Significance
A persistent theme that has emerged from the century-long experimental and theoretical study of ferroelectric crystals and liquid crystals is that ferroelectricity and antiferroelectricity go hand-in-hand, such that if one is found, the other will also be in the same or closely related families of materials. At the root of this behavior is the frustration inherent in steric and electrostatic dipolar interactions, exemplified by the latter, for which dipole pairs arranged end-to-end prefer to be parallel, while dipole side-by-side pairs prefer to be antiparallel. Here, we show that the unexpected recent discovery of the ferroelectric nematic phase and its nearly perfect ferroelectric polar order, is a pathway to antiferroelectricity in fluids, realized in a lamellar smectic array of fluid sheets of alternating polarity.
Keywords: liquid crystal, smectic, antiferroelectric, ferroelectric nematic
Abstract
We have structurally characterized the liquid crystal (LC) phase that can appear as an intermediate state when a dielectric nematic, having polar disorder of its molecular dipoles, transitions to the almost perfectly polar-ordered ferroelectric nematic. This intermediate phase, which fills a 100-y-old void in the taxonomy of smectic LCs and which we term the “smectic ZA,” is antiferroelectric, with the nematic director and polarization oriented parallel to smectic layer planes, and the polarization alternating in sign from layer to layer with a 180 Å period. A Landau free energy, originally derived from the Ising model of ferromagnetic ordering of spins in the presence of dipole–dipole interactions, and applied to model incommensurate antiferroelectricity in crystals, describes the key features of the nematic–SmZA–ferroelectric nematic phase sequence.
Proper ferroelectricity in liquids was predicted in the 1910s by P. Debye (1) and M. Born (2), who applied the Langevin–Weiss model of ferromagnetism to propose a liquid-state phase change in which the ordering transition is a spontaneous polar orientation of molecular electric dipoles. A century later, in 2017, two groups independently reported novel nematic phases in the polar molecules of Fig. 1A, the “splay nematic” in the molecule RM734 (3–5) and a “ferroelectric-like nematic” phase in the molecule DIO (6). These nematic phases were subsequently demonstrated to be ferroelectric in RM734 (7) and in DIO (8, 9), and a study of their binary mixtures (Fig. 1B) showed that these are the same phase, termed NF here, a uniaxially symmetric, spatially homogeneous, ferroelectric nematic liquid having >90% polar ordering of its longitudinal molecular dipoles (7, 9) (Fig. 1D). These mixtures also exhibit, at higher temperatures, a nonpolar, paraelectric nematic (N) phase (Fig. 1C). The separate molecular families which RM734 and DIO represent are based on distinctly different molecular structural themes. In the course of the remarkable developments around nematic ferroelectricity, homologous molecules with a variety of substitutions around these themes have been studied, and many of these have also been found to exhibit the NF phase (8, 10).
Fig. 1.
(A) Representative members DIO [(2,3',4',5'-tetrafluoro-[1,1'-biphenyl]-4-yl 2,6-difluoro-4-(5-propyl-1,3-dioxane-2-yl) benzoate] and RM734 [4-[(4-nitrophenoxy) carbonyl]phenyl2,4-dimethoxybenzoate] of nitro- and fluoro-based molecular families of molecules that exhibit the ferroelectric nematic phase. This behavior is attributable in part to their rod-like molecular shape, large longitudinal dipole moment, and a common tendency for head-to-tail molecular association. (B) Phase diagram of DIO and RM734 and their binary mixtures exhibiting: (C) Paraelectric nematic (N) phase, in which the molecular long axes are aligned in a common direction but with no net polarity, making a nonpolar phase with a large dielectric response. (D) Ferroelectric nematic (NF) phase, in which the polar orientation of dipoles is unidirectional and long-ranged, with a large polar order parameter (~0.9). The NF phase is uniaxial and spatially homogeneous, with a macroscopic polarization/director field P(r), shown here for antiparallel polar orienting surface alignment layers on the plates (9). The phase diagram shows that the transition temperature into the NF depends nearly linearly on concentration c, indicating that these molecules exhibit near ideal mixing behavior of the transition into the ferroelectric nematic (NF) phase. This transition is weakly first-order, with entropy ΔS ≈ 0.06R. (E) Intermediate phase observed in neat DIO and in DIO-rich mixtures. Originally called the M2 (6), we find this phase to be electron density-modulated into ~9-nm thick, polar layers, in which the director and polarization are parallel to the layer plane and alternate in direction from layer to layer, as sketched. We have termed this lamellar, antiferroelectric phase the smectic ZA (SmZA). The SmZA can be prepared with the layers either normal to the plates (bookshelf geometry) or parallel to the plates, or with the chevron texture shown in Figs. 2 and and SI Appendix Figs. 6-8.
The initial study of DIO reported the “M2” phase, a second, unidentified liquid crystal (LC) phase in a ΔT = 15 °C temperature range between the N and NF phases (Fig. 1B) (6). Here, we report that this phase exhibits an equilibrium, sinusoidal electron-density modulation of 8.8-nm periodicity. The M2 is, remarkably, a density-modulated, antiferroelectric phase, exhibiting lamellar (LAM) order with ~17.5-nm periodicity, comprising pairs of 8.8-nm-thick layers of opposite ferroelectric polarization (Fig. 1D), related by a reflection plane normal to the director. As in other layered LC phases, these two-dimensional (2D) layers tile three-dimensional (3D) space with fluid spatial periodicity, a structure that largely controls the optic and electrooptic (EO) behavior of bulk samples.
In one of the stunning achievements of condensed matter physics, Georges Friedel combined his optical microscopic observation of elliptical and hyperbolic focal conic defect lines in LCs with knowledge of the Dupin cyclides (complex, equally spaced, 3D-curved surfaces that envelop certain single-parameter families of spheres) to infer the existence of the fluid molecular layer ordering of smectic LCs, without the benefit of X-ray scattering (11). He termed such phases “smectic,” following the Greek, as similar structures were found in concentrated solutions of neat soaps (12). Friedel’s samples also exhibited what he termed “homeotropy,” where the preferred orientation in the smectic of its uniaxial optic axis, the local average long molecular axis, given by the director n, is normal to the inferred planes. In time, as understanding of the polymorphism of LCs advanced, it became necessary to subdivide the class of smectic-like phases by structure, and Friedel’s specific smectic geometry, the one having θ = 0°, where θ is the angle between the layer normal z and optic axis n, was appropriately named the “smectic A” (13). Here, following a path of textural analysis inspired by Friedel, and aided by synchrotron-based small and wide angle x-ray scattering (SAXS & WAXS), we report a fluid, layered smectic phase with θ = 90°, the structural antipode to Friedel’s smectic A, which we classify as the “smectic Z” (14).
Results
X-Ray Scattering.
DIO was synthesized using standard synthetic schemes (SI Appendix, Fig. S1). SAXS (Fig. 2) and WAXS images (SI Appendix, Figs. S2 and S3), obtained upon cooling DIO from the Iso phase with qz along the magnetically aligned nematic director, show the previously observed (3, 6), intense, diffuse scattering features at qz ~ 0.25 Å−1 and qy ~ 1.4 Å−1 from end-to-end and side-by-side molecular positional pair correlations, respectively. In the N phase, at small q there is only low-intensity scattering that varies smoothly with qz. However, upon cooling to the M2 phase, a pair of diffraction spots appear on this background, visible between T ≈ 83 °C and T ≈ 68 °C and with the integrated peak intensity a maximum at T ~ 76 °C. Scans of intensity through these peaks are shown in Fig. 2D and SI Appendix, Figs. S4 and S5. Simultaneously observed is a continuous reorientation, without change of shape, of the entire WAXS/SAXS scattering pattern, through an angle δ(T), starting at T ≈ 83 °C and reaching δ = 12° ± 1° at T = 75°, as seen in Fig. 2B. The diffraction peaks at T = 75 °C are at wavevector qM = 0.071Å−1 (Fig. 2D), indicating an electron-density modulation of wavelength wM = 8.8 nm, with qM along y, normal to the nematic director, the local mean molecular long-axis orientation, n, along z. The WAXS peak at qy ~ 1.4 Å−1 corresponds to a side-by-side molecular spacing in DIO of ~0.42 nm in the direction of the modulation, so the period of this modulation is supermolecular, ~20 molecules wide. The growth of such a periodic structure in a uniaxial, field-aligned nematic should give powder scattering in the (qx,qy) plane normal to n. That only a single pair of diffraction spots appear indicates that there is only a single diffraction ring in this plane intersecting the Ewald sphere (approximately the (x,z) plane), indicating that this ordering could be either 1D lamellar smectic-like, or 2D hexagonal columnar. The textural studies and electro-optic experiments presented below show definitively that this is a lamellar phase. Macroscopically, this phase is a fluid, as confirmed by its diffuse scattering features in the WAXS, which differ very little from those of the fluid nematic phase (SI Appendix, Fig. S2). This observed phase sequence is consistent with that of earlier reported DIO studies (6, 9, 15), indicating that it is the trans form of DIO that is being studied [SI Appendix, (15)].
Fig. 2.
