density can be identified in a closed analytic form when using the approximation of perfect uniaxial order.

Before proceeding in Sec. II.3.4 with a more accurate method to identify the onset of biaxial order, let us revisit alternative choices to align a biaxial particle in the body frame. As in the final paragraph of Sec. II.1.4, let us focus on an isosceles spherotriangle and suppose that we always align the triangle height with m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT. Now, in the special case of perfect uniaxial order, the manipulations necessary to describe a deviating director alignment can be intuitively illustrated by simply rotating the body frame instead of redefining its coordinates. Specifically, to achieve a perfect alignment of the director n𝑛\vec{n}over→ start_ARG italic_n end_ARG perpendicular to the triangle face (m2||n\vec{m}_{2}\,||\,\vec{n}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT | | over→ start_ARG italic_n end_ARG, compare the second case in Fig. 2a), we can rotate the body frame by ψ=π/2𝜓𝜋2\psi=\pi/2italic_ψ = italic_π / 2, θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2 and consider a shifted polar angle ϕϕ+π/2italic-ϕitalic-ϕ𝜋2\phi\rightarrow\phi+\pi/2italic_ϕ → italic_ϕ + italic_π / 2 in Eqs. (24)-(28), such that S=1/2𝑆12S=-1/2italic_S = - 1 / 2, U=3/2𝑈32U=-\sqrt{3}/2italic_U = - square-root start_ARG 3 end_ARG / 2, P=3S2d/2𝑃3subscript𝑆2𝑑2P=\sqrt{3}S_{2d}/2italic_P = square-root start_ARG 3 end_ARG italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT / 2 and F=S2d/2𝐹subscript𝑆2𝑑2F=-S_{2d}/2italic_F = - italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT / 2. This result is equivalent to making the substitutions in Eq. (8) and then choosing the order parameters according to Eq. (67). Moreover, a director perfectly aligned with the baseline of the triangle (m1||n\vec{m}_{1}\,||\,\vec{n}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT | | over→ start_ARG italic_n end_ARG, compare the third case in Fig. 2a) corresponds to ψ=0𝜓0\psi=0italic_ψ = 0, θ=π/2𝜃𝜋2\theta=\pi/2italic_θ = italic_π / 2 and a free ϕitalic-ϕ\phiitalic_ϕ in Eqs. (24)-(28), such that S=1/2𝑆12S=-1/2italic_S = - 1 / 2, U=3/2𝑈32U=\sqrt{3}/2italic_U = square-root start_ARG 3 end_ARG / 2, P=3S2d/2𝑃3subscript𝑆2𝑑2P=\sqrt{3}S_{2d}/2italic_P = square-root start_ARG 3 end_ARG italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT / 2 and F=S2d/2𝐹subscript𝑆2𝑑2F=S_{2d}/2italic_F = italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT / 2. This result is equivalent to making the substitutions in Eq. (10) and then using Eq. (67).

In summary, the above examples allow us to make sense of the altered interpretation of the order parameters upon drop** the convention that m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT denotes the particle axis which preferably aligned with the uniaxial director n𝑛\vec{n}over→ start_ARG italic_n end_ARG: in general, even perfect uniaxial order (S2d=0subscript𝑆2𝑑0S_{2d}=0italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT = 0) cannot be described by S𝑆Sitalic_S alone. However, the common convention, used in Sec. II.3.2, that the order parameter S𝑆Sitalic_S measures the degree of uniaxial alignment of the main particle axis (chosen parallel to m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT) with the director (chosen parallel to l3subscript𝑙3\vec{l}_{3}over→ start_ARG italic_l end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT), can always be restored by a redefinition of the body frame.

II.3.4 Biaxiality as perturbation of uniaxial order

While the assumption of perfectly uniaxial order made in Sec. II.3.3 is helpful to get a feeling for the relevant order parameters describing biaxiality, it may be a very crude approximation in practice. A more reliable calculation of the uniaxial–biaxial transition is by investigating the instability of the uniaxial solution, Eq. (62), under small perturbations related to the order parameter F𝐹Fitalic_F. This strategy will allow us for an exact location of the transition under the two assumptions that it is of second order (and can thus be identified as the limit of stability of the uniaxial phase) and that no other order parameter is relevant. To assess the latter assumption, we introduce for the sake of generality in the following presentation a dummy parameter Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F }, representing any of the three remaining order parameters and consider it as a perturbation to a phase with S𝑆Sitalic_S being the only nonzero order parameter, which we refer to in what follows as the simple uniaxial phase.

Assuming that there is only one relevant order parameter Y𝑌Yitalic_Y in addition to S𝑆Sitalic_S, let us first simplify the general orientational distribution (58) to (recalling that αS2=2α2/3superscriptsubscript𝛼𝑆22superscript𝛼23\alpha_{S}^{2}=2\alpha^{2}/3italic_α start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 3)

g(ϕ,θ,ψ)𝑔italic-ϕ𝜃𝜓\displaystyle g(\phi,\theta,\psi)italic_g ( italic_ϕ , italic_θ , italic_ψ ) =𝒩1g0(ϕ,θ,ψ),absentsuperscript𝒩1subscript𝑔0italic-ϕ𝜃𝜓\displaystyle=\mathcal{N}^{-1}g_{0}(\phi,\theta,\psi)\,,= caligraphic_N start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) , (69)
g0(ϕ,θ,ψ)subscript𝑔0italic-ϕ𝜃𝜓\displaystyle g_{0}(\phi,\theta,\psi)italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =e2α23fS(ϕ,θ,ψ)+αY2fY(ϕ,θ,ψ),absentsuperscript𝑒2superscript𝛼23subscript𝑓𝑆italic-ϕ𝜃𝜓superscriptsubscript𝛼𝑌2subscript𝑓𝑌italic-ϕ𝜃𝜓\displaystyle=e^{\frac{2\alpha^{2}}{3}\,f_{S}(\phi,\theta,\psi)+\alpha_{Y}^{2}% \,f_{Y}(\phi,\theta,\psi)}\,,= italic_e start_POSTSUPERSCRIPT divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG italic_f start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) + italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) end_POSTSUPERSCRIPT , (70)
𝒩𝒩\displaystyle\mathcal{N}caligraphic_N =d𝐎g0(ϕ,θ,ψ).absentdifferential-d𝐎subscript𝑔0italic-ϕ𝜃𝜓\displaystyle=\int\mathrm{d}\mathbf{O}\,g_{0}(\phi,\theta,\psi)\,.= ∫ roman_d bold_O italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) . (71)

Then, the free energy density (37) of the ideal gas can be rewritten in the explicit form

Φid=ρ(ln𝒩+2α23S+αY2Y+ln(ρΛ3)1),subscriptΦid𝜌𝒩2superscript𝛼23𝑆superscriptsubscript𝛼𝑌2𝑌𝜌superscriptΛ31\displaystyle\!\!\!\!\Phi_{\text{id}}=\rho\left(-\ln\mathcal{N}+\frac{2\alpha^% {2}}{3}\,S+\alpha_{Y}^{2}\,Y+\ln(\rho\Lambda^{3})-1\right)\,,roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT = italic_ρ ( - roman_ln caligraphic_N + divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG italic_S + italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_Y + roman_ln ( italic_ρ roman_Λ start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT ) - 1 ) , (72)

where 𝒩𝒩\mathcal{N}caligraphic_N, S𝑆Sitalic_S and Y𝑌Yitalic_Y are, in general, yet unknown functions of α𝛼\alphaitalic_α and αYsubscript𝛼𝑌\alpha_{Y}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT, compare Eq. (62) and Eq. (64) in the special case αY=0subscript𝛼𝑌0\alpha_{Y}=0italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0. Therefore, the total free energy density Φ(α,αY,S,Y,𝒩)Φ𝛼subscript𝛼𝑌𝑆𝑌𝒩\Phi(\alpha,\alpha_{Y},S,Y,\mathcal{N})roman_Φ ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT , italic_S , italic_Y , caligraphic_N ) depends on α𝛼\alphaitalic_α and αYsubscript𝛼𝑌\alpha_{Y}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT both explicitly and implicitly through 𝒩(α,αY)𝒩𝛼subscript𝛼𝑌\mathcal{N}(\alpha,\alpha_{Y})caligraphic_N ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ), S(α,αY)𝑆𝛼subscript𝛼𝑌S(\alpha,\alpha_{Y})italic_S ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ) and Y(α,αY)𝑌𝛼subscript𝛼𝑌Y(\alpha,\alpha_{Y})italic_Y ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ).

To investigate the stability of a uniaxial solution against perturbations due to the order parameter Y𝑌Yitalic_Y, we take a look at the minimum of ΦΦ\Phiroman_Φ as a function of αY2superscriptsubscript𝛼𝑌2\alpha_{Y}^{2}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, assuming that the value of α𝛼\alphaitalic_α, at a given density ρ𝜌\rhoitalic_ρ, is not affected by αYsubscript𝛼𝑌\alpha_{Y}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT at first order. For the ideal term (72), we get

dΦiddαY2dsubscriptΦiddsuperscriptsubscript𝛼𝑌2\displaystyle\frac{\mathrm{d}\Phi_{\text{id}}}{\mathrm{d}\alpha_{Y}^{2}}divide start_ARG roman_d roman_Φ start_POSTSUBSCRIPT id end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =ρ(1𝒩𝒩αY2+2α23SαY2+Y+αY2YαY2)absent𝜌1𝒩𝒩superscriptsubscript𝛼𝑌22superscript𝛼23𝑆superscriptsubscript𝛼𝑌2𝑌superscriptsubscript𝛼𝑌2𝑌superscriptsubscript𝛼𝑌2\displaystyle=\rho\left(-\frac{1}{\mathcal{N}}\frac{\partial\mathcal{N}}{% \partial\alpha_{Y}^{2}}+\frac{2\alpha^{2}}{3}\,\frac{\partial S}{\partial% \alpha_{Y}^{2}}+Y+\alpha_{Y}^{2}\,\frac{\partial Y}{\partial\alpha_{Y}^{2}}% \right)\!\!\!\!\!\!\!\!= italic_ρ ( - divide start_ARG 1 end_ARG start_ARG caligraphic_N end_ARG divide start_ARG ∂ caligraphic_N end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG divide start_ARG ∂ italic_S end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_Y + italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_Y end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (73)
=ρ(2α23SαY2+αY2YαY2),absent𝜌2superscript𝛼23𝑆superscriptsubscript𝛼𝑌2superscriptsubscript𝛼𝑌2𝑌superscriptsubscript𝛼𝑌2\displaystyle=\rho\left(\frac{2\alpha^{2}}{3}\,\frac{\partial S}{\partial% \alpha_{Y}^{2}}+\alpha_{Y}^{2}\,\frac{\partial Y}{\partial\alpha_{Y}^{2}}% \right)\,,= italic_ρ ( divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG divide start_ARG ∂ italic_S end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG + italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG ∂ italic_Y end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (74)

where we have used that the first term in the first line is equal to Y𝑌-Y- italic_Y, which can be verified by inserting the definition of 𝒩𝒩\mathcal{N}caligraphic_N from Eq. (71) with g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT from Eq. (70) and calculating the derivative within the integral. As we consider a perturbation of a simple uniaxial phase, we insert the equilibrium condition from Eq. (61), which must hold for αY=0subscript𝛼𝑌0\alpha_{Y}=0italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0, such that the derivative of the excess term (56) becomes

dΦexdαY2dsubscriptΦexdsuperscriptsubscript𝛼𝑌2\displaystyle\frac{\mathrm{d}\Phi_{\text{ex}}}{\mathrm{d}\alpha_{Y}^{2}}divide start_ARG roman_d roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG roman_d italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =ΦexYYαY2ρ2α23SαY2.absentsubscriptΦex𝑌𝑌superscriptsubscript𝛼𝑌2𝜌2superscript𝛼23𝑆superscriptsubscript𝛼𝑌2\displaystyle=\frac{\partial\Phi_{\text{ex}}}{\partial Y}\frac{\partial Y}{% \partial\alpha_{Y}^{2}}-\rho\,\frac{2\alpha^{2}}{3}\,\frac{\partial S}{% \partial\alpha_{Y}^{2}}\,.= divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_Y end_ARG divide start_ARG ∂ italic_Y end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG - italic_ρ divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG divide start_ARG ∂ italic_S end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (75)

Adding up Eqs. (74) and (75), we get

dΦdαY2dΦdsuperscriptsubscript𝛼𝑌2\displaystyle\frac{\mathrm{d}\Phi}{\mathrm{d}\alpha_{Y}^{2}}divide start_ARG roman_d roman_Φ end_ARG start_ARG roman_d italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG =(ραY2+ΦexY)YαY2.absent𝜌superscriptsubscript𝛼𝑌2subscriptΦex𝑌𝑌superscriptsubscript𝛼𝑌2\displaystyle=\left(\rho\alpha_{Y}^{2}+\frac{\partial\Phi_{\text{ex}}}{% \partial Y}\right)\frac{\partial Y}{\partial\alpha_{Y}^{2}}\,.= ( italic_ρ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_Y end_ARG ) divide start_ARG ∂ italic_Y end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (76)

Since the order parameter Y𝑌Yitalic_Y is a monotonous function of αYsubscript𝛼𝑌\alpha_{Y}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT (both increase for increasing order), i.e., Y/αY2>0𝑌superscriptsubscript𝛼𝑌20\partial Y/\partial\alpha_{Y}^{2}>0∂ italic_Y / ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0, the extremal condition dΦ/dαY2=0dΦdsuperscriptsubscript𝛼𝑌20\mathrm{d}\Phi/\mathrm{d}\alpha_{Y}^{2}=0roman_d roman_Φ / roman_d italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT = 0 consistently recovers Eq. (60) for X=Y𝑋𝑌X=Yitalic_X = italic_Y. This condition and the one for X=S𝑋𝑆X=Sitalic_X = italic_S, i.e., Eq. (61), must be mutually fulfilled in a stable system. In particular, the stability of the simple uniaxial solution has the necessary condition

Y(α):=(ΦexYYαY2)αY=0=0.assignsubscript𝑌𝛼subscriptsubscriptΦex𝑌𝑌superscriptsubscript𝛼𝑌2subscript𝛼𝑌00\displaystyle\mathcal{E}_{Y}(\alpha):=\left(\frac{\partial\Phi_{\text{ex}}}{% \partial Y}\frac{\partial Y}{\partial\alpha_{Y}^{2}}\right)_{\alpha_{Y}=0}=0\,.caligraphic_E start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) := ( divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_Y end_ARG divide start_ARG ∂ italic_Y end_ARG start_ARG ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) start_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0 end_POSTSUBSCRIPT = 0 . (77)

If it does not hold, the simple uniaxial phase is not stable for the given value of α𝛼\alphaitalic_α and we anticipate that αY>0subscript𝛼𝑌0\alpha_{Y}>0italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT > 0.

What is left to be determined in the case that Eq. (77) is fulfilled is whether the corresponding extremal point of the free energy in Eq. (76), is indeed a minimum, which is a sufficient condition for the stability of the simple uniaxial solution with αY=0subscript𝛼𝑌0\alpha_{Y}=0italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0 against additional order associated with the parameter Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F }. To be able to answer this question, we must calculate the second derivative of ΦΦ\Phiroman_Φ with respect to αY2superscriptsubscript𝛼𝑌2\alpha_{Y}^{2}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT and determine its sign. To do so, we take another derivative of Eq. (76), insert Eq. (60) for X=Y𝑋𝑌X=Yitalic_X = italic_Y (which removes the second derivative of Y𝑌Yitalic_Y), divide by ρ𝜌\rhoitalic_ρ and divide by Y/αY2𝑌superscriptsubscript𝛼𝑌2\partial Y/\partial\alpha_{Y}^{2}∂ italic_Y / ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This yields the condition

𝒜Y(α):=1+(ddαY21ρΦexY)αY=0=0,assignsubscript𝒜𝑌𝛼1subscriptddsuperscriptsubscript𝛼𝑌21𝜌subscriptΦex𝑌subscript𝛼𝑌00\displaystyle\mathcal{A}_{Y}(\alpha):=1+\left(\frac{\mathrm{d}}{\mathrm{d}% \alpha_{Y}^{2}}\frac{1}{\rho}\frac{\partial\Phi_{\text{ex}}}{\partial Y}\right% )_{\alpha_{Y}=0}=0\,,caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) := 1 + ( divide start_ARG roman_d end_ARG start_ARG roman_d italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG divide start_ARG 1 end_ARG start_ARG italic_ρ end_ARG divide start_ARG ∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT end_ARG start_ARG ∂ italic_Y end_ARG ) start_POSTSUBSCRIPT italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0 end_POSTSUBSCRIPT = 0 , (78)

for the value of α𝛼\alphaitalic_α, as determined from solving Eq. (61), at which the simple uniaxial phase becomes unstable if the free energy has an extremal point at Y=αY=0𝑌subscript𝛼𝑌0Y=\alpha_{Y}=0italic_Y = italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0. Then, the values of α𝛼\alphaitalic_α for which 𝒜Y(α)>0subscript𝒜𝑌𝛼0\mathcal{A}_{Y}(\alpha)>0caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) > 0 holds indicate the stability range of the simple uniaxial solution (if Y(α)=0subscript𝑌𝛼0\mathcal{E}_{Y}(\alpha)=0caligraphic_E start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) = 0 and the value of α𝛼\alphaitalic_α is beyond isotropic–uniaxial coexistence).

