You are on page 1of 11

PAPER

www.rsc.org/materials | Journal of Materials Chemistry

Supramolecular order of stilbenoid dendrons: importance of weak interactions{


Matthias Lehmann,*a Christiane Kohn,a Herbert Meier,b Sabine Renkerb and Annette Oehlhofb
Received 28th July 2005, Accepted 17th October 2005 First published as an Advance Article on the web 8th November 2005 DOI: 10.1039/b510713j Stilbenoid dendrons with various donor and acceptor groups on the focal unit were synthesised by a WittigHorner reaction, starting from an aldehyde functionalised dendron and various substituted phosphonic acid esters. The target molecules are composed of meta-branched arms, two of them with extended conjugation (distyrylbenzene) and three flexible dodecyloxy chains; the focal group consists of a donor or acceptor substituted styryl unit. The cross-conjugation of the arms prevents the strong electronic influence of substituents on the two extended oligophenylenevinylene chromophores. However, intermolecular interactions mediated by the focal unit allow control of the supramolecular stacking into liquid crystal phases. Simple weak acceptors stabilise the formation of columnar phases, whereas the additional propensity to build hydrogen bonds leads to a cubic mesophase. All acceptor substituted materials freeze at low temperature into a glassy state. Soft crystals are then formed upon heating the glassy material. Stilbenoid dendrons are photosensitive and degradation of the supramolecular order proceeds even in the glassy liquid crystal state.

Introduction
Stilbenoid molecules attract increasing attention due to their interesting photophysical, photochemical and electronic properties.1 Their propensity for charge transport and electroluminescence has led to applications in light emitting diodes,2 field effect transistors and photovoltaic cells.3 Recently, the conjugated stilbenoid scaffold has been combined with the concept of dendrimers.4 These molecules are highly symmetric, regularly branched and possess a welldefined size. Their modular design allows the synthesis of variable structures, i.e. the usage of different functional cores, branching and peripheral units.5 These features together with the good film-forming processability of dendrimers and the optoelectronic properties of the conjugated scaffold make such molecules eligible for application in electronic devices.5,6 The first two generations of stilbenoid dendrimers with long peripheral alkoxy chains form liquid crystalline (LC) phases.4 Self-assembly in mesophases has been shown to be favourable for electronic materials, because of a facile alignment of functional units and the possibility of structural self-healing.7 The molecular mobility in LC phases of stilbenoid mesogens is also correlated with photochemical and photophysical properties, e.g. molecular motion allows photoreactions to proceed even when the molecules are photostable in crystalline phases.8 The driving force of columnar mesophase formation
a

Non-Classical Synthetic Methods, Institute of Chemistry, Chemnitz University of Technology, Strasse der Nationen 62, 09111 Chemnitz, Germany. E-mail: matthias.lehmann@chemie.tu-chemnitz.de; Fax: +49 371 531 1839; Tel: +49 371 531 1205 b Institute of Organic Chemistry, University of Mainz, Duesbergweg 1014, 55099 Mainz, Germany { Electronic supplementary information (ESI) available: Comparison of 1H-NMR data, UV-Vis spectra, and the FT-IR spectrum of amide 3f. See DOI: 10.1039/b510713j

for stilbenoid dendrimers has been proposed to be microsegregation of the rigid conjugated scaffold and the flexible aliphatic chains, due to the preorganisation of these molecular units in the mesogen.5,9,10 If the number and the position of aliphatic chains change, mesomorphic behaviour changes too or is even lost. The mesophase is stabilised if a dipole11 or a pushpull character12 is introduced by the core unit. Dendrons are wedge-shaped building blocks from which dendrimers can be obtained in a convergent synthesis. They may show LC behaviour, if the contrast between the flexible chains and rigid scaffold allows micro-segregation.13,14 Different units can be easily attached to the focal position and thus functional groups can be arranged in the centre of a column by micro-segregation and hierarchical self-assembly.15 In the series of dendrons 1, made of stilbene building blocks, enantiotropic LC phases are formed in the third generation (Fig. 1).16 The larger distyrylbenzene scaffold already allows self-organisation in columnar mesophases starting with the second generation dendron 2. Thus, the size of the semi-rigid scaffold seems to be important for self-aggregation. In contrast to radially-symmetrical dendrimers, most dendrons can only form columnar phases with more than one mesogen placed in a columnar slice.13 Since the focal units will then meet in the centre of the columns, substituents at this position should strongly influence the mesogenic properties. Therefore, dendrons 3 with different small substituents on the focal unit have been designed based on precursor 2 (Fig. 2). The focal aldehyde group of 2 allows the facile modification of the molecular scaffold with donor or acceptor substituted phosphonic acid derivatives. Since 2 already possesses an enantiotropic mesophase, mesomorphic properties were also expected for 3, which is enlarged by a styryl group. Leaving the number of aliphatic chains constant, we studied the influence of weak interactions, i.e. dipoledipole interactions and H-bonds, on
J. Mater. Chem., 2006, 16, 441451 | 441

This journal is The Royal Society of Chemistry 2006

Fig. 1 Stilbenoid dendrons forming columnar mesophases.