SAXS scattering from the N and SmZA phases of DIO. These diffractograms are from a sample in a 1 mm-diameter capillaries in which the director, n, is aligned along the magnetic field B (red), studied on the SMI microbeam line at NSLS II. (A and B) Non-resonant diffractograms obtained on cooling show a diffuse scattering arc at qz = 0.25Å−1 from head-to-tail stacking of the molecules in the N and SmZA phases. In the SmZA phase, equatorial Bragg spots appear, indicating the presence of an electron density wave of wM = 8.8-nm periodicity, with wavevector qM normal to n. These peaks disappear upon transitioning to the NF phase (G, SI Appendix, Figs. S4 and S5). (C) The scattering pattern rotates from (A) to (B) because of layer reorientations caused by a small reduction in layer thickness upon cooling in the SmZA phase (see text). (D) Line scan through the non-resonant scattering peaks. The central part of the beam is blocked by the beam stop. (E) Since the sample is a powder in orientation about the B field, the observed scattering pattern could be indicative of either lamellar or hexagonal columnar positional ordering. Since non-resonant scattering is sensitive only to electron density, the layers (of thickness wM) of opposite polarization scatter identically. Layer boundaries (white stripes) have different electron density than the layer centers (gray stripes), making wM the period observed in nonresonant scattering. However, as shown in (H) and SI Appendix, Fig. S5, carbon-edge resonant scattering exhibits the half-order peak at qR = qM/2, showing conclusively that the SmZA is lamellar and bilayer with period 2wM. The DTOM experiments show that it is lamellar and antiferroelectric, with layers of alternating polarization P(y), sketched in (E1) for low polarization (ξP > w), and in (E2) for high polarization (ξP << w), the case relevant to the SmZA (see text). The tilt modulation θ(y) is suppressed at large P by the self-electric field due to the polarization charge, ρ = −∇·P, indicated in red. Other smectic structures (e.g., the SmC, sketched in E3) have spontaneous antiferroelectric polarization in the tilt plane as well as splay modulation SI Appendix, Section 8), but in different geometries. (F) X-ray scattering peaks from various smectics with the director aligned by the magnetic field B. (G–I) Dependence on T of the Bragg peak parameters from data as in SI Appendix, Figs. S4 and S5 [non-resonant peak intensity (Ip), non-resonant peak position (qM = 2π/wM), resonant peak position (qR = 2π/2wM), and non-resonant half-width at half-maximum (HWHMM)]. (H) Layers of opposite P have different carbon K-edge resonant scattering cross-sections, and therefore Bragg scatter at the full antiferroelectric period 2wM. (I) The HWHMM values of the mid-range peaks (T ~ 75 °C) are SAXS-resolution limited.
As indicated above, the diffraction pattern rotates on cooling, with the lateral diffraction spots remaining in the equatorial plane of the SAXS/WAXS patterns. This implies that the entire sample volume probed by the beam (a 1-mm × 2-µm × 20-µm cylinder), the director n, and the layering wavevector qM, have all uniformly reoriented together through δ(T), a rotation of n away from the orientation preferred by the magnetic field. This kind of temperature-dependent director/layer reorientation accompanying layer formation upon cooling of smectic LCs has been observed in magnetically aligned (16) and surface-aligned (17) samples undergoing nematic to smectic A (SmA) to smectic C (SmC) (18–20) or N to SmC phase transitions upon cooling (21), in the context of the study of surface-stabilized ferroelectric LCs (22). This behavior is understood to be a result of fluid smectic layers contracting everywhere upon cooling, in the SmC case due to an increasing tilt of the molecular long axes away from the smectic layer normal with decreasing temperature. The result of layers of thickness wM(T1) = 2π/qM(T1) growing into a bulk sample at temperature T1 with an aligned director and then homogeneously shrinking in thickness upon cooling to T2 is sketched in Fig. 2B. A local reorientation of the layers through the angle δ(T) = cos−1[wM(T2)/wM(T1)], under conditions of low compressive stress of the layering, preserves the original pitch, wM(T1) of the layering as it was initially formed. In a large (1-mm diameter) X-ray capillary, the pattern of such reorientation is complex, and different on each cooling run. On a larger scale, defects, such as parabolic focal conics (23) or the planar breaks of a chevron structure sketched in Fig. 2C, mediate changes in the sign of δ(T). In the experiment shown in Fig. 2, the scattering volume was fortuitously filled with a single such reorienting domain. In thin cells with alignment layers, layer shrinkage is typically accommodated by the chevron layer structure and its characteristic defects, discussed below.
These observations and, in particular, the simple appearance of the scattering pattern in Fig. 2B, make it clear that the M2 is a form of fluid, lamellar phase with layer planes normal to a density modulation wavevector qM, which is, in turn, normal to the mean molecular long-axis n, constraining n to be parallel to the (x,z) plane of the layers. Such a structure is orthorhombic biaxial, with a principal symmetry/optic axis triad (l, qM, n), where l is the auxiliary unit vector normal to the qM, n plane. Related biaxial lamellar phases have been achieved only in amphiphilic systems, such as the “LAMN” phase in bola-amphiphiles (24), and the hybrid DNA/cationic liposome lamellar phases (25), which employ strong nanophase segregation to isolate anisotropic 2D nematic-like molecular monolayers within a lamellar phase. We will show that in the M2 phase each layer is structurally and electrically polar, with alternating polarity from layer to layer, an antiferroelectric condition shared by some biaxial phases of bentcore molecules (26).
Finding similar behavior in a molecule as simple as DIO is truly remarkable, so we classify this phase to be among the thermotropic smectics of rod-shaped mesogens, as sketched in Fig. 2F, naming it the smectic Z, such that the smectic A, the phase having qM parallel to n (θ = 0°), and the smectic Z, the phase having qM normal to n (θ = 90°) represent the lower and upper limits of tilt angles, θ. We may also consider that the M2 could be classified as a kind of nematic, in analogy with the twist-bend nematic (NTB) phase. The NTB phase is characterized by a spontaneous, heliconical precession of the nematic director, a periodic modulation that is helically (C∞) glide symmetric and therefore has no accompanying electron density variation or nonresonant X-ray scattering: different positions along the helix are relatively rotated but otherwise physically equivalent, making the NTB a kind of nematic, with a modulation that requires the use of resonant scattering in order to be observable with X-rays. However, in the case of a finite-amplitude, periodic modulation of the director orientation involving splay and/or bend, as we report below and sketch in Fig. 2E, for example, there is no equivalent translational symmetry, as planes with maximum and zero director distortion, or with a particular polarization magnitude, are inequivalent and therefore have different electron density. Such structures are therefore smectic. In the literature, there are many proposals for, and examples of, polar-modulated nematic director fields based on structures that have director splay and bend (27), including the splay nematic model proposed for the NF phase of RM734 (4, 5), but all such model systems will have corresponding modulation in density. Those with 1D modulation are smectic, as is directly demonstrated here by the non-resonant SAXS in Fig. 2 and SI Appendix, Figs. S4 and S5, exhibiting the ~9-nm periodic density modulation. Additionally, resonant SAXS results in Fig. 2 and SI Appendix, Fig. S6 show that the SmZA is antiferroelectric, with adjacent 9-nm layers exhibiting opposite polarization along n, and delineated by planes of zero average polarization parallel to n.
The observation of a single pair of diffraction spots (no harmonics) for both the nonresonant (Fig. 2 and SI Appendix, Fig. S4) and resonant (SI Appendix, Fig. S5) scattering suggests that the density modulation of the layering at qM, and the polarization modulation at qM/2 may be nearly sinusoidal. However, this notion should be considered with caution, as the scattering peaks appear with substantial background noise, so that lower intensity harmonics may be present but not visible in these experiments.
Optical Textures and EOs.
The X-ray scattering suggests that the SmZ phase should be optically biaxial, with the principal axes (x,y,z) of its optical dielectric tensor (x ∥ l, y ∥ qM, z ∥ n, as in Figs. 3 and 4). Textures visualized using depolarized transmission optical microscopy (DTOM) in Fig. 3 show the planar alignment achieved on cooling a d = 3.5-µm antiparallel-rubbed cell with a 3° angle between the buffing direction and the in-plane electrode edges. In absence of applied field, the N–SmZA transition is difficult to observe optically, although there is a small increase in birefringence δΔnyz ≈ 0.008 across the phase transition (compare Fig. 3 A–D). Excellent extinction is obtained for optical polarization along the buffing direction in both the N and SmZA phases (compare Fig. 3 B–E). The subsequent transition to the NF , on the other hand, is quite dramatic, with no extinction observed for any sample orientation in the twisted NF state (28) (Fig. 3 G and H). Fig. 3 C–F show the same ~1.5 mm long section of this cell as in Fig. 3B but with a very slowly time-varying, effectively DC, field applied, at 30 V/mm. In the N phase (Fig. 3B), this field generates a twist Freedericksz transition, shown for E = 30 V/mm in the center of the electrode gap, whereas in the slowly growing-in SmZA phase (Fig. 3C), this transition is suppressed.
Fig. 3.
