Next Article in Journal
Novel Gluten-Free Bread with an Extract from Flaxseed By-Product: The Relationship between Water Replacement Level and Nutritional Value, Antioxidant Properties, and Sensory Quality
Next Article in Special Issue
Never Cared for What They Do: High Structural Stability of Guanine-Quadruplexes in the Presence of Strand-Break Damage
Previous Article in Journal
Evaluation of the Chemical Composition and Antioxidant Activity of Mulberry (Morus alba L.) Fruits from Different Varieties in China
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

A Ten-Year Perspective on Twist-Bend Nematic Materials

by
Richard J. Mandle
1,2
1
School of Physics and Astronomy, University of Leeds, Leeds LS2 9JT, UK
2
School of Chemistry, University of Leeds, Leeds LS2 9JT, UK
Molecules 2022, 27(9), 2689; https://doi.org/10.3390/molecules27092689
Submission received: 31 March 2022 / Revised: 20 April 2022 / Accepted: 20 April 2022 / Published: 21 April 2022
(This article belongs to the Special Issue Exclusive Feature Papers in Physical Chemistry)

Abstract

:
The discovery of the twist-bend nematic phase (NTB) is a milestone within the field of liquid crystals. The NTB phase has a helical structure, with a repeat length of a few nanometres, and is therefore chiral, even when formed by achiral molecules. The discovery and rush to understand the rich physics of the NTB phase has provided a fresh impetus to the design and characterisation of dimeric and oligomeric liquid crystalline materials. Now, ten years after the discovery of the NTB phase, we review developments in this area, focusing on how molecular features relate to the incidence of this phase, noting the progression from simple symmetrical dimeric materials towards complex oligomers, non-covalently bonded supramolecular systems.

1. Introduction

The term “liquid crystal” refers to a large number of states of matter that possess some degree of positional and/or orientational order, which is intermediate between isotropic liquids and crystalline solids. Much has been written about nematic liquid crystals and the twist-bend nematic phase, and so for the sake of brevity a short introduction to the topic suffices. The uniaxial nematic phase is arguably the simplest liquid crystal phase, with the constituent molecules (or particles) being, on average, oriented along a vector termed the director. Nematic liquid crystals are of special interest due to their role in display technology, and the discovery of new nematic ground states such as the twist-bend nematic is met with great enthusiasm. Biaxial nematics, in which the molecules are oriented along two orthogonal directors [1], are known to exist [2], but are outside the scope of this review. Similarly, although beyond the scope of this review, we note that nematic phases are almost exclusively apolar, that is, molecules orient both parallel and antiparallel to the director; very recently, the polar ferroelectric nematic phase has been shown to exist [3,4,5,6,7].
Introduction of chirality to a nematic liquid crystal leads to the formation of a chiral nematic phase, which has a helical superstructure. Dozov and Meyer independently suggested that bent shaped molecules could spontaneously form a heliconical nematic structure that is locally chiral, even when formed of achiral molecules [8]. This is termed the twist-bend nematic (NTB) phase, and was reported experimentally in a landmark work in 2011 [9]. The NTB phase has been described as the “structural link” between the uniaxial nematic phase and the helical chiral nematic mesophase [10]. A number of techniques have measured [10,11,12,13] (or inferred [14,15]) the repeat length of the NTB phase, with in situ resonant X-ray scattering being particularly noteworthy [16,17], and while the precise pitch length is material dependent, a value of around 10 nm is typical. This being said, other models have been proposed that merit further experimental investigation [18], but were outside the scope of this review.
The average conical angle between the mesogens and the helical axis can be measured by NMR [19] or birefringence [20], or by reconstructing the ODF using order parameter data from, for example, SAXS [21], polarised Raman spectroscopy [22], or NMR [23]. The conical angle of the NTB phase remains below the magic angle and the phase is uniaxial with positive birefringence, confirmed by conoscopic investigation [24]. Calorimetric studies show the NTB-N is typically first order, and close to tricritical [9,25,26,27,28]. The NTB phase has been shown to be strongly shear thinning; for the material KA(0.2), a 6-component mixture with a pitch length of 10.5 nm [29,30], it was shown that at low (<1 Pa) shear stress, the viscosity was ~1000× larger than the nematic phase. For the same material at high shear stress (>10 Pa), the viscosity of the NTB-phase dropped by two orders of magnitude as the helix underwent shear-induced realignment [31].
As with all mesophases, the formation of the NTB phase in a given material is intimately linked to molecular structure, and there has been significant effort in the design of new materials that exhibit the NTB phase [32]. The molecular structure of liquid crystalline dimers can be subdivided into distinct regions, as outlined in Figure 1B. In the simplest terms, a dimer consists of two rigid mesogenic units joined by a flexible spacer [33,34,35]. For a trimer, three mesogenic units are joined in a similar fashion, and so on. Today, in the region of 1000, materials are known to exhibit the NTB phase, and so in this review, we focused on systematic variations to key areas of molecular structure rather than making a futile attempt to cover all materials.

2. Materials

The CBnCB family are archetypal NTB materials, and a logical place to begin our review. These materials feature two cyanobiphenyl mesogenic units separated by n methylene units. As with all LC dimers, the CBnCB family displays a strong odd–even effect, with the even parity members displaying notably higher clearing points than those with odd spacer parity. With the exception of the shortest homologue (CB3CB), all odd parity CBnCBs displayed two nematic phases, the lower temperature nematic phase being identified as the NTB phase (Table 1). Even parity CBnCB materials displayed only a conventional nematic phase.
Due to their favourable working temperatures, the properties of the NTB phase of some members of the CBnCB family have been quite well explored. The helix pitch length of CB7CB was measured by freeze-fracture TEM by Chen et al. [11], who found a value of 8.3 nm. Later, Zhu et al. measured the NTB pitch length of CB9CB as a function of temperature by resonant X-ray scattering at the carbon K-edge [16]; the pitch length was largest close to the NTB–N transition (~9.8 nm), and decreased to around 8 nm with decreasing temperature. Yu and Wilson recently reported fully atomistic MD simulations of CB7CB, which yielded an NTB phase with a pitch length of 8.35 nm [41]. The conical tilt-angle within these simulations (~29°) agreed well with the experimental values [11,20].
The orientational order parameters of several members of the CBnCB family have been measured, with results from different methods generally being consistent with one another. For odd parity CBnCBs, the thermal evolution of orientational order within the nematic phase is unremarkable, however, at TNTB-N, there is a decrease in orientational ordering. This decrease results from the molecules tilting away from the helix, which manifests as a reduction in, or even negative value of <P4> [21,42]. This behaviour of the orientational order parameters has also been observed by NMR [43], and is reported to be consistent with the polar twisted nematic (NPT) model of the NTB phase. CB8CB, which has even spacer parity and does not show the NTB phase, displays unusually large nematic order parameters [22], as does CB10CB [43].
We next consider variations in the mesogenic units. Compound 11 belongs to a class of materials known as “PZP” dimers (P = phenyl, Z = carboxylate ester). The synthesis of these materials is trivial; the penultimate step is esterification of bis 1,9-(4-hydroxyphenyl)nonane, permitting the synthesis of a large number of variations in core structure via esterification. In Table 2, we present the transition temperatures of a set of PZP-9-PZP dimers with varying terminal groups. While cyano, isothiocyanato, and alkyl/alkoxy groups are found to support the formation of the NTB phase [44,45], various other polar units (nitro, fluoro, trifluoromethyl, pentafluorosulphanyl) render the resulting materials non-mesogenic [44]. For compound 15, the –NCS unit enables measurement of the NTB pitch length (~9 nm) using resonant X-ray scattering at the carbon K-edge and also the sulphur K-edge [46].
The synthetic flexibility afforded by the PZP dimers makes it possible to prepare dissymmetric materials such as those shown in Table 3 [47,48]. With a single phenyl 4-cyanobenzoate mesogenic unit, it is possible to obtain materials that display the NTB phase, even when the second mesogenic unit incorporates an ‘unfavourable’ terminal unit (e.g., NO2, SF5, etc.) [47,48]. This approach also lends itself to the synthesis of trimers, tetramers, and so on, as will be discussed later. A general trend in the materials presented in Table 3 is that the addition of additional fluorine atoms ortho to the terminal group leads to depressions in both TNTB-N and TN-Iso, mirroring the behaviour of calamitic materials [49]. One advantage enjoyed by unsymmetrical materials is that their melting points are generally lower than those of the corresponding symmetrical derivatives, which compensates for the more elaborate synthesis required.
Turning now to variations in the linking groups and spacer regions, while the NTB phase is most commonly associated with dimers incorporating a methylene spacer, it is also found in a large number of materials with imine linking units. The imine-linked (3-hydroxylphenyl) 4-alkylbenzoate dimers reported by Šepelj et al. (Table 4) showed a delicate balance between columnar, nematic, and NTB mesophases, with the specific phase type being dependent upon both the length of the central spacer as well as the peripheral alkyl chains [50].
Analogous in structure to the materials shown in Table 5, Šepelj et al. reported a family of imine linked phenyl 4-alkoxybenzoate dimers. Only one member displayed the twist-bend nematic phase, with m = 7 and n = 4, the majority of the materials displaying a B6 type mesophase.
We now explore the role of the chemical makeup of the central spacer beyond the methylene and imine systems already discussed. Archbold et al. reported a family of cyanobiphenyl dimers that are homologous to CB7CB in structure, having spacers of comparable length (seven methylene or equivalent units) but with different chemical makeup (Table 6) [54]. Materials incorporating two alkyne units were non-mesogenic. The onset temperature of the NTB phase was significantly reduced for the dipropyl ether spacer (55), while the NTB phase was absent for 61, which included a diethylegylcol spacer. The bis imine material 60 exhibited a direct NTB to isotropic transition temperature. Archbold et al. linked the observed transition temperatures to the average bend of the molecule, itself obtained as a probability weighted average of many conformers obtained with the rotational isomeric state (RIS) approximation, with the suggestion that an ‘optimal’ bend angle exists for the NTB phase, which leads to, inter alia, direct NTB–Iso transitions. This is exemplified by the high thermal stabilities of the NTB phases of the ketone-linked material (CBK-5-KCB) as well as the imine-linked material (CBI-3-ICB).
Refining the earlier approach of Archbold et al., Mandle and Goodby investigated the conformational preference of a series of homologues of CB9CB (Table 7) [55,56]. Again, the rotational isomeric state approximation was used to generate conformational libraries for each material; the average bend angle between the two mesogenic units then being calculated as a probability weighted average. Conformational ensembles were validated by comparison of average inter-proton distances with those obtained from 1H-1H NOESY NMR experiments. It is suggested that the stability of the NTB phase is related to the average bend angle, specifically, a high ratio of TNTB-N to TN-Iso is achieved by having an average bend angle in excess of 110°.
We now consider four families of related cyanobiphenyl dimers with varying linking groups and spacer lengths (Table 8) [59,60,61]. The chemical makeup of each family is evident from their names (e.g., the CBnOCB series feature one methylene and one ether linking unit, the CBnSCB series feature one methylene and one thioether linking unit, and so on). Further examples from each family are to be found in the given references. Generally, the transition temperatures of thioether containing materials are lower than the equivalent CBnOCB material, with the difference being most pronounced for shorter spacer lengths. A simple explanation suffices here, with a deeper understanding needing to draw on other relevant conformational effects. Consider that a methylene unit imparts a tetrahedral bond angle (109.5°), an ether gives an angle of 104.5°, but an arylthioether has a bond angle of ~90°. The influence of thioethers is therefore most pronounced for shorter chain lengths, whereas for longer spacers, they are somewhat offset. The thioether is somewhat unfavourable for the formation of the NTB phase as it tends to depress the average bend angle away from the apparent favoured value of >110 °C. However, for some materials (CBO7SCB, 67, and CBS7SCB, 66) the melting point is suppressed to such a degree that the materials are in the NTB phase at ambient temperature, which is a remarkable achievement that greatly simplifies experimentation. The helical pitch length of several compounds in Table 8 has been reported: CB6OCB ~10–15 nm [62], CBS7SCB 8.7 nm, CBO7SCB 18.4 nm, and CBO5SCB 14.8 nm [63]. Linking the NTB pitch length to molecular geometric parameters appears to be a logical future direction. Almost simultaneously with Arakawa et al. [61], the CBSnSCB and CBOnSCB materials were also synthesised and reported independently by Imrie et al. [64]. Imrie et al. also reported the pitch lengths of CBO5SCB (~8.9 nm), CBS7SB (~8.7 nm), and the mixed cyanoterphenyl/cyanobiphenyl material, CT6SCB (~9.7 nm); in all three cases, this corresponded to approximately four end-to-end molecular lengths.
Arakawa et al. subsequently demonstrated cyanobiphenyl dimers with mixed thioether/ketone linking units [65] (Table 9). Earlier, Archbold et al. found that ketone-linking units generated highly stable NTB phases due to their favourable bend angles (Table 6) [54]. Employing mixed ketone and thioether units showed a dramatic increase in the NTB onset temperature when compared to the equivalent materials employing either two thioethers (CBSnSCB), or one thioether and one other linking unit (CBnSCB, CBOnSCB, CBSnSCB). Again, a simple explanation suffices for the sake of this review: the ketone unit has a bond angle of ~120°, which offsets the unfavourable angle imposed by the thioether. Clearly, a detailed DFT study of the conformational landscape of these materials (and indeed, others) appears warranted, and presents one possible route to further understanding these intriguing materials.