Fig. 2 Structure of donor and acceptor substituted stilbenoid dendrons 3.

the supramolecular stacking of stilbenoid dendrons 3. Synthesis, mesomorphic, photophysical and photochemical properties of new stilbenoid materials 3 are presented.

Scheme 1 Synthesis of substituted stilbenoid dendrons 3.

Synthesis
The preparation of the target compounds 3, outlined in Scheme 1, was performed starting from dendron 2.17 The WittigHorner reaction of aldehyde 2 with the phosphonic acid diethyl esters 4a,18 4b,19 4c20 and 4d21 in THF in the presence of KOtBu furnishes the products 3ad in moderate isolated yields. The aldehyde 3g was obtained from 4f22 by a subsequent acidic treatment to cleave the acetal protecting group. The amide derivative 3f was prepared by the reaction with the phosphonic ester bearing the cyano group 4e23 and hydrolytic workup. In order to prevent the hydrolysis of the cyano function and thus to obtain 3e, NaH was used as the base for the WittigHorner coupling. All compounds are characterised by NMR-, IR-, mass-spectroscopy and elemental analysis. The different electronic influence of the substituents is evident from the 1H NMR spectra. Electron withdrawing groups (e.g. CHO, CN, CONH2) shift the aromatic signals of the focal benzene to lower field, compared to the signals of non-substituted (3c) or donor substituted
442 | J. Mater. Chem., 2006, 16, 441451

compounds (e.g. CH3 (3a), OCH3(3b)). Although bromine introduces a dipole in 3d, its electronic effect is very weak, since nearly no differences to the chemical shifts in 3c are apparent. The influence of the different substituents is strongest at the focal benzene ring; their effect decreases strongly at the central benzene and is not present for signals of the alkoxy substituted arms due to their cross-conjugation (meta-position).24 The vanishing electronic influence of the focal group upon the more extended oligophenylenevinylene arms can be also observed in the series of UV-Vis spectra. As in other meta-branched stilbenoid molecules,5,25 the absorption of 3 can be approximated as the sum of the absorptions of single arms (i.e. one stilbene and two distyrylbenzene units). Therefore the absorption maximum at 473 nm (CH2Cl2) corresponding to the distyrylbenzene unit remains unchanged within the series. Only a slight effect is observed on the high energy side of the long-wavelength band, which can be related to the differently substituted focal stilbene unit.24
This journal is The Royal Society of Chemistry 2006

Table 1 Thermotropic data obtained by DSC (heating rate 10u min21, phase transitions: onset [uC], enthalpies [kJ mol21]), POM and X-ray diffraction Compound 2 3a 3b 3c 3d 3e Phase transitions (Onset [uC]) and transition enthalpies [kJ mol21]a cooling: I 102/23 Colhd 77/249 Cr heating: Cr 89 Cr1 96/43 Colhd 104/4 I cooling: I 19/21 Colhd 215/216 Cr heating: Cr 242/11 Colhd 23/1 I 31/24 Cr1 40/4 Cr2 44/219 Cr3 86/36 I 1. cooling: I 210/226 Cr 2. heating: Cr 219/9 I 12/254 Cr1 46/44 I 1. cooling: I 16/21 Colhd 212/222 Cr 2. heating: Cr 221/15 Colhd 29/1 I 43/261 Cr1 80/70 I 1. cooling: I 64/21 Colhd 19(Tg) g 213/224 X 2. heating: X 248/15 g 12 (Tg) Colhd 65/1 I 1. cooling: I 88/21.6 Colhd 25(Tg) g 220/229 X1 2. heating: X1 241/20 g 28(Tg) Colhd/X2 92/1.9 I 3. heating after annealing at 50 uC: X2 92/1.9 I 4. heating after cooling only to 80 uC: Colhd 92/1.7 I 1. cooling: I 186/21 Cub 55(Tg) g 23/235 X 2. heating: X 242/22 g 55(Tg) Cub 189/1 I 1. cooling: I 96/21.6 Colhd 19(Tg) g 213/28 X1 2. heating: X1 235/11 g 24(Tg) Colhd/X2 98/1.7 I 3. heating after annealing at 50 uC: X2/Colhd 98/1.2 X2/I 103/4.5 I 1. 2. 1. 2.