Taking a closer look at the stability function in Eq. (78), we realize that its evaluation only requires the knowledge of all order parameters up to the first order in αY2superscriptsubscript𝛼𝑌2\alpha_{Y}^{2}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT. This result can be obtained analytically by expanding Eq. (70) according to

g0(ϕ,θ,ψ)subscript𝑔0italic-ϕ𝜃𝜓\displaystyle g_{0}(\phi,\theta,\psi)italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) =e2α23fS(ϕ,θ,ψ)(1+αY2fY(ϕ,θ,ψ))+𝒪(αY4)absentsuperscript𝑒2superscript𝛼23subscript𝑓𝑆italic-ϕ𝜃𝜓1superscriptsubscript𝛼𝑌2subscript𝑓𝑌italic-ϕ𝜃𝜓𝒪superscriptsubscript𝛼𝑌4\displaystyle=e^{\frac{2\alpha^{2}}{3}\,f_{S}(\phi,\theta,\psi)}(1+\alpha_{Y}^% {2}\,f_{Y}(\phi,\theta,\psi))+\mathcal{O}(\alpha_{Y}^{4})\,= italic_e start_POSTSUPERSCRIPT divide start_ARG 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 3 end_ARG italic_f start_POSTSUBSCRIPT italic_S end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) end_POSTSUPERSCRIPT ( 1 + italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_f start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_ϕ , italic_θ , italic_ψ ) ) + caligraphic_O ( italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) (79)

and performing an explicit normalization according to Eq. (71). Truncating the expansion of g0subscript𝑔0g_{0}italic_g start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT after the term of order αY2superscriptsubscript𝛼𝑌2\alpha_{Y}^{2}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT, we find that the nematic order parameter is unaffected for all Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F }, i.e., we have

S(α,αY)=S(α)+𝒪(αY4)𝑆𝛼subscript𝛼𝑌𝑆𝛼𝒪superscriptsubscript𝛼𝑌4\displaystyle S(\alpha,\alpha_{Y})=S(\alpha)+\mathcal{O}(\alpha_{Y}^{4})\,italic_S ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ) = italic_S ( italic_α ) + caligraphic_O ( italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) (80)

with S(α)𝑆𝛼S(\alpha)italic_S ( italic_α ) given by Eq. (64), while we obtain the leading terms

U(α,αU)𝑈𝛼subscript𝛼𝑈\displaystyle\!\!\!\!U(\alpha,\alpha_{U})italic_U ( italic_α , italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ) =(32α2)S(α)+2α28α2αU2+𝒪(αU4),absent32superscript𝛼2𝑆𝛼2superscript𝛼28superscript𝛼2superscriptsubscript𝛼𝑈2𝒪superscriptsubscript𝛼𝑈4\displaystyle=\frac{(-3-2\alpha^{2})S(\alpha)+2\alpha^{2}}{8\alpha^{2}}\,% \alpha_{U}^{2}+\mathcal{O}(\alpha_{U}^{4})\,,= divide start_ARG ( - 3 - 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_S ( italic_α ) + 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) , (81)
P(α,αP)𝑃𝛼subscript𝛼𝑃\displaystyle\!\!\!\!P(\alpha,\alpha_{P})italic_P ( italic_α , italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ) =(32α2)S(α)+2α28α2αP2+𝒪(αP4),absent32superscript𝛼2𝑆𝛼2superscript𝛼28superscript𝛼2superscriptsubscript𝛼𝑃2𝒪superscriptsubscript𝛼𝑃4\displaystyle=\frac{(-3-2\alpha^{2})S(\alpha)+2\alpha^{2}}{8\alpha^{2}}\,% \alpha_{P}^{2}+\mathcal{O}(\alpha_{P}^{4})\,,= divide start_ARG ( - 3 - 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_S ( italic_α ) + 2 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 8 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) , (82)
F(α,αF)𝐹𝛼subscript𝛼𝐹\displaystyle\!\!\!\!F(\alpha,\alpha_{F})italic_F ( italic_α , italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ) =(3+14α2)S(α)+10α248α2αF2+𝒪(αF4)absent314superscript𝛼2𝑆𝛼10superscript𝛼248superscript𝛼2superscriptsubscript𝛼𝐹2𝒪superscriptsubscript𝛼𝐹4\displaystyle=\frac{(-3+14\alpha^{2})S(\alpha)+10\alpha^{2}}{48\alpha^{2}}\,% \alpha_{F}^{2}+\mathcal{O}(\alpha_{F}^{4})\,= divide start_ARG ( - 3 + 14 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) italic_S ( italic_α ) + 10 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 48 italic_α start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + caligraphic_O ( italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT ) (83)

of the other order parameters. In each case, the coefficient of αY2superscriptsubscript𝛼𝑌2\alpha_{Y}^{2}italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT is positive for all finite values of α𝛼\alphaitalic_α, justifying the assumption Y/αY2>0𝑌superscriptsubscript𝛼𝑌20\partial Y/\partial\alpha_{Y}^{2}>0∂ italic_Y / ∂ italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT > 0 in deriving Eq. (78). Putting everything together, an analytic expression for 𝒜Y(α)subscript𝒜𝑌𝛼\mathcal{A}_{Y}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) in Eq. (78) with Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F } can be explicitly found by inserting, (i), the explicit uniaxial solution ρ(α)𝜌𝛼\rho(\alpha)italic_ρ ( italic_α ) of Eq. (61) with Y=0𝑌0Y=0italic_Y = 0, (ii), S(α,αY)𝑆𝛼subscript𝛼𝑌S(\alpha,\alpha_{Y})italic_S ( italic_α , italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ) from Eq. (80), where only the leading term is relevant, and, (iii), either of Eqs. (81)-(83).

The whole calculation is analogous for the general orientational distribution from Eq. (58), where all four order parameters are taken into account simultaneously. In this case, all generalized expansions (80)-(83) would contain constant terms in α𝛼\alphaitalic_α, reflecting the rotational invariance of the problem. This symmetry is broken by choosing appropriately aligned coordinate frames. In particular, for our common choice described in Sec. II.1.4, we argue that we can assume P=0𝑃0P=0italic_P = 0 in our perturbation analysis without loss of generality.

III Results for ordering behavior

Next we corroborate the theoretical conclusions drawn in Sec. II by investigating the ordering behavior of the hard bodies illustrated in Fig. 1. As uniaxial shapes we consider a cylinder (left) with diameter D𝐷Ditalic_D and height H𝐻Hitalic_H (which equals its mantle length L𝐿Litalic_L), as well as a cone (middle) with circular base area of diameter D𝐷Ditalic_D and height H𝐻Hitalic_H. To define in each case an aspect ratio l𝑙litalic_l measuring the particle anisotropy, we choose the convention l=H/D𝑙𝐻𝐷l=H/Ditalic_l = italic_H / italic_D. As a biaxial shape we consider an isosceles spherotriangle (right), defined as the parallel set at distance D/2𝐷2D/2italic_D / 2 of an isosceles triangle with base length A𝐴Aitalic_A and two side lengths B𝐵Bitalic_B (or, in other words, a triangular prism whose three side faces are capped by cylindrical halves connected by spherical parts). For the spherotriangle, we define the aspect ratio l=(A+2B)/(2D)𝑙𝐴2𝐵2𝐷l=(A+2B)/(2D)italic_l = ( italic_A + 2 italic_B ) / ( 2 italic_D ) such that it is consistent with the typical convention l=L/D𝑙𝐿𝐷l=L/Ditalic_l = italic_L / italic_D for a spherocylinder with cylindrical mantle length L𝐿Litalic_L and diameter D𝐷Ditalic_D. Specifically, the spherocylindrical shape is recovered in two limits: upon setting either A=L𝐴𝐿A=Litalic_A = italic_L and B=L/2𝐵𝐿2B=L/2italic_B = italic_L / 2 or A=0𝐴0A=0italic_A = 0 and B=L𝐵𝐿B=Litalic_B = italic_L. Hence, in addition to l𝑙litalic_l, we need a second dimensionless parameter to fully describe the shape of a spherotriangle. By defining the shape ratio x𝑥xitalic_x as

x=A2B=sinγ,𝑥𝐴2𝐵𝛾\displaystyle x=\frac{A}{2B}=\sin\gamma\,,italic_x = divide start_ARG italic_A end_ARG start_ARG 2 italic_B end_ARG = roman_sin italic_γ , (84)

we describe isosceles triangles of all possible opening angles 2γ2𝛾2\gamma2 italic_γ, which reduce to spherocylinders in the limiting cases x=0𝑥0x=0italic_x = 0 and x=1𝑥1x=1italic_x = 1. The corresponding weighted densities required to construct the functional (56) are stated in appendix B for each particle shape.

Our general strategy is to first justify in Sec. III.1 that the order parameters U𝑈Uitalic_U and P𝑃Pitalic_P can be disregarded in our treatment of the homogeneous bulk phases, where we relate the onset of biaxial order to a nonzero value of F𝐹Fitalic_F. Then, we discuss the phase diagrams of the different hard-body fluids, where we only focus on homogeneous phases. In doing so we neglect the expected transition to positionally ordered liquid crystal phases and the solid state. Therefore, we draw our phase diagrams in Figs. 4 and 5 with backgrounds fading into gray for an increasing packing fraction, indicating the increasing probability that the presented states are only metastable, as other phases might be predicted by the functional. In Sec. III.2, we compare the isotropic–nematic transition of two uniaxial shapes which possess the same limiting behavior for extreme aspect ratios: a polar cylinder and an apolar cone. In Sec. III.3, we turn to biaxial isosceles spherotriangles. The onset of biaxial order is discussed further in appendix C.

Refer to caption
Figure 3: Stability of the uniaxial nematic phase under different perturbations, investigated here for hard spherotriangles with x=0.18𝑥0.18x=0.18italic_x = 0.18 and l=5𝑙5l=5italic_l = 5. The results are representative for all dominant director orientations at all shape ratios considered, compare Fig. 5. According to the legend, we show U(α)subscript𝑈𝛼\mathcal{E}_{U}(\alpha)caligraphic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_α ) from Eq. (77) and 𝒜P(α)subscript𝒜𝑃𝛼\mathcal{A}_{P}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_α ) and 𝒜F(α)subscript𝒜𝐹𝛼\mathcal{A}_{F}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) from Eq. (78) as a function of α𝛼\alphaitalic_α (not shown are the trivial results P(α)=F(α)=0subscript𝑃𝛼subscript𝐹𝛼0\mathcal{E}_{P}(\alpha)=\mathcal{E}_{F}(\alpha)=0caligraphic_E start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_α ) = caligraphic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) = 0 and the meaningless function 𝒜U(α)subscript𝒜𝑈𝛼\mathcal{A}_{U}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_α )). As annotated, simple uniaxial order, solely specified by α𝛼\alphaitalic_α and thus S𝑆Sitalic_S, is stable against a perturbation in Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F } if Y(α)=0subscript𝑌𝛼0\mathcal{E}_{Y}(\alpha)=0caligraphic_E start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) = 0 and 𝒜Y(α)>0subscript𝒜𝑌𝛼0\mathcal{A}_{Y}(\alpha)>0caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) > 0, where the horizontal black dotted line serves as a guide to the eye. Hence, the value of α𝛼\alphaitalic_α for which 𝒜F(α)=0subscript𝒜𝐹𝛼0\mathcal{A}_{F}(\alpha)=0caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) = 0 indicates the uniaxial–biaxial transition. This stability analysis of uniaxial order is only meaningful beyond coexistence with the isotropic phase for ααIU𝛼subscript𝛼IU\alpha\geq\alpha_{\text{IU}}italic_α ≥ italic_α start_POSTSUBSCRIPT IU end_POSTSUBSCRIPT, as indicated by the dotted vertical line.

III.1 Relevant order parameters

To illustrate the relevance of the different order parameters, we perform the perturbation analysis of simple uniaxial nematic order, as described in Sec. II.3.4, to identify whether an appropriate description of the state point of interest requires to account for the order parameters U𝑈Uitalic_U, P𝑃Pitalic_P or F𝐹Fitalic_F in addition to the standard nematic order parameter S𝑆Sitalic_S. We choose to work here with an isosceles spherotriangle (whose height specifies the main axis) since it is the most general shape considered in our investigation and because its weighted densities in appendix B.3 depend on all four order parameters. Moreover, this biaxial shape reduces to a uniaxial spherocylinder when either the length of the base line becomes zero (x=0𝑥0x=0italic_x = 0) or its height vanishes (x=1𝑥1x=1italic_x = 1). The results of both limits differ due to the different direction of the assumed symmetry axis within the body frame, as further discussed in appendix B.4. In the first case, the only remaining order parameters are S𝑆Sitalic_S and P𝑃Pitalic_P, as expected for uniaxial particles, while, in the second case, the spherocylinder is formally treated as being a biaxial particle.

As a first step of our perturbation analysis, we must check the first derivatives of the free energy for the simple uniaxial reference case αY=0subscript𝛼𝑌0\alpha_{Y}=0italic_α start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT = 0 with Y{U,P,F}𝑌𝑈𝑃𝐹Y\in\{U,P,F\}italic_Y ∈ { italic_U , italic_P , italic_F } according to Eq. (77). Indeed, we find that the free energy has an extremal point for αP=0subscript𝛼𝑃0\alpha_{P}=0italic_α start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT = 0 and αF=0subscript𝛼𝐹0\alpha_{F}=0italic_α start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT = 0, i.e., P(α)=0subscript𝑃𝛼0\mathcal{E}_{P}(\alpha)=0caligraphic_E start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_α ) = 0 and F(α)=0subscript𝐹𝛼0\mathcal{E}_{F}(\alpha)=0caligraphic_E start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) = 0 for all values of α𝛼\alphaitalic_α. Regarding a perturbation in terms of the molecular biaxiality order parameter U𝑈Uitalic_U, the free energy has a negative slope at αU=0subscript𝛼𝑈0\alpha_{U}=0italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = 0, i.e., U(α)<0subscript𝑈𝛼0\mathcal{E}_{U}(\alpha)<0caligraphic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_α ) < 0 for 0<α<0𝛼0<\alpha<\infty0 < italic_α < ∞, as shown in Fig. 3. This suggests that U𝑈Uitalic_U affects all ordered phases of biaxial particles. Accordingly, it was shown in Ref. Cuetos et al., 2017 by minimizing a modified Onsager functional with respect to a trial orientational distribution that (a parameter closely related to) U𝑈Uitalic_U is nonzero at the isotropic–uniaxial coexistence of hard cuboids. However, ignoring this parameter affects the calculated transition densities only marginally Cuetos et al. (2017). Moreover, the decreasing absolute value |U(α)|subscript𝑈𝛼|\mathcal{E}_{U}(\alpha)|| caligraphic_E start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_α ) | of the slope for large α𝛼\alphaitalic_α in Fig. 3 suggests that the effect of U𝑈Uitalic_U should become smaller and smaller with increasingly strong uniaxial order and eventually turn fully irrelevant in the limit α𝛼\alpha\rightarrow\inftyitalic_α → ∞ (or S1𝑆1S\rightarrow 1italic_S → 1), which is consistent with the prediction U0𝑈0U\rightarrow 0italic_U → 0 in Eq. (67). As biaxial phases imply a large degree of uniaxial order and because U𝑈Uitalic_U does not measure phase biaxiality, we assume that U=0𝑈0U=0italic_U = 0 as a presumably good approximation for making analytic progress.

As a second step, we take a closer look at the stability functions (78) 𝒜Y(α)subscript𝒜𝑌𝛼\mathcal{A}_{Y}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) with Y{P,F}𝑌𝑃𝐹Y\in\{P,F\}italic_Y ∈ { italic_P , italic_F } (𝒜U(α)subscript𝒜𝑈𝛼\mathcal{A}_{U}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT ( italic_α ) is only meaningful at α=0𝛼0\alpha=0italic_α = 0, since we found that there is otherwise no extremal point of the free energy at αU=0subscript𝛼𝑈0\alpha_{U}=0italic_α start_POSTSUBSCRIPT italic_U end_POSTSUBSCRIPT = 0). As detailed in Sec. II.3.4, simple uniaxial order, where only S𝑆Sitalic_S has a nonzero value, does not appropriately characterize the system if 𝒜Y(α)<0subscript𝒜𝑌𝛼0\mathcal{A}_{Y}(\alpha)<0caligraphic_A start_POSTSUBSCRIPT italic_Y end_POSTSUBSCRIPT ( italic_α ) < 0, which indicates that the stable state should have a nonzero value of Y𝑌Yitalic_Y. Representative results for the stability functions are shown in Fig. 3. In general, 𝒜Psubscript𝒜𝑃\mathcal{A}_{P}caligraphic_A start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT increases with increasing α𝛼\alphaitalic_α, while the opposite trend is observed for 𝒜Fsubscript𝒜𝐹\mathcal{A}_{F}caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT. Again, the limit α𝛼\alpha\rightarrow\inftyitalic_α → ∞ (or S1𝑆1S\rightarrow 1italic_S → 1) of perfect uniaxial order can be directly understood from the order parameters in Eq. (67). As P0𝑃0P\rightarrow 0italic_P → 0, we always find that 𝒜P1subscript𝒜𝑃1\mathcal{A}_{P}\rightarrow 1caligraphic_A start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT → 1, because then the term in Eq. (78) involving the derivatives of ΦexsubscriptΦex\Phi_{\text{ex}}roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT does not contribute. In contrast, the limiting behavior of 𝒜Fsubscript𝒜𝐹\mathcal{A}_{F}caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT depends on the particular system, since F𝐹Fitalic_F does not vanish.