3. Chiral NTB Materials

The NTB phase has been described as the “structural link” between the conventional nematic phase and the helical chiral nematic (cholesteric) phase [10]. As discussed, the NTB phase has a helical structure and when formed from achiral molecules, there is no preference for left- or right-handed helices. Our focus in this review was on molecular structure and the materials that generate the NTB phase, so our focus was on examples whereby chirality results from the molecular structure of the dimer itself, rather than systems in which chirality is introduced via an additive [66].
Gorecka et al. reported a number of ester-linked unsymmetrical dimers that incorporated cholesterol as a mesogenic unit; these materials are chiral, and thus they are the first reported chiral NTB materials. The length of the central spacer and its parity dictate the balance between exhibiting NTB (odd parity) or SmA (even parity) mesophases in these materials. The mesophase behaviour of these materials is more complex than shown in Table 10, with the materials exhibiting multiple nematic phases and/or blue phases. The NTB pitch length of 90 was measured to be 50 nm by in situ AFM, notably larger than that of CB7CB, but still about a few molecular lengths [24,67]. Later, the helical pitch length of 99 was measured by resonant carbon K-edge X-ray scattering and found to take a value of ~11 nm [68]. The pitch length of the chiral nematic phase was determined to be 224 nm by the same method.
Gorecka et al. also reported another family of cholesterol containing dimers that exhibited the NTB phase (Table 11), with some members also exhibiting a smectic phase of unknown structure (SmX) [24]. The pitch length of 104 in the NTB phase was measured by the resonant X-ray scattering method, with a temperature dependent value of 13.3–20.3 nm. Conversely, the chiral nematic pitch length for this material was measured by the same technique to be 220 nm [68]. Although not reported upon, the incorporation of an azo unit presumably enables isothermal NTB transitions in these materials (via photoisomerisation), as first reported by Paterson et al. [69].
Mandle and Goodby reported the synthesis of symmetrical LC dimers in which the central spacer is itself chiral (Table 12); starting from (R)-2-methylglutaric acid, four synthetic steps telescoped into two reactions afforded the key (R)-bis-1,5-(4-hydroxyphenyl)-2-methylpentane intermediate, which was elaborated using standard esterification protocols, affording compounds 107113. The measured helical twisting power of 107 was rather low (0.36 mm−1 wt%−1) due to the large degree of conformational freedom experienced by the lateral methyl unit. The NTB phase was only exhibited by materials whose mesogenic units had a large aspect ratio due to the unfavourable conformational effects of the lateral methyl unit within the central spacer. However, this also gave rise to the unusual SmA–NTB transition in compounds 111 and 112, which has previously only been observed for a small handful of materials [70].
Walker et al. reported a family of unsymmetrical dimers that are terminated by either butyl, racemic 2-methylbutyl, or (S)-2-methylbutyl chains (Table 13) [71]. The NTB–N transition temperature was marginally higher for chiral materials than the achiral analogues, in agreement with the theoretical predictions [72]. Notably, the NTB phase formed by 119 was found to be miscible with that of the achiral material CB6OCB (77).

4. Bent-Core Systems

Compared to dimers comprising rod-like (calamitic) mesogenic units, there are relatively few examples of bent-core liquid crystals that exhibit the NTB phase. A family of no symmetrical bent-core materials with a central phenyl piperazine group were prepared by Schroder et al. [73] with shorter homologues exhibiting two nematic phases, denoted as NX. Longer chain homologues exhibited the SmCP phase. Later, compound 126 was studied by FFTEM, and the lower temperature nematic was shown to be a NTB phase with a pitch length of 14 nm—larger than CB7CB, but of the order of a few molecular lengths [12]. Based on this, it seems probable that the ‘NX’ phase of compounds 124 and 125 is also the NTB phase.
Tamba et al. reported an ether-linked dimer, comprising a bent-core unit as well as a calamitic unit, which exhibited the twist-bend nematic phase as well as an unidentified ‘M2’ mesophase [74]. Homologues employing other spacer lengths (trimethyleneoxy or hexamethyleneoxy) or a dodecyloxy terminal chain in lieu of the nitrile unit employed in 134 do not exhibit the twist-bend nematic phase. To date, compound 134 (Figure 2), along with those in Table 14, are the only known examples of bent-core materials that exhibit the twist-bend nematic phase, although others have been previously suggested [75].