3f 3g

Fig. 3 Texture of 2 at 88 uC between crossed polarisers. Three different aligned domains are visible: (a) pseudo-focal-conic domains; (b) domains with homogenous aligned material and (c) a homeotropic aligned region. The inset shows a hexagonal growing germ at the I-LC phase transition.

Colhd columnar hexagonal disordered phase; Cub cubic phase; g glassy state; I isotropic liquid; Cr crystalline phase; X partially crystalline or plastic crystalline phases with unknown phase structure.

Thermotropic behaviour
The thermotropic properties of 2 and 3 were studied by means of differential scanning calorimetry (DSC), polarised optical microscopy (POM) and X-ray diffraction and are summarised in Table 1. DSC and POM investigations Aldehyde 2, the starting material for the preparation of the target molecules 3, forms a liquid crystalline phase within a small temperature range in the second heating trace. In the cooling curve this interval increases since the LC phase can be supercooled by almost 20 uC. However, when the sample is annealed during the heating scan at 96 uC in the LC phase a different crystal phase forms. Subsequent melting of the crystal and clearing of the LC phase are very close and cannot be separately evaluated. POM studies show mosaic and pseudofocal-conic textures typically found for columnar phases (Fig. 3). The hexagonal growing germ at 104 uC points to a hexagonal two-dimensional order of columns in the mesophase. When the aldehyde group is exchanged with the styrene building block to yield compounds 3, a decreased crystallisation tendency can be observed. Molecules 3a and 3c, either with or without an electron donor substituent, do not crystallise upon cooling but form a monotropic mesophase at low temperature instead. Obviously, 3b does not show a supramolecular aggregation to a liquid crystal phase, which presumable has its origin in the additional steric interaction of the methyl group in ortho-position to the double bond. In contrast, all compounds with electron withdrawing substituents assemble in, at first sight, enantiotropic mesophases. Even the small, local dipole of the bromine substituent
This journal is The Royal Society of Chemistry 2006

stabilises mesophase formation. At low temperature the columnar liquid crystals do not crystallise but freeze into a glassy state. Decreasing further the temperature results in another first order transition. It is assumed that this is related to a partial crystallisation of side chains, which can still be mobile above the transition.6 The interval of the LC phase can be further increased by stronger acceptors or dipoles. The largest mesophase range of 74 uC is observed for the aldehyde substituted compound 3g. The pseudo-focal-conic textures of 3d, 3e and 3g point to columnar stacking of mesogens (Fig. 4).26 In the case of 3g, conoscopy on a homeotropic aligned sample indicates an uniaxial phase with a negative optical anisotropy, which is additional evidence for a columnar

Fig. 4 Pseudo-focal-conic texture of 3g at 89 uC between crossed polarisers. The inset A shows a homeotropic aligned region. The conoscopic picture B taken with a l-wave plate proofs the optically negative, uniaxial nature of the mesophase.

J. Mater. Chem., 2006, 16, 441451 | 443

Table 2

X-Ray data obtained for stilbenoid dendrons 2 and 3 T [uC] 99 hkl 100 110 200 Halo 100 200 Halo 100 Halo 100 110 200 Halo 100 110 200 Halo 100 200 Halo dexp/A 40.1 23.2 20.5 4.7 42.8 21.8 4.5 41.4 20.6 4.5 39.1 22.3 19.9 4.4 48.8 28.0 24.6 4.4 47.3 23.9 4.4 dcalc/A 23.2 20.1 21.4 20.7 22.6 19.6 28.2 24.4 23.7 ahexa/A 46.3

Compound 2

3a 3c 3d

1 5 50

49.4 47.8 45.1

3e Fig. 5 Right: pseudo-focal-conic texture of 3e at 35 uC between crossed polarisers. Left: phase transition upon heating with 5 uC min21 to 60 uC.

