Introduction

Colloids form a platform for scalable fabrication of mesostructured composite materials and provide a framework for testing theoretical descriptions of crystals and glasses, albeit they are also commonly encountered in daily life in forms of milk, paints, coffee, fog and so on1,2,3,4,5. Nematic liquid crystal (NLC) colloids6,7,8,9,10,11,12,13,14,15,16,17 attract particularly strong interest because they can be tuned by weak external stimuli, such as low-voltage fields and light, similar to the NLC host fluids themselves12,17. Colloidal particles locally perturb the uniform ground-state alignment of the anisotropic NLC molecules described by the so-called ‘director’ n≡−n with non-polar symmetry, producing spatial elastic distortions of the director field n(r) that resemble electric field configurations around dipolar and quadrupolar charge distributions7,8,9,10,11,12. The type of distortions depends on surface boundary conditions for n(r), which can be tangential or perpendicular to the particle surfaces and can induce both bulk and surface line and point defects7,8,9,10,11,12, topological singularities along which n(r) and NLC order cannot be defined1. Such NLC colloidal particles tend to arrange themselves so that they can share energetically costly n(r) distortions to minimize free energy, exhibiting highly anisotropic elasticity-mediated interactions that resemble interactions of electrostatic dipoles and quadrupoles6,7,8,9,10,11,12,15,16. This electrostatics analogy7 provides a framework for understanding, predicting and engineering the ensuing colloidal self-assembly. However, for many years of very active research efforts6,7,8,9,10,11,12,15,16, only colloidal elastic dipoles and quadrupoles have been found in experiments and theories alike, which limits diversity of the accessible colloidal bonds and self-assembled structures6,7,8,9,10,11,12,13,14,15,16,17.

In this work, we describe colloidal elastic hexadecapoles that spontaneously form around solid microspheres immersed in a uniformly aligned NLC. The unusual symmetry of elastic distortions arises from conically degenerate18 boundary conditions for n(r) at the colloidal surfaces, which induce surface point and line defects at the same time. Using a combination of holographic optical tweezers (HOT)19, nonlinear optical imaging20 and polarizing optical microscopy (POM)1,6, we probe the n(r)-distortions and quantify colloidal pair interactions by measuring distance and angular dependencies of elastic potentials, demonstrating relations between the director structure and medium-mediated inter-particle forces. Finally, we explain our findings using a model based on elastic multipole expansion and discuss how the experimental framework we have developed may enable colloidal self-assembly into novel forms of tunable pre-engineered matter without known atomic analogues.

Results

Hexadecapolar elastic multipole

When dispersed in a uniformly aligned NLC fluid host, polystyrene microspheres (PSMs) of a radius r0 locally distort n(r), which is manifested by eight bright lobes around the particle perimeter seen in POM between the crossed polarizer and analyser (Fig. 1a). These bright lobes are separated by eight dark regions within which n(r) at the particle’s perimeter is parallel to polarizer or analyser. Using a phase retardation plate and interference of polarized light propagating through the particle-distorted structure, we reveal that n(r) tilting away from the far-field director n0 switches between clockwise and counterclockwise directions (corresponding to the blue and yellow colours in the micrograph) eight times as one circumnavigates the sphere (Fig. 1b). Bright-field micrographs obtained at different depth of focus reveal presence of weakly scattering surface point defects (boojums) at the particle poles along n0 as well as a circular loop of a defect line (often called ‘Saturn ring’)9,10 at the particle’s equator (Fig. 1c,e). Based on POM and three-dimensional nonlinear optical imaging (Fig. 1a–c,e and Supplementary Figs 1–3), we uncover the structure of n(r) distortions schematically shown in Fig. 1d. This structure is consistent with conically degenerate surface boundary conditions for n(r) with respect to PSM’s local surface normals s, which were previously demonstrated for NLCs at flat polystyrene-coated surfaces18. The director’s easy axis orientation lies on a cone of equilibrium polar angle ψe (Fig. 1f). To minimize the free-energy cost of bulk elastic distortions, interaction of conically degenerate surface boundary conditions on the microsphere with the uniform n0 lifts this conical degeneracy and yields an axially symmetric n(r) depicted in Fig. 1d. Extending this analysis to three dimensions, the projection nx of n(r) onto the x-axis orthogonal to n0 can be visualized around PSM using colours that highlight positive, near zero and negative nx (Fig. 1g). The black points at the poles and ring at the equator of the sphere are regions of discontinuity of n(r) at the NLC-PSM interface and correspond to the boojum and Saturn ring topological defects, respectively. Away from the particle surface and these topological singularities, the experimentally reconstructed n(r) is continuous (Fig. 1d) and consistent with the theoretical model for a hexadecapolar director distortions presented in a similar way in Fig. 1h.