Textures of DIO in a d = 3.5-µm cell with in-plane electrodes spaced by 1 mm and antiparallel buffing parallel to the electrode edges, cooled at −1 °C/min and viewed in DTOM with polarizer and analyzer as indicated. (A–C) Planar-aligned nematic monodomain with n parallel to the plates and to the electrode edges, showing (A) a birefringence color corresponding to d = 3.5 and Δn = 0.18, (B) excellent extinction with n parallel to the polarizer, and (C) the transmission due to twist of n during a field-induced twist Freedericksz transition. (D and E) Bookshelf-aligned (BK) smectic ZA monodomain formed on cooling. The smectic layers are parallel to n and normal to the cell plates. The birefringence is slightly larger than in the nematic phase and excellent extinction is again obtained with n parallel to the polarizer. (F) Optical response of coexisting nematic and SmZA phases to an applied 200-Hz square wave field of peak amplitude, E. The nematic region undergoes a twist Freedericksz transition but in the smectic ZA phase, rotation of the director is suppressed by the layering. (G and H) NF phase growing into the SmZA upon further cooling. The antiparallel buffing stabilizes a π-twist state in the polarization/director field of the NF, which does not extinguish between crossed polarizer and analyzer at any cell orientation. (I and J) Optical intensity scans along the magenta dashed lines across the bottom of the images in (C and F), probing the electric field-induced in-plane twist Freedericksz reorientation in the NF phase and showing that there is no observable electro-optic effect in the SmZA phase. In this geometry, where the layers have bookshelf (BK) alignment, an in-plane field would tend to rotate the director out of the smectic layer planes, a deformation that is locally resisted by the SmZA structure.
Fig. 4.
Zig-zag defects in a SmZA cell with chevron layering confirm the lamellar structure of the SmZA. Zig-zag walls appear in smectic cells originally with the layers substantially normal to the cell plates (bookshelf geometry) when the layers are caused to shrink in thickness. A common response to accommodate layer shrinkage in thin cells where there is surface pinning is for the layers to tilt, forming the chevron variation (CH) in bookshelf alignment, and zig-zag walls are the layering defects which mediate binary change in the pointing direction of the chevron tip. (A–C) Zig-zags generated in the SmZA phase as it is grown by cooling in a weak electric field (1 Hz square-wave, ~10 V/mm). Zig-zag walls are of two types, sketched in (C): “diamond” walls [〈〈〈〈〉〉〉〉], detailed in (D), where the chevrons point out from the wall center, and which run nearly along qM, i.e., nearly normal to the layers; and “broad” walls [〉〉〉〉〈〈〈〈], where the chevrons point in toward the defect wall, which run parallel to the layers and to the director, n. These zig-zag walls are analogous to those found in planar-aligned SmC and SmC* cells (SI Appendix, Fig. S7). (D) Detailed structure of a diamond wall in which the normal to the diamond-shaped layer elements is rotated to make an angle with the cell normal that is larger than the layer tilt angle, δ, of the chevrons, enabling the wall to space out the chevron tips in opposite directions (21). The orientation and birefringence of the diamonds walls causes them to transmit light when the overall chevron structure is at or near extinction, as in (A–C). The orientation of the diamond elements can be determined from their extinction angle (SI Appendix, Fig. S8C). The broad wall structure is less visible in the SmZA than in the SmC cells shown in SI Appendix, Fig. S6 because in the SmZA the rotations of layer elements in the wall do not result in reorientation of n. The presence of fringing fields near the electrodes [marked with white dashed lines in (B and C)] enables E fields to be applied either parallel or and normal to n in separate regions of the cell. The difference in optical saturation observed in these regions for different signs of field applied parallel to n (B and C) reveals that the polarity of the chevron structure generates an interfacial polarization that is indicated by the white arrows: the chevron areas are polar along y, with the white arrows rotating in opposite directions on opposite sides of a zig-zag wall, indicating a polar rotation of the director near the chevron interface when a field Ez is applied. In contrast, with the field along y there is no polar reorientational response of the director to change of field sign (B and C), indicating that the AF cancels the net polarization in the bulk (magenta, cyan arrows). The red twist and splay schematics are the possible director modulations considered in calculating the biaxial birefringence Δnbiax ≡ nq – nl = ΔnPA – ΔnBK (see text) of the SmZA. Measurement shows that Δnbiax > 0, indicating a splay modulation of n.
The in-plane electric field can induce a dielectric twist Freedericksz transition in the N phase at a threshold field as small as ~10 V/mm. The 3° bias of the buffing direction means that the director n initially makes an angle of 87° with the field E, so that the initial dielectric field-induced twist reorientation of n is everywhere in the same azimuthal direction (the direction that reduces the angle between the director and the applied field). The initially uniform cell shows good extinction at E = 0, but the applied electric field causes a twist distortion of the director field that results in transmission of light that increases with field strength, as illustrated by the white-light DTOM intensity profiles plotted in Fig. 3I. Under the same field conditions in the SmZA phase, however, there is no discernable optical response, as is evident from Fig. 3J, indicating that the director reorientation is suppressed in the SmZA phase, either by elimination of nematic dielectric torques, or by structure that strongly limits reorientation. The former possibility can be ruled out because the nematic dielectric torques are proportional only to the dielectric anisotropy Δε, which does not change substantially at the N–SmZ transition in DIO (4, 5). However, the latter phenomenon is quite familiar in smectic A phases. For example, the splay-bend Freedericksz reorientation observed when an in-plane electric field is applied to a homeotropically oriented cell of a nematic with Δε > 0 is eliminated at the N–SmA phase transition by the appearance of a local potential energy that keeps n along the smectic layer normal (29, 30). The restoring torque maintaining n normal to the plates in the SmA phase is not that of Frank elasticity over a micron-scale cell gap, d, but rather that of a local potential energy well [Friedel homeotropy (11)] that is typically stronger than the Frank elasticity by a large factor ~(d/l)2, where l is the nanometer-scale smectic layer thickness. This, combined with the condition that the smectic layering is essentially immovable at typical nematic Freedericksz threshold fields, suppresses molecular reorientation except under conditions of catastrophic layer reorganization at very high fields. In the SmZ phase, similar considerations explain why rotation of n out of the layer plane is suppressed.
The EO response to applied field can provide definitive information on how the SmZA layers organize upon cooling from the N phase. With n planar-aligned along the buffing direction, the layers will tend to order spontaneously either parallel to the plates (PA), or in “bookshelf” geometry (BK), as depicted in Fig. 1E, with the layers normal to the plates or nearly so. In the parallel geometry, an in-plane field readily induces in-plane reorientation of n, with, since E is normal to n, dielectric-like linear induced polarization response of the antiferroelectric structure driving twist deformation opposed by Frank elasticity. In the bookshelf geometry, on the other hand, molecular reorientation in the plane of the cell would force molecules out of the plane of the SmZ layers, a locally resisted, high energy deformation (28, 29). Suppression of the twist Freedericksz response in a particular cell therefore indicates that the layer normal y ∥ qM is essentially parallel to the plates.
Bookshelf and Chevron Layer Structures.
Further DTOM observations of the grown-by-cooling SmZA cell textures show that the layers typically adopt bookshelf (BK) geometry at the N–SmZA transition, with the layers normal to the plates. However, as illustrated in Fig. 2C, the layers contract slightly with decreasing temperature, and the layers respond by buckling (Fig. 2C), leading to the “chevron” (CH) variant of the lamellar bookshelf geometry (Fig. 4D and SI Appendix, Figs. S6–S10), in which the layer normal exhibits a step reorientation at the chevron interface (18–21), a planar surface parallel to the plates where the layer tilt δ changes sign. This structure maintains the locations of the contact lines of the layers at the surfaces and the uniform |δ| maintains the bulk layering pitch along y (in the plane of the cell), p = 2π/qM(TNZ), that was established on the surfaces and in the bulk at the N-SmZA transition.
The chevron structure and the associated zig-zag walls stand on the same footing as Friedel’s focal conics as being fundamental layering defects of smectics: focal conics result from step orientational discontinuities along lines, whereas chevrons result from step orientational discontinuities on sheets. Zig-zags are higher energy defects that appear when the requirement in buffed cells to satisfy the constraint of uniform orientation at the surfaces does not permit focal conics. Zig-zag walls have been extensively studied in systems undergoing the N–SmA–SmC (18) and N–SmC (20) transitions, where the layer shrinkage is due to tilt of the molecular long-axis away from the layer normal in the SmC phase. Typical SmC zig-zag wall textures are shown in SI Appendix, Figs. S6 and S7–S10. An analogous texture of zig-zag walls obtained on cooling DIO from the N to the SmZA phase is shown in Fig. 4. In this case, a small, in-plane square-wave electric field was applied along qM to alternate the direction of the nascent chevrons as they formed, giving a texture having an array of zig-zag walls.
We probed the polar characteristics of the as-grown chevron domains by applying small, in-plane electric fields both parallel to the layers (along n) and perpendicular to them (along qM), the latter achieved using fringing fields near the upper edge of one of the rectangular electrodes, as shown in Fig. 4 B and C. As can be seen, the field generated parallel to n by the upper electrode edge induces a difference in hue between the chevron domains on opposite sides of zig-zag walls, indicating field-induced director rotation that is uniform over the domain, but in opposite directions in domains with opposite chevron direction. This implies a polarization of the chevron domains as indicated by the white arrows in Fig. 4 B–D, leading to field-induced director reorientation near the chevron interface. Such polarity, arising from the layer structure, is consistent with the directionality of the chevron. On the other hand, where the applied field is normal to n (at the right electrode edge), there is no observable director reorientation, indicating that as the SmZA is grown, it has no net component of polarization parallel to n (cyan, magenta arrows), evidence for antiferroelectricity (AF).
Observation of AF of the SmZ A Phase.