15

56.3

3g

23

54.6

self-assembly of mesogens 3 from POM investigations. However, the thermotropic properties of the acceptor substituted mesogens are more complex when investigated at different heating rates in POM and DSC. Upon slow heating from room temperature (RT) above the glass transition, the mesophase crystallises, as it is evidenced by a texture change (Fig. 5). Note that this observation depends on the thermal history of the sample and is only monitored when the sample has been pre-frozen to the columnar glass. The smooth pseudo-focal-conic texture typically obtained for the mesophases is stable upon cooling from the isotropic phase. Annealing at 50 uC for 1 h does not lead to any transformation. Compound 3f behaves differently compared to the column forming mesogens. The particularity of its neat phases is the optical isotropy when observed under POM. However, DSC clearly indicates a first order transition between two optical isotropic phases. The phase below the transition is viscous but fluid and becomes liquid-like above 198 uC. These observations point to the presence of a cubic LC phase for 3f. X-Ray diffraction X-Ray diffraction was performed on cooling the samples from the isotropic liquid to the mesophase. The results are collected in Table 2. All compounds show diffraction patterns typically observed for mesophases in the temperature range of LC phases and frozen, glassy mesophases. The wide angle scattering powder diffraction patterns of 3a,c,d,e,g present up to three reflections at small angles and only a halo corresponding to the mean distance of liquid-like aliphatic chains and stilbenoid cores. Fig. 6 shows the results, obtained for compound 3e. The diffraction pattern A from an oriented fibre of 3e exhibits reflections of the small angle region only on the equator. A detailed investigation of this angular region in panel D presents signals in the ratio 1 : !3 : 2, thus demonstrating the hexagonal symmetry of the two-dimensional lattice. At the meridian of pattern A, a broad halo can be observed. Comparison of the integration curves B (integration along the meridian) and C (integration along the equator)
444 | J. Mater. Chem., 2006, 16, 441451

a The parameter ahex is calculated with ahex = 2d100/!3 and corresponds to the diameter of a column.

illustrates that the halo consists of two diffuse signals at 4.50 and 4.05 A, which are partially superimposed. The halo corresponding to the mean distance of liquid-like aliphatic chains is distributed over the whole angular range. The corresponding d-value is acquired from curve C and amounts to 4.50 A. The second signal with a maximum at d = 4.05 A is related to the mean distance of chromophores along the column. Thus, the diffraction pattern of the extruded fibre aligned with the meridian shows intercolumnar distances only on the equator and intracolumnar distances exclusively on the meridian. This is clear evidence for the columnar nature of the phase. The absence of reflections attributed to mixed indices

Fig. 6 X-Ray scattering of 3e in the columnar phase. A: Wide angle X-ray scattering pattern of an extruded fibre of 3e at 23 uC. B Integrated intensity (q over azimuthal angle). C Integration along the equator of the pattern. D Small angle region from a sample quenched from the isotropic liquid to 15 uC.

This journal is The Royal Society of Chemistry 2006

hkl with l ? 0 demonstrates the two-dimensional correlation of columns and thus the liquid crystalline character. From the half width of the broad halo attributed to the mean intracolumnar separation of chromophores 3e, a correlation length of 23.5 A is determined.27 This low value indicates the relative high disorder of mesogens along the columnar axis. For compounds 2 and 3d similar results where obtained, identifying the formation of Colh phases. The diffraction patterns of 3a, 3c and 3g show, besides the fundamental 100 reflection, only a broad signal indexed as 200. Consequently, their symmetry can not be directly defined since such patterns may be observed for hexagonal, tetragonal columnar and lamellar mesophases. However, the closely related molecular structure and similar POM textures compared with those of 3e or 3d suggest the formation of Colh phases. A model for the columnar assembly of 3e based on nanosegregated different molecular units (stilbenoid scaffold and aliphatic chains) gives more insight in the self-organisation of mesogens. Information on the number of molecules forming a columnar unit can be obtained from the X-ray density, given by eqn (1); z|M (1) r~ NA |A|h where r = density, z = number of molecules in the columnar unit, M = molecular weight, NA = Avogadros constant, A = columnar cross section and h = height of the columnar unit. The height of a columnar slice and the density are in principal not known. However, the height can be estimated by the mean distance given by the halo attributed to intracolum nar stacking (4.05 A) and the density can be set to 1 g cm23, a value typically found for organic material. The columnar cross-section filled by molecules 3e at RT can be calculated by A = a2 6 sin 60 = 2696 A2, which is the area of the hexagonal two-dimensional unit cell. By simple transformation of eqn (1), the number of molecules z can be calculated and amounts to 3.8 molecules per columnar slice. Thus, four molecules form a columnar unit. If antiparallel orientation of local dipoles is assumed, two pairs of molecules fill the space given by A 6 h as shown in Fig. 7.28 Further details can be obtained considering the uniaxial nature of the hexagonal lattice which presumes a circular cross section of the columnar core, occupied by stilbenoid chromophores. With this model the volume fraction of the stilbenoid core Vcore can be calculated if the volume fraction of the aliphatic chains VCH is known; then Vcore = A 6 h 2 VCH (eqn (2)). The calculation of VCH was carried out according to data from dilatometry investigations.29 At 23 uC VCH amounts to 7801 A3, which is the volume of the dodecyl chains of four molecules, and thus Vcore = 3120 A3. In this model Vcore is assumed to be of cylindrical shape. Consequently, the radius of such a cylinder can be calculated by rcore = !Vcore/ (h 6 p) = 15.7 A (eqn (3)). The value compares excellently with the extension of the stilbenoid scaffold from the central benzene ring to the middle oxygen of a distyrylbenzene arm (see Fig. 7), which is 16.4 A.28 A close inspection of the hexagonal parameter ahex in Table 2 clearly shows a large difference in column diameters of donor and acceptor substituted mesogens of up to 10 A. This is surprising since all molecules are based on the same scaffold.
This journal is The Royal Society of Chemistry 2006

Fig. 7 Premilinary apparent model of the supramolecular stacking of 3e in a columnar slice.28 In the space filling representation alkyl chains are omitted for clarity.