Most importantly, Fig. 3 reveals that the central order parameter to characterize phase biaxiality in our approach is F𝐹Fitalic_F and we can use the criterion 𝒜F(α)=0subscript𝒜𝐹𝛼0\mathcal{A}_{F}(\alpha)=0caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) = 0 to predict the onset of biaxial order. Indeed, 𝒜F(α)subscript𝒜𝐹𝛼\mathcal{A}_{F}(\alpha)caligraphic_A start_POSTSUBSCRIPT italic_F end_POSTSUBSCRIPT ( italic_α ) is generally found to change its sign at values of α𝛼\alphaitalic_α that are larger than αIUsubscript𝛼IU\alpha_{\text{IU}}italic_α start_POSTSUBSCRIPT IU end_POSTSUBSCRIPT at the isotropic–uniaxial transition. In contrast, we find 𝒜P(α)>0subscript𝒜𝑃𝛼0\mathcal{A}_{P}(\alpha)>0caligraphic_A start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( italic_α ) > 0 for all α>0𝛼0\alpha>0italic_α > 0, which implies that the simple uniaxial phase is always stable against perturbations in P𝑃Pitalic_P, such that the order parameter P𝑃Pitalic_P is not relevant for identifying the onset of biaxial order in our setup. Only within the biaxial phase, we find by generalizing Eq. (77) that P𝑃Pitalic_P can take nonzero values, because Φex/P<0subscriptΦex𝑃0\partial\Phi_{\text{ex}}/\partial P<0∂ roman_Φ start_POSTSUBSCRIPT ex end_POSTSUBSCRIPT / ∂ italic_P < 0 at P=0𝑃0P=0italic_P = 0 does no longer vanish if F>0𝐹0F>0italic_F > 0, but this does not affect the phase boundaries. We also note that the result 𝒜P(0)=0subscript𝒜𝑃00\mathcal{A}_{P}(0)=0caligraphic_A start_POSTSUBSCRIPT italic_P end_POSTSUBSCRIPT ( 0 ) = 0 suggests that the isotropic phase can equally be destabilized by a nonzero value of P𝑃Pitalic_P instead of S𝑆Sitalic_S (which corresponds to a different director alignment in the lab frame), but neither by U𝑈Uitalic_U nor F𝐹Fitalic_F.

III.2 Uniaxial bodies: effect of shape polarity

Refer to caption
Figure 4: Isotropic–nematic transition of hard cones (solid lines) and hard cylinders (dotted lines) depending on the packing fraction η𝜂\etaitalic_η and the rescaled aspect ratio l/(1+l)𝑙1𝑙l/(1+l)italic_l / ( 1 + italic_l ). The transition is always found to be of first order, but the density difference between coexisting phases is not always visible. The limits l/(1+l)=0𝑙1𝑙0l/(1+l)=0italic_l / ( 1 + italic_l ) = 0 and l/(1+l)=1𝑙1𝑙1l/(1+l)=1italic_l / ( 1 + italic_l ) = 1 correspond to hard disks (oblate nematic phase) and rods (prolate nematic phase), respectively. As detailed at the beginning of Sec. III, our functional is expected to predict phases with positional order at larger packing fractions. Hence, we indicate the region η0.45greater-than-or-equivalent-to𝜂0.45\eta\gtrsim 0.45italic_η ≳ 0.45 of the phase diagram, where positional order is expected when applying the used functional to hard spherocylinders Wittmann et al. (2014), by shading the background in light gray.
Refer to caption
Refer to caption
Figure 5: Phase diagram depicting the spatially homogeneous phases of hard spherotriangles at aspect ratios l=5𝑙5l=5italic_l = 5 (left) and l=25𝑙25l=25italic_l = 25 (right) in dependence of the packing fraction η𝜂\etaitalic_η and the shape ratio x𝑥xitalic_x. The red lines indicate isotropic–uniaxial coexistence, the blue lines indicate uniaxial–biaxial coexistence and the orange lines indicate uniaxial–uniaxial coexistence. The latter becomes metastable within the biaxial regime and is thus shown as dotted lines to give an idea of what might be the dominant particle axis in the biaxial phase. As illustrated in Fig. 2a, there are three stable orientations of the nematic director, which are visualized in the left diagram and marked with Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT, Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT and Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT in the right one. Once again, we shaded the area with η0.45greater-than-or-equivalent-to𝜂0.45\eta\gtrsim 0.45italic_η ≳ 0.45, where the used functional predicts stable phases with positional order in the limit of hard spherocylinders (x=0𝑥0x=0italic_x = 0 or x=1𝑥1x=1italic_x = 1), in light gray.

The homogeneous phase diagrams of hard cones and cylinders are shown in Fig. 4, depending on the packing fraction η𝜂\etaitalic_η and the rescaled aspect ratio l/(1+l)𝑙1𝑙l/(1+l)italic_l / ( 1 + italic_l ). This rescaling of l𝑙litalic_l allows us to display all relevant information within a range from zero to one. In the phase diagram, the dotted lines represent the isotropic-uniaxial coexistence densities for cylinders and the solid lines represent the coexistence densities for cones. Both results are qualitatively similar. For small (large) aspect ratios l1much-less-than𝑙1l\ll 1italic_l ≪ 1 (l1much-greater-than𝑙1l\gg 1italic_l ≫ 1) the nematic phase becomes more stable upon further decreasing (increasing) l𝑙litalic_l, such that the shape becomes more and more oblate (prolate). For the more isotropic shapes at l1𝑙1l\approx 1italic_l ≈ 1 in between, no nematic order is possible and the predicted transition region exceeds densities of presumed crystallization and even close packing, with an unphysical maximum at η=1𝜂1\eta=1italic_η = 1 in the extreme case. Moreover, both transitions are found to be of first order with comparable differences ΔηINΔsubscript𝜂IN\Delta\eta_{\text{IN}}roman_Δ italic_η start_POSTSUBSCRIPT IN end_POSTSUBSCRIPT of the packing fractions at coexistence, e.g., we find ΔηIN0.001Δsubscript𝜂IN0.001\Delta\eta_{\text{IN}}\approx 0.001roman_Δ italic_η start_POSTSUBSCRIPT IN end_POSTSUBSCRIPT ≈ 0.001 for cones and ΔηIN0.002Δsubscript𝜂IN0.002\Delta\eta_{\text{IN}}\approx 0.002roman_Δ italic_η start_POSTSUBSCRIPT IN end_POSTSUBSCRIPT ≈ 0.002 for cylinders with the same aspect ratio l=0.1𝑙0.1l=0.1italic_l = 0.1. As further expected, the phase behavior becomes asymptotically equal in the limits l0𝑙0l\rightarrow 0italic_l → 0 and l𝑙l\rightarrow\inftyitalic_l → ∞, where both shapes reduce to hard thin disks and Onsager rods Onsager (1949), respectively, as studied with the present functional in Ref. Wittmann et al., 2014.

However, there are quantitative differences of the phase transition for the two shapes. For example, the transition densities of the cylinder fluid peak at the unphysical value η=1𝜂1\eta=1italic_η = 1 of the packing fraction for an aspect ratio of l*=1superscript𝑙1l^{*}=1italic_l start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT = 1, while this happens for l*1.4superscript𝑙1.4l^{*}\approx 1.4italic_l start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT ≈ 1.4 in the case of cones. This behavior is a direct mathematical consequence of the particle geometry: we can identify a most isotropic shape (characterized by l=l*𝑙superscript𝑙l=l^{*}italic_l = italic_l start_POSTSUPERSCRIPT * end_POSTSUPERSCRIPT) by noticing that the order parameter S𝑆Sitalic_S drops out of the chosen functional when putting together all tensorial components, such that no ordered phase can be described at all, compare appendices B.1 and B.2. We are more interested in the regions, where the transition densities are lower and can thus be expected to describe the actual physical behavior. For comparison, the functional used here predicts the onset of smectic order in a fluid of hard spherocylinders at packing fractions η0.45greater-than-or-equivalent-to𝜂0.45\eta\gtrsim 0.45italic_η ≳ 0.45 Wittmann et al. (2014) and we indicate this lower bound by using a gray background in Fig. 4. For all relevant aspect ratios, we predict that the isotropic–nematic transition occurs at a larger packing fraction for cylinders than for cones. As the volume of a cone is only one third of that of a cylinder, the particle number at a given η𝜂\etaitalic_η is larger, which results in a stronger drive towards orientational order (in turn, the transition occurs at a lower particle number for cylinders). This behavior is consistent with that shown in Fig. 2 of Ref. Kubala et al., 2023 for hard connected spheres with different radii (mind that the aspect ratios of the bodies compared in this figure are not equal).

III.3 Biaxial bodies: different directors

The homogeneous phase diagram of hard isosceles spherotriangles in Fig. 5 depicts the different transition densities as a function of the shape ratio x𝑥xitalic_x for two aspect ratios l=5𝑙5l=5italic_l = 5 (left) and l=25𝑙25l=25italic_l = 25 (right). We find that three distinct uniaxial phases, denoted by Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT, Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT and Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT, which we model independently by choosing the main axis in the molecular frame to be the triangle height, the face normal and the triangle base, respectively, are stable, compare Fig. 2.

The transition from the isotropic phase to a (uniaxial) nematic is always of first order. In the two limiting cases x=0𝑥0x=0italic_x = 0 and x=1𝑥1x=1italic_x = 1, we recover the known result for hard spherocylinders, which have been demonstrated to agree well with simulation results Wittmann et al. (2014). It is thus reassuring for our predictions to remain meaningful at intermediate values of x𝑥xitalic_x. For increasing biaxiality of the particles (larger shape ratio when forming a Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT or smaller shape ratio when forming Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT) the packing fraction at which the transition occurs increases, as the main axis becomes shorter, thus destabilizing (prolate) nematic order. When the particle shape is sufficiently oblate, the stability limit of the isotropic phase is found at higher packing fractions than in the more prolate case and we eventually predict a transition to the Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT phase. Here, the corresponding transition densities are largely independent of the shape ratio. At higher densities, we accordingly find two first-order transitions between uniaxial phases when increasing the shape ratio: first from Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT to Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT and then from Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT to Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT. In both cases, the transition region is slightly bent, such that the oblate phase destabilizes upon increasing the density. Hence, there is a small range of shape ratios at which we observe two uniaxial phases following the sequence isotropic to uniaxial oblate nematic to uniaxial prolate nematic and, finally, to biaxial nematic.

The biaxial phase, which we identify as a perturbation to simple uniaxial order (compare Sec. II.3.4), is most stable for shape ratios close to a uniaxial–uniaxial transition, which points to an equal weight of two distinct directors. Moreover, the metastable uniaxial–uniaxial transition within the biaxial region (continued dotted lines) approximately indicates the dominant axis in the case of biaxial order (which we did not determine explicitly). Our calculations are based on the assumption that the uniaxial–biaxial transition is of second order while we cannot fully rule out the possibility of a first-order transition within our current analytic treatment. It is, however, quite reassuring that the biaxial perturbations of two distinct uniaxial phases at the uniaxial–uniaxial transition agree closely, suggesting that our calculations are consistent. In contrast, we discuss in appendix C that the more simplistic method to assume that one axis is perfectly ordered (compare Sec. II.3.3) is not consistent for the different director orientations: the stability of the biaxial phase is overestimated for taking Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT or Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT as the reference state but underestimated for Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT. On the other hand, these more approximate transition lines (not included in Fig. 5, but detailed in Fig. 6 below) are fully analytic and the second-order nature can explicitly be verified.

The qualitative behavior described above does not differ strongly between the two aspect ratios l=5𝑙5l=5italic_l = 5 and l=25𝑙25l=25italic_l = 25 considered. However, we find that the range of shape ratios x𝑥xitalic_x which give rise to a stable Noo{}^{\text{o}}start_FLOATSUPERSCRIPT o end_FLOATSUPERSCRIPT phase increases due to the generally more oblate shape of spherotriangles with larger aspect ratio l𝑙litalic_l. More strikingly, the onset of biaxiality occurs at lower densities for these thinner particles, as the shape also becomes less isotropic (in the opposite limit, l=0𝑙0l=0italic_l = 0, a spherotriangle reduces to a perfect sphere). This observation is quite important when contemplating the global stability of the biaxial phase, as we must take into account that other phases with positional order will probably preempt our predicted transitions at packing fractions η0.45greater-than-or-equivalent-to𝜂0.45\eta\gtrsim 0.45italic_η ≳ 0.45 (as indicated by the fading backgrounds in Fig. 5). Hence, our results suggest that biaxial order should become stable for larger aspect ratios and, therefore, shape ratios closer to zero or one (for which oblate and prolate uniaxial phases coexist). This conclusion is consistent with previous observations of biaxial nematic order in a system of extremely anisotropic biaxial particles Dussi et al. (2018).

IV Conclusions

In this work we have provided and applied a general recipe to investigate the homogeneous phase behavior of biaxial hard particle fluids within fundamental measure theory (FMT). This framework allows us to determine all phase boundaries by solving independent algebraic equations. In addition, we have demonstrated that treating biaxiality as a perturbation to uniaxial order constitutes a much more consistent procedure to determine the onset of the biaxial phase than the simple assumption of perfect uniaxial order. Specifically, we have considered four order parameters that are established measures for orientational order in systems of biaxial particles. While, in principle, even more order parameters will become important for more general shapes, we found here that, upon choosing appropriate coordinate frames, all transitions between homogeneous phases can be consistently identified when taking into account only two parameters: the standard uniaxial order parameter S𝑆Sitalic_S and the order parameter F𝐹Fitalic_F, which measures biaxiality. We have investigated different particle shapes with a relatively low symmetry, which still allowed for a detailed characterization of the phase behavior using the order parameters at hand. As exemplified here for hard cones, uniaxial polar shapes typically require only one order parameter S𝑆Sitalic_S to describe bulk nematic order if the chosen coordinate systems are properly aligned with the symmetry axis. The formation of a phase with global polar order can be most likely ruled out in our hard-core system for entropic reasons. For biaxial shapes with D2hsubscript𝐷2hD_{2\mathrm{h}}italic_D start_POSTSUBSCRIPT 2 roman_h end_POSTSUBSCRIPT symmetry (three mutually orthogonal symmetry planes) the four order parameters considered here are a standard choice Rosso (2007). We have demonstrated here that no additional order parameters are needed to describe the biaxial nematic phase as long as all relevant axes are polar, which is the case for hard isosceles spherotriangles, but not for general hard spherotriangles.

Biaxial order can also emerge in mixtures of uniaxial bodies, specifically those involving both prolate and oblate species Stroobants and Lekkerkerker (1984); Camp and Allen (1996); Do Carmo et al. (2010). From the point of view of FMT, which naturally applies to mixtures without any conceptual complication, we expect that the phase biaxiality order parameter P𝑃Pitalic_P will play an important role in such a scenario. A detailed investigation would be an interesting perspective for future work. In turn, for arbitrary (convex) hard particles, FMT offers a straightforward way of identifying additional relevant order parameters through averages of products of distinct components of the rotation matrix (11) between the body frame and the lab frame, following the recipe outlined in Sec. II.2.4. A comprehensive investigation of possible polar order using FMT would potentially require adding tensorial Wittmann et al. (2014) or mixed Wittmann et al. (2015a) weighted densities to the current functional, or generalizing its expansion into spherical harmonics Marechal et al. (2017), and performing a numerical minimization.

Although our present investigation is restricted to spatially homogeneous systems, FMT can also be applied to inhomogeneous situations, which are encountered in the presence of external walls Hansen-Goos and Mecke (2009); Marechal and Löwen (2013); Schönhöfer et al. (2018), for free interfaces between coexisting homogeneous phases Wittmann and Mecke (2014) or when more complex liquid crystal phases with positional order emerge Wittmann et al. (2014); Marechal et al. (2017). In view of polar shapes, it would further be interesting to investigate the possibility of local polar order in adjacent splay domains Mertelj et al. (2018); Sebastián et al. (2020); Kubala et al. (2023) or twist-bend or splay-bend structures Chaturvedi and Kamien (2019); Chiappini and Dijkstra (2021); Kotni et al. (2022), scenarios in which the global orientational order remains nematic. It is also worthwhile to calculate the Frank elastic coefficients Wittmann et al. (2015b); De Gregorio et al. (2016) for different particles and investigate the effect of shape polarity and biaxiality on the elastic behavior. An important issue concerns the global stability of the biaxial order predicted here: in Ref. Cuetos et al., 2017 the transition to a biaxial nematic was found to be preempted by the onset of a smectic phase. While we expect that the same happens for spherotriangles with l=5𝑙5l=5italic_l = 5, flatter particles, like those with l=25𝑙25l=25italic_l = 25, were predicted here to exhibit biaxial behavior at lower packing fractions, such that it is more likely that this phase is actually stable for moderate densities (in particular for even larger values of l𝑙litalic_l). An elegant possibility to increase the stability range of the spatially homogeneous phases is to introduce slight modifications to the system, such as polydispersity Belli et al. (2011); Chiappini et al. (2019) or depletion interactions Belli et al. (2012), which can destabilize positional order. Also considering rounder shapes, which enhance the chance of a particle to slide out of a smectic layer, can favor the stability of a biaxial phase Chiappini et al. (2019). Hence, one may suspect that a similar mechanism could be at work for smectics formed by spherotriangles (which require the particles to align in layers with an alternating up-down configuration), such that these could be less stable than those formed by spheroplatelets, which would be worthwhile to investigate in future work.