Figure 1: Elastic hexadecapole induced by a colloidal PSM.
figure 1

(ad) Optical micrographs obtained with (a,b) POM and (c,e) bright field microscopy. P, A and γ mark the crossed polarizer, analyser and a slow axis of a 530 nm retardation plate (aligned at 45° to P and A), respectively. (d) Schematic diagram of induced n(r) (green rods) satisfying the tilted boundary conditions at the PSM surface (red rods), with the ‘easy axis’ at a constant angle ψe to a local normal s to the surface (black rods). (f) Schematic of conic degenerate surface boundary conditions. (g,h) Three-dimensional visualization of the x-component nx of n(r) (g) at the surface of PSM for ψe=45° and (h) at a spherical surface of radius 1.2r0 shown using a dashed red circle in d. Blue, yellow and magenta colours correspond, respectively, to a positive, near-zero and negative nx according to the colour scheme shown in g. Dashed equatorial line in d and a black solid line in g depict the ‘Saturn ring’ surface defect loop at the particle’s equator visible in c. Black hemispheres in d and g show surface point defects boojums at the poles of the particle visible in e. Scale bars, 2 μm.

Elastic interactions between hexadecapolar colloids

Elasticity-mediated interactions between PSMs (Fig. 2) differ from all NLC colloids studied so far. To get insights into the strength and direction-dependence of these interactions, we utilize HOT to optically trap one ‘stationary’ PSM and then release another particle at different centre-to-centre vector R orientations with respect to n0 as well as at different separation distances. Using video-microscopy, we track the ensuing particle motions, which result from a combination of Brownian jiggling and elasticity-mediated interactions. The submicron waist and relatively low power (10 mW) of a focused trapping beam allow us to avoid the influence of the trapping on the measurements21. Furthermore, the laser tweezers are used only to bring the particles to the desired initial conditions and are turned off within the time when the pair-interaction is probed with video microscopy, allowing us to avoid possible artefacts associated with the complex effects of the laser trapping light at small inter-particle distances. The colour-coded time-coordinate trajectories of particles released from optical traps at different R are shown in Fig. 2a. Unlike in the case of dipolar and quadrupolar NLC colloids7,8,15,16, elastic forces are relatively short-ranged and exceed the strength of thermal fluctuations only at distances of four-to-five particle radii r0. However, the angular dependence of these forces is very rich, with eight angular segments of inter-particle attractions separated by eight angular segments of repulsions, with the intermediate angular sectors within which particles move sideways as the inter-particle elastic forces are orthogonal or at large angles to R (Fig. 2a). These angular sectors of attraction and repulsion correlate with the bright and dark regions of POM micrographs (Fig. 1a) as well as with the predictions of our model based on elastic multipole expansion (Fig. 2).

Figure 2: Elastic interactions of colloidal hexadecapoles.
figure 2

(a) Angular dependence of interactions is probed by tracking motion of a particle released from the optical trap and moving with respect to the ‘stationary’ trapped particle in the centre depending on the orientation of R with respect to n0. The elapsed time is coded according to the colour scale (inset) and the maximum elapsed times tmaxt0 are marked next to the corresponding trajectories. R|| and R denote the centre-to-centre distance R components along and perpendicular to n0, respectively. The PSM of radius r0 is surrounded by the spherical volume of radius 2r0 that is excluded for centres of other PSMs. The nonlinear zone is shown by a dashed circle at R=2.4r0. (b) Pair-potential Uint versus θ for two particles at R=2.4r0. Angular zones of repulsion and attraction are highlighted using magenta and green arrows and colouring, respectively.