Increasing the field to ~10X, the Freedericksz threshold in SI Appendix, Fig. S9F reorients the director toward being normal to the plates. If a field in this range is applied to the sandwich cell filled with zig-zags as in SI Appendix, Fig. S9 B and C or if bookshelf alignment is obtained by cooling from the N phase with a comparable field applied, the chevron layering defects, which rarely disappear on their own, can be forced out, and monodomain bookshelf alignment with the layers uniformly oriented normal to the plates achieved (SI Appendix, Fig. S9E). These monodomains also exhibit the splay-bend Freedericksz transition at low applied fields (SI Appendix, Fig. S9F), but at higher fields, they provide key evidence for the antiferroelectric nature of the SmZA phase. The I-V characteristics of a d = 100 µm, bookshelf indium-tin oxide (ITO)-sandwich cell with a 5 Hz, 50 V/100-µm amplitude triangle wave electric field applied along n during an N–SmZA–NF cooling scan are shown in Fig. 5A. At high temperatures, in the N phase (T > 84 °C), the current shows the usual cell capacitance step and an ion bump. In the SmZA phase (84 °C > T > 68 °C), new polarization peaks appear at the highest voltages, growing in area and with their peak center voltages VFA moving to lower magnitudes as T decreases. This is typical antiferroelectric behavior, the peaks marking the transition at finite voltage between the field induced ferroelectric (F) state at high fields, and the low-field antiferroelectric (A) state. The time integral of this current, Q = ∫i(t)dt gives the magnitude of polarization reversal in the antiferroelectic–ferroelectric transition, PFA(T) = Q(T)/A, where A is the electrode area. Both the polarization P(T) and the transition voltage, VFA(T) are plotted in Fig. 5B. The open squares give the first-order SmZA–NF phase boundary EFA(T) = VFA(T)/100µm in the (Ez–T) plane, and the open circles the polarization change PFA at this transition (SI Appendix, Fig. S13). The transition entropy ΔS decreases along this line to zero at the maximum Ez, according to the Clausius–Clapyron equation. Also shown is the product PFA(T) × VFA(T), proportional to the stabilization energy of the antiferroelectric state, which is maximum in the middle of the SmZA temperature range. These observations clearly establish that the SmZA phase is antiferroelectric. The SmZA to NF transition occurs between 69 °C and 68 °C, with the current transitioning to a single peak near V = 0, as seen in Fig. 5A, corresponding to the field-induced, Goldstone-mode reorientation of a macroscopic polarization, PF. The measured polarization in the NF phase is comparable to values previously obtained for DIO (7). Recently, Brown et al. have reported evidence for “local antiferroelectric” behavior in the M2 phase of DIO (31), and Nacke et al. have observed antiferroelectric polarization behavior in an intermediate phase between N and NF in a different fluorinated material (32).
Fig. 5.
(A) Measured I(t)-V(t) characteristics of DIO as a function of temperature, revealing a large field-induced shift in the SmZA–NF phase transition temperature for electric field Ez = V(t)/100µm, applied along n. The plot shows the current response I(t) of DIO in a d = 100 µm sandwich cell with bare ITO electrodes generating bookshelf layering, to a 5 Hz, 50 V amplitude triangle wave [V(t), white curve] applied during an N–SmZA–NF cooling scan. In the N phase (T > 84 °C), the current shows the expected cell capacitance step and an ion bump. In the SmZA phase (84 °C > T > 68 °C), new polarization peaks appear at the highest voltages, growing in area, with their peak center voltages VFA becoming smaller on cooling. This is typical antiferroelectric behavior, the peaks marking the transition at finite voltage between the field-induced ferroelectric (F) state and the equilibrium antiferroelectric (A) state. (B) Polarization values PFA(T) [open circles] and ferroelectric-to-antiferroelectric depolarization voltage VFA(T) [open squares] obtained from the depolarization current response (in the time range t < 0), where there is the least interference from ion flow). The corresponding field EFA(T) = VFA(T)/100µm. The open squares give the first order SmZA–NF phase boundary in the (Ez–T) plane, and the open circles the polarization change PFA at this transition. According to the Clausius–Clapyron equation, the transition entropy, ΔS, also decreases with increasing Ez along this line, to zero at the maximum Ez. The blue curve is proportional to the product PFA(T)VFA(T), and therefore proportional to the stabilization energy of the antiferroelectric state, which is maximum at T = 73 °C. The SmZA to NF transition occurs between T = 69 °C and 68 °C, with the current transforming to a single peak centered around V = 0, typical of field-induced, Goldstone-mode reorientation of macroscopic polarization, PF [red solid circles]. This peak area corresponds to a net polarization comparable to that measured in the NF phase of RM734 (7).
Bookshelf to-Parallel Layering Transformation.
If a low-frequency (20 mHz) in-plane triangle electric field is applied to the bookshelf SmZA cell and its amplitude increased above about 80 V/mm, ferroelectrohydrodynamic flow is generated (7), which disrupts the SmZA layering. The field-induced polarization tends to orient along the field direction, and in doing so restructures the layers in some areas, switching them from the bookshelf (BK) orientation to be parallel to the plates (PA), a process shown in SI Appendix, Figs. S11 and S15. When the field is reduced, the PA areas anneal into highly ordered monodomains filling the thickness of the cell with the SmZA layers parallel to the plates as sketched for the PA geometry in Fig. 1D. The director n remains along the buffing direction, so that the layers have effectively been rotated by 90° about n from the original bookshelf geometry. Near the N–SmZA transition, a typical cell treated in this way can simultaneously exhibit N, SmZA bookshelf, and SmZA parallel domains, as seen in SI Appendix, Figs. S11 and S15 B–D.
As in the N phase, the SmZA parallel domains respond to in-plane applied fields by exhibiting a field-induced azimuthal reorientation of n(x) about x, the normal to the cell plates. EO observations discussed in SI Appendix, sectionS8 with in-plane fields normal to n, are summarized in SI Appendix, Fig. S15 showing this orientational response to be a strictly dielectric-driven, twist Freedericksz transition as in the N phase. This confirms that the bulk polarization density along n is zero, i.e., that, at fields well below those required to induce the AF to F transition, the field-free SmZA phase has zero net ferroelectric polarization, and its linear response at small field is purely dielectric.
Biaxiality and the Modulation Structure of the SmZA Phase.
The simultaneous presence of bookshelf and parallel domains in planar aligned cells enables a measurement of features of the optical biaxiality of the SmZA phase, as detailed in SI Appendix, section S9 and Fig. S16.
Discussion: Modeling The N–SmZA–NF Phase Diagram
A persistent theme that has emerged from the century-long experimental and theoretical study of ferroelectric materials, both crystals and LCs, is that ferroelectricity (F) and AF go hand-in-hand, such that if one is found, the other will be in related materials, and even phase coexistence of the two in equilibrium in a given material is a common phenomenon. At the root of this behavior is the inherent frustration of dipole–dipole steric and electrostatic interactions, e.g., for the latter, dipole pairs arranged head-to-tail prefer relative parallel orientation, whereas side-by-side dipole pairs prefer relative antiparallel orientation. This frustration is almost a recipe for generating modulated AF phases of the SmZA type, having ferroelectric stripes of uniform P in the end-to-end direction, and antiferroelectric ordering of adjacent stripes in the side-by-side direction. Thus, in trying to understand the first AF phase of the ferroelectric nematic realm, we are moved to explore the basic physics of this F/AF relationship in pursuit of theoretical modeling of the SmZA.
Our observations identify the SmZA to be a kind of spontaneously modulated phase, of which there are many examples in soft materials (33, 34), LCs (26, 33–36), and ferroelectric (or ferromagnetic) solids (37–41). Such modulated phases have been explored by theory and simulation of equilibrium states in systems of molecular-scale (e.g., spin) variables, as well as with coarse-grained and Landau theoretical modeling of their mesoscopic averages as fields. The observation of both the N–NF and N–SmZA–NF phase sequences in the RM734/DIO binary phase diagram of Fig. 1 presents a new paraelectric–antiferroelectric–ferroelectric experimental landscape, one that ideally is viewed theoretically as a feature of some common ordering process. Among modulated states found in ferroelectric or ferromagnetic materials are the antipolar-ordered stripe phases, which appear in a wide variety of theoretical and experimental systems exhibiting the above-mentioned frustration arising from competing ferroelectric short- and antiferroelectric-long-range (power law) interactions in 2D and 3D (35, 36, 40–44). These modulated phases exhibit various periodic patterns of alternating polarity (44), and at finite temperature they are observed to order in patterns that can be either multi-unit-cellular or incommensurate, the latter having periods that are not rational fractions of the unit cell period of an underlying crystal lattice (42, 45).
An important theoretical approach to understanding the physics of such modulated systems are models where macroscopic polarization is calculated as a local average of molecular-scale dipoles of number density n, interacting with a Ising Hamiltonian,
, typically employing variables of binary sign, having local ferro-like interactions , and long-range dipole–dipole interactions (40, 41, 44–47). The frustration introduced by the anisotropy of the dipolar interaction, can stabilize transverse periodic modulation of the macroscopic polarization Pz(y), while at the same time suppressing spatial variations Pz(z) that separate opposite signs of polarization space charge ρP = ∇·P(r) and increase Coulomb energy.
We have found that a promising Landau-based direction for modeling the N–SmZA–NF system is the simple free energy shown in Eq. 1, proposed by Shiba and Ishibashi (48, 49), and applied to describe incommensurate AF in thiourea (SC(NH2)2) (50) and sodium nitrite (NaNO2) (51, 52):
[1] |
Here ( > 0, ,η > 0), (B < 0, C > 0), and the polarization inverse susceptibility is χP(T)−1 = A(T) = a(T−ToF). This free-energy functional can be obtained as a mean-field approximation to the ferroelectric Ising Hamiltonian with dipolar interactions, with the key feature that the—α(∂Pz/∂y)2 term is negative, which arises from the above mentioned F/AF frustration (53). FIS has been used extensively in the analysis of soft matter modulation instabilities (36) and a variety of structural and dynamic experiments on modulated antiferroelectric phases (54–62).