However, results from temperature-dependent powder X-ray investigations, summarised in Table 3, may explain the unexpected large variation of cell parameters. For the molecules with the largest temperature interval of the hexagonal phase (3e, 3g), columnar diameters decrease with increasing temperature. In the same temperature range the position of the halo remains almost constant between 4.4 and 4.5 A. This would imply an increasing density of the material with increasing temperature, which is not reasonable. Data corresponding to the columnar extension along the axis could not be obtained from the X-ray patterns. However, an increase in the height of the columnar slice can be expected to be the reason for the decreasing spatial requirements perpendicular to the columnar axis. The height hcol can be calculated by using eqn (2) and (3) when the volume fraction of the aliphatic chains is known and a constant core diameter of 16.4 A is assumed.30 The acceptor substituted mesogens 3e and 3g exhibit an increasing height hcol with increasing temperature
Table 3 X-Ray parameters of 3 in the columnar liquid crystal phase T [uC] 1 5 30 50 23 50 70 90 23 45 90 ahexa/A 49.4 47.8 45.7 45.1 55.8 53.6 51.3 49.8 54.6 53.0 49.5 VCHb/A3 7669 7693 7845 7976 7801 7976 8115 8263 7801 7943 8263 hcolc/A 6.04 6.78 8.12 8.67 4.22 4.86 5.64 6.33 4.49 5.00 6.45 rd/g cm23 0.90 0.84 0.80 0.77 1.00 0.95 0.89 0.84 0.99 0.94 0.83

Compound 3a 3c 3d 3e

3g

Diameter of columns in the hexagonal columnar phase (Colh). Volume fraction of alkyl chains of four molecules 3 at temperature T, calculated as described in ref. 29a. c Calculated columnar height hcol = Vch/(A 2 r2core 6 p) with rcore = 16.4 A obtained from a molecular model. d Density calculated according to eqn (1) with h = hcol.
b

J. Mater. Chem., 2006, 16, 441451 | 445

(Table 3, column 5), e.g. in the mesophase of 3e at 90 uC, four mesogens occupy the space of a columnar unit with hcol = 7.3 A; at 23 uC this value decreases to 4.2 A. The latter number is in good agreement with the intracolumnar distance obtained from the diffraction pattern of an oriented fibre (Fig. 6) and monitors the quality of the estimation. In contrast, all donor substituted mesogens show a significantly higher value for the height of the columnar unit even at temperatures as low as 1 uC. This behaviour can be explained by interactions of local dipoles at the focal unit of the dendrons. Dipole moments are relatively large for the cyanobenzene (4.18 D)31a,32 and benzaldehyde units (2.97 D)31b,32 in 3e and 3g, respectively. Dipoledipole interactions are most attractive in antiparallel arrangements. According to the model in Fig. 7, it is reasonable to assume antiparallel orientation of adjacent mesogens. In such a configuration, attractive dipoledipole interactions should hold mesogens closely together, which should lead to decreasing intracolumnar distances with increasing dipole strength. This is in agreement with the calculated values, where 3e and 3g show the smallest distances hcol compared with donor substituted mesogens 3a and 3c. Note, that although the cyano derivative 3e possesses the larger dipole moment, the clearing temperature for aldehyde 3g is higher by 6 uC. This cannot be rationalised only with dipole moments, but may have its origin in additional weak contributions, e.g. CHp interactions.33 The dipoles do not only force the molecules in columnar units of smaller extension along the columns compared to donor substituted mesogens, but also facilitate the nano-segregation and, thus, the formation of mesophases over a large temperature range. This is evidence that the interaction introduced by the different focal substituents play a key role on structural parameters of the mesophases. The last column in Table 3 lists calculated densities, based on the measured diameters and the heights hcol, assuming four molecules in a columnar unit. The range of the values and the decreasing densities with increasing temperature are reasonable for organic molecules, thus they support the proposed model for columnar self-organisation. A different situation is observed for 3f. Although the amide should have a dipole moment similar to benzamide (3.77 D)31c, a value between those of benzaldehyde and cyanobenzene, the clearing temperature is much higher than clearing temperatures for mesophases of 3e and 3g. POM observations point to a cubic arrangement of mesogens. X-Ray investigations at small angles were performed in the isotropic phase at 220 uC and after cooling to 150 uC in the mesophase (Fig. 8). In the isotropic phase only a halo at 47 A is detected, which can be related to the molecular size. In the mesophase, many reflections appear at small angles. The shoulder at 426 A is attributed to the 100 reflection. The other signals can then be indexed according to a cubic phase. The different thermotropic behaviour cannot be explained by simple dipole interactions, but is attributed to the formation of H-bonds between the amide functions, which manifests in the position of the NH vibrations at 3356 and 3196 cm21 in the FT-IR-spectrum of a thin film at RT.34 Shoulders at 3480 and 3400 cm21, however, indicate, that not all NH functions are involved in hydrogen bonding.24
446 | J. Mater. Chem., 2006, 16, 441451