To quantify elastic interactions, we first probe the centre-to-centre separation R=|R| versus time t for particles released at different angles between R and n0 within the angular sectors of attraction (Fig. 3a) and then calculate particle velocities =d R/d t. Because the system is highly over-damped and the inertia effects can be neglected22, the experimentally measured R(t) and the simplified equation of motion 0≈−ξ+Fint yield the pair interaction potential Uint (inset of Fig. 3a), where ξ is a drag coefficient measured separately by characterizing Brownian motion of PSMs (Methods and Supplementary Fig. 4) and Fint=−Uint is the elastic interaction force. The attractively interacting particles stop short of physically touching each other, instead forming stable dimer assemblies with typical R≈(2.05-2.2)r0 and stable R orientations with respect to n0 within one of the two angular sectors of assembly in each quadrant dependent on ψe: θ1≈22°-26° or θ2≈64°-75° (Fig. 3b–f). Multi-particle assemblies with different combinations of angles θ1 and θ2 are also observed (Fig. 3d). The inter-particle binding energies are measured to be in the range of hundreds of kBT (inset of Fig. 3a), making them robust with respect to thermal fluctuations. Although Uint versus θ has eight minima, only four of them can be occupied simultaneously in one plane by particles of the same size because of the ‘excluded volume’ effects, yielding two-dimensional colloidal crystals with rhombic elementary cells (Supplementary Fig. 5 and Supplementary Note 1). Following similar considerations, a large number of low-symmetry three-dimensional colloidal structures can be envisaged too. Since the elastic interactions potential is hundreds of kBT and the particle assemblies can be entrapped in metastable states, the assembly of two- and three-dimensional colloidal lattices from micrometer-sized particles requires the use of optical tweezers for guiding colloidal particles. Alternatively, the elastic interaction potentials between colloidal particles of smaller size or with weaker surface anchoring can be brought to the order of 10 kBT and lower, so that the crystal self-assembly can occur without the assistance of optical tweezers, which will be explored elsewhere.

Figure 3: Self-assembly of hexadecapolar colloids.
figure 3

(a) Separation R versus time t for attractive interactions at different θ (measured upon the dimer formation). The solid red line shows a least-squares fit with R(t) obtained from the simplified equation of motion with (b2, b4, b6)=(−0.017, −0.092, 0.003); it is impossible to reproduce such a dependence R(t) with the parameter b2 only since the quadrupole–quadrupole interaction is repulsive along θ≈23°. Inset shows the corresponding Uint versus R. (b) and (c) POM micrographs of self-assembled colloidal dimers of hexadecapoles with R aligned at θ1 and θ2 to n0, respectively. (d) POM micrograph of a kinked colloidal chain of hexadecapoles. (e,f) Schematic diagrams of n(r) (green lines) for colloidal dimers shown in c,d, respectively. Dashed lines and black hemispheres in c and d depict surface defect loops and surface point defects at the poles of PSMs, respectively. Scale bars, 2 μm.

To model experiments, we exploit the electrostatic analogy of the far-field director distortions because of a colloidal sphere that can be represented as elastic multipoles23, albeit our colloids dramatically differ from elastic dipoles and quadrupoles studied so far7,8,9,10,11,12,23,24,25,26. Far from the colloidal sphere, the director deviations nμ (μ=x; y) from n0=(0, 0, 1) are small. Assuming n(r)≈(nx, ny, 1), the NLC elastic free-energy reads7,8,23

where K is an average Frank elastic constant1. Euler–Lagrange equations arising from the functional (1) are of Laplace type, Δnμ=0, with solutions expanded into multipoles

where al=blr0l+1 is the elastic multipole moment of the lth order (2l-pole)22,23,24,25,26 and one can find coefficients bl from exact solutions for n(r) or from relevant experiments (Figs 2a and 3a). Odd moments vanish because n(r) is symmetric about the particle centre, similar to analogous electrostatic charge distributions that have no dipole or octupole electric moments22,23,24,25,26. The multipole expansion of the induced n(r) shown in Fig. 1d is characterized by coefficients b2, b4 and b6, which also determine the colloidal pair-interaction energy24