5. Beyond Dimers: Trimers, Tetramers and Oligomers

Now, our focus shifts beyond covalent dimers to equivalent trimers, tetramers, and so on [76,77,78,79,80,81,82,83,84,85]. We will focus first on materials of special note before considering the behaviour of families of materials. CB6OBA, a hydrogen bonded LC-trimer, was the first oligomeric material shown to exhibit the NTB phase (Table 15) [86]. The hexamethyleneoxy spacer of CB6OBA imparts a gross bent shape to the hydrogen bonded dimer, permitting the formation of the NTB phase, which is absent for the even parity homologue, CB5OBA. The terminal carboxylic acid group contains a hydrogen bond donor and acceptor, thus, both open and closed forms can be observed [87,88].
Wang et al. reported a symmetrical trimer that features two cyanobiphenyl units appended to a bent-core [13]; curiously, spacers in this system have even parity, with the requisite bent shape resulting from the 1,3-disubstituted phenyl ring employed within the bent-core unit. The pitch length of the NTB phase of 137 (Figure 3) was measured to be 19 nm using the FFTEM method, which was roughly four molecular lengths. X-ray scattering showed that the nematic phase is intercalated, with the d-spacing of the diffuse small angle peak being at ~1/3 of the molecular length. The orientational order parameters <P2> and <P4> were measured by X-ray scattering, both being found to decrease on entering the NTB phase. Replacement of a single cyanobiphenyl with a decyloxy chain was shown to eliminate the NTB phase [89].
In 2016, Mandle and Goodby reported a methylene linked tetramer comprising phenyl benzoate mesogenic units (Figure 4) [47,90]. The heliconical tilt angle of the NTB phase of 138 was estimated via the X-ray scattering method, and was found to be comparable to the parent dimer, compound 11 (Table 2) [42]. The synthetic strategy used to prepare this tetramer was further refined to deliver a twist-bend nematic hexamer, with six mesogenic units connected in a linear manner.
So far, all of the materials encountered have had a linear sequence of mesogenic units. The hexamer 139 features two trimers appended via a central heptamethylenedioxy spacer (Figure 5) [91]. The individual trimers are themselves mesogenic and display the NTB phase; however, the NTB–N transition occurs at a significantly higher temperature in the duplexed hexamer [91].
Jákli et al. reported a homologous family of 2′,3′-difluoroterphenyls [92], from the simple monomer (n = 0, 140, Table 16) to the homologous tetramer 143. Some homologous dimers, with varying terminal/spacer chain length, have also been reported [93,94,95], and a wealth of investigations have been performed on this family of materials [96,97,98].
The monomeric material displays only a nematic phase whereas those with two or more mesogenic units display nematic and NTB phases, with the transition temperature increasing as the number of mesogenic units is increased. So far, we have considered odd–even effects as being restricted to those resulting from the parity of the central spacer. However, for the ‘DTC5-C9’ family, Jákli et al. showed remarkable odd–even effects in birefringence, bend elastic constants, and X-ray scattering, which resulted from the number of mesogenic units [92].
Arakawa et al. reported two families of symmetrical liquid crystalline trimers featuring ether (CBOnOBOnOCB, Table 17) [99] and mixed ether/thioether (CBSnOBOnSCB, Table 18) [100] linking groups, but with varying spacer lengths. For a given chain length, the former family exhibited higher transition temperatures than the latter. In both families, materials of even parity displayed nematic and smectic A phases, whereas those of odd parity showed nematic and NTB phases.
Al-Janabi and Mandle reported a set of liquid crystalline trimers that incorporated various saturated hydrocarbon rings, isosteric with 1,4-disubstituted benzene (Table 19) [101]. Only the 2,6-cuneane material did not exhibit the NTB phase, and this was attributed to the unfavourable bend angles imposed by this non-linear motif. For the materials that exhibited the NTB phase (X = 1,4-benzene, 1,4-cyclohexane, and 1,4-cubane), the heliconical tilt angle was found to be effectively independent of the chemical makeup of the central ring.
The relationship between the stability of the NTB phase and oligomer shape was studied in detail for a family of cyanobiphenyl/benzylideneaniline oligomers (Table 20) [102]. Due to the explored variations in the composition and parity of the spacer units, neighbouring mesogenic units within this family can have both ‘bent’ or ‘linear’ configurations. While all materials exhibit the NTB phase, the heliconical pitch length has a strong dependence on the gross molecular shape; the all bent material 166 has three odd-parity spacers and a pitch of 7 nm, the bent-linear-bent tetramer 170 has a pitch of 12 nm, and the linear-bent-linear tetramer 172 has a pitch length of ~17 nm.

6. Supramolecular NTB Materials

We now consider supramolecular LC dimers that result from non-covalent bonds, both achiral and chiral, which exhibit the NTB phase. The hydrogen bonded-LC trimer CB6OBA was the first reported supramolecular NTB material, however, as the complex incorporates two identical molecules, there is no scope for tunability of structure. Walker et al. demonstrated a remarkable pair of supramolecular LC dimers that incorporates dissimilar hydrogen bond acceptors and donors. The parent 4-methoxybiphenyl/stilbazole dimer (173) is non-mesogenic; complexation with either 4-butoxy- or 4-pentyloxybenzoic acid affords the isolable supramolecular complexes shown in Table 21, both of which exhibit the NTB phase [103]. While many materials are known to exhibit transitions from the NTB phase to tilted (SmC, vide infra) or heliconical (SmCTB) [104] phases, complex 175 is unusual in that it exhibits a transition from the NTB phase to an orthogonal smectic A phase, as do compounds 111 and 112.
This design strategy was also used to deliver chiral supramolecular NTB materials by use of the cyanobiphenyl/stilbazole system with an appropriate chiral benzoic acid [105]. The transition temperatures for this chiral supramolecular complex are somewhat lower than those of the linear analogue (shown in Table 22), and this depression of transition temperatures by branched alkyl chains is a general phenomenon in LC dimers. Resonant soft X-ray scattering at the carbon K-edge was used to measure the pitch length of 176, which was found to take a temperature dependent value of 8.1–8.4 nm, or around two complex lengths.
Walker et al. subsequently demonstrated the CB6OCB:nOS series of materials, utilising the benzoic acid/stilbazole system [106]. A weak odd–even effect was seen for both the NTB–N and N–Iso transition temperatures. Homologues with longer terminal chain length (n ≥ 4) also exhibited smectic C phases. For homologues with longer chains still (n ≥ 8, not shown in Table 23), the NTB phase was absent, instead, the materials showed SmA and SmAB phases [106].
In the same paper, further elaboration of the cyanobiphenyl/stilbazole system to unsymmetrical supramolecular trimers comprising CB6OBA and a stilbazole dimer was reported (Table 24). Although the two materials differed in their melting point, the NTB–N and N–Iso transitions were only slightly different. We note that, although examples are presently limited to hydrogen bonded systems, there is no obvious reason why other types of non-covalent interactions could not be employed in the design of NTB materials (e.g., halogen bonds [107]), and presents a logical avenue for future research.

7. Summary and Outlook

The decade since the experimental discovery of the NTB phase by Cestari et al. has seen a resurgence of interest in liquid crystalline dimers and oligomers. Recently, there has been a notable move from symmetrical methylene-linked dimers to more complex forms: chiral systems, photoresponsive dimers, supermolecular materials, higher and non-linear oligomers, and polymers. The recent development of room temperature materials greatly facilitates the exploration of the rich physics of these systems.
General design principles for NTB materials are always evolving; with the ability to tune molecular bend/shape through synthetic chemistry, the ability to prepare oligomers, and supramolecular systems, it is possible to obtain twist-bend nematic materials through rational design rather than through ad hoc experimentation. This being said, the majority of twist-bend nematic materials follow the tried-and-tested formula of end-to-end appended mesogenic units, with only a handful of bent-core materials and a single non-linear oligomer falling outside of this description. It is interesting to speculate as to whether more unconventional molecular geometries (e.g., lambda shaped trimers, mixed rod-disk architectures) are capable of supporting twist-bend nematic order.
Different materials display rather different helix pitches that themselves evolve differently over temperature; an understanding of this from a molecular perspective is currently elusive, but appears a reasonable proposition for future work. The potential for incorporating stimuli responsive groups, coupled with the remarkable physical properties of this phase of matter, suggests that interest in this area will continue for some time.

Funding

R.J.M. acknowledges funding from UK Research and Innovation (UKRI).