Of particular relevance here is its successful application to the paraelectric–incommensurate modulated antiferroelectric–ferroelectric (P–A–F) phase sequences in SC(NH2)2 (63–65) and NaNO2 (60, 61), as these exhibit key features similar to those of our observations of the binary N–NF /N–SmZA–NF phase behavior of Fig. 1, and of the SmZA phase. Applied to the RM734/DIO system, the dipolar Ising system and its FIS mean field appear to contain the essential physics as described below, with the transition a phenomenon of fluctuations in Pz, acting under competition between isotropic local ferroelectric ordering, and conditions of strong anisotropy and frustration introduced by dipole–dipole interactions.
Although the FIS model is limited to treating the 1D behavior of the single polarization component Pz(y), there is good reason to believe that it can be usefully applied to the SmZA. The experiments show that the SmZA is lamellar and antiferroelectric, with layers of alternating polarization P(y), as sketched in Fig. 2E(1,2). Some flexoelectrically induced tilt (splay) modulation θ(y) must be present in such a modulated phase by symmetry [as drawn in Fig. 2E(1)], and the coupling of splay and P has been invoked as a potential driving force for the transition to a modulated state (5, 66–69) However, as also indicated in Fig. 2E(1), and not considered in previous theory in the ferroelectric nematic realm, splay of P(r) generates polarization charge
ρ(r) = −∇·P(r), making the gray stripes in Fig. 2E(1) negatively charged and the white stripes positively charged, which, in turn, generates a field (in red), which acts to reduce θ(y) and the splay deformation. The strength of this effect is parameterized by the length ξP = √(εLCK/P2), which in the characteristic distance from defects or surfaces within which splay deformation of n can remain (69). At the lower temperatures in the SmZA range, with Frank elastic constant K ~ 2 pN, polarization within the SmZA layers P > 3 µC/cm2 (Fig. 5), and εLC ~ 5, we find ξP < 0.3 nm. The resulting condition, ξP << w, means that electrostatic interactions dominate the texture of the phase, overwhelming flexoelectricity and effectively forcing Py to zero, as in Fig. 2E(2). The only charge-free variations in the resulting Pz are ∂Pz/∂x and ∂Pz/∂y, the latter stabilized by the leading gradient term in the FIS, which enables the modulated antiferroelectric SmZA phase as an ordered state of a paraelectric free energy well, as follows:
(i) The RM734/DIO binary phase diagram vs. concentration—Applied to the N–SmZA–NF system, FIS describes the first-order mean-field transition into the NF phase as one from a paraelectric free energy minimum at Pz = 0 to a ferroelectric nematic minimum at finite Pz. This transition is controlled primarily by the A(T,c) = a(c)[T−ToF (c)], B(c), C(c) terms of FIS which, because it is first-order, have B(c) < 0 and C(c) > 0, and ToF is the temperature at which the paraelectric minimum of FIS disappears. The transition to the NF phase exhibits ideal mixing behavior (9), such that ToF(c) extrapolates linearly between the DIO and RM734 values, and the transition entropy ΔS = a(c)B(c)/C(c) and polarization in the NF phase at T = To are nearly constant, indicating that a, B, and C change little across the phase diagram (9). The divergence of the small-signal dielectric constant approaching the NF transition in the paraelectric minimum given by FIS is Δε ≈ ε|| ∝ χP(T) ∝ 1/ a(T−ToF). Both RM734 and DIO exhibit this behavior, as seen in SI Appendix, Fig. S14. In RM734 this divergence takes place in the paraelectric N phase, while in DIO it occurs entirely within the SmZA phase, manifest in Fig. 5 as the decreasing VFA required to induce the NF as it is approached. The first-order transition to the NF occurs where the paraelectric and ferroelectric minima of have the same energy, at TZF = ToF + B2/4aC = ToF + ~2 °C.
(ii) The NF is a single unmodulated phase across the binary phase diagram—This weak dependence of A, B, and C on c is consistent with NF being the same phase across the RM734/DIO phase diagram (9). Our observations indicate this to be the case, showing the NF being a homogeneous (unmodulated) fluid ferroelectric nematic having: a) a nematic director, n(r), which is locally free to reorient in any direction, subject only to the mutual, field-induced, and surface interactions (no underlying lattice as in solid-state ferroelectrics); b) polarization P(r), locally parallel to n(r), forming a coupled n(r)-P(r) orientation variable. P(r) does not differ observably in direction from n(r) in the NF phase, but is observed to vary spatially in magnitude, for example in pure polarization-reversal walls (7); c) Frank elastic behavior with the addition of polarization-generated space–charge interactions.
(iii) Modulation—The paraelectric Pz = 0 minimum of FIS exhibits a transition from the N to the SmZA phase where the spatially modulated Pz(y) = PqcosqMy about 〈Pz〉 = 0 appears, as in Fig. 2E(1,2). This modulation is driven by the energy-reducing – (∂Pz/∂y)2 term of FIS, its negative sign arising from the anisotropy of the dipole–dipole interaction in H. The finite-Pz free energy minimum of FIS describes the NF ferroelectric, low-temperature state, which in principle could have the magnitude Pz spatially modulated about . However, with minimized at FIS predicts that neither ferroelectric or antiferroelectric modulation of Pz(y) is possible in the NF for parameters that give the other features of the phase diagram, i.e., FIS predicts that the SmZA modulation will disappear in the NF, as observed (60, 66). Additionally, director orientational modulation has been proposed to appear in the NF phase, driven by the coupling between Pz and director splay (4, 5), such that where Pz is nonzero it induces preferred local splay, |∇·n|= ∂ny/∂y ∝Pz, which can stabilize polarization/splay modulated states. However, as in Fig. 2E(2), space charge expels splay and in the NF minimum, where the positive coeffiecient (∂Pz/∂y)2 term dominates, the uniform ferroelectric is the ground state.
(iv) Role of long-range dipolar interaction (Polarization space charge self energy)—Longitudinal polarization variations, ∂Pz/∂z, are also strongly opposed by polarization space charge, leading, as previously pointed out (7), to the polarization correlation fluctuations in nematic RM734 that are extremely extended in the z direction, and become more so upon approaching the transition (SI Appendix, Fig. S12). This is an important feature of the paranematic pretransitional behavior of RM734, that provides evidence for a combination of short-range ferroelectric interactions and long-range (dipole–dipole) electrostatic interactions establishing the correlations. The observation of the SmZA provides further evidence for this view. Aharony, in the detailed renormalization group analysis of polar ordering systems with Ising/dipolar interactions undergoing a paraelectric/ferroelectric phase transition, derived the polarization pair correlation function 〈Pz(0)Pz(r)〉 of the paraelectric N phase and its static structure factor, χP(q) = 〈Pz(q)Pz(q)*〉, shown in SI Appendix, Fig. S12 (70). This form was also obtained from a mean-field model (71). χP(q) is written in SI Appendix, Eq. S1 and plotted as a contour in SI Appendix, Fig. S12, showing that its principal feature is the energetic stretching of longitudinal fluctuations of Pz(r) (those in which ρP = ∂Pz/∂z is nonzero), in order to expel polarization space-charge ρP. These effects are evident in the polarization fluctuations of paranematic RM734, as shown in SI Appendix, Fig. S12, which also points out the additional important Aharony findings that the resulting anisotropy also makes the transition mean-field-like, rather than fluctuation-dominated with 3D Ising universality, and that the growth of correlation in y is the principal soft mode of the transition (72–74). The corresponding Landau mean field free energy description of the transition can be derived from the Ising/dipolar Hamiltonian (55, 75), and applying it to the SmZA phase results in the FIS model of Eq. 1: a) the SmZA is the extreme limit of polarization space-charge expulsion where the ordering of the Pz(r) into stripe or layer arrays having both 〈Pz(r)〉 = 0 and ∂Pz/∂z = 0 induces the complete expulsion of longitudinal space charge from the paraelectric minimum state. The FIS model, by treating Pz(y) as a function only of y, is by construction an embodiment of this limit. b) The Aharony polarization correlation function and the corresponding negative gradient term, — (∂Pz/∂y)2, in the Landau free energies like FIS, express the frustration arising from the tendencies for head-to-tail (side-by-side) dipoles to be parallel (antiparallel), and make possible the periodic modulation of Pz(y). This is predominately a Pz effect. A useful generalization of FIS would be explicit treatment of the z dependence of P.
(v) The SmZA is an ordered antiferroelectric state of the paraelectric minimum—Experimentally, the linearity vs. c of the NF upper phase boundary, and its transition entropy that is weakly varying across the phase diagram and large compared with that of the N–SmZA transition, suggests that the N and SmZA phases are both features of the paraelectric minimum and that the N–SmZA phase change is a transition within the paraelectric minimum. This is the FIS prediction: a paraelectric phase with a susceptibility for polarization diverging at the transition to the NF in which the typical pretransition scenario approaching the transition to a ferroelectric phase is interrupted by an instability to an ordered (modulated antiferroelectric) state, in our case the SmZA. This instability occurs because of the decreasing energy cost of finite Pz as the transition to the NF is approached.