where Pl+l(cosθ) are the Legendre polynomials. For colloidal quadrupoles, b2 dominates and b4 and b6 can play a role only at small25 R. For our particles (Fig. 1), the induced n(r) can be qualitatively understood as a superposition of configurations of two separate quadrupoles, one with the Saturn-ring and one with boojums9,10,11,16,22,23,24,25,26, having opposite signs of quadrupole moments (compare Fig. 4a,d,g,j and 4b,e,h,k). Therefore, the net quadrupole moment is small and the high-order multipoles manifest themselves in a wide range of R. Fitting experimental R(t) within different angular sectors of pair interaction with the corresponding theoretical predictions yields a unique set of parameters b2, b4 and b6 (Fig. 3a). The quadrupole moment a2=−0.017r03 is about two orders of magnitude smaller than that of elastic quadrupoles16,24, consistent with the hexadecapolar symmetry of n(r) that is found playing a dominant role, and even the higher-order term (64-pole) plays a detectable role at relatively small R (Supplementary Fig. 6 and Supplementary Note 2). The rich angular dependence of elastic pair-interactions predicted by equation 3 for coefficients b2, b4 and b6 obtained from fitting is consistent with our experimental characterization of the hexadecapolar nature of PSM colloids in the NLC host (Fig. 2).

Figure 4: Elastic multipoles induced by spherical particles with different anchoring.
figure 4

(a,b) Polarizing and (d,e) bright field micrographs and (g,h) schematic diagrams of a director field n(r) (magenta lines) around elastic quadrupoles, respectively, formed by spheres with a strong (a,d,g) tangential and (b,e,h) homeotropic anchoring. Black filled semicircles at the poles of the particle in g show surface point defects dubbed ‘boojums.’ A line at the particle equator in h shows a closed surface disclination loop dubbed ‘Saturn ring.’ (c) Polarizing micrograph and (f) a schematic diagram of n(r) around an elastic dipole formed by a sphere with a strong homeotropic surface anchoring. A black point in the front of the particle shown schematically in f depicts a bulk point defect ‘hedgehog.’ (i) Schematic diagram of a quadrupolar n(r) around a sphere with weak surface anchoring. A dashed loop inside the particle shows a virtual disclination ring. (jl) Colour-coded visualizations of the nx component of n(r), which is caused by the tilt of n(r) away from n0, at the spherical surface enveloping the particles shown, respectively, in f,g,h. The spherical surfaces of colour-coded visualization of director distortions have radius of (1.2-1.4)r0, as shown by dashed circles in fh. Blue, yellow and magenta colours correspond, respectively, to a positive, zero and negative nx according to a colour scale bar provided in k. Scale bars, 4 μm.