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Freiser, M.J. Ordered states of a nematic liquid. Phys. Rev. Lett. 1970, 24, 1041–1043. [Google Scholar] [CrossRef]
  2. Yu, L.J.; Saupe, A. Observation of a biaxial nematic phase in potassium laurate-1-decanol-water mixtures. Phys. Rev. Lett. 1980, 45, 1000–1003. [Google Scholar] [CrossRef] [Green Version]
  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. 2017, 19, 11429–11435. [Google Scholar] [CrossRef] [Green Version]
  4. Nishikawa, H.; Shiroshita, K.; Higuchi, H.; Okumura, Y.; Haseba, Y.; Yamamoto, S.I.; Sago, K.; Kikuchi, H. A fluid liquid-crystal material with highly polar order. Adv. Mater. 2017, 29, 1702354. [Google Scholar] [CrossRef]
  5. Mertelj, A.; Cmok, L.; Sebastián, N.; Mandle, R.J.; Parker, R.R.; Whitwood, A.C.; Goodby, J.W.; Čopič, M.J.P.R.X. Splay nematic phase. Phys. Rev. X 2018, 8, 041025. [Google Scholar] [CrossRef] [Green Version]
  6. Sebastian, N.; Cmok, L.; Mandle, R.J.; de la Fuente, M.R.; Drevensek Olenik, I.; Copic, M.; Mertelj, A. Ferroelectric-ferroelastic phase transition in a nematic liquid crystal. Phys. Rev. Lett. 2020, 124, 037801. [Google Scholar] [CrossRef] [Green Version]
  7. Chen, X.; Korblova, E.; Dong, D.; Wei, X.; Shao, R.; Radzihovsky, L.; Glaser, M.A.; Maclennan, J.E.; Bedrov, D.; Walba, D.M.; et al. First-principles experimental demonstration of ferroelectricity in a thermotropic nematic liquid crystal: Polar domains and striking electro-optics. Proc. Natl. Acad. Sci. USA 2020, 117, 14021–14031. [Google Scholar] [CrossRef]
  8. Dozov, I. On the spontaneous symmetry breaking in the mesophases of achiral banana-shaped molecules. Europhys. Lett. 2001, 56, 247–253. [Google Scholar] [CrossRef]
  9. Cestari, M.; Diez-Berart, S.; Dunmur, D.A.; Ferrarini, A.; de la Fuente, M.R.; Jackson, D.J.; Lopez, D.O.; Luckhurst, G.R.; Perez-Jubindo, M.A.; Richardson, R.M.; et al. Phase behavior and properties of the liquid-crystal dimer 1″,7″-bis(4-cyanobiphenyl-4′-yl) heptane: A twist-bend nematic liquid crystal. Phys. Rev. E Stat. Nonlinear Biol. Soft Matter Phys. 2011, 84, 031704. [Google Scholar] [CrossRef] [Green Version]
  10. Borshch, V.; Kim, Y.K.; Xiang, J.; Gao, M.; Jakli, A.; Panov, V.P.; Vij, J.K.; Imrie, C.T.; Tamba, M.G.; Mehl, G.H.; et al. Nematic twist-bend phase with nanoscale modulation of molecular orientation. Nat. Commun. 2013, 4, 2635. [Google Scholar] [CrossRef] [Green Version]
  11. Chen, D.; Porada, J.H.; Hooper, J.B.; Klittnick, A.; Shen, Y.; Tuchband, M.R.; Korblova, E.; Bedrov, D.; Walba, D.M.; Glaser, M.A.; et al. Chiral heliconical ground state of nanoscale pitch in a nematic liquid crystal of achiral molecular dimers. Proc. Natl. Acad. Sci. USA 2013, 110, 15931–15936. [Google Scholar] [CrossRef] [Green Version]
  12. Chen, D.; Nakata, M.; Shao, R.; Tuchband, M.R.; Shuai, M.; Baumeister, U.; Weissflog, W.; Walba, D.M.; Glaser, M.A.; Maclennan, J.E.; et al. Twist-bend heliconical chiral nematic liquid crystal phase of an achiral rigid bent-core mesogen. Phys. Rev. E Stat. Nonlinear Biol. Soft Matter Phys. 2014, 89, 022506. [Google Scholar] [CrossRef] [Green Version]
  13. Wang, Y.; Singh, G.; Agra-Kooijman, D.M.; Gao, M.; Bisoyi, H.K.; Xue, C.M.; Fisch, M.R.; Kumar, S.; Li, Q. Room temperature heliconical twist-bend nematic liquid crystal. Crystengcomm 2015, 17, 2778–2782. [Google Scholar] [CrossRef]
  14. Beguin, L.; Emsley, J.W.; Lelli, M.; Lesage, A.; Luckhurst, G.R.; Timimi, B.A.; Zimmermann, H. The chirality of a twist-bend nematic phase identified by nmr spectroscopy. J. Phys. Chem. B 2012, 116, 7940–7951. [Google Scholar] [CrossRef]
  15. Meyer, C.; Luckhurst, G.R.; Dozov, I. Flexoelectrically driven electroclinic effect in the twist-bend nematic phase of achiral molecules with bent shapes. Phys. Rev. Lett. 2013, 111, 067801. [Google Scholar] [CrossRef]
  16. Zhu, C.; Tuchband, M.R.; Young, A.; Shuai, M.; Scarbrough, A.; Walba, D.M.; Maclennan, J.E.; Wang, C.; Hexemer, A.; Clark, N.A. Resonant carbon k-edge soft x-ray scattering from lattice-free heliconical molecular ordering: Soft dilative elasticity of the twist-bend liquid crystal phase. Phys. Rev. Lett. 2016, 116, 147803. [Google Scholar] [CrossRef] [Green Version]
  17. Stevenson, W.D.; Ahmed, Z.; Zeng, X.B.; Welch, C.; Ungar, G.; Mehl, G.H. Molecular organization in the twist–bend nematic phase by resonant x-ray scattering at the se k-edge and by saxs, waxs and gixrd. Phys. Chem. Chem. Phys. 2017, 19, 13449–13454. [Google Scholar] [CrossRef] [Green Version]
  18. Samulski, E.T.; Vanakaras, A.G.; Photinos, D.J. The twist bend nematic: A case of mistaken identity. Liq. Cryst. 2020, 47, 2092–2097. [Google Scholar] [CrossRef]
  19. Jokisaari, J.P.; Luckhurst, G.R.; Timimi, B.A.; Zhu, J.F.; Zimmermann, H. Twist-bend nematic phase of the liquid crystal dimer cb7cb: Orientational order and conical angle determined by xe-129 and h-2 nmr spectroscopy. Liq. Cryst. 2015, 42, 708–721. [Google Scholar]
  20. Meyer, C.; Luckhurst, G.R.; Dozov, I. The temperature dependence of the heliconical tilt angle in the twist-bend nematic phase of the odd dimer cb7cb. J. Mater. Chem. C 2015, 3, 318–328. [Google Scholar] [CrossRef]
  21. Singh, G.; Fu, J.X.; Agra-Kooijman, D.M.; Song, J.K.; Vengatesan, M.R.; Srinivasarao, M.; Fisch, M.R.; Kumar, S. X-ray and raman scattering study of orientational order in nematic and heliconical nematic liquid crystals. Phys. Rev. E 2016, 94, 060701. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Zhang, Z.P.; Panov, V.P.; Nagaraj, M.; Mandle, R.J.; Goodby, J.W.; Luckhurst, G.R.; Jones, J.C.; Gleeson, H.F. Raman scattering studies of order parameters in liquid crystalline dimers exhibiting the nematic and twist-bend nematic phases. J. Mater. Chem. C 2015, 3, 10007–10016. [Google Scholar] [CrossRef] [Green Version]
  23. Barnes, P.J.; Douglass, A.G.; Heeks, S.K.; Luckhurst, G.R. An enhanced odd even effect of liquid-crystal dimers orientational order in the alpha,omega-bis(4′-cyanobiphenyl-4-yl)alkanes. Liq. Cryst. 1993, 13, 603–613. [Google Scholar] [CrossRef]
  24. Gorecka, E.; Vaupotic, N.; Zep, A.; Pociecha, D.; Yoshioka, J.; Yamamoto, J.; Takezoe, H. A twist-bend nematic (n-tb) phase of chiral materials. Angew. Chem. Int. Ed. 2015, 54, 10155–10159. [Google Scholar] [CrossRef]
  25. Tripathi, C.S.P.; Losada-Perez, P.; Glorieux, C.; Kohlmeier, A.; Tamba, M.G.; Mehl, G.H.; Leys, J. Nematic-nematic phase transition in the liquid crystal dimer cbc9cb and its mixtures with 5cb: A high-resolution adiabatic scanning calorimetric study. Phys. Rev. E 2011, 84, 041707. [Google Scholar] [CrossRef] [Green Version]
  26. Mandle, R.J.; Davis, E.J.; Archbold, C.T.; Cowling, S.J.; Goodby, J.W. Microscopy studies of the nematic NTB phase of 1,11-di-(1″-cyanobiphenyl-4-yl)undecane. J. Mater. Chem. C 2014, 2, 556–566. [Google Scholar] [CrossRef]
  27. Robles-Hernandez, B.; Sebastian, N.; de la Fuente, M.R.; Lopez, D.O.; Diez-Berart, S.; Salud, J.; Ros, M.B.; Dunmur, D.A.; Luckhurst, G.R.; Timimi, B.A. Twist, tilt, and orientational order at the nematic to twist-bend nematic phase transition of 1″,9″-bis(4-cyanobiphenyl-4′-yl) nonane: A dielectric, h-2 nmr, and calorimetric study. Phys. Rev. E 2015, 92, 062505. [Google Scholar] [CrossRef] [Green Version]
  28. Sebastian, N.; Lopez, D.O.; Robles-Hernandez, B.; de la Fuente, M.R.; Salud, J.; Perez-Jubindo, M.A.; Dunmur, D.A.; Luckhurst, G.R.; Jackson, D.J. Dielectric, calorimetric and mesophase properties of 1″-(2′,4-difluorobiphenyl-4′-yloxy)-9″-(4-cyanobiphenyl-4′-yloxy) nonane: An odd liquid crystal dimer with a monotropic mesophase having the characteristics of a twist-bend nematic phase. Phys. Chem. Chem. Phys. 2014, 16, 21391–21406. [Google Scholar] [CrossRef]
  29. Adlem, K.; Copic, M.; Luckhurst, G.R.; Mertelj, A.; Parri, O.; Richardson, R.M.; Snow, B.D.; Timimi, B.A.; Tuffin, R.P.; Wilkes, D. Chemically induced twist-bend nematic liquid crystals, liquid crystal dimers, and negative elastic constants. Phys. Rev. E Stat. Nonlinear Biol. Soft Matter Phys. 2013, 88, 022503. [Google Scholar] [CrossRef]
  30. Ribeiro de Almeida, R.R.; Zhang, C.; Parri, O.; Sprunt, S.N.; Jákli, A. Nanostructure and dielectric properties of a twist-bend nematic liquid crystal mixture. Liq. Cryst. 2014, 41, 1661–1667. [Google Scholar] [CrossRef]