The free energy of Eq.1 thus gives either an N–(TNF)–NF or an N–(TNZ)–SmZA–(TZF)–NF phase sequence, depending on parameters, where the transition to the NF phase is first-order, and the N – SmZA second-order or weakly first-order, as confirmed experimentally by the optical observations in Fig. 3F and by DSC (6) in DIO. An important feature of the N–SmZA transition is the temperature independence of qM, having a value qM ~ 0.07Å−1 even as a small pretransitional scattering bump in the N phase, and as scattering from a few remnant domains (Fig. 2H and SI Appendix, Figs. S4 and S5). In the FIS energy, the antiferroelectric incommensurate (SmZA) phase is produced by frustration between the term forcing slope ( < 0) and that resisting curvature ( > 0) of Pz(y), producing a nearly temperature-independent modulation wavevector qM. Upon cooling, near T ~ TNZ the initially weak modulation of Pz(y) is sinusoidal with wavevector , determined mainly by and , and independent of the magnitude of Pz. Eq.1 correctly predicts the temperature dependence of qM, which comprises the 5% decrease in qM with T through the SmZA range (Fig. 2H) contributed by the higher order η term. Specifically worth noting is the lack of a anomaly at high temperatures in the SmZA, a feature of Landau models in which qM is controlled directly by the coefficient of the P2 term (4, 5, 76).
The weak variation in qM(T) over the entire SmZA range in Fig. 2H is rather remarkable, given a) the decrease with T of the antiferroelectric–ferroelectric transition threshold field EFA(T) (Fig. 5B); b) the increase in the induced polarization PFA (Fig. 5B); and (iii) the strong variation in scattering intensity (Fig. 2G), which combined indicate a significant change in layer structure over the same range. The SAXS scans exhibit only the fundamental peak of the scattering electron density modulation, which would generally be interpreted in smectics to indicate a sinusoidal electron density modulation. However, although the peak intensity Ip(T) decreases for both large and small T (Fig. 2G), the layer structures must be very different near the two limits, in a way not describable by simply varying a sine-wave amplitude. If higher sinusoidal harmonics of Pz(y) are included in the FIS model, then Pz(y) evolves on cooling from sinusoidal to approaching a solitonic or square waveform for T ~ TZF, in which alternating +/− plateaus of maximum |Pz(y)| and high electron-density [based on the higher mass density of the NF relative to the N (77) are separated by narrow solitonic domain walls of width d(T) where the sign of Pz(y) changes and electron density is lower (56–59). The decrease in the non-resonant SAXS peak intensity as T approaches TZF in Fig. 2G will depend on d(T) approximately as Ip(T) ∝ Ipmaxsin2(π d(T)/wM). The question of the constancy of qM(T) has also been considered by Kats, who perceptively speculated that the M2 phase could be a modulated antiferroelectric (78) and suggested that the N to M2 transition would then be a “weak-crystallization” phenomenon, with qM(T) constrained as in weak-crystallization theory (79).
(vi) The temperature–electric field (T, Ez) phase diagram—Fig. 5A shows that the transition to the NF phase can be induced by an electric field applied along the director, with Ez(T)ZF, the field required to induce the transition, shown as a function of temperature in Fig. 5B. A similar field-induced shift in the transition to the ferroelectric phase has been observed in thiourea and sodium nitrite, and is a prominent feature of the P–A–F experimental phase diagrams obtained by Shiba and Ishibashi and the theoretical fits calculated from Eq.1 (60, 61) and shown in SI Appendix, Fig. S13. The (T, E) thiourea phase diagram (61), obtained using Eq.1 with B > 0 and C = 0, matches the optical (4) and neutron scattering (66) data for thiourea, and describes well the DIO (T, Ez) phase behavior with appropriate temperature and field scaling. A remarkable feature of both the IC data in NaNO2 and the predictions of Eq.1 is that in the entire IC area of the (T, E) phase diagram, the modulation wavevector is nearly constant at , changing by only 20% over the entire (T, Ez) range (60, 61). Application of the FIS model to the SmZA case would therefore suggest that in the NF – SmZA phase coexistence condition obtained upon reducing the field in Fig. 5, the SmZA domains have a nearly field-independent periodicity qM ~ 2π/(9 nm). This has not yet been tested experimentally.
(vii) Concentration dependence of the SmZA phase— (c) and (c) also determine the temperature range of the SmZA according to TNZ – TZF = 2/2a , (60, 61) suggesting that (c) changes sign at the SmZA endpoint at c ~ 40%, and that measurement of will be a key experiment for determining the dependence of these parameters on c. In any case, given that non-zero ∂Pz/∂y is favored in DIO, it may be that ∂Pz/∂y will generally at least not be strongly suppressed in the paranematic phase of ferroelectric nematic materials. Such a tendency is already evident in RM734, in the dominance of ∂Pz/∂y fluctuations in its paranematic phase (SI Appendix, Fig. S12).
Supplementary Material
Appendix 01 (PDF)
Acknowledgments
This work was supported by NSF Condensed Matter Physics Grants Division of Materials Research (DMR) 1710711 and DMR 2005170, and by Materials Research Science and Engineering Center Grant DMR 1420736. This research used the microfocus Soft Matter Interfaces beamline 12-ID of the National Synchrotron Light Source II, a U.S. Department of Energy (DOE) Office of Science User Facility operated for the DOE Office of Science by Brookhaven National Laboratory under Contract No. DE-SC0012704. We also acknowledge the use of beam line 11.0.1.2 of the Advanced Light Source at the Lawrence Berkeley National Laboratory supported by the Director of the Office of Science, Office of Basic Energy Sciences, of the U.S. DOE under Contract No. DE-AC02-05CH11231. Substantial portions of this work overlap with the PhD dissertation of X.C. (“New Phases and Phenomena of the Ferroelectric Nematic Realm,” University of Colorado, 2022).
Author contributions
X.C., V.M., G.F., M.Z., M.A.G., C.W., C.Z., L.R., and N.A.C. designed research; X.C., V.M., E.K., G.F., M.Z., M.A.G., C.W., C.Z., L.R., J.E.M., D.M.W., and N.A.C. performed research; E.K. contributed new reagents/analytic tools; X.C., V.M., M.A.G., L.R., and N.A.C. analyzed data; and X.C., J.E.M., D.M.W., and N.A.C. wrote the paper.
Competing interests
The authors have stock ownership to disclose, N.A.C., M.A.G., J.E.M., D.M.W. have partial ownership of Polaris Electro-Optics, the authors have patent filings to disclose, patent filing through University of Colorado Venture Partners, the authors have research support to disclose, research support from Polaris Electro-Optics.
Footnotes
Reviewers: O.D.L., Kent State University; and P.P., Universite Paris-Saclay.