As a next step, the Brownian dynamics of the hard particles considered here could be explored by employing the present functional in a dynamical DFT (DDFT) Marconi and Tarazona (1999); Archer and Evans (2004); te Vrugt et al. (2020); te Vrugt and Wittkowski (2022). For example, in orientationally ordered phases, there is an anisotropic long-time diffusion which has been explored for uniaxial particles Löwen (1999), but not yet for biaxial particles. As a computationally cheaper alternative for spatially inhomogeneous problems, phase field crystal (PFC) models Elder et al. (2002); Emmerich et al. (2012) are commonly used to study the dynamics in complex systems. Here, the numerical effort is reduced by considering the dynamics of an orientation-averaged density field and different orientational order parameters rather than that of the full orientation-resolved density Wittkowski et al. (2010); Emmerich et al. (2012); te Vrugt et al. (2022, 2021). Consequently, the results obtained here, which allow to express DFT functionals as a function of orientational order parameters in the biaxial case, are an excellent starting point to derive a PFC model for biaxial particles. Finally, a further extension would be to investigate active biaxial particles Wittkowski and Löwen (2012), which, also owing to the less symmetric particle shape, may exhibit circle swimming behavior Kümmel et al. (2013). Being a nonequilibrium system, this would again require the use of a dynamical theory. By combining an appropriate FMT functional with the DDFT for biaxial active particles developed in Ref. Wittkowski and Löwen, 2011, the strategy presented in this work could be generalized to investigate the dynamics of the biaxial order parameters in the active system.

Acknowledgments

The authors would like to thank Michael A. Klatt and Paul A. Monderkamp for helpful suggestions. Funding by the Deutsche Forschungsgemeinschaft (DFG) under Project-ID 525063330 (MtV) and through the SPP 2265 under grant numbers LO 418/25-1 (HL) and WI 5527/1-1 (RW) is gratefully acknowledged. AEM wants to thank the Studienstiftung des deutschen Volkes for financial support.

Appendix A Calculation of weighted densities

In this appendix, we provide additional details required to follow the general procedure, outlined in Sec. II.2.4, to calculate the weighted densities of the hard bodies shown in Fig. 1.

A.1 Body parts

We begin by presenting an appropriate parameterization for all body parts of hard cones, cylinders and spherotriangles, required as the second step in Sec. II.2.4. This allows us to replace the integral d𝐫subscriptdifferential-d𝐫\int_{\partial\mathcal{B}}\mathrm{d}{\mathbf{r}}∫ start_POSTSUBSCRIPT ∂ caligraphic_B end_POSTSUBSCRIPT roman_d bold_r over the surface \partial\mathcal{B}∂ caligraphic_B of the body in Eqs. (41)-(46) by a sum of integrals corresponding to all contributing body parts, with respect to two parameters each, as specified below. Here, we choose the parameterization, such that the symmetry axis of the cone and the cylinder points in the z𝑧zitalic_z-direction (i.e., it is parallel to m3subscript𝑚3\vec{m}_{3}over→ start_ARG italic_m end_ARG start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT). For the isosceles spherotriangles, we choose the height of the triangle, compare the left picture in Fig. 2a.

A.1.1 Cone mantle

We parameterize the surface of the cone’s mantle in the form

𝐫(t,φ)=(R(1tL)cos(φ)R(1tL)sin(φ)HtL),𝐫𝑡𝜑𝑅1𝑡𝐿𝜑𝑅1𝑡𝐿𝜑𝐻𝑡𝐿{\mathbf{r}}(t,\varphi)=\left(\begin{array}[]{c}R(1-\frac{t}{L})\cos(\varphi)% \\ R(1-\frac{t}{L})\sin(\varphi)\\ \frac{Ht}{L}\end{array}\right),bold_r ( italic_t , italic_φ ) = ( start_ARRAY start_ROW start_CELL italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_H italic_t end_ARG start_ARG italic_L end_ARG end_CELL end_ROW end_ARRAY ) , (85)

where L𝐿Litalic_L is defined as

L:=H2+R2.assign𝐿superscript𝐻2superscript𝑅2L:=\sqrt{H^{2}+R^{2}}.italic_L := square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG . (86)

Here, R=D/2𝑅𝐷2R=D/2italic_R = italic_D / 2 is the radius of the cone and H𝐻Hitalic_H is its height. We use φ𝜑\varphiitalic_φ and t𝑡titalic_t to parameterize the cone surface, where φ𝜑\varphiitalic_φ is the polar angle and t𝑡titalic_t gives the distance along the cone mantle from the base of the cone. Therefore, t[0,L]𝑡0𝐿t\in[0,L]italic_t ∈ [ 0 , italic_L ] and φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ].

A.1.2 Disk

The surface of a disk of radius R=D/2𝑅𝐷2R=D/2italic_R = italic_D / 2 is parameterized by polar coordinates given by two parameters r[0,R]𝑟0𝑅r\in[0,R]italic_r ∈ [ 0 , italic_R ] and φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ]. Therefore,

𝐫(r,φ)=(rcos(φ)rsin(φ)0).𝐫𝑟𝜑𝑟𝜑𝑟𝜑0{\mathbf{r}}(r,\varphi)=\left(\begin{array}[]{c}r\cos\left(\varphi\right)\\ r\sin\left(\varphi\right)\\ 0\end{array}\right).bold_r ( italic_r , italic_φ ) = ( start_ARRAY start_ROW start_CELL italic_r roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_r roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARRAY ) . (87)

The cone only requires one disk with the normal vector pointing down, while the cylinder requires two disks with same parameterizations but opposite normal vectors.

A.1.3 Circular ring (torus)

A small complication arises for the cone and the cylinder, as these bodies contain points with infinite curvature, where the normal vector is not well defined. To resolve this issue at the circular ring between mantle and cap** disks, we consider appropriate parts of a torus with width r𝑟ritalic_r and radius R=D/2𝑅𝐷2R=D/2italic_R = italic_D / 2. Then, after defining all necessary geometric quantities, we take the limit r0𝑟0r\to 0italic_r → 0. For the calculation of the weighted densities, we use standard torus coordinates, which are defined as

𝐫(φ,ξ)=((R+rcos(ξ))cos(φ)(R+rcos(ξ))sin(φ)rsin(ξ))𝐫𝜑𝜉𝑅𝑟𝜉𝜑𝑅𝑟𝜉𝜑𝑟𝜉{\mathbf{r}}\left(\varphi,\xi\right)=\left(\begin{array}[]{c}(R+r\cos(\xi))% \cos(\varphi)\\ (R+r\cos(\xi))\sin(\varphi)\\ r\sin(\xi)\end{array}\right)bold_r ( italic_φ , italic_ξ ) = ( start_ARRAY start_ROW start_CELL ( italic_R + italic_r roman_cos ( italic_ξ ) ) roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL ( italic_R + italic_r roman_cos ( italic_ξ ) ) roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_r roman_sin ( italic_ξ ) end_CELL end_ROW end_ARRAY ) (88)

where φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ] and, in the case of the cone, ξ[0,arccos(R/L)]𝜉0𝑅𝐿\xi\in[0,\arccos\left(-R/L\right)]italic_ξ ∈ [ 0 , roman_arccos ( - italic_R / italic_L ) ]. For hard cylinders, we need two tori with ξ[0,π/2]𝜉0𝜋2\xi\in[0,\pi/2]italic_ξ ∈ [ 0 , italic_π / 2 ] and ξ[π/2,π]𝜉𝜋2𝜋\xi\in[\pi/2,\pi]italic_ξ ∈ [ italic_π / 2 , italic_π ]. When the corresponding weighted densities are calculated (according to the six steps in Sec. II.2.4), we finally set r0𝑟0r\to 0italic_r → 0.

A.1.4 Sphere

The surface of a sphere can be parameterized by using the spherical coordinates

𝐫(ϑ,φ)=(Rsin(ϑ)cos(φ)Rsin(ϑ)sin(φ)Rcos(ϑ))𝐫italic-ϑ𝜑𝑅italic-ϑ𝜑𝑅italic-ϑ𝜑𝑅italic-ϑ{\mathbf{r}}\left(\vartheta,\varphi\right)=\left(\begin{array}[]{c}R\sin(% \vartheta)\cos(\varphi)\\ R\sin(\vartheta)\sin(\varphi)\\ R\cos(\vartheta)\end{array}\right)bold_r ( italic_ϑ , italic_φ ) = ( start_ARRAY start_ROW start_CELL italic_R roman_sin ( italic_ϑ ) roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R roman_sin ( italic_ϑ ) roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R roman_cos ( italic_ϑ ) end_CELL end_ROW end_ARRAY ) (89)

and fixing the radial coordinate R𝑅Ritalic_R. For the cone, we require a part of a sphere in the limit of R0𝑅0R\to 0italic_R → 0 on its tip, which allows us to calculate well-defined geometrical quantities, as explained above for the circular ring. In this case, we have ϑ[0,πarccos(R/L)]italic-ϑ0𝜋𝑅𝐿\vartheta\in[0,\pi-\arccos\left(-R/L\right)]italic_ϑ ∈ [ 0 , italic_π - roman_arccos ( - italic_R / italic_L ) ] and φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ]. In practice, a sphere in the limit R0𝑅0R\to 0italic_R → 0 only contributes to the integrated Gaussian curvature and thus only to the scalar weighted density n0subscript𝑛0n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT (together with the circular ring), which can be set directly to n0=ρsubscript𝑛0𝜌n_{0}=\rhoitalic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT = italic_ρ, as the one is a simply connected body. For spherotriangles, there are three parts of a sphere with radius R=D/2𝑅𝐷2R=D/2italic_R = italic_D / 2 on their edges. As these parts will always add up to a full sphere, we can simply assume φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ] and ϑ[0,π]italic-ϑ0𝜋\vartheta\in[0,\pi]italic_ϑ ∈ [ 0 , italic_π ] for all contributions at once.

A.1.5 Cylinder Mantle

In our calculations for a hard cylinder, we parameterize the cylinder mantle by choosing a parameter h[0,H]0𝐻h\in[0,H]italic_h ∈ [ 0 , italic_H ], where H𝐻Hitalic_H is the full height of the cylinder and φ[0,2π]𝜑02𝜋\varphi\in[0,2\pi]italic_φ ∈ [ 0 , 2 italic_π ] is an angle. This equivalent to using standard cylinder coordinates with a fixed radial coordinate R=D/2𝑅𝐷2R=D/2italic_R = italic_D / 2. This corresponds to the parameterization

𝐫(h,φ)=(Rcos(φ)Rsin(φ)h)𝐫𝜑𝑅𝜑𝑅𝜑\displaystyle{\mathbf{r}}\left(h,\varphi\right)=\left(\begin{array}[]{c}R\cos% \left(\varphi\right)\cr R\sin\left(\varphi\right)\cr h\end{array}\right)bold_r ( italic_h , italic_φ ) = ( start_ARRAY start_ROW start_CELL italic_R roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_h end_CELL end_ROW end_ARRAY ) (93)

For spherotriangles, we must consider three parts of cylinders with h[0,A]0𝐴h\in[0,A]italic_h ∈ [ 0 , italic_A ] or h[0,B]0𝐵h\in[0,B]italic_h ∈ [ 0 , italic_B ] and always φ[π/2,π/2]𝜑𝜋2𝜋2\varphi\in[-\pi/2,\pi/2]italic_φ ∈ [ - italic_π / 2 , italic_π / 2 ]. Moreover, for each part, 𝐫(h,φ)𝐫𝜑{\mathbf{r}}\left(h,\varphi\right)bold_r ( italic_h , italic_φ ) needs to be rotated in the xz𝑥𝑧xzitalic_x italic_z-plane. This rotation angle is π/2𝜋2-\pi/2- italic_π / 2 for the base line of length A𝐴Aitalic_A, and γ𝛾\gammaitalic_γ or πγ𝜋𝛾\pi-\gammaitalic_π - italic_γ for the other two sides of length B𝐵Bitalic_B.

A.1.6 Triangle

Flat isosceles triangles are parameterized with Cartesian coordinates, therefore

𝐫(x,z)=(x0z),𝐫𝑥𝑧𝑥0𝑧{\mathbf{r}}\left(x,z\right)=\left(\begin{array}[]{c}x\\ 0\\ z\end{array}\right),bold_r ( italic_x , italic_z ) = ( start_ARRAY start_ROW start_CELL italic_x end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW start_ROW start_CELL italic_z end_CELL end_ROW end_ARRAY ) , (94)

where x[A/2,A/2]𝑥𝐴2𝐴2x\in[-A/2,A/2]italic_x ∈ [ - italic_A / 2 , italic_A / 2 ] and z[0,B2A2/4(12|x|/A)]𝑧0superscript𝐵2superscript𝐴2412𝑥𝐴z\in[0,\sqrt{B^{2}-A^{2}/4}(1-2|x|/A)]italic_z ∈ [ 0 , square-root start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT / 4 end_ARG ( 1 - 2 | italic_x | / italic_A ) ] for x[0,A/2]𝑥0𝐴2x\in[0,A/2]italic_x ∈ [ 0 , italic_A / 2 ]. We require in total two triangles with opposite normal vectors.

A.2 Geometrical measures

Using the different parameterizations of the relevant body parts, we can calculate all geometric quantities in Eqs. (41)-(46), i.e., the two principal curvature directions 𝐯1subscript𝐯1{\mathbf{v}}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝐯2subscript𝐯2{\mathbf{v}}_{2}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the surface unit normal vector 𝐧𝐧{\mathbf{n}}bold_n, the two principal curvatures κ1subscript𝜅1\kappa_{1}italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and κ2subscript𝜅2\kappa_{2}italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the Gaussian curvature 𝒢𝒢\mathcal{G}caligraphic_G and the mean curvature \mathcal{H}caligraphic_H. This third step in Sec. II.2.4, is carried out below explicitly for the exemplary case of a cone mantle, whose surface is parameterized in Eq. (85).

A.2.1 Unit vectors

To calculate the unit vectors, we take the derivative of the vector 𝐫(t,φ)𝐫𝑡𝜑{\mathbf{r}}(t,\varphi)bold_r ( italic_t , italic_φ ), given in Eq. (85), with respect to t𝑡titalic_t and φ𝜑\varphiitalic_φ, which yields

𝐫φ=(R(1tL)sin(φ)R(1tL)cos(φ)0)𝐫𝜑𝑅1𝑡𝐿𝜑𝑅1𝑡𝐿𝜑0\frac{\partial{\mathbf{r}}}{\partial\varphi}=\left(\begin{array}[]{c}-R(1-% \frac{t}{L})\sin(\varphi)\\ R(1-\frac{t}{L})\cos(\varphi)\\ 0\end{array}\right)divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG = ( start_ARRAY start_ROW start_CELL - italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARRAY ) (95)

and

𝐫t=(RLcos(φ)RLsin(φ)HL).𝐫𝑡𝑅𝐿𝜑𝑅𝐿𝜑𝐻𝐿\frac{\partial{\mathbf{r}}}{\partial t}=\left(\begin{array}[]{c}-\frac{R}{L}% \cos(\varphi)\\ -\frac{R}{L}\sin(\varphi)\\ \frac{H}{L}\end{array}\right).divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_t end_ARG = ( start_ARRAY start_ROW start_CELL - divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG end_CELL end_ROW end_ARRAY ) . (96)

Both of these two vectors are tangential to the cone’s surface and perpendicular to each other. Therefore, these vectors, after being normalized, yield the expressions

𝐯1=(sin(φ)cos(φ)0).subscript𝐯1𝜑𝜑0{\mathbf{v}}_{1}=\left(\begin{array}[]{c}-\sin(\varphi)\\ \cos(\varphi)\\ 0\end{array}\right).bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = ( start_ARRAY start_ROW start_CELL - roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARRAY ) . (97)

and

𝐯2=(RLcos(φ)RLsin(φ)HL),subscript𝐯2𝑅𝐿𝜑𝑅𝐿𝜑𝐻𝐿{\mathbf{v}}_{2}=\left(\begin{array}[]{c}-\frac{R}{L}\cos(\varphi)\\ -\frac{R}{L}\sin(\varphi)\\ \frac{H}{L}\end{array}\right),bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = ( start_ARRAY start_ROW start_CELL - divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL - divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG end_CELL end_ROW end_ARRAY ) , (98)

and are precisely the vectors 𝐯1subscript𝐯1{\mathbf{v}}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT and 𝐯2subscript𝐯2{\mathbf{v}}_{2}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT we have been looking for.