  31. Salili, S.M.; Kim, C.; Sprunt, S.; Gleeson, J.T.; Parri, O.; Jakli, A. Flow properties of a twist-bend nematic liquid crystal. RSC Adv. 2014, 4, 57419–57423. [Google Scholar] [CrossRef]
  32. Mandle, R.J. The dependency of twist-bend nematic liquid crystals on molecular structure: A progression from dimers to trimers, oligomers and polymers. Soft Matter 2016, 12, 7883–7901. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Imrie, C.T.; Henderson, P.A. Liquid crystal dimers and oligomers. Curr. Opin. Colloid Interface Sci. 2002, 7, 298–311. [Google Scholar] [CrossRef]
  34. Reddy, R.A.; Tschierske, C. Bent-core liquid crystals: Polar order, superstructural chirality and spontaneous desymmetrisation in soft matter systems. J. Mater. Chem. 2006, 16, 907–961. [Google Scholar] [CrossRef]
  35. Henderson, P.A.; Seddon, J.M.; Imrie, C.T. Methylene- and ether-linked liquid crystal dimers ii. Effects of mesogenic linking unit and terminal chain length. Liq. Cryst. 2005, 32, 1499–1513. [Google Scholar] [CrossRef]
  36. Ramou, E.; Welch, C.; Hussey, J.; Ahmed, Z.; Karahaliou, P.K.; Mehl, G.H. The induction of the Ntb phase in mixtures. Liq. Cryst. 2018, 45, 1929–1935. [Google Scholar] [CrossRef]
  37. Emsley, J.W.; Luckhurst, G.R.; Shilstone, G.N. The orientational order of nematogenic molecules with a flexible core—A dramatic odd even effect. Mol. Phys. 1984, 53, 1023–1028. [Google Scholar] [CrossRef]
  38. Emsley, J.W.; Luckhurst, G.R.; Shilstone, G.N.; Sage, I. The preparation and properties of the alpha,omega-bis(4,4′-cyanobiphenyloxy)alkanes—Nematogenic molecules with a flexible core. Mol. Cryst. Liq. Cryst. 1984, 102, 223–233. [Google Scholar] [CrossRef]
  39. Panov, V.P.; Vij, J.K.; Mehl, G.H. Twist-bend nematic phase in cyanobiphenyls and difluoroterphenyls bimesogens. Liq. Cryst. 2017, 44, 147–159. [Google Scholar]
  40. Miglioli, I.; Bacchiocchi, C.; Arcioni, A.; Kohlmeier, A.; Mehl, G.H.; Zannoni, C. Director configuration in the twist-bend nematic phase of cb11cb. J. Mater. Chem. C 2016, 4, 9887–9896. [Google Scholar] [CrossRef]
  41. Yu, G.; Wilson, M.R. All-atom simulations of bent liquid crystal dimers: The twist-bend nematic phase and insights into conformational chirality. Soft Matter 2022, 18, 3087–3096. [Google Scholar] [CrossRef] [PubMed]
  42. Mandle, R.J.; Goodby, J.W. Order parameters, orientational distribution functions and heliconical tilt angles of oligomeric liquid crystals. Phys. Chem. Chem. Phys. 2019, 21, 6839–6843. [Google Scholar] [CrossRef] [PubMed]
  43. Heist, L.M.; Samulski, E.T.; Welch, C.; Ahmed, Z.; Mehl, G.H.; Vanakaras, A.G.; Photinos, D.J. Probing molecular ordering in the nematic phases of para-linked bimesogen dimers through nmr studies of flexible prochiral solutes. Liq. Cryst. 2020, 47, 2058–2073. [Google Scholar] [CrossRef] [Green Version]
  44. Mandle, R.J.; Davis, E.J.; Voll, C.C.A.; Archbold, C.T.; Goodby, J.W.; Cowling, S.J. The relationship between molecular structure and the incidence of the NTB phase. Liq. Cryst. 2015, 42, 688–703. [Google Scholar]
  45. Mandle, R.J.; Davis, E.J.; Archbold, C.T.; Voll, C.C.; Andrews, J.L.; Cowling, S.J.; Goodby, J.W. Apolar bimesogens and the incidence of the twist-bend nematic phase. Chem. Eur. J. 2015, 21, 8158–8167. [Google Scholar] [CrossRef]
  46. Salamończyk, M.; Mandle, R.J.; Makal, A.; Liebman-Peláez, A.; Feng, J.; Goodby, J.W.; Zhu, C. Double helical structure of the twist-bend nematic phase investigated by resonant x-ray scattering at the carbon and sulfur k-edges. Soft Matter 2018, 14, 9760–9763. [Google Scholar] [CrossRef]
  47. Mandle, R.J.; Goodby, J.W. Progression from nano to macro science in soft matter systems: Dimers to trimers and oligomers in twist-bend liquid crystals. RSC Adv. 2016, 6, 34885–34893. [Google Scholar] [CrossRef] [Green Version]
  48. Mandle, R.J.; Goodby, J.W. Does topology dictate the incidence of the twist-bend phase? Insights gained from novel unsymmetrical bimesogens. Chem.–A Eur. J. 2016, 22, 18456–18464. [Google Scholar] [CrossRef]
  49. Hird, M. Fluorinated liquid crystals—Properties and applications. Chem. Soc. Rev. 2007, 36, 2070–2095. [Google Scholar] [CrossRef]
  50. Sepelj, M.; Lesac, A.; Baumeister, U.; Diele, S.; Bruce, D.W.; Hamersak, Z. Dimeric salicylaldimine-based mesogens with flexible spacers: Parity-dependent mesomorphism. Chem. Mater. 2006, 18, 2050–2058. [Google Scholar] [CrossRef]
  51. Ivsic, T.; Vinkovic, M.; Baumeister, U.; Mikleusevic, A.; Lesac, A. Retraction. Milestone in the NTB phase investigation and beyond: Direct insight into molecular self-assembly. Soft Matter 2015, 11, 6716. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  52. Ivsic, T.; Vinkovic, M.; Baumeister, U.; Mikleusevic, A.; Lesac, A. Towards understanding the NTB phase: A combined experimental, computational and spectroscopic study. RSC Adv. 2016, 6, 5000–5007. [Google Scholar] [CrossRef]
  53. Sepelj, M.; Lesac, A.; Baumeister, U.; Diele, S.; Nguyen, H.L.; Bruce, D.W. Intercalated liquid-crystalline phases formed by symmetric dimers with an alpha,omega-diiminoalkylene spacer. J. Mater. Chem. 2007, 17, 1154–1165. [Google Scholar] [CrossRef]
  54. Archbold, C.T.; Mandle, R.J.; Andrews, J.L.; Cowling, S.J.; Goodby, J.W. Conformational landscapes of bimesogenic compounds and their implications for the formation of modulated nematic phases. Liq. Cryst. 2017, 44, 2079–2088. [Google Scholar] [CrossRef] [Green Version]
  55. Mandle, R.J.; Archbold, C.T.; Sarju, J.P.; Andrews, J.L.; Goodby, J.W. The dependency of nematic and twist-bend mesophase formation on bend angle. Sci. Rep. 2016, 6, 36682. [Google Scholar] [CrossRef] [Green Version]
  56. Mandle, R.J.; Goodby, J.W. Molecular flexibility and bend in semi-rigid liquid crystals: Implications for the heliconical nematic ground state. Chem.–A Eur. J. 2019, 25, 14454–14459. [Google Scholar] [CrossRef]
  57. Arakawa, Y.; Komatsu, K.; Tsuji, H. Twist-bend nematic liquid crystals based on thioether linkage. New J. Chem. 2019, 43, 6786–6793. [Google Scholar] [CrossRef]
  58. Arakawa, Y.; Tsuji, H. Selenium-linked liquid crystal dimers for twist-bend nematogens. J. Mol. Liq. 2019, 289, 111097. [Google Scholar] [CrossRef]
  59. Paterson, D.A.; Gao, M.; Kim, Y.K.; Jamali, A.; Finley, K.L.; Robles-Hernandez, B.; Diez-Berart, S.; Salud, J.; de la Fuente, M.R.; Timimi, B.A.; et al. Understanding the twist-bend nematic phase: The characterisation of 1-(4-cyanobiphenyl-4′-yloxy)-6-(4-cyanobiphenyl-4′-yl)hexane (CB6OCB) and comparison with CB7CB. Soft Matter 2016, 12, 6827–6840. [Google Scholar] [CrossRef] [Green Version]
  60. Paterson, D.A.; Abberley, J.P.; Harrison, W.T.A.; Storey, J.M.D.; Imrie, C.T. Cyanobiphenyl-based liquid crystal dimers and the twist-bend nematic phase. Liq. Cryst. 2017, 44, 127–146. [Google Scholar] [CrossRef]
  61. Arakawa, Y.; Komatsu, K.; Shiba, T.; Tsuji, H. Methylene- and thioether-linked cyanobiphenyl-based liquid crystal dimers cbnscb exhibiting room temperature twist-bend nematic phases and glasses. Mater. Adv. 2021, 2, 1760–1773. [Google Scholar] [CrossRef]
  62. Tuchband, M.R.; Paterson, D.A.; Salamończyk, M.; Norman, V.A.; Scarbrough, A.N.; Forsyth, E.; Garcia, E.; Wang, C.; Storey, J.M.D.; Walba, D.M.; et al. Distinct differences in the nanoscale behaviors of the twist–bend liquid crystal phase of a flexible linear trimer and homologous dimer. Proc. Natl. Acad. Sci. USA 2019, 116, 10698–10704. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  63. Cao, Y.; Feng, J.; Nallapaneni, A.; Arakawa, Y.; Zhao, K.; Zhang, H.; Mehl, G.H.; Zhu, C.; Liu, F. Deciphering helix assembly in the heliconical nematic phase via tender resonant x-ray scattering. J. Mater. Chem. C 2021, 9, 10020–10028. [Google Scholar] [CrossRef]
  64. Cruickshank, E.; Salamończyk, M.; Pociecha, D.; Strachan, G.J.; Storey, J.M.D.; Wang, C.; Feng, J.; Zhu, C.; Gorecka, E.; Imrie, C.T. Sulfur-linked cyanobiphenyl-based liquid crystal dimers and the twist-bend nematic phase. Liq. Cryst. 2019, 46, 1595–1609. [Google Scholar] [CrossRef] [Green Version]
  65. Arakawa, Y.; Komatsu, K.; Ishida, Y.; Igawa, K.; Tsuji, H. Carbonyl- and thioether-linked cyanobiphenyl-based liquid crystal dimers exhibiting twist-bend nematic phases. Tetrahedron 2021, 81, 131870. [Google Scholar] [CrossRef]
  66. Archbold, C.T.; Davis, E.J.; Mandle, R.J.; Cowling, S.J.; Goodby, J.W. Chiral dopants and the twist-bend nematic phase–induction of novel mesomorphic behaviour in an apolar bimesogen. Soft Matter 2015, 11, 7547–7557. [Google Scholar] [CrossRef]