Data, Materials, and Software Availability
Data and videos data have been deposited in an Open Science Framework repository named “Smectic ZA”. The link is https://osf.io/9fev2/ and the DOI is DOI: 10.17065/OSF.IO/9fev2/. (80)
Supporting Information
References
- 1.Debye P., Einige Resultate einer kinetischen Theorie der Isolatoren. Physikalische Zeitschrift 13, 97–100 (1912). [Google Scholar]
- 2.Born M., Über Anisotrope Flüssigkeiten. Versuch Einer Theorie Der Flüssigen Kristalle Und Des Elektrischen Kerr-Effekts In Flüssigkeiten. Sitzungsber. Preuss. Akad Wiss. 30, 614–650 (1916). [Google Scholar]
- 3.Mandle R. J., Cowling S. J., Goodby J. W., A nematic to nematic transformation exhibited by a rod-like liquid crystal. Phys. Chem. Chem. Phys. 19, 11429–11435 (2017), 10.1039/C7CP00456G. [DOI] [PubMed] [Google Scholar]
- 4.Mertelj A., et al. , Splay nematic phase. Phys. Rev. X 8, 041025 (2018), 10.1103/PhysRevX.8.041025. [DOI] [Google Scholar]
- 5.Sebastian N., et al. , Ferroelectric-ferroelastic phase transition in a nematic liquid crystal. Phys. Rev. Lett. 124, 037801 (2020), 10.1103/PhysRevLett.124.037801. [DOI] [PubMed] [Google Scholar]
- 6.Nishikawa H., et al. , A fluid liquid-crystal material with highly polar order. Adv. Mater. 29, 1702354 (2017), 10.1002/adma.201702354. [DOI] [PubMed] [Google Scholar]
- 7.Chen X., et al. , First-principles experimental demonstration of ferroelectricity in a thermotropic nematic liquid crystal: Spontaneous polar domains and striking electro-optics. Proc. Natl. Acad. Sci. U. S. A. 117, 14021–14031 (2020), 10.1073/pnas.2002290117. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 8.Li J., et al. , Development of ferroelectric nematic fluids with giant-ε dielectricity and nonlinear optical properties. Sci. Adv. 7, eabf5047 (2021), 10.1126/sciadv.abf5047. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 9.Chen X., et al. , Ideal mixing of paraelectric and ferroelectric nematic phases in liquid crystals of distinct molecular species. Liq. Cryst. 49, 1531–1544 (2022). [Google Scholar]
- 10.Mandle R. J., Cowling S. J., Goodby J. W.. Structural variants of RM734 in the design of splay nematic materials, Liq. Cryst. 48, 1780-1790 (2021), 10.1080/02678292.2021.1934740. [DOI] [Google Scholar]
- 11.Friedel G., Les états mésomorphes de la matière. Ann. Phys. (Paris) 18, 273–474 (1922), 10.1051/anphys/192209180273. [DOI] [Google Scholar]
- 12.Friedel G., Grandjean F., Cristaux liquides. Réponse à MM. G. Friedel, et F. Grandjean a M. O. Lehmann. Bull. de Minéral. 33, 300–317 (1910), 10.3406/bulmi.1910.3440. [DOI] [Google Scholar]
- 13.Sackmann H., Demus D., The polymorphism of liquid crystals. Mol. Cryst. Liq. Cryst. 2, 81–102 (1966), 10.1080/15421406608083062. [DOI] [Google Scholar]
- 14.Clark N. A., Antiferroelectric Smectic Ordering as a Prelude to the Ferroelectric Nematic: Introducing the Smectic Z Phase. Plenary Lecture 2, 18th International Conference on Ferroelectric Liquid Crystals (FLC2021, Ljubljana, Slovenia, 2021). [Google Scholar]
- 15.Nishikawa H., et al. , Nano-clustering mediates phase transitions in a diastereomerically-stabilized ferroelectric nematic system. Commun. Mater. 3, 89 (2022), 10.1038/s43246-022-00312-9. [DOI] [Google Scholar]
- 16.Safinya C. R., et al. , High-resolution x-ray study of a smectic-A – smectic-C phase transition. Phys. Rev. B 21, 4149–4153 (1980), 10.1103/PhysRevB.21.4149o. [DOI] [Google Scholar]
- 17.Rieker T. P., et al. , Chevron’ local layer structure in surface-stabilized ferroelectric smectic-C cells. Phys. Rev. Lett. 59, 2658–2661 (1987), 10.1103/PhysRevLett.59.2658. [DOI] [PubMed] [Google Scholar]
- 18.Clark N. A., Rieker T. P., Smectic C “chevron”, a planar liquid-crystal defect: Implications for the surface-stabilized ferroelectric liquid-crystal geometry. Phys. Rev. A. Rapid Commun. 37, 1053–1056 (1988), 10.1103/physreva.37.1053. [DOI] [PubMed] [Google Scholar]
- 19.Clark N. A., Rieker T. P., Maclennan J. E., Director and layer structure of SSFLC cells. Ferroelectrics 85, 79–97 (1988), 10.1080/00150198808007647. [DOI] [Google Scholar]
- 20.Rieker T. P., Clark N. A., Smith G. S., Safinya C. R., Layer and director structure in surface stabilized ferroelectric liquid crystal cells with non-planar boundary conditions. Liq. Cryst. 6, 565–576 (1989), 10.1080/02678298908034176. [DOI] [Google Scholar]
- 21.Rieker T. P., Clark N. A., Safinya C. R., Chevron layer structures in surface stabilized ferroelectric liquid crystal (SSFLC) cells filled with a material which exhibits the chiral nematic to smectic C* phase transition. Ferroelectrics 113, 245–256 (1991), 10.1080/00150199108014067. [DOI] [Google Scholar]
- 22.Lagerwall S. T., Ferroelectric and Antiferroelectric Liquid Crystals (Wiley VCH, Weinheim, ISBN 3-527-2983 1-2, 1999). [Google Scholar]
- 23.Takanishi Y., Ouchi Y., Takezoe H., Chevron layer structure in the smectic-A phase of 8CB. Jpn. J. Appl. Phys. Part 2-Lett. 28, L487–L489 (1989), 10.1143/JJAP.28.L487. [DOI] [Google Scholar]
- 24.Cheng X., et al. , Calamitic bolaamphiphiles with (semi)perfluorinated lateral chains: Polyphilic block molecules with new liquid crystalline phase structures. J. Am. Chem. Soc. 125, 10977–10996 (2003), 10.1021/ja036213g. [DOI] [PubMed] [Google Scholar]
- 25.Rädler J. O., Koltover I., Salditt T., Safinya C. R., Structure of DNA-cationic liposome complexes: dna intercalation in multilamellar membranes in distinct interhelical packing regimes. Science 275, 810–814 (1997), 10.1126/science.275.5301.810. [DOI] [PubMed] [Google Scholar]
- 26.Reddy R. A., Tschierske C., Bent-core liquid crystals: Polar order, superstructural chirality and spontaneous desymmetrisation in soft matter systems. J. Mater. Chem. 16, 907 (2006), 10.1039/B504400F. [DOI] [Google Scholar]
- 27.Selinger J. V., Director deformations, geometric frustration, and modulated phases in liquid crystals. Annual Review of Condensed Matter Physics 13, 49-71 (2022) 10.1146/annurev-conmatphys-031620-105712. [DOI] [Google Scholar]
- 28.Chen X., et al. , Polar in-plane surface orientation of a ferroelectric nematic liquid crystal: Polar monodomains and twisted state electro-optics. Proc. Natl. Acad. Sci. U.S.A. 118, e2104092118 (2021), 10.1073/pnas.2104092118. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 29.de Gennes P. G., An analogy between superconductors and smectics A. Solid State Commun. 10, 753–756 (1972), 10.1016/0038-1098(93)90291-T. [DOI] [Google Scholar]
- 30.Birecki H., Schaetzing R., Rondelez F., J. D., Litster light-scattering study of a smectic-A phase near the smectic-A – nematic transition. Phys. Rev. Lett. 36, 1376–1379 (1976), 10.1103/PhysRevLett.36.1376. [DOI] [Google Scholar]
- 31.Brown S., et al. , Multiple polar and non-polar nematic phases. Chemphyschem 22 2506–2510 (2021), 10.1002/cphc.202100644. [DOI] [PubMed] [Google Scholar]
- 32.Nacke P., Manabe A., Klasen-Memmer M., Bremer M., Giesselmann F., “New example of a ferroelectric nematic phase material” in Poster P2 in 18th International Conference on Ferroelectric Liquid Crystals (FLC2021) Ljubljana, Slovenia, (2021) Institute Josef Stefan, Ljubliana. [Google Scholar]
- 33.Seul M., Andelman D., Domain shapes and patterns: The phenomenology of modulated phases. Science 267, 476–483 (1995), 10.1126/science.267.5197.476. [DOI] [PubMed] [Google Scholar]
- 34.Marder M., Frisch H. L., Langer J. S., McConnell H. M., Theory of the intermediate rippled phase of phospholipid bilayers. Proc. Natl. Acad. Sci. U.S.A. 81, 6559–6561 (1984), 10.1073/pnas.81.20.6559. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Jacobs A. E., Goldner G., Mukamel D., Modulated structures in tilted chiral smectic films. Phys. Rev. A. 45, 5783–5788 (1992), 10.1103/PhysRevA.45.5783. [DOI] [PubMed] [Google Scholar]
- 36.Hinshaw G. A. Jr., Petschek R. G., Transitions and modulated phases in centrosymmetric ferroelectrics: Mean-field and renormalization-group predictions. Phys. Rev. B. 37, 2133–2155 (1988), 10.1103/PhysRevB.37.2133. [DOI] [PubMed] [Google Scholar]
- 37.Felix J. W., Mukamel D., Hornreich R. M., Novel class of continuous phase transitions to incommensurate structures. Phys. Rev. Lett. 57, 2180–2183 (1986), 10.1103/PhysRevLett.57.2180. [DOI] [PubMed] [Google Scholar]
- 38.De'Bell K., MacIsaac A.B., Whitehead J.P., Dipolar effects in magnetic thin films and quasi-two-dimensional systems. Rev. Mod. Phys. 72, 225–257 (2000), 10.1103/RevModPhys.72.225. [DOI] [Google Scholar]
- 39.Szabo G., Kadar G., Magnetic hysteresis in an ising-like dipole-dipole model. Phys. Rev. B. 58, 5584–5587 10.1103/PhysRevB.58.5584(1998), 10.1103/PhysRevB.58.5584. [DOI] [Google Scholar]
- 40.Cummins H. Z., Experimental studies of structurally incommensurate crystal phases. Phys. Rep. 185, 211–409 (1980), 10.1016/0370-1573(90)90058-A. [DOI] [Google Scholar]
- 41.Blinc R., Levanyuk A. P., Incommensurate Phases in Dielectrics: Volume 1 Fundamentals; Volume 2 Materials (Elsevier Science Publishers, Amsterdam, 1986). [Google Scholar]
- 42.MacIsaac A. B., Whitehead J. P., Robinson M. C., De’Bell K., Striped phases in two-dimensional dipolar ferromagnets. Phys. Rev. B. 51, 16033–16045 10.1103/PhysRevB.51.16033(1995), 10.1103/PhysRevB.51.16033. [DOI] [PubMed] [Google Scholar]
- 43.Grousson M., Tarjus G., Viot P., Phase diagram of an Ising model with long-range frustrating interactions: A theoretical analysis. Phys. Rev. E. 62, 7781–7792 (2000), 10.1103/physreve.62.7781. [DOI] [PubMed] [Google Scholar]
- 44.Grousson M., Tarjus G., Viot P., Monte Carlo study of the three-dimensional Coulomb frustrated Ising ferromagnet. Phys. Rev. E. 64, 036109 (2000), 10.1103/PhysRevE.64.036109. [DOI] [PubMed] [Google Scholar]
- 45.Hornreich R. M., Luban M., Shtrikman S., Critical behavior at the onset of k-Space instability on the l line. Phys. Rev. Lett. 35, 1678–1681 (1975), 10.1103/PhysRevLett.35.1678. [DOI] [Google Scholar]
- 46.Pighín S. A., Cannas S. A., Phase diagram of an Ising model for ultrathin magnetic films: Comparing mean field and Monte Carlo predictions. Phys. Rev. E. 75, 224433 (2007), 10.1103/PhysRevB.75.224433. [DOI] [Google Scholar]
- 47.Choi M. Y., Domain-wall pinning in the incommensurate phase of sodium nitrite. Phys. Rev. B. 37, 5874–5876 (1988), 10.1103/PhysRevB.37.5874. [DOI] [PubMed] [Google Scholar]
- 48.Ishibashi Y., Shiba H., Successive phase-transitions in ferroelectric NaNO2 and SC(NH2)2. J. Phys. Soc. Japan 45, 409–413 (1978), 10.1143/JPSJ.45.409. [DOI] [Google Scholar]
- 49.Ishibashi Y., Shiba H., Incommensurate-commensurate phase-transitions in ferroelectric substances. J. Phys. Soc. Japan 45, 1592–1599 10.1143/JPSJ.44.159(1978), 10.1143/JPSJ.44.159. [DOI] [Google Scholar]
- 50.Shiozaki Y., Satellite x-ray scattering and structural modulation of thiourea. Ferroelectrics 2, 245–260 (1978), 10.1080/00150197108234099. [DOI] [Google Scholar]
- 51.Tanisaki S., Shibuya I., Microdomain structure in paraelectric phase of NaNO2. J. Phys. Soc. Japan 16, 579 (1961), 10.1143/JPSJ.16.579. [DOI] [Google Scholar]
- 52.Yamada Y., Hoshino S., Shibuya I., Phase transition in NaNO2. J. Phys. Soc. Japan 18, 1594–1603 (1963), 10.1143/JPSJ.18.1594. [DOI] [Google Scholar]
- 53.Choi M. Y., Domain-wall pinning in the incommensurate phase of sodium nitrite. Phys. Rev. B 37, 5874–5876 (1988), 10.1103/PhysRevB.37.5874. [DOI] [PubMed] [Google Scholar]
- 54.Iwata M., Orihara H., Ishibashi Y., Phenomenological theory of the linear and nonlinear dielectric susceptibilities in the type-II incommensurate phase. J. Phys. Soc. Japan 67, 3130–3136 (1998), 10.1143/jpsj.67.3130. [DOI] [Google Scholar]
- 55.Berezovsky S. V., Soliton regime in the model with no Lifshitz invariant. arXiv:cond-mat/9909079 (1999) Accessed 6 July 2022.