Fig. 8 Small angle X-ray scattering of a powder sample 3f in the isotropic phase at 220 uC and the LC phase at 150 uC.

Fig. 9 Panel A: X-ray scattering from compound 3g at 23 uC extruded from the soft crystal phase (85 uC). Panel B: Pattern obtained when annealing the LC phase of 3g at 40 uC. Positions of reflections at the equator correspond well with diffraction rings of the annealed sample.

In contrast to the cubic phase of mesogen 3f, the columnar phases are all monotropic, as emphasised earlier. Even the acceptor stabilised mesophases transform to a different phase when heated from RT above the glass transition. A sample of 3g extruded at 80 uC shows an X-ray pattern with many signals at small angles for such a phase (Fig. 9). All these reflections are found at the equator of the pattern, thus the molecules are aligned along the fibre axis. Since the material in the new phase remains soft and can be oriented by extrusion, a soft columnar crystalline nature is proposed for this phase. However, the still relatively diffuse reflections at the meridian of the X-ray pattern point to a considerably large intracolumnar disorder. The two-dimensional order of the columns, i.e. the 2D space group, will be studied in more detail by SAXS measurements and will be published elsewhere. A similar structural change in the vicinity of the glass transition was recently observed for star-shaped oligobenzoates, where the columnar liquid crystalline phase serves as a template for the formation of columnar crystals.35

Photophysical and photochemical properties


As mentioned earlier, the meta substitution of stilbenoid molecules 3 decouples the three chromophores electronically in
This journal is The Royal Society of Chemistry 2006

Fig. 10 Absorption and emission spectra of 3e in hexane and CH2Cl2. For comparison, the absorption spectrum of 3e in hexane is normalised to the maximum of the spectrum in CH2Cl2. Absorption maxima: 363 nm (hexane), 366 nm (CH2Cl2); emission maxima: 419 nm (hexane), 474 nm (CH2Cl2), 500 nm (liquid crystal at RT).

the first approximation. Therefore, the small influence of substituents on the stilbene chromophore at the focal position of the dendron can be considered as small variations of the high energy side of the long-wavelength absorption band. Fig. 10 presents the absorption and emission spectra of 3e in hexane, CH2Cl2 and solid state. The absorption band of 3e in hexane is normalised to the intensity of the absorption band in CH2Cl2. The hexane spectrum is only slightly hypsochromically shifted by about 3 nm. Considerably larger effects are observed for the emission spectra. The fluorescence spectrum of 3e in CH2Cl2 shows a maximum at 474 nm, which is bathochromically shifted to 500 nm in the glassy liquid crystal state at RT for a thin film of a neat sample. In contrast, the spectrum of 3e in hexane shows not only a much smaller band width, but also a hypsochromically shifted emission maximum

at 419 nm with a shoulder at 440 nm. Similar observations were made for stilbenoid dendrimers with C3-symmetry and were attributed to the formation of weak aggregates.4 The simple exciton theory proposed by Kasha36 predicts a hypsochromic shift of the emission maximum if the transition dipole moments are aligned in parallel. The formation of aggregates is supported by investigating the photodegradation upon irradiation. Fig. 11 depictes the fast photodegradation of 3e in hexane. The long-waved band corresponding to the distyrylbenzene chromophores at 366 nm disappears rapidly. After 20 s a new maximum at 352 nm emerges, which may be related with an initial cistrans isomerisation of the stilbene unit. The prolonged irradiation leads to an irreversible formation of CC bonds and, consequently, a new band with a maximum at 324 nm corresponding to remaining stilbene

Fig. 11 Irradiation of 3e in hexane with l > 300 nm (light of a Xn lamp was filtered by a 1 M NiSO4 solution).