Going on, we find that

𝐫φ×𝐫t=RL(H(1tL)cos(φ)H(1tL)sin(φ)R(1tL)).𝐫𝜑𝐫𝑡𝑅𝐿𝐻1𝑡𝐿𝜑𝐻1𝑡𝐿𝜑𝑅1𝑡𝐿\frac{\partial{\mathbf{r}}}{\partial\varphi}\times\frac{\partial{\mathbf{r}}}{% \partial t}=\frac{R}{L}\left(\begin{array}[]{c}H\left(1-\frac{t}{L}\right)\cos% (\varphi)\\ H\left(1-\frac{t}{L}\right)\sin(\varphi)\\ R\left(1-\frac{t}{L}\right)\end{array}\right).divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG × divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_t end_ARG = divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG ( start_ARRAY start_ROW start_CELL italic_H ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_H ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) end_CELL end_ROW end_ARRAY ) . (99)

This vector is necessarily perpendicular to the cone’s surface. After normalizing it, we get our required normal vector, which is given by

𝐧=(HLcos(φ)HLsin(φ)RL).𝐧𝐻𝐿𝜑𝐻𝐿𝜑𝑅𝐿{\mathbf{n}}=\left(\begin{array}[]{c}\frac{H}{L}\cos(\varphi)\\ \frac{H}{L}\sin(\varphi)\\ \frac{R}{L}\end{array}\right).bold_n = ( start_ARRAY start_ROW start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_R end_ARG start_ARG italic_L end_ARG end_CELL end_ROW end_ARRAY ) . (100)

A.2.2 Weingarten map and curvatures

Next, we determine the components of the Weingarten map, which allows us to calculate the principal curvatures of the cone. In order to do so, the first step is to calculate the metric tensor g𝑔gitalic_g of the cone’s mantle. The ij𝑖𝑗ijitalic_i italic_jth component of the metric tensor is defined as

gij:=𝐫xi𝐫xj,assignsubscript𝑔𝑖𝑗𝐫subscript𝑥𝑖𝐫subscript𝑥𝑗g_{ij}:=\frac{\partial{\mathbf{r}}}{\partial x_{i}}\cdot\frac{\partial{\mathbf% {r}}}{\partial x_{j}}\,,italic_g start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT := divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG , (101)

where xi,xj{φ,t}subscript𝑥𝑖subscript𝑥𝑗𝜑𝑡x_{i},x_{j}\in\{\varphi,t\}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∈ { italic_φ , italic_t }. Following this definition, we arrive at

g11=𝐫t𝐫t=R2L2(cos(φ)2+sin(φ)2)+H2L2=1g_{11}=\frac{\partial{\mathbf{r}}}{\partial t}\cdot\frac{\partial{\mathbf{r}}}% {\partial t}=\frac{R^{2}}{L^{2}}\left(\cos(\varphi)^{2}+\sin(\varphi)^{2}% \right)+\frac{H^{2}}{L^{2}}=1italic_g start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_t end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_t end_ARG = divide start_ARG italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( roman_cos ( italic_φ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_sin ( italic_φ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) + divide start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG = 1 (102)

and

g22=𝐫φ𝐫φsubscript𝑔22𝐫𝜑𝐫𝜑\displaystyle g_{22}=\frac{\partial{\mathbf{r}}}{\partial\varphi}\cdot\frac{% \partial{\mathbf{r}}}{\partial\varphi}italic_g start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT = divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG =R2(1tL)2(cos(φ)2+sin(φ)2)\displaystyle=R^{2}\left(1-\frac{t}{L}\right)^{2}\left(\cos(\varphi)^{2}+\sin(% \varphi)^{2}\right)\!\!\!\!\!\!\!\!= italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( roman_cos ( italic_φ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + roman_sin ( italic_φ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) (103)
=R2(1tL)2,absentsuperscript𝑅2superscript1𝑡𝐿2\displaystyle=R^{2}\left(1-\frac{t}{L}\right)^{2},= italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (104)

while the cross terms

g12=g21=𝐫t𝐫φ=0subscript𝑔12subscript𝑔21𝐫𝑡𝐫𝜑0g_{12}=g_{21}=\frac{\partial{\mathbf{r}}}{\partial t}\cdot\frac{\partial{% \mathbf{r}}}{\partial\varphi}=0italic_g start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = italic_g start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_t end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG = 0 (105)

vanish. Therefore, the metric tensor of the cone’s mantle can be represented by the matrix

g=(100R2(1tL)2),𝑔10missing-subexpression0superscript𝑅2superscript1𝑡𝐿2missing-subexpressiong=\left(\begin{array}[]{rrr}1&0\\ 0&R^{2}\left(1-\frac{t}{L}\right)^{2}\\ \end{array}\right),italic_g = ( start_ARRAY start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_CELL start_CELL end_CELL end_ROW end_ARRAY ) , (106)

whose inverse is given by

g1=(1001R2(1tL)2).superscript𝑔110missing-subexpression01superscript𝑅2superscript1𝑡𝐿2missing-subexpressiong^{-1}=\left(\begin{array}[]{rrr}1&0\\ 0&\frac{1}{R^{2}\left(1-\frac{t}{L}\right)^{2}}\\ \end{array}\right).italic_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT = ( start_ARRAY start_ROW start_CELL 1 end_CELL start_CELL 0 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL divide start_ARG 1 end_ARG start_ARG italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_CELL start_CELL end_CELL end_ROW end_ARRAY ) . (107)

In order to calculate the components of the Weingarten map, we need a second matrix. Its elements are defined by

Bij=𝐧xi𝐫xj,subscript𝐵𝑖𝑗𝐧subscript𝑥𝑖𝐫subscript𝑥𝑗B_{ij}=\frac{\partial{\mathbf{n}}}{\partial x_{i}}\cdot\frac{\partial{\mathbf{% r}}}{\partial x_{j}},italic_B start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT = divide start_ARG ∂ bold_n end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT end_ARG , (108)

where xi,xj{φ,t}subscript𝑥𝑖subscript𝑥𝑗𝜑𝑡x_{i},x_{j}\in\{\varphi,t\}italic_x start_POSTSUBSCRIPT italic_i end_POSTSUBSCRIPT , italic_x start_POSTSUBSCRIPT italic_j end_POSTSUBSCRIPT ∈ { italic_φ , italic_t }. We calculate

𝐧φ=(HLsin(φ)HLcos(φ)0),nt=0.formulae-sequence𝐧𝜑𝐻𝐿𝜑𝐻𝐿𝜑0𝑛𝑡0\frac{\partial{\mathbf{n}}}{\partial\varphi}=\left(\begin{array}[]{c}-\frac{H}% {L}\sin(\varphi)\\ \frac{H}{L}\cos(\varphi)\\ 0\end{array}\right)\,,\ \ \ \frac{\partial\vec{n}}{\partial t}=\vec{0}\,.divide start_ARG ∂ bold_n end_ARG start_ARG ∂ italic_φ end_ARG = ( start_ARRAY start_ROW start_CELL - divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG roman_sin ( italic_φ ) end_CELL end_ROW start_ROW start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG roman_cos ( italic_φ ) end_CELL end_ROW start_ROW start_CELL 0 end_CELL end_ROW end_ARRAY ) , divide start_ARG ∂ over→ start_ARG italic_n end_ARG end_ARG start_ARG ∂ italic_t end_ARG = over→ start_ARG 0 end_ARG . (109)

As we have calculated all required derivatives, we are now ready to calculate the whole matrix. Both B11=0subscript𝐵110B_{11}=0italic_B start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = 0 and B12=0subscript𝐵120B_{12}=0italic_B start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT = 0 vanish immediately, since they contain the vanishing factor n/t𝑛𝑡\partial\vec{n}/\partial t∂ over→ start_ARG italic_n end_ARG / ∂ italic_t. Moreover, B21=0subscript𝐵210B_{21}=0italic_B start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT = 0 vanishes since n/ϕ𝑛italic-ϕ\partial\vec{n}/\partial\phi∂ over→ start_ARG italic_n end_ARG / ∂ italic_ϕ is perpendicular to t/t𝑡𝑡\partial\vec{t}/\partial t∂ over→ start_ARG italic_t end_ARG / ∂ italic_t from Eq. (96). Thus, only

B22=𝐧φ𝐫φ=HLR(1tL)subscript𝐵22𝐧𝜑𝐫𝜑𝐻𝐿𝑅1𝑡𝐿B_{22}=\frac{\partial{\mathbf{n}}}{\partial\varphi}\cdot\frac{\partial{\mathbf% {r}}}{\partial\varphi}=\frac{H}{L}R\left(1-\frac{t}{L}\right)\,italic_B start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT = divide start_ARG ∂ bold_n end_ARG start_ARG ∂ italic_φ end_ARG ⋅ divide start_ARG ∂ bold_r end_ARG start_ARG ∂ italic_φ end_ARG = divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) (110)

contributes a nontrivial result, such that

B=(000HLR(1tL)).𝐵00missing-subexpression0𝐻𝐿𝑅1𝑡𝐿missing-subexpressionB=\left(\begin{array}[]{rrr}0&0\\ 0&\frac{H}{L}R\left(1-\frac{t}{L}\right)\\ \end{array}\right).italic_B = ( start_ARRAY start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L end_ARG italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) end_CELL start_CELL end_CELL end_ROW end_ARRAY ) . (111)

Finally, the matrix representation of the Weingarten map is given by

W=g1B=(000HLR(1tL)),𝑊superscript𝑔1𝐵00missing-subexpression0𝐻𝐿𝑅1𝑡𝐿missing-subexpressionW=g^{-1}\cdot B=\left(\begin{array}[]{rrr}0&0\\ 0&\frac{H}{LR\left(1-\frac{t}{L}\right)}\\ \end{array}\right),italic_W = italic_g start_POSTSUPERSCRIPT - 1 end_POSTSUPERSCRIPT ⋅ italic_B = ( start_ARRAY start_ROW start_CELL 0 end_CELL start_CELL 0 end_CELL start_CELL end_CELL end_ROW start_ROW start_CELL 0 end_CELL start_CELL divide start_ARG italic_H end_ARG start_ARG italic_L italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) end_ARG end_CELL start_CELL end_CELL end_ROW end_ARRAY ) , (112)

where we have inserted Eqs. (107) and (111).

From differential geometry, it is known that the principal curvatures are just the eigenvalues of the Weingarten map. Since W𝑊Witalic_W is a diagonal matrix, the eigenvalues are κ2=0subscript𝜅20\kappa_{2}=0italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 with principal direction 𝐯2subscript𝐯2{\mathbf{v}}_{2}bold_v start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT and κ1=HLR(1t/L)subscript𝜅1𝐻𝐿𝑅1𝑡𝐿\kappa_{1}=\frac{H}{LR\left(1-t/L\right)}italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT = divide start_ARG italic_H end_ARG start_ARG italic_L italic_R ( 1 - italic_t / italic_L ) end_ARG with principal direction 𝐯1subscript𝐯1{\mathbf{v}}_{1}bold_v start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT. Therefore, the cone mantle has the Gaussian curvature

𝒦=κ1κ2=0𝒦subscript𝜅1subscript𝜅20\mathcal{K}=\kappa_{1}\kappa_{2}=0caligraphic_K = italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = 0 (113)

and the mean curvature

=12(κ1+κ2)=12HLR(1tL),12subscript𝜅1subscript𝜅212𝐻𝐿𝑅1𝑡𝐿\mathcal{H}=\frac{1}{2}(\kappa_{1}+\kappa_{2})=\frac{1}{2}\frac{H}{LR\left(1-% \frac{t}{L}\right)}\,,caligraphic_H = divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( italic_κ start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT + italic_κ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) = divide start_ARG 1 end_ARG start_ARG 2 end_ARG divide start_ARG italic_H end_ARG start_ARG italic_L italic_R ( 1 - divide start_ARG italic_t end_ARG start_ARG italic_L end_ARG ) end_ARG , (114)

which depends on the position t𝑡titalic_t on the cone’s mantle. As expected, the curvatures are identical to those of a cylinder in the limit H𝐻H\to\inftyitalic_H → ∞, L𝐿L\to\inftyitalic_L → ∞ with H/L1𝐻𝐿1H/L\to 1italic_H / italic_L → 1 and finite R𝑅Ritalic_R.

Appendix B Weighted densities

B.1 Hard cones

We now present the full set of relevant weighted densities for hard cones with height H𝐻Hitalic_H and base diameter D=2R𝐷2𝑅D=2Ritalic_D = 2 italic_R. For simplicity, we neglect the order parameter P𝑃Pitalic_P by assuming P=0𝑃0P=0italic_P = 0, as it is irrelevant for the homogeneous bulk phase behavior for our choice of coordinates (compare the discussion in Sec. II.1.4). The contribution of P𝑃Pitalic_P and other order parameters will be explicitly shown in appendix B.4 for hard spherocylinders.

We begin by giving the scalar weighted densities

n0subscript𝑛0\displaystyle n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =ρ,absent𝜌\displaystyle=\rho\,,= italic_ρ , (115)
n1subscript𝑛1\displaystyle n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρ4(H+Rarccos(RH2+R2)),absent𝜌4𝐻𝑅𝑅superscript𝐻2superscript𝑅2\displaystyle=\frac{\rho}{4}\left(H+R\,\arccos\!\left(\frac{-R}{\sqrt{H^{2}+R^% {2}}}\right)\right),= divide start_ARG italic_ρ end_ARG start_ARG 4 end_ARG ( italic_H + italic_R roman_arccos ( divide start_ARG - italic_R end_ARG start_ARG square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) ) , (116)
n2subscript𝑛2\displaystyle n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρπR(R+H2+R2),absent𝜌𝜋𝑅𝑅superscript𝐻2superscript𝑅2\displaystyle=\rho\,\pi R(R+\sqrt{H^{2}+R^{2}})\,,= italic_ρ italic_π italic_R ( italic_R + square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (117)
n3subscript𝑛3\displaystyle n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =ρ3πR2H.absent𝜌3𝜋superscript𝑅2𝐻\displaystyle=\frac{\rho}{3}\pi R^{2}H\,.= divide start_ARG italic_ρ end_ARG start_ARG 3 end_ARG italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_H . (118)

For our choice of coordinates, only the third component

(n1)3subscriptsubscript𝑛13\displaystyle(\overrightarrow{n}_{1})_{3}( over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =ρRHRH2+R24H2+R2cos(θ)absent𝜌𝑅𝐻𝑅superscript𝐻2superscript𝑅24superscript𝐻2superscript𝑅2delimited-⟨⟩𝜃\displaystyle=\rho\,R\,\frac{H-R-\sqrt{H^{2}+R^{2}}}{4\sqrt{H^{2}+R^{2}}}\,% \langle\cos(\theta)\rangle= italic_ρ italic_R divide start_ARG italic_H - italic_R - square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG start_ARG 4 square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ⟨ roman_cos ( italic_θ ) ⟩ (119)

of the first vector weighted density is nonzero, where cos(θ)=d𝐎g(ϕ,θ,ψ)cos(θ)delimited-⟨⟩𝜃differential-d𝐎𝑔italic-ϕ𝜃𝜓𝜃\langle\cos(\theta)\rangle=\int\mathrm{d}\mathbf{O}\,g(\phi,\theta,\psi)\,\cos% (\theta)⟨ roman_cos ( italic_θ ) ⟩ = ∫ roman_d bold_O italic_g ( italic_ϕ , italic_θ , italic_ψ ) roman_cos ( italic_θ ) denotes the orientational average of cos(θ)𝜃\cos(\theta)roman_cos ( italic_θ ), as in Eq. (22), which is zero for (apolar) nematic order. Since n2=0subscript𝑛20\overrightarrow{n}_{2}=\overrightarrow{0}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = over→ start_ARG 0 end_ARG, the vectors do not contribute overall, such that there is no order parameter like cos(θ)delimited-⟨⟩𝜃\langle\cos(\theta)\rangle⟨ roman_cos ( italic_θ ) ⟩ that is sensitive to polar order in the functional employed here. In view of other shapes, we learn here that the vectorial contributions do not necessarily support polar order even if the particle shape is polar, such that it appears to be a safe approximation to generally neglect these terms in our study.