Discussion

It is instructive to compare hexadecapolar NLC colloids to other known elastic multipoles and their electrostatic analogues. Colloidal spheres with strong tangential anchoring at their surface form n(r)-distortions of quadrupolar configuration16,24,27,28 and are commonly called elastic quadrupoles (Fig. 4a,d,g,j). They have two surface point defects at the two poles of the spherical particle, which are called ‘boojums’. Spherical particles with homeotropic surface anchoring can induce two different elastic multipoles, a dipole23,28 (Fig. 4c,f,l) and a quadrupole16,23,27,28,29,30,31,32 (Fig. 4b,e,h,k). The elastic colloidal dipole has a bulk point defect called ‘hedgehog’ near one of the poles and lacks mirror symmetry with respect to the plane going through its equator orthogonally to n0 (Fig. 4f,l). Such a colloidal dipole has a small octupole moment as well33. The elastic quadrupole around the particle with a strong homeotropic anchoring has a closed disclination loop around the equator called ‘Saturn ring’ and, in all four quadrants, has an opposite tilt of the deformed director with respect to n0 as compared with the tangential elastic quadrupole16 (compare Fig. 4g,j,h,k). In the case when anchoring is weak, an elastic multipole with weak distortions of quadrupolar configuration can be formed somewhat similar to the elastic quadrupole with ‘Saturn ring’, but its disclination loop is ‘virtual’ (within particle’s volume) as the director is allowed to deviate away from the easy axis orientation31,32. The maps of x-component of the director, nx, calculated using methods described in refs 23, 24 and plotted on spherical surfaces encompassing the particles and the defects, clearly illustrate the multipolar nature of elastic distortions similar to that in the electrostatics analogues34,35. A direct comparison of the symmetry of the director distortions around elastic dipoles and quadrupoles with that of hexadecapolar PSMs in NLCs helps to highlight the very different nature of these high-order NLC colloidal multipoles studied in our work (compare Figs 4j–l and 1h). On the other hand, the hexadecapolar NLC colloids further expand the analogy between electrostatic and elastic colloidal multipoles. In fact, the colour presentations of elastic distortions induced by colloidal dipoles (Fig. 4l), two different quadrupoles (Fig. 4j,k) and our hexadecapoles (Fig. 1g,h) resemble very closely the corresponding dipolar, quadrupolar and hexadecapolar electrostatic charge distributions on a sphere described by a spherical harmonic function35 σlm(Θ, ϕ)=Ncos()Plm(cosΘ) with (l, m)=(1, 1), (2, ±1) and (4, 1), respectively, where N is a normalization constant, Θ is a polar and ϕ is an azimuthal angles, Plm(cosΘ) is the associated Legendre polynomial; the constant l determines the order of a multipole (that is, 2l-th pole) and –lml. This close analogy between electrostatic and elastic colloidal multipoles may help devising approaches for self-assembly of colloidal mesostructured composite materials.

To conclude, we describe hexadecapolar NLC colloids with unusual field configurations, highly anisotropic elastic interactions, and versatile forms of self-assembly. Our findings pose a challenge of realizing pure NLC colloidal octupoles, lower order multipoles that have not been observed so far, and provide the means of probing the role of hexadecapolar moments in inter-molecular interactions using colloidal model systems. Self-assembly of hexadecapolar NLC colloids is expected to yield a diverse family of two- and three-dimensional low-symmetry crystal lattices. Beyond the rich experimental platform for fundamental studies, our colloids have potential uses in designing and realizing reconfigurable ‘soft’ photonic crystals36 and other NLC-based mesostructured composites37 with properties that can be pre-engineered through controlling tilted boundary conditions on nanoparticle surfaces38,39,40, particle shape, composition and topology,2,6,8,41 as well as the NLC host properties, particle active behaviour42 and the use of external stimuli12,17.

Methods

Preparation of nematic colloidal dispersions

Colloidal PSM particles of radii r0≈2.05, 2.65 and 4 μm and with conically degenerate surface boundary conditions were prepared using one of two different methods as follows43,44. In the first method, asymmetric colloidal polystyrene dimers43,45 were synthesized using a modified seeded polymerization technique44,45,46 (Supplementary Fig. 1a). Resulting dimers in ethanol were sonicated for tens of minutes using an ultrasonic bath 8891 (Cole-Parmer) to break apart large and small spherical lobes forming the asymmetric colloidal dimer (Supplementary Fig. 1b,c). Colloidal spheres originating from these dimers were then re-dispersed in pentyl-cyano biphenyl (5CB) obtained from Frinton Laboratories, Inc. The larger (diameter D=2r0≈4.1 μm) colloidal spheres, which were used as PSMs in our experiments, intrinsically provide conic degenerate surface boundary conditions for the NLC director. In the second preparation method, we used polystyrene divinylbenzene spherical particles PC06N (Bang Laboratories, Inc.) of D≈8 μm and DC-05 (Thermo Scientific) of D≈5.3 μm dispersed in 5CB. As dispersed, these particles provide tangential alignment for 5CB molecules and the director field n(r). However, after keeping these colloidal dispersions in 5CB at the elevated temperature of ≈100 °C for about 12 h, the alignment at the surface of these polymer particles changed to a conical degenerate (Supplementary Fig. 1d–f), similar to that reported in ref. 18 for flat surfaces coated by polystyrene. The tilt angle of the easy axis ψe was varying approximately within 20°-75°, depending on details of preparation. The comparison of the surface free-energy density Fs=Wc(cos2ψ−cos2ψe)2/2 costs (ref. 18) for deviating n(r) to a polar angle ψ away from the easy axis at ψe relative to the bulk elastic energy density is characterized by the so-called conical anchoring extrapolation length K/Wc(50-500) nm <<r0, where18 Wc10−5 J m−2. Since K/Wc<<r0, one can assume that the tilted conically degenerate boundary conditions are rigid with respect to changing ψ. To prepare samples with NLC dispersions in a uniformly aligned host, dry PSMs were added to 5CB. The dilute NLC colloidal dispersions were then filled into the cells comprised of two glass plates separated by glass spacers defining the cell gap in the range 15-30 μm, which were sealed at all edges using fast-setting epoxy glue after the filling process. The inner surfaces of cell substrates were treated with polyimide PI2555 (obtained from HD Microsystems) for in-plane homogeneous alignment of the far-field director n0 defined by unidirectional rubbing.