This journal is The Royal Society of Chemistry 2006

J. Mater. Chem., 2006, 16, 441451 | 447

Fig. 12 Irradiation of 1e in the glassy LC state at RT. A before irradiation; B after irradiation for 6 h 15 min; C irradiated area at 84 uC; D irradiated area after cooling from the isotropic liquid to the glassy LC state at RT.

that focal substituents only play a minor role for the chemical shifts and the absorption of the two extended oligophenylenevinylene units. However, the weak interactions mediated by the focal unit have a large impact on the supramolecular assembly of the mesogens. Donor substituents destabilise the mesophases compared to the initial aldehyde. Acceptor groups initiate the formation of more stable liquid crystal phases with an increased mesophase range. A possible model of the hexagonal columnar mesophases proposes four molecules with antiparallel dipoles in a columnar slice, which can be regarded as the smallest columnar unit. The local dipoles at the focal position of the stilbenoid scaffold affect the intracolumnar distances. The amide, as hydrogen bond donor and acceptor, allows the creation of hydrogen bonds and, instead of a columnar assembly, the amide substituted mesogens stack into a cubic phase. All acceptor substituted mesogens freeze at low temperature into a glassy state. Despite the stabilisation effect of large dipoles, all columnar mesophases are monotrope; annealing above the glass transitions transforms the mesophases into crystalline materials, which are reminiscent of phases formed by different star-shaped oligobenzoates. Investigations are in progress to study the crystallisation process and the packing of stilbenoid mesogens in these soft crystalline phases.

units develops. All chromophores have been degraded after additional irradiation for 26 min. This process is considerably slower in a solution of 3e in CH2Cl2. Thus hexane, which is a poor solvent for the more polar aromatic scaffold of 3e, assists the formation of aggregates, which then undergo fast photooligomerisation, as observed previously for different stilbenoid denrimers.4,5,37 Photopolymerisation and crosslinking of stilbenoid compounds in thin films could be nicely pursued by AFM measurements.38 The photosensitivity has also been investigated in the neat condensed phases of mesogen 3e by means of POM (Fig. 12). The liquid crystal phase at a temperature close to the LCI transition already transforms after a few minutes to an isotropic phase when the material is not protected from the UV part of the microscope light. Interestingly, 6 h of irradiation of 3e in the glassy LC state only led to a slight texture change. However, when heated, the sample melts at a temperature below the clearing point of the LC phase and does not return to an LC state in the irradiated areas. Thus, the frozen glassy state does not prevent photoreactions of compounds 3, as opposed to the crystalline state of a first generation stilbenoid dendrimer.8

Experimental section
General methods Middle pressure liquid chromatography (MPLC) was performed using a Buchi apparatus with silica (J. T. Baker H2272, 5 6 40 cm). Differential scanning calorimetry (DSC) was performed on a Perkin Elmer DSC 7 instrument. Polarised optical microscopy (POM) observations were made with a Zeiss Axioscop 40 equiped with a Linkam THMS600 hot stage. PFT 1H and 13C NMR spectra were recorded in CDCl3 with Bruker AM400, AC200 and ARX400 spectrometers. Mass spectra were obtained on Finnigan MAT95 (FD MS). UV/Vis spectra were recorded with a Zeiss MCS 320/340 spectrometer and fluorescence spectra were obtained with a Perkin Elmer LS 50B instrument. The X-ray diffraction was measured on a Siemens D500 diffractometer or a Kratky Compact Camara with a Braun detector (Cu Ka radiation, l = 0.154 nm). The WAXS measurements on aligned samples obtained by extrusion were made by using a rotating anode (Rigaku 18 kW) source with pinhole collimation equipped with a graphite double monochromator (l = 0.154 nm) and a Siemens area detector with 1024 6 1024 pixels. (E,E,E,E,E)-1,3-Bis{2-[4-(2-{3,4,5-tridodecyloxyphenyl}ethenyl)phenyl]ethenyl}-5-[2-(4-methoxyphenyl)ethenyl]benzene (3a) Potassium-tert-butylat (3.00 g, 26 mmol) was dissolved in 60 ml THF under argon and cooled with an ice bath to 0 uC. A mixture of aldehyde 2 (1.50 g, 0.93 mmol) and diethyl methoxybenzylphosphonate 4a18 (0.25 g, 0.97 mmol) in THF was added dropwise. The solution was then stirred at room temperature (RT) for 48 h, poured on 100 g crushed ice and 50 ml HCl (18%) were added. The crude solid product was
This journal is The Royal Society of Chemistry 2006