The diagonal tensor weighted densities read

(n1)11subscriptsubscript𝑛111\displaystyle(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =(n1)22absentsubscriptsubscript𝑛122\displaystyle=(\overleftrightarrow{n}_{1})_{22}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT (120)
=ρ24(H5H22R2(H2+R2)3Rx^)S,absent𝜌24𝐻5superscript𝐻22superscript𝑅2superscript𝐻2superscript𝑅23𝑅^𝑥𝑆\displaystyle=\frac{\rho}{24}\left(H\,\frac{5H^{2}-2R^{2}}{\left(H^{2}+R^{2}% \right)}-3R\hat{x}\right)S,= divide start_ARG italic_ρ end_ARG start_ARG 24 end_ARG ( italic_H divide start_ARG 5 italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG - 3 italic_R over^ start_ARG italic_x end_ARG ) italic_S , (121)
(n1)33subscriptsubscript𝑛133\displaystyle(\overleftrightarrow{n}_{1})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ24(H4R210H2(H2+R2)+6Rx^)S,absent𝜌24𝐻4superscript𝑅210superscript𝐻2superscript𝐻2superscript𝑅26𝑅^𝑥𝑆\displaystyle=\frac{\rho}{24}\left(H\,\frac{4R^{2}-10H^{2}}{\left(H^{2}+R^{2}% \right)}+6R\hat{x}\right)S,= divide start_ARG italic_ρ end_ARG start_ARG 24 end_ARG ( italic_H divide start_ARG 4 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 10 italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG ( italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) end_ARG + 6 italic_R over^ start_ARG italic_x end_ARG ) italic_S , (122)

where

x^:=12(arccos(RH2+R2)HRH2+R2)assign^𝑥12𝑅superscript𝐻2superscript𝑅2𝐻𝑅superscript𝐻2superscript𝑅2\hat{x}:=\frac{1}{2}\left(\arccos\!\left(-\frac{R}{\sqrt{H^{2}+R^{2}}}\right)-% \frac{HR}{H^{2}+R^{2}}\right)over^ start_ARG italic_x end_ARG := divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( roman_arccos ( - divide start_ARG italic_R end_ARG start_ARG square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) - divide start_ARG italic_H italic_R end_ARG start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) (123)

and

(n2)11subscriptsubscript𝑛211\displaystyle(\overleftrightarrow{n}_{2})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =(n2)22absentsubscriptsubscript𝑛222\displaystyle=(\overleftrightarrow{n}_{2})_{22}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT (124)
=ρ6((2πR2+πRH22R2H2+R2)S\displaystyle=\frac{\rho}{6}\left(\left(-2\pi R^{2}+\pi R\frac{H^{2}-2R^{2}}{% \sqrt{H^{2}+R^{2}}}\right)S\right.= divide start_ARG italic_ρ end_ARG start_ARG 6 end_ARG ( ( - 2 italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_π italic_R divide start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) italic_S (125)
+2πR2+2πRH2+R2),\displaystyle\quad\quad\quad\ +2\pi R^{2}+2\pi R\sqrt{H^{2}+R^{2}}\bigg{)}\,,+ 2 italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_π italic_R square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (126)
(n2)33subscriptsubscript𝑛233\displaystyle(\overleftrightarrow{n}_{2})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ3((2πR2πRH22R2H2+R2)S\displaystyle=\frac{\rho}{3}\left(\left(2\pi R^{2}-\pi R\frac{H^{2}-2R^{2}}{% \sqrt{H^{2}+R^{2}}}\right)S\right.= divide start_ARG italic_ρ end_ARG start_ARG 3 end_ARG ( ( 2 italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_π italic_R divide start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 2 italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG end_ARG ) italic_S (127)
+πR2+πRH2+R2),\displaystyle\quad\quad\quad\ \ +\pi R^{2}+\pi R\sqrt{H^{2}+R^{2}}\bigg{)}\,,+ italic_π italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_π italic_R square-root start_ARG italic_H start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_R start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (128)

while all nondiagonal elements vanish.

B.2 Hard cylinders

Here, we give the full set of weighted densities for hard cylinders Wittmann et al. (2014) with height H𝐻Hitalic_H and diameter D𝐷Ditalic_D, setting again P=0𝑃0P=0italic_P = 0. We start by giving the scalar weighted densities

n0subscript𝑛0\displaystyle n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =ρ,absent𝜌\displaystyle=\rho\,,= italic_ρ , (129)
n1subscript𝑛1\displaystyle n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρ8(2H+πD),absent𝜌82𝐻𝜋𝐷\displaystyle=\frac{\rho}{8}\left(2H+\pi D\right),= divide start_ARG italic_ρ end_ARG start_ARG 8 end_ARG ( 2 italic_H + italic_π italic_D ) , (130)
n2subscript𝑛2\displaystyle n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρ2(2πHD+πD2),absent𝜌22𝜋𝐻𝐷𝜋superscript𝐷2\displaystyle=\frac{\rho}{2}\left(2\pi HD+\pi D^{2}\right),= divide start_ARG italic_ρ end_ARG start_ARG 2 end_ARG ( 2 italic_π italic_H italic_D + italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (131)
n3subscript𝑛3\displaystyle n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =ρ4πHD2,absent𝜌4𝜋𝐻superscript𝐷2\displaystyle=\frac{\rho}{4}\,\pi HD^{2}\,,= divide start_ARG italic_ρ end_ARG start_ARG 4 end_ARG italic_π italic_H italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (132)

while the vectors vanish for this apolar shape. The diagonal components of the (n1)subscript𝑛1(\overleftrightarrow{n}_{1})( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) tensor are

(n1)11subscriptsubscript𝑛111\displaystyle(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =(n1)22=12(n1)33absentsubscriptsubscript𝑛12212subscriptsubscript𝑛133\displaystyle=(\overleftrightarrow{n}_{1})_{22}=-\frac{1}{2}(% \overleftrightarrow{n}_{1})_{33}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT = - divide start_ARG 1 end_ARG start_ARG 2 end_ARG ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT (133)
=ρ32(4HπD)S.absent𝜌324𝐻𝜋𝐷𝑆\displaystyle=\frac{\rho}{32}\left(4H-\pi D\right)S.= divide start_ARG italic_ρ end_ARG start_ARG 32 end_ARG ( 4 italic_H - italic_π italic_D ) italic_S . (134)

The diagonal components of the (n2)subscript𝑛2(\overleftrightarrow{n}_{2})( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) tensor are

(n2)11subscriptsubscript𝑛211\displaystyle(\overleftrightarrow{n}_{2})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =(n2)22absentsubscriptsubscript𝑛222\displaystyle=(\overleftrightarrow{n}_{2})_{22}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT (135)
=ρ6(πHD(2+S)+πD2(1S)),absent𝜌6𝜋𝐻𝐷2𝑆𝜋superscript𝐷21𝑆\displaystyle=\frac{\rho}{6}\left(\pi HD(2+S)+\pi D^{2}(1-S)\right),= divide start_ARG italic_ρ end_ARG start_ARG 6 end_ARG ( italic_π italic_H italic_D ( 2 + italic_S ) + italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 1 - italic_S ) ) , (136)
(n2)33subscriptsubscript𝑛233\displaystyle(\overleftrightarrow{n}_{2})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ6(2πHD(1S)+πD2(2S+1)).absent𝜌62𝜋𝐻𝐷1𝑆𝜋superscript𝐷22𝑆1\displaystyle=\frac{\rho}{6}\left(2\pi HD(1-S)+\pi D^{2}(2S+1)\right).= divide start_ARG italic_ρ end_ARG start_ARG 6 end_ARG ( 2 italic_π italic_H italic_D ( 1 - italic_S ) + italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( 2 italic_S + 1 ) ) . (137)

Again, all nondiagonal elements vanish. The total contribution of the tensors to ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in Eq. (35) is

Tr[n1n2]=ρ2πD(HD)(4HπD)32S2,Trdelimited-[]subscript𝑛1subscript𝑛2superscript𝜌2𝜋𝐷𝐻𝐷4𝐻𝜋𝐷32superscript𝑆2\displaystyle\mbox{Tr}\!\left[\overleftrightarrow{n}_{1}\overleftrightarrow{n}% _{2}\right]=\rho^{2}\pi D\frac{(H-D)(4H-\pi D)}{32}S^{2}\,,Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] = italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π italic_D divide start_ARG ( italic_H - italic_D ) ( 4 italic_H - italic_π italic_D ) end_ARG start_ARG 32 end_ARG italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT , (138)

which vanishes for H=D𝐻𝐷H=Ditalic_H = italic_D (and 4H=πD4𝐻𝜋𝐷4H=\pi D4 italic_H = italic_π italic_D). It can also be shown that the expression in Eq. (36) vanishes for H=D𝐻𝐷H=Ditalic_H = italic_D, such that the dependence on S𝑆Sitalic_S drops out of the functional for such a shape.

B.3 Hard spherotriangles

The scalar weighted densities of hard isosceles spherotriangles with diameter D𝐷Ditalic_D, base length A𝐴Aitalic_A and two sides of length B𝐵Bitalic_B read

n0subscript𝑛0\displaystyle n_{0}italic_n start_POSTSUBSCRIPT 0 end_POSTSUBSCRIPT =ρ,absent𝜌\displaystyle=\rho,= italic_ρ , (139)
n1subscript𝑛1\displaystyle n_{1}italic_n start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT =ρ8(A+2B+4D),absent𝜌8𝐴2𝐵4𝐷\displaystyle=\frac{\rho}{8}\left(A+2B+4D\right),= divide start_ARG italic_ρ end_ARG start_ARG 8 end_ARG ( italic_A + 2 italic_B + 4 italic_D ) , (140)
n2subscript𝑛2\displaystyle n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT =ρ2(πD(A+2B)+2πD2+A4B2A2),absent𝜌2𝜋𝐷𝐴2𝐵2𝜋superscript𝐷2𝐴4superscript𝐵2superscript𝐴2\displaystyle=\frac{\rho}{2}\left(\pi D(A+2B)+2\pi D^{2}+A\sqrt{4B^{2}-A^{2}}% \right),= divide start_ARG italic_ρ end_ARG start_ARG 2 end_ARG ( italic_π italic_D ( italic_A + 2 italic_B ) + 2 italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_A square-root start_ARG 4 italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (141)
n3subscript𝑛3\displaystyle n_{3}italic_n start_POSTSUBSCRIPT 3 end_POSTSUBSCRIPT =ρ24(3πD2(A+2B)+4πD3+12DA4B2A2),absent𝜌243𝜋superscript𝐷2𝐴2𝐵4𝜋superscript𝐷312𝐷𝐴4superscript𝐵2superscript𝐴2\displaystyle=\frac{\rho}{24}\left(3\pi D^{2}(A+2B)+4\pi D^{3}+12DA\sqrt{4B^{2% }-A^{2}}\right),= divide start_ARG italic_ρ end_ARG start_ARG 24 end_ARG ( 3 italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ( italic_A + 2 italic_B ) + 4 italic_π italic_D start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT + 12 italic_D italic_A square-root start_ARG 4 italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ) , (142)

where the contributions which are quadratic, linear and constant in D𝐷Ditalic_D originate from the spherical caps, parts of cylinder mantles and triangles, respectively. In general, the vectors do yield nonzero contributions, which vanish for the apolar phases of interest (see the discussion in appendix B.1) and are omitted here due to their lengthy general form. Only in the special case A=B𝐴𝐵A=Bitalic_A = italic_B of equilateral spherotriangles, we have n2=0subscript𝑛20\overrightarrow{n}_{2}=\overrightarrow{0}over→ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT = over→ start_ARG 0 end_ARG such that the vectors are truly irrelevant.

The tensor weighted densities explicitly depend on the shape ratio x𝑥xitalic_x of the spherotriangles through γ𝛾\gammaitalic_γ, which is the half opening angle of the triangle, see Fig. 1. The nonvanishing diagonal components of n1subscript𝑛1\overleftrightarrow{n}_{1}over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT have only contributions from the cylindrical parts and read

(n1)11subscriptsubscript𝑛111\displaystyle(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =(n1)22(S,U,P,F)absentsubscriptsubscript𝑛122𝑆𝑈𝑃𝐹\displaystyle=(\overleftrightarrow{n}_{1})_{22}(S,U,-P,-F)= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_S , italic_U , - italic_P , - italic_F ) (144)
=ρ32(A(S+3U+3P3F)\displaystyle=\frac{\rho}{32}\left(A(-S+\sqrt{3}U+\sqrt{3}P-3F)\right.= divide start_ARG italic_ρ end_ARG start_ARG 32 end_ARG ( italic_A ( - italic_S + square-root start_ARG 3 end_ARG italic_U + square-root start_ARG 3 end_ARG italic_P - 3 italic_F ) (145)
+2B((S+3U+3P3F)sin(γ)2\displaystyle\quad\ \ \ \ \ +2B\left((-S+\sqrt{3}U+\sqrt{3}P-3F)\sin(\gamma)^{% 2}\right.+ 2 italic_B ( ( - italic_S + square-root start_ARG 3 end_ARG italic_U + square-root start_ARG 3 end_ARG italic_P - 3 italic_F ) roman_sin ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (146)
+2(S3P)cos(γ)2))\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left.\left.+2(S-\sqrt{3}P)\cos% (\gamma)^{2}\right)\right)+ 2 ( italic_S - square-root start_ARG 3 end_ARG italic_P ) roman_cos ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) (147)

and

(n1)33subscriptsubscript𝑛133\displaystyle(\overleftrightarrow{n}_{1})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ16(A(S3U)\displaystyle=\frac{\rho}{16}\left(A(S-\sqrt{3}U)\right.= divide start_ARG italic_ρ end_ARG start_ARG 16 end_ARG ( italic_A ( italic_S - square-root start_ARG 3 end_ARG italic_U ) (148)
+2B((S3U)sin(γ)2+2Scos(γ)2)).\displaystyle\quad\ \ \ \ \ \left.+2B\left((S-\sqrt{3}U)\sin\left(\gamma\right% )^{2}+2S\cos\left(\gamma\right)^{2}\right)\right).+ 2 italic_B ( ( italic_S - square-root start_ARG 3 end_ARG italic_U ) roman_sin ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_S roman_cos ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) . (149)

The relation between (n1)11subscriptsubscript𝑛111(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT and (n1)22subscriptsubscript𝑛122(\overleftrightarrow{n}_{1})_{22}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT in Eq. (147) stems from the fact that these tensor components are related by a polar rotation of the body in the lab frame, achieved by a redefinition ϕϕ±π/2italic-ϕplus-or-minusitalic-ϕ𝜋2\phi\rightarrow\phi\pm\pi/2italic_ϕ → italic_ϕ ± italic_π / 2 of the polar angle ϕitalic-ϕ\phiitalic_ϕ. Just like n2subscript𝑛2n_{2}italic_n start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT, the tensor n2subscript𝑛2\overleftrightarrow{n}_{2}over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT has contributions from all body parts. Its relevant diagonal components read

(n2)11=(n2)22(S,U,P,F)subscriptsubscript𝑛211subscriptsubscript𝑛222𝑆𝑈𝑃𝐹\displaystyle(\overleftrightarrow{n}_{2})_{11}=(\overleftrightarrow{n}_{2})_{2% 2}(S,U,-P,-F)( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT = ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT ( italic_S , italic_U , - italic_P , - italic_F ) (151)
=ρ24(πAD(S+3U+3P3F+2)\displaystyle=\frac{\rho}{24}\left(\pi AD(-S+\sqrt{3}U+\sqrt{3}P-3F+2)\right.= divide start_ARG italic_ρ end_ARG start_ARG 24 end_ARG ( italic_π italic_A italic_D ( - italic_S + square-root start_ARG 3 end_ARG italic_U + square-root start_ARG 3 end_ARG italic_P - 3 italic_F + 2 ) (152)
+2πBD((S+3U+3P3F)sin(γ)2\displaystyle\quad\quad\quad+2\pi BD\left((-S+\sqrt{3}U+\sqrt{3}P-3F)\sin(% \gamma)^{2}\right.+ 2 italic_π italic_B italic_D ( ( - italic_S + square-root start_ARG 3 end_ARG italic_U + square-root start_ARG 3 end_ARG italic_P - 3 italic_F ) roman_sin ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (153)
+2(S3P)cos(γ)2+4)\displaystyle\quad\quad\quad\quad\quad\quad\quad\ +\left.2\,(S-\sqrt{3}P)\cos(% \gamma)^{2}+4\right)+ 2 ( italic_S - square-root start_ARG 3 end_ARG italic_P ) roman_cos ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 4 ) (154)
+8πD2+2A4B2A2(S+3P+1))\displaystyle\quad\quad\quad+\left.8\pi D^{2}+2A\sqrt{4B^{2}-A^{2}}(-S+\sqrt{3% }P+1)\right)\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ + 8 italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 italic_A square-root start_ARG 4 italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( - italic_S + square-root start_ARG 3 end_ARG italic_P + 1 ) ) (155)

and

(n2)33subscriptsubscript𝑛233\displaystyle(\overleftrightarrow{n}_{2})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ12(πAD(2+S3U)\displaystyle=\frac{\rho}{12}\left(\pi AD(2+S-\sqrt{3}U)\right.= divide start_ARG italic_ρ end_ARG start_ARG 12 end_ARG ( italic_π italic_A italic_D ( 2 + italic_S - square-root start_ARG 3 end_ARG italic_U ) (156)
+2πBD((S3U)sin(γ)2\displaystyle\ \ \ \ \ \ \ \ \ +2\pi BD\left((S-\sqrt{3}U)\sin(\gamma)^{2}\right.+ 2 italic_π italic_B italic_D ( ( italic_S - square-root start_ARG 3 end_ARG italic_U ) roman_sin ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (157)
2Scos(γ)2+2)\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left.-2S\cos(\gamma)% ^{2}+2\right)- 2 italic_S roman_cos ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 2 ) (158)
+4πD2+A4B2A2(2S+1)).\displaystyle\ \ \ \ \ \ \ \ \ \left.+4\pi D^{2}+A\sqrt{4B^{2}-A^{2}}(2S+1)% \right).\ \ \ \ \ \ \ \ \ \ \ \ + 4 italic_π italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_A square-root start_ARG 4 italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG ( 2 italic_S + 1 ) ) . (159)

To determine the tensorial weighted densities for other orientations of the primary director, we can use the substitutions from Eq. 8 or Eq. 10, as detailed in Sec. II.1.4. For example, aligning the z𝑧zitalic_z-axis of the body frame with the base line of the triangle, compare the right picture in Fig. 2b, we get the first tensor component

(n1)11subscriptsubscript𝑛111\displaystyle(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =ρ16(A(S3P)\displaystyle=\frac{\rho}{16}\left(A\,(S-\sqrt{3}P)\right.= divide start_ARG italic_ρ end_ARG start_ARG 16 end_ARG ( italic_A ( italic_S - square-root start_ARG 3 end_ARG italic_P ) (160)
+B(2(S3P)sin(γ)2\displaystyle\quad\ \ \ \ \ +B\left(2(S-\sqrt{3}P)\sin(\gamma)^{2}\right.+ italic_B ( 2 ( italic_S - square-root start_ARG 3 end_ARG italic_P ) roman_sin ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT (161)
+(S+3U+3P3F)cos(γ)2))\displaystyle\ \ \ \ \ \ \ \ \ \ \ \ \ \ \ \left.\left.+(-S+\sqrt{3}U+\sqrt{3}% P-3F)\cos(\gamma)^{2}\right)\right)+ ( - italic_S + square-root start_ARG 3 end_ARG italic_U + square-root start_ARG 3 end_ARG italic_P - 3 italic_F ) roman_cos ( italic_γ ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) ) (162)

upon applying Eq. 10 to Eq. 147.