- 56.Aramburu I., Madariaga G., Perez-Mato J. M., Phenomenological model for type-II incommensurate phases having a soliton regime: Thiourea case. Phys. Rev. B. 49, 802–814 10.1103/PhysRevB.49.802(1994), 10.1103/PhysRevB.49.802. [DOI] [PubMed] [Google Scholar]
- 57.Aramburu I., Madariaga G., Perez-Mato J. M., A structural viewpoint on the sine-Gordon equation in incommensurate phases. J. Phys. Condens. Matter 7, 6187–6196 (1995), 10.1088/0953-8984/7/31/003. [DOI] [Google Scholar]
- 58.Durand D., Denoyer F., Lefur D., Neutron diffraction study of sodium nitrite in an applied electric field. J. de Phys. 44, L207–L216 (1983), 10.1051/jphyslet:01983004405020700. [DOI] [Google Scholar]
- 59.Durand D., Denoyer F., Currat R., Lambert M., “Chapter 13 - Incommensurate phase in NaNO2” in Incommensurate Phases in Dielectrics: 2 Materials, Blinc R., Levanyuk A. P., Eds. (Elsevier Science Publishers, Amsterdam, 1986), 10.1016/B978-0-444-86970-8.50010-X. [DOI] [Google Scholar]
- 60.van Raaij G. H. F., van Bemmel K. J. H., Janssen T., Lattice models and Landau theory for type-II incommensurate crystals. Phys. Rev. B. 62, 3751–3765 (2000), 10.1103/PhysRevB.62.3751. [DOI] [Google Scholar]
- 61.Jacobs A. E., Grein C., Marsiglio F., Rippled commensurate state: A possible new type of incommensurate state. Phys. Rev. B. 29, 4179–4181 (1984), 10.1103/PhysRevB.29.4179. [DOI] [Google Scholar]
- 62.Jacobs A. E., Intrinsic domain-wall pinning and spatial chaos in continuum models of one-dimensionally incommensurate systems. Phys. Rev. B. 33, 6340–6345 (1986), 10.1103/PhysRevB.33.6340. [DOI] [PubMed] [Google Scholar]
- 63.Lederer P., Chaves C. M., Phase diagram of thiourea at atmospheric pressure under electric field: A theoretical analysis. J. de Phys. Lett. 42, L127–L130 (1981), 10.1051/jphyslet:01981004206012700. [DOI] [Google Scholar]
- 64.Jamet J. P., Electric field phase diagram of thiourea determined by optical birefringence. J. Phys. Lett. 1981, L123–L12 (1981), 10.1051/jphyslet:01981004206012300. [DOI] [Google Scholar]
- 65.Denoyer F., Currat R., Chapter 14 - Currat, Modulated Phases in Thiourea - Incommensurate phases in dielectrics: 2 Materials, Blinc R., Levanyuk A.P., Ed. (Elsevier Science Publishers, Amsterdam, 1986), 10.1016/B978-0-444-86970-8.50010-X. [DOI] [Google Scholar]
- 66.Rosseto M. P., Selinger J. V., Theory of the splay nematic phase: Single versus double splay, Phys. Rev. E. 101, 052707 (2020). [DOI] [PubMed] [Google Scholar]
- 67.Copic M., Mertelj A., Q-tensor model of twist-bend and splay nematic phases. Phys. Rev. E. 101, 022704 (2020), 10.1103/PhysRevE.101.022704. [DOI] [PubMed] [Google Scholar]
- 68.Sebastian N., Copic M., Mertelj A., Ferroelectric nematic liquid crystalline phases. Phys. Rev. E. 106, 021001 (2022), 10.1103/PhysRevE.106.021001. [DOI] [PubMed] [Google Scholar]
- 69.Zhuang Z., Maclennan J. E., Clark N. A., Device applications of ferroelectric liquid crystals: Im-portance of polarization charge interactions. Proc. Soc. Photo-Optical Instrum. Eng. 1080, 110–114 (1989), 10.1117/12.976407. [DOI] [Google Scholar]
- 70.Aharony A., Critical behavior of magnets with dipolar interactions. V. Uniaxial magnets in d-dimensions. Phys. Rev. B 8, 3363–3370 (1973), 10.1103/PhysRevB.8.3363. [DOI] [Google Scholar]
- 71.Larkin A. I., Khmel’nitskii D. E., Phase transition in uniaxial ferroelectrics. Soviet Phys. JETP 29, 1123–1128 (1969). [Google Scholar]
- 72.Kotzler J., Critical phenomena in dipolar magnets. J. Magn. Magn. Mater. 54–57, 649–654 (1986), 10.1016/0304-8853(86)90197-6. [DOI] [Google Scholar]
- 73.Als-Nielsen J., Experimental test of renormalization group theory on the uniaxial, dipolar coupled ferromagnet LiTbF4. Phys. Rev. Lett. 37, 1161–1164 (1976), 10.1103/PhysRevLett.37.1161. [DOI] [Google Scholar]
- 74.Als-Nielsen J., Birgeneau R. J., Mean field theory, the Ginzburg criterion, and marginal dimensionality of phase transitions. Am. J. Phys. 45, 554–560 (1977), 10.1119/1.11019. [DOI] [Google Scholar]
- 75.Belim S. M., Critical behavior of isotropic three dimensional systems with dipole-dipole interactions. J. Exp. Theoret. Phys. 116, 963–974 (2013), 10.1134/S1063776113050178. [DOI] [Google Scholar]
- 76.Rosseto M. P., Selinger J. V., Theory of the splay nematic phase: Single versus double splay nematic phase. Phys. Rev. E. 101, 052707 (2020). [DOI] [PubMed] [Google Scholar]
- 77.Mandle R. J., Sebastián N., Martinez-Perdiguero J., Mertelj A., On the molecular origins of the ferroelectric splay nematic phase. Nat. Commun. 12, 4962 (2021), 10.1038/s41467-021-25231-0. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 78.Kats E. I., Stability of the uniform ferroelectric nematic phase. Phys. Rev. E. 103, 012704 (2021), 10.1103/PhysRevE.103.012704. [DOI] [PubMed] [Google Scholar]
- 79.Kats E. I., Lebedev V. V., Muratov A. R., Weak crystallization theory. Phys. Rep. 228, 1–91 (1993), 10.1016/0370-1573(93)90119-X. [DOI] [Google Scholar]
- 80.Glaser M. A., MacLennan J. E., Chen X., Smectic ZA. Open Science Framework (OSF) https://osf.io/PFTGR/. Deposited 11 June 2022. [Google Scholar]
Associated Data
This section collects any data citations, data availability statements, or supplementary materials included in this article.
Supplementary Materials
Appendix 01 (PDF)
Data Availability Statement
Data and videos data have been deposited in an Open Science Framework repository named “Smectic ZA”. The link is https://osf.io/9fev2/ and the DOI is DOI: 10.17065/OSF.IO/9fev2/. (80)