  67. Zep, A.; Aya, S.; Aihara, K.; Ema, K.; Pociecha, D.; Madrak, K.; Bernatowicz, P.; Takezoe, H.; Gorecka, E. Multiple nematic phases observed in chiral mesogenic dimers. J. Mater. Chem. C 2013, 1, 46–49. [Google Scholar] [CrossRef]
  68. Salamończyk, M.; Vaupotič, N.; Pociecha, D.; Wang, C.; Zhu, C.; Gorecka, E. Structure of nanoscale-pitch helical phases: Blue phase and twist-bend nematic phase resolved by resonant soft x-ray scattering. Soft Matter 2017, 13, 6694–6699. [Google Scholar] [CrossRef] [Green Version]
  69. Paterson, D.A.; Xiang, J.; Singh, G.; Walker, R.; Agra-Kooijman, D.M.; Martínez-Felipe, A.; Gao, M.; Storey, J.M.D.; Kumar, S.; Lavrentovich, O.D.; et al. Reversible isothermal twist–bend nematic–nematic phase transition driven by the photoisomerization of an azobenzene-based nonsymmetric liquid crystal dimer. J. Am. Chem. Soc. 2016, 138, 5283–5289. [Google Scholar] [CrossRef]
  70. Mandle, R.J.; Davis, E.J.; Lobato, S.A.; Vol, C.C.A.; Cowling, S.J.; Goodby, J.W. Synthesis and characterisation of an unsymmetrical, ether-linked, fluorinated bimesogen exhibiting a new polymorphism containing the NTB or ‘twist-bend’ phase. Phys. Chem. Chem. Phys. 2014, 16, 6907–6915. [Google Scholar] [CrossRef]
  71. Walker, R.; Pociecha, D.; Storey, J.M.D.; Gorecka, E.; Imrie, C.T. The chiral twist-bend nematic phase (N*TB). Chem.–A Eur. J. 2019, 25, 13329–13335. [Google Scholar] [CrossRef] [PubMed]
  72. Meyer, C. Nematic twist-bend phase under external constraints. Liq. Cryst. 2016, 43, 2144–2162. [Google Scholar] [CrossRef]
  73. Schroder, M.W.; Diele, S.; Pelzl, G.; Dunemann, U.; Kresse, H.; Weissflog, W. Different nematic phases and a switchable smcp phase formed by homologues of a new class of asymmetric bent-core mesogens. J. Mater. Chem. 2003, 13, 1877–1882. [Google Scholar] [CrossRef]
  74. Tamba, M.G.; Baumeister, U.; Pelzl, G.; Weissflog, W. Banana-calamitic dimers: Further variations of the bent-core mesogenic unit. Ferroelectrics 2014, 468, 52–76. [Google Scholar] [CrossRef]
  75. Gortz, V.; Southern, C.; Roberts, N.W.; Gleeson, H.F.; Goodby, J.W. Unusual properties of a bent-core liquid-crystalline fluid. Soft Matter 2009, 5, 463–471. [Google Scholar] [CrossRef]
  76. Yoshizawa, A. Unconventional liquid crystal oligomers with a hierarchical structure. J. Mater Chem. 2008, 18, 2877–2889. [Google Scholar] [CrossRef]
  77. Keller, P. Synthesis of new mesomorphic polyesters by polymerization of bifunctional monomers. Mol. Cryst. Liq. Cryst. 1985, 123, 247–256. [Google Scholar] [CrossRef]
  78. Imrie, C.T.; Luckhurst, G.R. Liquid crystal trimers. The synthesis and characterisation of the4,4′-bis[ω-(4-cyanobiphenyl-4′-yloxy)alkoxy]biphenyls. J. Mater. Chem. 1998, 8, 1339–1343. [Google Scholar] [CrossRef]
  79. Andersch, J.; Diele, S.; Lose, D.; Tschierske, C. Synthesis and liquid crystalline properties of novel laterally connected trimesogens and tetramesogens. Liq. Cryst. 1996, 21, 103–113. [Google Scholar] [CrossRef]
  80. Kreuder, W.; Ringsdorf, H.; Herrmannschonherr, O.; Wendorff, J.H. The wheel of mainz as a liquid-crystal—Structural variation and mesophase properties of trimeric discotic compounds. Angew. Chem.-Int. Ed. Engl. 1987, 26, 1249–1252. [Google Scholar] [CrossRef]
  81. Griffin, A.C.; Sullivan, S.L.; Hughes, W.E. Effect of molecular structure on mesomorphism xxi. Monodisperse tetrameric model compounds for liquid crystalline polymers. Liq. Cryst. 1989, 4, 677–684. [Google Scholar] [CrossRef]
  82. Imrie, C.T.; Stewart, D.; Remy, C.; Christie, D.W.; Hamley, I.W.; Harding, R. Liquid crystal tetramers. J. Mater Chem. 1999, 9, 2321–2325. [Google Scholar] [CrossRef]
  83. Yelamaggad, C.V.; Nagamani, S.A.; Hiremath, U.S.; Rao, D.S.S.; Prasad, S.K. The first examples of monodispersive liquid crystalline tetramers possessing four non-identical anisometric segments. Liq. Cryst. 2002, 29, 231–236. [Google Scholar] [CrossRef]
  84. Henderson, P.A.; Imrie, C.T. Semiflexible liquid crystalline tetramers as models of structurally analogous copolymers. Macromolecules 2005, 38, 3307–3311. [Google Scholar] [CrossRef]
  85. Yelamaggad, C.V.; Achalkumar, A.S.; Rao, D.S.S.; Prasad, S.K. Monodispersive linear supermolecules stabilizing unusual fluid layered phases. Org. Lett. 2007, 9, 2641–2644. [Google Scholar] [CrossRef]
  86. Jansze, S.M.; Martinez-Felipe, A.; Storey, J.M.; Marcelis, A.T.; Imrie, C.T. A twist-bend nematic phase driven by hydrogen bonding. Angew. Chem. Int. Ed. Engl. 2015, 54, 643–646. [Google Scholar] [CrossRef]
  87. Martinez-Felipe, A.; Imrie, C.T. The role of hydrogen bonding in the phase behaviour of supramolecular liquid crystal dimers. J. Mol. Struct. 2015, 1100, 429–437. [Google Scholar] [CrossRef]
  88. Paterson, D.A.; Martinez-Felipe, A.; Jansze, S.M.; Marcelis, A.T.M.; Storey, J.M.D.; Imrie, C.T. New insights into the liquid crystal behaviour of hydrogen-bonded mixtures provided by temperature-dependent ftir spectroscopy. Liq. Cryst. 2015, 42, 928–939. [Google Scholar] [CrossRef]
  89. Wang, Y.; Yoon, H.G.; Bisoyi, H.K.; Kumar, S.; Li, Q. Hybrid rod-like and bent-core liquid crystal dimers exhibiting biaxial smectic a and nematic phases. J. Mater. Chem. 2012, 22, 20363–20367. [Google Scholar] [CrossRef]
  90. Mandle, R.J.; Goodby, J.W. A liquid crystalline oligomer exhibiting nematic and twist-bend nematic mesophases. Chemphyschem 2016, 17, 967–970. [Google Scholar] [CrossRef]
  91. Mandle, R.J.; Goodby, J.W. A nanohelicoidal nematic liquid crystal formed by a non-linear duplexed hexamer. Angew. Chem. Int. Ed. Engl. 2018, 57, 7096–7100. [Google Scholar] [CrossRef] [PubMed]
  92. Saha, R.; Babakhanova, G.; Parsouzi, Z.; Rajabi, M.; Gyawali, P.; Welch, C.; Mehl, G.H.; Gleeson, J.; Lavrentovich, O.D.; Sprunt, S.; et al. Oligomeric odd–even effect in liquid crystals. Mater. Horiz. 2019, 6, 1905–1912. [Google Scholar] [CrossRef]
  93. Panov, V.P.; Nagaraj, M.; Vij, J.K.; Panarin, Y.P.; Kohlmeier, A.; Tamba, M.G.; Lewis, R.A.; Mehl, G.H. Spontaneous periodic deformations in nonchiral planar-aligned bimesogens with a nematic-nematic transition and a negative elastic constant. Phys. Rev. Lett. 2010, 105, 167801. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  94. Panov, V.P.; Balachandran, R.; Nagaraj, M.; Vij, J.K.; Tamba, M.G.; Kohlmeier, A.; Mehl, G.H. Microsecond linear optical response in the unusual nematic phase of achiral bimesogens. Appl. Phys. Lett. 2011, 99, 261903. [Google Scholar] [CrossRef] [Green Version]
  95. Sebastian, N.; Tamba, M.G.; Stannarius, R.; de la Fuente, M.R.; Salamonczyk, M.; Cukrov, G.; Gleeson, J.; Sprunt, S.; Jakli, A.; Welch, C.; et al. Mesophase structure and behaviour in bulk and restricted geometry of a dimeric compound exhibiting a nematic-nematic transition. Phys. Chem. Chem. Phys. 2016, 18, 19299–19308. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Merkel, K.; Loska, B.; Welch, C.; Mehl, G.H.; Kocot, A. The role of intermolecular interactions in stabilizing the structure of the nematic twist-bend phase. RSC Adv. 2021, 11, 2917–2925. [Google Scholar] [CrossRef] [PubMed]
  97. Cukrov, G.; Mosaddeghian Golestani, Y.; Xiang, J.; Nastishin, Y.A.; Ahmed, Z.; Welch, C.; Mehl, G.H.; Lavrentovich, O.D. Comparative analysis of anisotropic material properties of uniaxial nematics formed by flexible dimers and rod-like monomers. Liq. Cryst. 2017, 44, 219–231. [Google Scholar] [CrossRef]
  98. Saha, R.; Feng, C.; Welch, C.; Mehl, G.H.; Feng, J.; Zhu, C.; Gleeson, J.; Sprunt, S.; Jákli, A. The interplay between spatial and heliconical orientational order in twist-bend nematic materials. Phys. Chem. Chem. Phys. 2021, 23, 4055–4063. [Google Scholar] [CrossRef]
  99. Arakawa, Y.; Komatsu, K.; Shiba, T.; Tsuji, H. Phase behaviors of classic liquid crystal dimers and trimers: Alternate induction of smectic and twist-bend nematic phases depending on spacer parity for liquid crystal trimers. J. Mol. Liq. 2021, 326, 115319. [Google Scholar] [CrossRef]
  100. Arakawa, Y.; Komatsu, K.; Ishida, Y.; Shiba, T.; Tsuji, H. Thioether-linked liquid crystal trimers: Odd–even effects of spacers and the influence of thioether bonds on phase behavior. Materials 2022, 15, 1709. [Google Scholar] [CrossRef]
  101. Al-Janabi, A.; Mandle, R.J. Utilising saturated hydrocarbon isosteres of para benzene in the design of twist-bend nematic liquid crystals. ChemPhysChem 2020, 21, 697–701. [Google Scholar] [CrossRef] [PubMed]