Refer to caption
Refer to caption
Refer to caption
Figure 6: Phase diagrams of hard spherotriangles upon imposing the three different orientations of the uniaxial director illustrated in Fig. 2, as indicated by the drawn bodies. In each plot, the red lines indicate the isotropic–uniaxial coexistence densities calculated by free minimization, following Sec. II.3.2, and the blue lines indicate the uniaxial–biaxial transition identified by a perturbation of simple uniaxial order as outlined in Sec. II.3.4. In addition to these results, which are also shown in Fig. 5, we compare the predictions of the perfect uniaxial approximation (dotted purple line) to locate the uniaxial–biaxial transition, as outlined in Sec. II.3.3. The black dots mark the triple points, where two uniaxial phases with different director orientations coexist with either the isotropic or the biaxial phase, compare Fig. 5. The phase diagrams presented here thus also include the regions beyond these points, where the given director orientation is unstable.

B.4 Limits of hard spherocylinders

In the limits x=0𝑥0x=0italic_x = 0 or x=1𝑥1x=1italic_x = 1 of extreme shape ratios, a spherotriangle turns into a spherocylinder. Therefore, by evaluating the weighted densities of hard spherotriangles from appendix B.3 for these special cases, we can directly obtain the full set of weighted densities of hard spherocylinders for two different choices of the body frame, which results in a different dependence on the order parameters S𝑆Sitalic_S, U𝑈Uitalic_U, P𝑃Pitalic_P and F𝐹Fitalic_F, as we elaborate below. Recall that we have used the convention, illustrated in the left picture in Fig. 2a, that the height of the triangle points in the z𝑧zitalic_z-direction of the body frame.

In the first case, x=0𝑥0x=0italic_x = 0, we recover the standard convention used for uniaxial bodies. By setting A=0𝐴0A=0italic_A = 0 and L=2B𝐿2𝐵L=2Bitalic_L = 2 italic_B in Eqs. (147)-(159), we obtain the tensor components

(n1)11subscriptsubscript𝑛111\displaystyle(\overleftrightarrow{n}_{1})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =ρDL(S83P8),absent𝜌𝐷𝐿𝑆83𝑃8\displaystyle=\rho DL\left(\frac{S}{8}-\frac{\sqrt{3}P}{8}\right),= italic_ρ italic_D italic_L ( divide start_ARG italic_S end_ARG start_ARG 8 end_ARG - divide start_ARG square-root start_ARG 3 end_ARG italic_P end_ARG start_ARG 8 end_ARG ) , (164)
(n1)22subscriptsubscript𝑛122\displaystyle(\overleftrightarrow{n}_{1})_{22}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT =ρDL(S8+3P8),absent𝜌𝐷𝐿𝑆83𝑃8\displaystyle=\rho DL\left(\frac{S}{8}+\frac{\sqrt{3}P}{8}\right),= italic_ρ italic_D italic_L ( divide start_ARG italic_S end_ARG start_ARG 8 end_ARG + divide start_ARG square-root start_ARG 3 end_ARG italic_P end_ARG start_ARG 8 end_ARG ) , (165)
(n1)33subscriptsubscript𝑛133\displaystyle(\overleftrightarrow{n}_{1})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρDL(S4),absent𝜌𝐷𝐿𝑆4\displaystyle=\rho DL\left(-\frac{S}{4}\right),= italic_ρ italic_D italic_L ( - divide start_ARG italic_S end_ARG start_ARG 4 end_ARG ) , (166)
(n2)11subscriptsubscript𝑛211\displaystyle(\overleftrightarrow{n}_{2})_{11}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 11 end_POSTSUBSCRIPT =ρ(D2π3+DLπ6(2+S3P)),absent𝜌superscript𝐷2𝜋3𝐷𝐿𝜋62𝑆3𝑃\displaystyle=\rho\left(\frac{D^{2}\pi}{3}+\frac{DL\pi}{6}\left(2+S-\sqrt{3}P% \right)\right),= italic_ρ ( divide start_ARG italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π end_ARG start_ARG 3 end_ARG + divide start_ARG italic_D italic_L italic_π end_ARG start_ARG 6 end_ARG ( 2 + italic_S - square-root start_ARG 3 end_ARG italic_P ) ) , (167)
(n2)22subscriptsubscript𝑛222\displaystyle(\overleftrightarrow{n}_{2})_{22}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 22 end_POSTSUBSCRIPT =ρ(D2π3+DLπ6(2+S+3P)),absent𝜌superscript𝐷2𝜋3𝐷𝐿𝜋62𝑆3𝑃\displaystyle=\rho\left(\frac{D^{2}\pi}{3}+\frac{DL\pi}{6}\left(2+S+\sqrt{3}P% \right)\right),= italic_ρ ( divide start_ARG italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π end_ARG start_ARG 3 end_ARG + divide start_ARG italic_D italic_L italic_π end_ARG start_ARG 6 end_ARG ( 2 + italic_S + square-root start_ARG 3 end_ARG italic_P ) ) , (168)
(n2)33subscriptsubscript𝑛233\displaystyle(\overleftrightarrow{n}_{2})_{33}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 33 end_POSTSUBSCRIPT =ρ(D2π3+DLπ6(1S)),absent𝜌superscript𝐷2𝜋3𝐷𝐿𝜋61𝑆\displaystyle=\rho\left(\frac{D^{2}\pi}{3}+\frac{DL\pi}{6}\left(1-S\right)% \right)\,,= italic_ρ ( divide start_ARG italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π end_ARG start_ARG 3 end_ARG + divide start_ARG italic_D italic_L italic_π end_ARG start_ARG 6 end_ARG ( 1 - italic_S ) ) , (169)

while the total contribution of the tensors to ϕ2subscriptitalic-ϕ2\phi_{2}italic_ϕ start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT in Eq. (35) is

Tr[n1n2]=ρ2DL2π8(S2+3P2).Trdelimited-[]subscript𝑛1subscript𝑛2superscript𝜌2𝐷superscript𝐿2𝜋8superscript𝑆23superscript𝑃2\displaystyle\mbox{Tr}\!\left[\overleftrightarrow{n}_{1}\overleftrightarrow{n}% _{2}\right]=\rho^{2}\frac{DL^{2}\pi}{8}\left(S^{2}+3P^{2}\right).Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] = italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_D italic_L start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_π end_ARG start_ARG 8 end_ARG ( italic_S start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 3 italic_P start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) . (170)

As we can see, within our general treatment, there is a dependence on P𝑃Pitalic_P in addition to the standard nematic order parameter S𝑆Sitalic_S. Setting P=0𝑃0P=0italic_P = 0, these weighted densities reduce to the common expressions previously reported in the literature (we also recover the orientation-independent scalar weighted densities which are not repeated here) Hansen-Goos and Mecke (2010); Wittmann et al. (2014, 2016); Wittmann et al. (2015b). Hence, neglecting P𝑃Pitalic_P has already been demonstrated to be well justified when one is only interested in the bulk phase behavior, which we have explicitly verified in Sec. III.1. The form of the weighted densities in Eq. (169) is, however, helpful to understand physical scenarios where the uniaxial symmetry of the phase is broken, for example due to external fields. Most importantly, we anticipate the explicit importance of P𝑃Pitalic_P for the bulk phase behavior of (uniaxial) rod-disk mixtures.

While we have seen that P𝑃Pitalic_P contributes in general to the weighted densities of uniaxial bodies, U𝑈Uitalic_U and F𝐹Fitalic_F do not appear in Eq. (170) because a spherocylinder is not a biaxial particle. However, in the second case, x=1𝑥1x=1italic_x = 1, the symmetry axis of the resulting spherocylinder does not coincide with the orientation in the body frame, which means that we formally treat the body as if it was biaxial. Indeed, if we set L𝐿Litalic_L=A𝐴Aitalic_A=2B2𝐵2B2 italic_B in Eqs. (147)-(159), we arrive at the following contribution of the tensorial weighted densities

Tr[n1n2]=ρ2D2Lπ32((3US)2+3(3FP)2),Trdelimited-[]subscript𝑛1subscript𝑛2superscript𝜌2superscript𝐷2𝐿𝜋32superscript3𝑈𝑆23superscript3𝐹𝑃2\displaystyle\mbox{Tr}\!\left[\overleftrightarrow{n}_{1}\overleftrightarrow{n}% _{2}\right]=\rho^{2}\frac{D^{2}L\pi}{32}\left((\sqrt{3}U-S)^{2}+3(\sqrt{3}F-P)% ^{2}\right),Tr [ over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ] = italic_ρ start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT divide start_ARG italic_D start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT italic_L italic_π end_ARG start_ARG 32 end_ARG ( ( square-root start_ARG 3 end_ARG italic_U - italic_S ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 3 ( square-root start_ARG 3 end_ARG italic_F - italic_P ) start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT ) , (171)

which depends on all four order parameters (the same result can be obtained from Eq. 170 by performing the substitutions in Eq. 10). These weighted densities still contain the same physics, but one has to give up the interpretations of the order parameters outlined in Sec. II.1.4. In our particular example, this means that S𝑆Sitalic_S does not appropriately describe the uniaxial order of hard spherocylinders when setting U=P=F=0𝑈𝑃𝐹0U=P=F=0italic_U = italic_P = italic_F = 0 in Eq. 171.

To show the difference between the two descriptions of hard spherocylinders, we show in Fig. 6 the predicted isotropic–uniaxial transition lines (only allowing for nonzero values of S𝑆Sitalic_S) of hard spherotriangles for the full range of shape ratios x𝑥xitalic_x of each chosen director orientation. This compilation includes results which are unstable with respect to a different director orientation and thus not contained in Fig. 5. The left plot in Fig. 6 shows that the stable spherocylinder limit of isosceles spherotriangles aligned along the direction of their height is obtained for x=0𝑥0x=0italic_x = 0 with Eq. (170), while the transition for x=1𝑥1x=1italic_x = 1 with Eq. (171) is located at much higher densities and in the unstable regime. The same unphysical result is found in both limits x=0𝑥0x=0italic_x = 0 and x=1𝑥1x=1italic_x = 1 for isosceles spherotriangles aligned along the normal direction to their face (central plot in Fig. 6) and in the limit x=0𝑥0x=0italic_x = 0 for isosceles spherotriangles aligned along their triangle base (right plot in Fig. 6). In the latter case, taking x=1𝑥1x=1italic_x = 1 again yields the stable spherocylinder limit.

B.5 Comments on general hard spherotriangles

For previous calculations, due to the symmetry of isosceles spherotriangles, all nondiagonal terms of the tensor weighted densities were automatically zero. However, for arbitrary spherotriangles with side lengths A𝐴Aitalic_A, B𝐵Bitalic_B and CB𝐶𝐵C\neq Bitalic_C ≠ italic_B, there may be both such nonzero cross terms and additional order parameters resulting from orientational averages of mixed terms ^ij^kldelimited-⟨⟩subscript^𝑖𝑗subscript^𝑘𝑙\langle\hat{\mathcal{R}}_{ij}\hat{\mathcal{R}}_{kl}\rangle⟨ over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_i italic_j end_POSTSUBSCRIPT over^ start_ARG caligraphic_R end_ARG start_POSTSUBSCRIPT italic_k italic_l end_POSTSUBSCRIPT ⟩ with ik𝑖𝑘i\neq kitalic_i ≠ italic_k and/or jl𝑗𝑙j\neq litalic_j ≠ italic_l of the rotation matrix, which are not included in Eq. (55), since these correspond to cross terms of the Saupe matrix.

To give an example of such cross terms, we consider the limit of perfect uniaxial order, as introduced in Sec. II.3.3, for a general spherotriangle with the nematic director parallel to the face normal of the triangle, compare the central picture in Fig. 2a. Then, the nondiagonal elements of the tensorial weighted densities can be given in the compact form

(n1)12subscriptsubscript𝑛112\displaystyle(\overleftrightarrow{n}_{1})_{12}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT =(n1)21absentsubscriptsubscript𝑛121\displaystyle=(\overleftrightarrow{n}_{1})_{21}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 1 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT (172)
=ρ48(Bsin(μ)cos(μ)(4S2d1)\displaystyle=\frac{\rho}{48}\bigg{(}B\sin\left(\mu\right)\cos\left(\mu\right)% (4S_{2d}-1)= divide start_ARG italic_ρ end_ARG start_ARG 48 end_ARG ( italic_B roman_sin ( italic_μ ) roman_cos ( italic_μ ) ( 4 italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT - 1 ) (173)
+Csin(η)cos(η)(4S2d1))\displaystyle\quad\ \ \ \ \ \ +C\sin(\eta)\cos(\eta)(4S_{2d}-1)\bigg{)}\ \ \ % \ \ \ + italic_C roman_sin ( italic_η ) roman_cos ( italic_η ) ( 4 italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT - 1 ) ) (174)

and

(n2)12subscriptsubscript𝑛212\displaystyle(\overleftrightarrow{n}_{2})_{12}( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 12 end_POSTSUBSCRIPT =(n2)21absentsubscriptsubscript𝑛221\displaystyle=(\overleftrightarrow{n}_{2})_{21}= ( over↔ start_ARG italic_n end_ARG start_POSTSUBSCRIPT 2 end_POSTSUBSCRIPT ) start_POSTSUBSCRIPT 21 end_POSTSUBSCRIPT (175)
=ρ6(πDBsin(μ)cos(μ)(14S2d)\displaystyle=\frac{\rho}{6}\bigg{(}\pi DB\sin(\mu)\cos(\mu)(1-4S_{2d})= divide start_ARG italic_ρ end_ARG start_ARG 6 end_ARG ( italic_π italic_D italic_B roman_sin ( italic_μ ) roman_cos ( italic_μ ) ( 1 - 4 italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT ) (176)
+πDCsin(η)cos(η)(14S2d)),\displaystyle\quad\ \ \ \ \ \,+\pi DC\sin(\eta)\cos(\eta)(1-4S_{2d})\bigg{)},% \ \ \ \ \ \ \ + italic_π italic_D italic_C roman_sin ( italic_η ) roman_cos ( italic_η ) ( 1 - 4 italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT ) ) , (177)

where S2dsubscript𝑆2𝑑S_{2d}italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT, as defined in Eq. (68), is identified according to Eq. (67) and the angles μ𝜇\muitalic_μ and η𝜂\etaitalic_η are defined as

μ𝜇\displaystyle\muitalic_μ =arccos(B2+A2C22BA),absentsuperscript𝐵2superscript𝐴2superscript𝐶22𝐵𝐴\displaystyle=\arccos\bigg{(}\frac{B^{2}+A^{2}-C^{2}}{2BA}\bigg{)},= roman_arccos ( divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_B italic_A end_ARG ) , (178)
η𝜂\displaystyle\etaitalic_η =arccos(B2+C2A22BC)+μ.absentsuperscript𝐵2superscript𝐶2superscript𝐴22𝐵𝐶𝜇\displaystyle=\arccos\bigg{(}\frac{B^{2}+C^{2}-A^{2}}{2BC}\bigg{)}+\mu.= roman_arccos ( divide start_ARG italic_B start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + italic_C start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - italic_A start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT end_ARG start_ARG 2 italic_B italic_C end_ARG ) + italic_μ . (179)

Note that the cross terms presented here vanish for isosceles spherotriangles, i.e., if we set B=C𝐵𝐶B=Citalic_B = italic_C.