Imaging techniques and data acquisition

POM (Fig. 1 and Supplementary Figs 1,2), nonlinear optical imaging (Supplementary Fig. 3) and laser manipulations of PSM particles were performed using a single integrated setup capable of simultaneous conventional optical and nonlinear three-photon excitation fluorescence polarizing microscopy (3PEF-PM)20 and HOT optical control operating at 1,064 nm (ref. 19). The setup was built around an inverted microscope IX81 (Olympus). For 3PEF-PM imaging20, we have employed a tunable (680-1,080 nm) Ti-Sapphire oscillator (Chameleon Ultra II, Coherent) emitting 140 fs pulses at a repetition rate of 80 MHz. The laser wavelength was tuned to 870 nm for the three-photon excitation of 5CB molecules and the 3PEF-PM signals were collected in epi-detection mode with a photomultiplier tube (H5784-20, Hamamatsu). An Olympus × 100 oil-immersion objective with a high numerical aperture of 1.4 was used for both optical imaging and laser trapping. This experimental setup is described in details elsewhere19,20. Motion of colloidal particles was recorded with a charge-coupled camera (Flea, PointGrey) at a rate of 15 fps and their lateral positions versus time were determined from captured sequences of images using motion tracking plugins of ImageJ software (obtained from National Institutes of Health) with the accuracy8 of 7-10 nm.

Measuring anisotropic diffusion of elastic NLC hexadecapoles

PSMs with conic anchoring distort the homogeneous n(r) and form elastic hexadecapoles in the aligned NLC (Fig. 1a–e and Supplementary Figs 1b,d–f, 2,3b). These hexadecapoles stay suspended in the bulk of NLC (Supplementary Fig. 3b), which is facilitated by elastic repulsion of PSMs from the confining substrates and their Brownian motion because of thermal fluctuations. Using bright field microscopy and video tracking, one can determine a position of the particle within each frame and then analyse its translational displacements at regular time steps corresponding to the video frame rates. Following the well-established methodology29 and using the histograms of displacements shown in Supplementary Fig. 4, it is possible to experimentally determine two independent diffusion coefficients D||=1.63 × 10−3 μm2 s−1 and D=1.05 × 10−3 μm2 s−1, which characterize, respectively, the diffusion of elastic hexadecapoles along and perpendicular to the far-field director n0. As D||>D, it is easier for the NLC hexadecapoles to move along n0 than in other directions. The ratio D||/D=1.55 is not much different from values determined for other elastic multipoles studied previously29,30 (Fig. 4). These coefficients are also used to determine friction coefficients from the Stokes–Einstein relation, which are then subsequently used in the calculation of anisotropic interaction forces27 between hexadecapoles.

Additional information

How to cite this article: Senyuk, B. et al. Hexadecapolar colloids. Nat. Commun. 7:10659 doi: 10.1038/ncomms10659 (2016).