  102. Majewska, M.M.; Forsyth, E.; Pociecha, D.; Wang, C.; Storey, J.M.D.; Imrie, C.T.; Gorecka, E. Controlling spontaneous chirality in achiral materials: Liquid crystal oligomers and the heliconical twist-bend nematic phase. Chem. Commun. 2022. [Google Scholar] [CrossRef] [PubMed]
  103. Walker, R.; Pociecha, D.; Abberley, J.P.; Martinez-Felipe, A.; Paterson, D.A.; Forsyth, E.; Lawrence, G.B.; Henderson, P.A.; Storey, J.M.D.; Gorecka, E.; et al. Spontaneous chirality through mixing achiral components: A twist-bend nematic phase driven by hydrogen-bonding between unlike components. Chem. Commun. 2018, 54, 3383–3386. [Google Scholar] [CrossRef] [PubMed]
  104. Abberley, J.P.; Killah, R.; Walker, R.; Storey, J.M.D.; Imrie, C.T.; Salamończyk, M.; Zhu, C.; Gorecka, E.; Pociecha, D. Heliconical smectic phases formed by achiral molecules. Nat. Commun. 2018, 9, 228. [Google Scholar] [CrossRef] [Green Version]
  105. Walker, R.; Pociecha, D.; Salamończyk, M.; Storey, J.M.D.; Gorecka, E.; Imrie, C.T. Supramolecular liquid crystals exhibiting a chiral twist-bend nematic phase. Mater. Adv. 2020, 1, 1622–1630. [Google Scholar] [CrossRef]
  106. Walker, R.; Pociecha, D.; Martinez-Felipe, A.; Storey, J.M.; Gorecka, E.; Imrie, C.T. Twist-bend nematogenic supramolecular dimers and trimers formed by hydrogen bonding. Crystals 2020, 10, 175. [Google Scholar] [CrossRef] [Green Version]
  107. Nguyen, H.L.; Horton, P.N.; Hursthouse, M.B.; Legon, A.C.; Bruce, D.W. Halogen bonding:  A new interaction for liquid crystal formation. J. Am. Chem. Soc. 2004, 126, 16–17. [Google Scholar] [CrossRef]
Figure 1. (A) Cartoon depiction of the heliconical director precession in the twist-bend nematic phase; mesogenic units are shown as cylinders, and are coloured according to their position along the helix axis. (B) The general structure of terminally appended liquid crystalline dimers and oligomers, subdivided into regions of interest to this review. For each subdivision example, chemical fragments that have been utilised in NTB materials are given. (C) Schematic depiction of the relationship between dimers, trimers, and tetramers in terms of their subunit composition.
Figure 1. (A) Cartoon depiction of the heliconical director precession in the twist-bend nematic phase; mesogenic units are shown as cylinders, and are coloured according to their position along the helix axis. (B) The general structure of terminally appended liquid crystalline dimers and oligomers, subdivided into regions of interest to this review. For each subdivision example, chemical fragments that have been utilised in NTB materials are given. (C) Schematic depiction of the relationship between dimers, trimers, and tetramers in terms of their subunit composition.
Molecules 27 02689 g001
Figure 2. Chemical structure and transition temperatures (°C) of the hybrid bent-core/calamitic dimer reported by Tamba et al. [74]. Phase transitions are presented in parenthesis are monotropic.
Figure 2. Chemical structure and transition temperatures (°C) of the hybrid bent-core/calamitic dimer reported by Tamba et al. [74]. Phase transitions are presented in parenthesis are monotropic.
Molecules 27 02689 g002
Figure 3. Chemical structure and transition temperatures (°C) of the hybrid bent-core/calamitic trimer (137) reported by Wang et al. [13]. The melting point was not reported.
Figure 3. Chemical structure and transition temperatures (°C) of the hybrid bent-core/calamitic trimer (137) reported by Wang et al. [13]. The melting point was not reported.
Molecules 27 02689 g003
Figure 4. Chemical structure and transition temperatures (°C) of the linear tetramer reported by Mandle and Goodby.
Figure 4. Chemical structure and transition temperatures (°C) of the linear tetramer reported by Mandle and Goodby.
Molecules 27 02689 g004
Figure 5. Chemical structure and transition temperatures (°C) of the non-linear hexamer reported by Mandle and Goodby.
Figure 5. Chemical structure and transition temperatures (°C) of the non-linear hexamer reported by Mandle and Goodby.
Molecules 27 02689 g005
Table 1. Transition temperatures (TA-B, °C) of the CBnCB family of materials [9,23,26,36,37,38,39,40].
Table 1. Transition temperatures (TA-B, °C) of the CBnCB family of materials [9,23,26,36,37,38,39,40].
Molecules 27 02689 i001
No.NamenTMPTNTB-NTN-Iso
1CB3CB3142.1--
2CB5CB51509297
3CB6CB6183-230
4CB7CB7102104.5116
5CB8CB8175-195.9
6CB9CB983105.0119.8
7CB10CB10140-174.1
8CB11CB1199.9108.6125.5
9CB12CB12139-157
10CB13CB13106105122
Table 2. Transition temperatures (TA-B, °C) of some symmetric (PZP)-9 materials with varying terminal unit [44,45].
Table 2. Transition temperatures (TA-B, °C) of some symmetric (PZP)-9 materials with varying terminal unit [44,45].
Molecules 27 02689 i002
No.XTMPTNTB-NTN-Iso
11–CN157.6114.5146.6
12–NO2105.4--
13–F97.6--
14–CF3102.4--
15–NCS97.7103.7127.4
16–SF5123.0--
17–C5H1172.858.866.2
18–OC5H1171.876.086.6
Table 3. Transition temperatures (TA-B, °C) of some dissymmetric (PZP)-9 materials with varying terminal unit [47,48].
Table 3. Transition temperatures (TA-B, °C) of some dissymmetric (PZP)-9 materials with varying terminal unit [47,48].
Molecules 27 02689 i003
No.XTMPTNTB-NTN-Iso
11 Molecules 27 02689 i004157.6114.5146.6
19 Molecules 27 02689 i005112.695.0120.7
20 Molecules 27 02689 i006115.4100.5124.9
21 Molecules 27 02689 i00792.678.897.6
22 Molecules 27 02689 i00886.278.295.9
23 Molecules 27 02689 i00983.263.879.0
24 Molecules 27 02689 i01093.846.060.0
25 Molecules 27 02689 i011110.069.678.3
26 Molecules 27 02689 i01295.6100.0123.8
27 Molecules 27 02689 i013102.261.272.8
28 Molecules 27 02689 i01488.780.795.1
29 Molecules 27 02689 i01591.985.2110.0
Table 4. Transition temperatures (TA-B, °C) of some imine-linked phenyl 4-alkoxybenzoate dimers [51,52,53].
Table 4. Transition temperatures (TA-B, °C) of some imine-linked phenyl 4-alkoxybenzoate dimers [51,52,53].
Molecules 27 02689 i016
No.nmTMPTCol-N/IsoTNTB-NTN-Iso
3045106--102
3165113-8797
328510399-102
33105105111--
34125107116--
35145105120--
3647119--121
376798-93113
3887103--110
3910797106-108
40127100112--
4114799117--
Table 5. Transition temperatures (TA-B, °C) of salicylaldimine dimers with odd spacer parity [50].
Table 5. Transition temperatures (TA-B, °C) of salicylaldimine dimers with odd spacer parity [50].
Molecules 27 02689 i017
No.nmTMPTB6-N/IsoTNTB-NTN-Iso
4245114.099.1-102.0
4365123.5116.9--
448594.2121.0--
4510588.5109.5--
4612596.296.1--
47145101.2---
4847112.284.496.6115.0
496796.4114.7--
5087111.2119.5--
51107100.1110.1--
5212790.999.3--
Table 6. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with varying central spacer composition, equivalent to heptamethylene [54].
Table 6. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with varying central spacer composition, equivalent to heptamethylene [54].
Molecules 27 02689 i018
No.NameχTMPTNTB-NTN-IsoAve. Bend/°
53CBT3TCB Molecules 27 02689 i019160.5--n/r
54CBT1O1TCB Molecules 27 02689 i020>225--n/r
4CB7CB Molecules 27 02689 i021104.4105.5118.9103.5
55CB3O3CB Molecules 27 02689 i022100.546.068.091.0
56CBT4OCB Molecules 27 02689 i023132.897.0145.2100.5
57CB6OCB Molecules 27 02689 i024102.1110.5154.2104.4
58CBO5OCB Molecules 27 02689 i025137.981.3189.2102.9
59CBK5KCB Molecules 27 02689 i026158.1145.1189.4108.2
60CBI3ICB Molecules 27 02689 i027170.8114.9-115.1
61CBO2O2OCB Molecules 27 02689 i028150.5-157.8104.5
Table 7. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with varying central spacer composition, equivalent to nonamethylene [54,55,56,57,58]. * Glass to NTB transition.
Table 7. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with varying central spacer composition, equivalent to nonamethylene [54,55,56,57,58]. * Glass to NTB transition.
Molecules 27 02689 i029
No.NameχTMPTNTB-NTN-IsoAve. Bend/°
6CB9CB Molecules 27 02689 i03083.3105.4121.5103.1
62CBI7ICB Molecules 27 02689 i031140.8114.7138.7111.5
63CB8KCB Molecules 27 02689 i032127.8128.1153.998.5
64CBT6OCB Molecules 27 02689 i033137.1102.0153.698.5
65CB8OCB Molecules 27 02689 i034110.6109.9153.3100.7
66CBS7SCB Molecules 27 02689 i03515.9 *88.3115.299.2
67CBS7OCB Molecules 27 02689 i03655.095.9146.796.8
68CBSe7SeCB Molecules 27 02689 i03780.843.171.998.8
69CBcZ5OCB Molecules 27 02689 i03895.139.891.393.0