Appendix C Details on the uniaxial–biaxial transition

In the main text, we outline two different methods to determine the location of the uniaxial–biaxial transition. First, in the perfect uniaxial approximation, we assume S=1𝑆1S=1italic_S = 1 and P=U=0𝑃𝑈0P=U=0italic_P = italic_U = 0, which leaves us with a single order parameter FS2dsimilar-to-or-equals𝐹subscript𝑆2𝑑F\simeq S_{2d}italic_F ≃ italic_S start_POSTSUBSCRIPT 2 italic_d end_POSTSUBSCRIPT and we can analytically minimize the basic equations of DFT as for the functional in two spatial dimensions. We explain this in greater detail in Sec. II.3.3. We also use a perturbative approach, where we locate the transition as the state point where simple uniaxial order (solely described by a nonzero value of S𝑆Sitalic_S) becomes unstable for a small perturbation by a finite value of the full biaxiality order parameter F𝐹Fitalic_F, while setting P=U=0𝑃𝑈0P=U=0italic_P = italic_U = 0. More details are given in Sec. II.3.3 and the results of this method are included in our phase diagrams of hard isosceles spherotriangles in Fig. 5, where we argue that it is quite reliable, since the uniaxial–uniaxial–biaxial triple points are consistently approached from both sides, where uniaxial order is assumed with respect to different main axes. We thus expect that the assumption that one particle axis is perfectly aligned yields only reliable results at unrealistically high packing fractions.

To compare our two approaches, we present in Fig. 6 three phase diagrams of hard isosceles spherotriangles, each obtained upon imposing a different type of uniaxial order, as illustrated in Fig. 2a, but here over the full range 0x10𝑥10\leq x\leq 10 ≤ italic_x ≤ 1 of shape ratios. Compared to the perturbative approach, the perfect uniaxial approximation systematically overestimates the packing fraction at the uniaxial–biaxial transition for the prolate uniaxial phases Nphph{}^{\text{ph}}start_FLOATSUPERSCRIPT ph end_FLOATSUPERSCRIPT and Npbpb{}^{\text{pb}}start_FLOATSUPERSCRIPT pb end_FLOATSUPERSCRIPT, while it is underestimated in the oblate case. Therefore, when only considering the most stable ordered state (where the transition is predicted at the lowest density among the results for the three different uniaxial orientations), the perfect uniaxial approximation does not result in a consistent location of the triple points, in contrast to the perturbation result shown in Fig. 5.

However, one noteworthy advantage of the perfect uniaxial approximation, as opposed to the perturbation approach, is its ability to produce closed analytic expressions for the transition densities (although these are, in general, too long to be stated here). Only in the limits x0𝑥0x\rightarrow 0italic_x → 0 and x1𝑥1x\rightarrow 1italic_x → 1 of hard spherocylinders, the results can be presented in a compact form (recall the discussion in appendix B.4). In the limit where the spherocylinder is perfectly aligned along its long axis, no biaxial order is possible as F𝐹Fitalic_F drops out of the functional, compare Eq. (170). Hence, the results of our two methods in Fig. 6 become comparable when this limit is approached. In the (unphysical) opposite limit of the rod axis being perpendicular to its perfect orientation, F𝐹Fitalic_F remains in the functional, compare Eq. (171), and the uniaxial–biaxial transition can be formally located at

ηIU=31536l2+225l4+2304l345l2192l12818l2192l128,subscript𝜂IU31536superscript𝑙2225superscript𝑙42304superscript𝑙345superscript𝑙2192𝑙12818superscript𝑙2192𝑙128\eta_{\text{IU}}=\frac{3\sqrt{1536l^{2}+225l^{4}+2304l^{3}}-45l^{2}-192l-128}{% 18l^{2}-192l-128}\,,italic_η start_POSTSUBSCRIPT IU end_POSTSUBSCRIPT = divide start_ARG 3 square-root start_ARG 1536 italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT + 225 italic_l start_POSTSUPERSCRIPT 4 end_POSTSUPERSCRIPT + 2304 italic_l start_POSTSUPERSCRIPT 3 end_POSTSUPERSCRIPT end_ARG - 45 italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 192 italic_l - 128 end_ARG start_ARG 18 italic_l start_POSTSUPERSCRIPT 2 end_POSTSUPERSCRIPT - 192 italic_l - 128 end_ARG , (180)

which is given as an explicit function of the aspect ratio l𝑙litalic_l. For l=5𝑙5l=5italic_l = 5, we get ηIU=0.255subscript𝜂IU0.255\eta_{\text{IU}}=0.255italic_η start_POSTSUBSCRIPT IU end_POSTSUBSCRIPT = 0.255, which lies below the (unstable) isotropic–uniaxial transition and is thus not shown in Fig. 6.

References

  • de Gennes and Prost (1993) P.-G. de Gennes and J. Prost, The physics of liquid crystals, 83 (Oxford University Press, Oxford, United Kingdom, 1993).
  • Sebastián et al. (2022) N. Sebastián, M. Čopič, and A. Mertelj, Phys. Rev. E 106, 021001 (2022).
  • Bruce (2004) D. W. Bruce, Chem. Rec. 4, 10 (2004).
  • Chen et al. (2020) X. Chen, E. Korblova, D. Dong, X. Wei, R. Shao, L. Radzihovsky, M. A. Glaser, J. E. Maclennan, D. Bedrov, D. M. Walba, et al., Proc. Natl. Acad. Sci. U.S.A. 117, 14021 (2020).
  • Lavrentovich (2020) O. D. Lavrentovich, Proc. Natl. Acad. Sci. U.S.A. 117, 14629 (2020).
  • Freiser (1970) M. J. Freiser, Phys. Rev. Lett. 24, 1041 (1970).
  • Madsen et al. (2004) L. A. Madsen, T. J. Dingemans, M. Nakata, and E. T. Samulski, Phys. Rev. Lett. 92, 145505 (2004).
  • Luckhurst and Sluckin (2015) G. R. Luckhurst and T. J. Sluckin, eds., Biaxial nematic liquid crystals: theory, simulation and experiment (Wiley, Chichester, UK, 2015).
  • Vroege and Lekkerkerker (1992) G. J. Vroege and H. N. W. Lekkerkerker, Rep. Prog. Phys. 55, 1241 (1992).
  • Smalyukh (2018) I. I. Smalyukh, Annu. Rev. Condens. Matter Phys. 9, 207 (2018).
  • Sacanna et al. (2013) S. Sacanna, D. J. Pine, and G.-R. Yi, Soft Matter 9, 8096 (2013).
  • Hueckel et al. (2021) T. Hueckel, G. M. Hocky, and S. Sacanna, Nat. Rev. Mater. 6, 1053 (2021).
  • Mederos et al. (2014) L. Mederos, E. Velasco, and Y. Martínez-Ratón, J. Phys. Condens. Matter 26, 463101 (2014).
  • McGrother et al. (1996) S. C. McGrother, D. C. Williamson, and G. Jackson, J. Chem. Phys. 104, 6755 (1996).
  • Bolhuis and Frenkel (1997) P. Bolhuis and D. Frenkel, J. Chem. Phys. 106, 666 (1997).
  • Wensink and Vroege (2003) H. H. Wensink and G. J. Vroege, J. Chem. Phys. 119, 6868 (2003).
  • Pfleiderer and Schilling (2007) P. Pfleiderer and T. Schilling, Phys. Rev. E 75, 020402 (2007).
  • Marechal et al. (2011) M. Marechal, A. Cuetos, B. Martínez-Haya, and M. Dijkstra, J. Chem. Phys. 134, 094501 (2011).
  • Odriozola (2012) G. Odriozola, J. Chem. Phys. 136, 134505 (2012).
  • Kubala et al. (2023) P. Kubala, M. Cieśla, and L. Longa, Phys. Rev. E 108, 054701 (2023).
  • Ellison et al. (2006) L. J. Ellison, D. J. Michel, F. Barmes, and D. J. Cleaver, Phys. Rev. Lett. 97, 237801 (2006).
  • Schönhöfer et al. (2017) P. W. A. Schönhöfer, L. J. Ellison, M. Marechal, D. J. Cleaver, and G. E. Schröder-Turk, Interface Focus 7, 20160161 (2017).
  • Monderkamp et al. (2023) P. A. Monderkamp, R. S. Windisch, R. Wittmann, and H. Löwen, J. Chem. Phys. 158, 164505 (2023).
  • Martínez-Ratón and Velasco (2021) Y. Martínez-Ratón and E. Velasco, Phys. Rev. E 104, 054132 (2021).
  • Martínez-Ratón and Velasco (2022) Y. Martínez-Ratón and E. Velasco, Phys. Fluids 34, 037110 (2022).
  • Alben (1973) R. Alben, Phys. Rev. Lett. 30, 778 (1973).
  • Taylor and Herzfeld (1991) M. P. Taylor and J. Herzfeld, Phys. Rev. A 44, 3742 (1991).
  • Vanakaras et al. (2003) A. G. Vanakaras, M. A. Bates, and D. J. Photinos, Phys. Chem. Chem. Phys. 5, 3700 (2003).
  • Martínez-Ratón et al. (2011) Y. Martínez-Ratón, S. Varga, and E. Velasco, Phys. Chem. Chem. Phys. 13, 13247 (2011).
  • Belli et al. (2011) S. Belli, A. Patti, M. Dijkstra, and R. van Roij, Phys. Rev. Lett. 107, 148303 (2011).
  • Belli et al. (2012) S. Belli, M. Dijkstra, and R. van Roij, J. Phys. Condens. Matter 24, 284128 (2012).
  • Peroukidis and Vanakaras (2013) S. D. Peroukidis and A. G. Vanakaras, Soft Matter 9, 7419 (2013).
  • Cuetos et al. (2017) A. Cuetos, M. Dennison, A. Masters, and A. Patti, Soft Matter 13, 4720 (2017).
  • Van den Pol et al. (2009) E. Van den Pol, A. V. Petukhov, D. M. E. Thies-Weesie, D. V. Byelov, and G. J. Vroege, Phys. Rev. Lett. 103, 258301 (2009).
  • Rosenfeld (1989) Y. Rosenfeld, Phys. Rev. Lett. 63, 980 (1989).
  • Roth (2010) R. Roth, J. Phys. Condens. Matter 22, 063102 (2010).
  • Roth et al. (2012) R. Roth, K. Mecke, and M. Oettel, J. Chem. Phys. 136, 081101 (2012).
  • Hansen-Goos and Mecke (2009) H. Hansen-Goos and K. Mecke, Phys. Rev. Lett. 102, 018302 (2009).
  • Hansen-Goos and Mecke (2010) H. Hansen-Goos and K. Mecke, J. Phys. Condens. Matter 22, 364107 (2010).
  • Wittmann et al. (2014) R. Wittmann, M. Marechal, and K. Mecke, J. Chem. Phys. 141, 064103 (2014).
  • Wittmann et al. (2015a) R. Wittmann, M. Marechal, and K. Mecke, EPL 109, 26003 (2015a).
  • Wittmann (2015) R. Wittmann, Density functional theory for liquid Crystals: Refining fundamental measure theory for anisotropic bodies (FAU University Press, 2015).
  • Wittmann et al. (2016) R. Wittmann, M. Marechal, and K. Mecke, J. Phys. Condens. Matter 28, 244003 (2016).
  • Marechal et al. (2017) M. Marechal, S. Dussi, and M. Dijkstra, J. Chem. Phys. 146, 124905 (2017).
  • Marechal and Löwen (2013) M. Marechal and H. Löwen, Phys. Rev. Lett. 110, 137801 (2013).
  • Wittmann and Mecke (2014) R. Wittmann and K. Mecke, J. Chem. Phys. 140, 104703 (2014).
  • Wittmann et al. (2015b) R. Wittmann, M. Marechal, and K. Mecke, Phys. Rev. E 91, 052501 (2015b).
  • Schönhöfer et al. (2018) P. W. A. Schönhöfer, G. E. Schröder-Turk, and M. Marechal, J. Chem. Phys. 148, 124104 (2018).
  • Rosso (2007) R. Rosso, Liq. Cryst. 34, 737 (2007).
  • te Vrugt and Wittkowski (2020a) M. te Vrugt and R. Wittkowski, AIP Adv. 10, 035106 (2020a).
  • Wittmann et al. (2017) R. Wittmann, C. E. Sitta, F. Smallenburg, and H. Löwen, J. Chem. Phys. 147, 134908 (2017).
  • Gray and Gubbins (1984) C. G. Gray and K. E. Gubbins, Theory of Molecular Fluids: Fundamentals, vol. 1 of International Series of Monographs on Chemistry 9 (Oxford University Press, Oxford, 1984), 1st ed.
  • Cremer et al. (2012) P. Cremer, M. Marechal, and H. Löwen, EPL 99, 38005 (2012).
  • te Vrugt and Wittkowski (2020b) M. te Vrugt and R. Wittkowski, Ann. Phys. (Berlin) 532, 2000266 (2020b).
  • Andrienko (2018) D. Andrienko, J. Mol. Liq. 267, 520 (2018).
  • Allender and Lee (1984) D. W. Allender and M. A. Lee, Mol. Cryst. Liq. Cryst. 110, 331 (1984).
  • Turzi (2011) S. S. Turzi, J. Math. Phys. 52, 053517 (2011).
  • Hohenberg and Kohn (1964) P. Hohenberg and W. Kohn, Phys. Rev. 136, B864 (1964).
  • Evans (1979) R. Evans, Adv. Phys. 28, 143 (1979).
  • Tarazona (2000) P. Tarazona, Phys. Rev. Lett. 84, 694 (2000).
  • Hansen-Goos and Roth (2006) H. Hansen-Goos and R. Roth, J. Phys. Condens. Matter 18, 8413 (2006).
  • Tarazona and Rosenfeld (1997) P. Tarazona and Y. Rosenfeld, Phys. Rev. E 55, R4873 (1997).
  • Onsager (1949) L. Onsager, Ann. N. Y. Acad. Sci 51, 627 (1949).
  • Dussi et al. (2018) S. Dussi, N. Tasios, T. Drwenski, R. Van Roij, and M. Dijkstra, Phys. Rev. Lett. 120, 177801 (2018).
  • Stroobants and Lekkerkerker (1984) A. Stroobants and H. N. W. Lekkerkerker, J. Phys. Chem. 88, 3669 (1984).
  • Camp and Allen (1996) P. J. Camp and M. P. Allen, Physica A 229, 410 (1996).
  • Do Carmo et al. (2010) E. do Carmo, D. B. Liarte, and S. R. Salinas, Phys. Rev. E 81, 062701 (2010).
  • Mertelj et al. (2018) A. Mertelj, L. Cmok, N. Sebastián, R. J. Mandle, R. R. Parker, A. C. Whitwood, J. W. Goodby, and M. Čopič, Phys. Rev. X 8, 041025 (2018).
  • Sebastián et al. (2020) N. Sebastián, L. Cmok, R. J. Mandle, M. R. de la Fuente, I. D. Olenik, M. Čopič, and A. Mertelj, Phys. Rev. Lett. 124, 037801 (2020).
  • Chaturvedi and Kamien (2019) N. Chaturvedi and R. D. Kamien, Phys. Rev. E 100, 022704 (2019).
  • Chiappini and Dijkstra (2021) M. Chiappini and M. Dijkstra, Nat. Commun. 12, 2157 (2021).
  • Kotni et al. (2022) R. Kotni, A. Grau-Carbonell, M. Chiappini, M. Dijkstra, and A. van Blaaderen, Nat. Commun. 13, 7264 (2022).
  • De Gregorio et al. (2016) P. De Gregorio, E. Frezza, C. Greco, and A. Ferrarini, Soft Matter 12, 5188 (2016).
  • Chiappini et al. (2019) M. Chiappini, T. Drwenski, R. van Roij, and M. Dijkstra, Phys. Rev. Lett. 123, 068001 (2019).
  • Marconi and Tarazona (1999) U. Marini Bettolo Marconi and P. Tarazona, J. Chem. Phys. 110, 8032 (1999).
  • Archer and Evans (2004) A. J. Archer and R. Evans, J. Chem. Phys. 121, 4246 (2004).
  • te Vrugt et al. (2020) M. te Vrugt, H. Löwen, and R. Wittkowski, Adv. Phys. 69, 121 (2020).
  • te Vrugt and Wittkowski (2022) M. te Vrugt and R. Wittkowski, J. Phys. Condens. Matter 35, 041501 (2022).
  • Löwen (1999) H. Löwen, Phys. Rev. E 59, 1989 (1999).
  • Elder et al. (2002) K. R. Elder, M. Katakowski, M. Haataja, and M. Grant, Phys. Rev. Lett. 88, 245701 (2002).
  • Emmerich et al. (2012) H. Emmerich, H. Löwen, R. Wittkowski, T. Gruhn, G. I. Tóth, G. Tegze, and L. Gránásy, Adv. Phys. 61, 665 (2012).
  • Wittkowski et al. (2010) R. Wittkowski, H. Löwen, and H. R. Brand, Phys. Rev. E 82, 031708 (2010).
  • te Vrugt et al. (2022) M. te Vrugt, M. P. Holl, A. Koch, R. Wittkowski, and U. Thiele, Model. Simul. Mater. Sci. Eng. 30, 084001 (2022).
  • te Vrugt et al. (2021) M. te Vrugt, J. Jeggle, and R. Wittkowski, New J. Phys. 23, 063023 (2021).
  • Wittkowski and Löwen (2012) R. Wittkowski and H. Löwen, Phys. Rev. E 85, 021406 (2012).
  • Kümmel et al. (2013) F. Kümmel, B. ten Hagen, R. Wittkowski, I. Buttinoni, R. Eichhorn, G. Volpe, H. Löwen, and C. Bechinger, Phys. Rev. Lett. 110, 198302 (2013).
  • Wittkowski and Löwen (2011) R. Wittkowski and H. Löwen, Mol. Phys. 109, 2935 (2011).