70CBcO5OCB Molecules 27 02689 i039122.471.3129.996.8
71CBO7OCB Molecules 27 02689 i040120.0---
72CBT5CBT Molecules 27 02689 i041169.1---
Table 8. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with methylene, ether, or thioether linking groups and varying spacer length: selected members of the CBnOCB [59,60], CBnSCB [61], CBOnSCB [61], and CBSnSCB [61] series. * Glass to NTB transition.
Table 8. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with methylene, ether, or thioether linking groups and varying spacer length: selected members of the CBnOCB [59,60], CBnSCB [61], CBOnSCB [61], and CBSnSCB [61] series. * Glass to NTB transition.
Molecules 27 02689 i042
No.NamenXYTMPTNTB-NTN-Iso
73CB4OCB3–CH2–O–121103143
74CB4SCB3–CH2–S–126.270.386.8
75CBO3SCB3–O––S–101.147137.5
76CBS3SCB3–S––S–70.144.083.2
77CB6OCB5–CH2–O–99109155
78CB6SCB5–CH2–S–99.089.6113.2
79CBO5SCB5–O––S–59.590.1143.8
80CBS5SCB5–S––S–68.978.0107.8
65CB8OCB7–CH2–O–112108154
81CB8SCB7–CH2–S–92.693.9117.8
67CBO7SCB7–O––S–16.0 *95.9146.7
66CBS7SCB7–S––S–15.9 *88.3115.2
82CB9OCB9–CH2–O–116107148
83CB9SCB9–CH2–S–103.495.5119.0
84CBO9SCB9–O––S–98.095143.0
85CBS9SCB9–S––S–100.889116.7
Table 9. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with mixed ketone/thioether linking units [65].
Table 9. Transition temperatures (TA-B, °C) of cyanobiphenyl derivatives with mixed ketone/thioether linking units [65].
Molecules 27 02689 i043
No.NamenTMPTNTB-NTN-Iso
86CBK3SCB3114.6104.6132.6
87CBK5SCB5120.9110.0152.3
88CBK7SCB7121.8113.6152.1
89CBK9SCB9155.8-145.0
Table 10. Transition temperatures (TA-B, °C) of unsymmetrical cholesterol containing dimers.
Table 10. Transition temperatures (TA-B, °C) of unsymmetrical cholesterol containing dimers.
Molecules 27 02689 i044
No.nmTMPTSmA-N*TNTB-N*TN*-Iso
9031112.3-55.167.7
9141127.9146.9-191.8
925192.5-67.9102.3
9361152.3120.0-158.4
947183.0-75.7110.8
959181.2-75.7113.6
9610182.7--137.5
9715162.9-63.2107.9
985274.0-60.693.6
995377.5-62.298.8
1005444.271.9-99.3
Table 11. Transition temperatures (TA-B, °C) of unsymmetrical dimers comprising cholesterol and azobenzene units.
Table 11. Transition temperatures (TA-B, °C) of unsymmetrical dimers comprising cholesterol and azobenzene units.
Molecules 27 02689 i045
No.nRTMPTSmX-N*TNTB-N*TN*-Iso
1015–CH383.4-67.1105.5
1025–OCH384.3-80.8)124.9
1035–OC2H5107.097.6-134.4
1047–CH363.9-74.0111.5
1059–CH369.6-75.8113.6
10615–CH360.560.3-100.3
Table 12. Transition temperatures (TA-B, °C) of symmetrical dimers containing an (R)-2-methylpentamethylene spacer.
Table 12. Transition temperatures (TA-B, °C) of symmetrical dimers containing an (R)-2-methylpentamethylene spacer.
Molecules 27 02689 i046
No.RTMPTSmA-NTB*TNTB-N*TN*-Iso
107 Molecules 27 02689 i047134.1--71.1
108 Molecules 27 02689 i04861.6---
109 Molecules 27 02689 i04969.1---
110 Molecules 27 02689 i05066.6--233.6
111 Molecules 27 02689 i051117.1110.8123.5219.8
112 Molecules 27 02689 i052119.3161.0167.2236.4
113 Molecules 27 02689 i053123.3--237.4
Table 13. Transition temperatures (TA-B, °C) of unsymmetrical CBnOPZPZP dimers with a terminal butyl, (S)-2-methylbutyl, or rac 2-methylbutyl chain [71].
Table 13. Transition temperatures (TA-B, °C) of unsymmetrical CBnOPZPZP dimers with a terminal butyl, (S)-2-methylbutyl, or rac 2-methylbutyl chain [71].
Molecules 27 02689 i054
NonmTMP/°CTSmA-NTB*TNTB-N*TN*-Iso
1144 Molecules 27 02689 i055154--249
1156 Molecules 27 02689 i056154-96248
1168 Molecules 27 02689 i057144-93241
11710 Molecules 27 02689 i058135-89231
1184 Molecules 27 02689 i059162--209
1196 Molecules 27 02689 i060153-89214
1208 Molecules 27 02689 i061132-93212
12110 Molecules 27 02689 i06213188-203
1226 Molecules 27 02689 i063153-85214
1238 Molecules 27 02689 i064131-92207
Table 14. Transition temperatures (TA-B, °C) of N-phenyl piperazine derived bent-core compounds [12,73].
Table 14. Transition temperatures (TA-B, °C) of N-phenyl piperazine derived bent-core compounds [12,73].
Molecules 27 02689 i065
No.nTMPTSmCP-N/IsoTColX-NTNTB-NTN-Iso
1244201--193212
1255187--172192
1266176-157169188
1277169---186
1288177180--180.5
1299167185---
13010162191---
13111161194---
13212165201---
13316142193---
Table 15. Transition temperatures (TA-B, °C) of the hydrogen-bonded trimers CB5OCB (linear) and CB6OBA (bent) [86].
Table 15. Transition temperatures (TA-B, °C) of the hydrogen-bonded trimers CB5OCB (linear) and CB6OBA (bent) [86].
Molecules 27 02689 i066
No.NamenTMPTNTB-NTN-Iso
135CB5OBA5196-209
136CB6OCB6160159197
Table 16. Transition temperatures (TA-B, °C) of compounds 140143, the ‘DTC5-C9’ family [92].
Table 16. Transition temperatures (TA-B, °C) of compounds 140143, the ‘DTC5-C9’ family [92].
Molecules 27 02689 i067
No.nTMPTSmX-NTNTB-NTN-Iso
140034--116.5
14117785124162
1422127-145192
1433142-168205
Table 17. Transition temperatures (TA-B, °C) of ether-linked trimers 144151 [99].
Table 17. Transition temperatures (TA-B, °C) of ether-linked trimers 144151 [99].
Molecules 27 02689 i068
No.nTMPTSmA-NTNTB-NTN-Iso
1444230.1197-297.4
1455176.6-122215.4
1466217.4192-257.8
1477162.9-132207.2
1488201.0175-231.9
1499155.0-135195.8
15010198.0150-210.8
15111147.4-130.5185.9
Table 18. Transition temperatures (TA-B, °C) of mixed ether/thioether linked timers 152160 [100].
Table 18. Transition temperatures (TA-B, °C) of mixed ether/thioether linked timers 152160 [100].
Molecules 27 02689 i069
No.nTMPTSmA-NTNTB-NTN-Iso
1523186.4--142.0
1534235.3--253.2
1545171.8-117154.2
1556219.0180-217.4
1567149.4-126162.4
1578195.6167-198.8
1589131.3-122.7158.8
15910184.8146-176.2
16011135.4-121154.4
Table 19. Transition temperatures (TA-B, °C) of compounds 161164; # sample decomposes at and above 230 °C [101].
Table 19. Transition temperatures (TA-B, °C) of compounds 161164; # sample decomposes at and above 230 °C [101].
Molecules 27 02689 i070
No.XTMPTNTB-NTN-Iso
161 Molecules 27 02689 i071178.8166.2304.6
162 Molecules 27 02689 i072169.3159.8285.5
163 Molecules 27 02689 i073174.4163.9>230 #
164 Molecules 27 02689 i074162.2->230 #
Table 20. Transition temperatures (TA-B, °C) of compounds 77, 165172. Link Seq. refers to the shape (B = bent, L = linear) shape of the all trans conformation of the linkage units [102].
Table 20. Transition temperatures (TA-B, °C) of compounds 77, 165172. Link Seq. refers to the shape (B = bent, L = linear) shape of the all trans conformation of the linkage units [102].
Molecules 27 02689 i075
No.nχpqTMPTNTB-NTN-IsoLink Seq.
776-0099109155B
1656-10153142185BB
1666–(CH2)711140159180BBB
1676–(CH2)6O–11151158195BBB
1686–O(CH2)5O–11156162215BBB
1696–O(CH2)6O–11166166235BLB
1706–(CH2)811131152212BLB
1717–(CH2)711145132225LBL
172O6–(CH2)711131143242LBL
Table 21. Transition temperatures (TA-B, °C) of compound 173 and complexes 174 and 175 [103].
Table 21. Transition temperatures (TA-B, °C) of compound 173 and complexes 174 and 175 [103].
Molecules 27 02689 i076
No.XTMPTSmXTSmA-NTBTNTB-NTN-Iso
173none142.8----
174 Molecules 27 02689 i077121.986.2-109.4166.4
175 Molecules 27 02689 i078112.085.593.598.0157.7
Table 22. Transition temperatures (TA-B, °C) of complexes 176 and 177 [105].
Table 22. Transition temperatures (TA-B, °C) of complexes 176 and 177 [105].
Molecules 27 02689 i079
No.XTMP/°CTNTB-N/°CTN-Iso/°C
176 Molecules 27 02689 i08010795167
177 Molecules 27 02689 i08110388145
Table 23. Transition temperatures (TA-B, °C) of complexes 178184 [106].
Table 23. Transition temperatures (TA-B, °C) of complexes 178184 [106].
Molecules 27 02689 i082
No.nTMPTSmC-NTBTNTB-NTN-Iso
1781130-110182
1792119-115190
1803109-108173
181410886113180
182512192107165
183610596106158
1847100100104155
Table 24. Transition temperatures (TA-B, °C) of complexes 185 and 186 [106].
Table 24. Transition temperatures (TA-B, °C) of complexes 185 and 186 [106].
Molecules 27 02689 i083
No.XTMPTNTB-NTN-Iso
185-CN127143186
186-OCH3140142184
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Mandle, R.J. A Ten-Year Perspective on Twist-Bend Nematic Materials. Molecules 2022, 27, 2689. https://doi.org/10.3390/molecules27092689

AMA Style

Mandle RJ. A Ten-Year Perspective on Twist-Bend Nematic Materials. Molecules. 2022; 27(9):2689. https://doi.org/10.3390/molecules27092689

Chicago/Turabian Style

Mandle, Richard J. 2022. "A Ten-Year Perspective on Twist-Bend Nematic Materials" Molecules 27, no. 9: 2689. https://doi.org/10.3390/molecules27092689

Article Metrics

Back to TopTop