Summary and conclusions


C2-Symmetric dendrons 3 have been synthesised starting from aldehyde 2 by a WittigHorner reaction with appropriate phosphonic acid diethyl esters. They are composed of a 1,3,5substituted benzene core, two linear stilbenylethenylene arms bearing three peripheral dodecyloxy chains each and one styryl group with various electron withdrawing or electron donating substituents (OCH3, CH3, H, Br, CN, CONH2, CHO). The different arms are cross-conjugated, thus in first approximation the UV-spectra are superpositions of absorptions from individual chromophores. NMR and UV data give evidence
448 | J. Mater. Chem., 2006, 16, 441451

collected, dissolved in CHCl3 and precipitated by addition of ethanol. A further purification step by column chromatography (petrol ether (4070 uC)acetone = 50 : 1) furnished 456 mg (29%) of a light yellow solid, mp 79 uC; 1H NMR (400 MHz, CDCl3): d = 0.87 (m, 18H; CH3), 1.261.85 (m, 120H; CH2), 3.97, 4.02 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.92 (AA9BB9, 2H; aromat. H), 6.96 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.01, 7.15 (2d, 2H, 3J = 16.4 Hz; olefin. H), 7.03 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.14 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.19 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.488 (AA9BB9, 2H; aromat. H), 7.493, 7.53 (AA9BB9, 8H; aromat. H), 7.52 (s, 3H; 2H, 4H, 6H); 13C NMR (100 MHz, CDCl3): d = 14.1 (CH3), 22.731.9 (CH2), 55.3 (OCH3), 69.3, 73.6 (OCH2), 105.6, 114.3, 123.6, 123.8 (aromat. CH), 126.3, 128.9 (olefin. CH), 126.7, 126.9 (aromat. CH), 127.3 (olefin. CH), 127.8 (aromat. CH), 128.3, 128.9, 129.0 (olefin. CH), 130.1, 132.6, 136.5, 137.0, 138.1, 138.5, 138.7 (Cq, CqO), 153.4 (CqO); FD MS: m/z (%): 1723.7 (76, M+?), 1724.7 (100), 1725.7 (82), 1726.7 (41); elemental analysis: calcd for C119H182O7: C 82.87, H 10.64; found C 83.09, H 10.70. (E,E,E,E,E)-1,3-Bis{2-[4-(2-{3,4,5-tridodecyloxyphenyl}ethenyl)phenyl]ethenyl}-5-[2-(2-methylphenyl)ethenyl]benzene (3b) Preparation analogous to 3a using diethyl 2-methylbenzylphosphonate 4b.19 The precipitated product was purified by column chromatography (hexaneacetone = 30 : 1). Yield 436 mg (28%) of a yellow solid, mp 46 uC; 1H NMR (400 MHz, CDCl3): d = 0.87 (t, 18H, CH3), 1.251.85 (m, 120H, CH2), 2.47 (s, 3H, ArCH3), 3.96, 4.02 (t, 12H, OCH2), 6.71 (s, 4H; aromat. H), 6.96 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.03 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.16, 7.20 (2d, 3J = 16.1 Hz, 4H; olefin. H), 7.177.25 (m, 3H; aromat. H), 7.03, 7.41 (2d, 3J = 16.2 Hz, 2H; olefin. H), 7.50, 7.53 (AA9BB9, 8H; aromat. H), 7.54 (s, 2H; 4H, 6H), 7.58 (s, 1H; 2H), 7.62 (m, 1H; aromat. H); 13C NMR (100 MHz, CDCl3): d = 14.1 (CH3), 20.0 (ArCH3), 22.731.9 (CH2), 69.4, 73.6 (OCH2), 105.6, 123.7, 124.1, 125.5, 126.2, 126.7, 126.9 (aromat. CH), 126.7, 129.8, 127.3 (olefin. CH), 127.7 (aromat. CH), 128.2, 128.96, 129.02 (olefin. CH), 130.5 (aromat. CH), 132.5, 135.9, 136.37, 136.43, 137.1, 138.2, 138.5, 138.7 (Cq, CqO), 153.4 (CqO); FD MS: m/z (%): 1708.1 (56, M+?), 1709.1 (100), 1710.2 (79), 1711.1 (16); elemental analysis: calcd for C119H182O6: C 83.65, H 10.74; found C 83.38, H 10.82. (E,E,E,E,E)-1,3-Bis{2-[4-(2-{3,4,5-tridodecyloxyphenyl}ethenyl)phenyl]ethenyl}-5-(2-phenylethenyl)benzene (3c) Preparation analogous to 3a using diethyl benzylphosphonate 4c.20 The precipitated product was purified by column chromatography on basic alumina (petrol ether (4070 uC) acetone = 95 : 1) and subsequent middle pressure liquid chromatography (MPLC) on silica (petrol ether (4070 uC) acetone = 50 : 1), which afforded 568 mg (36%) of a yellow solid, mp 80 uC; 1H NMR (400 MHz, CDCl3): d = 0.88 (t, 18H; CH3), 1.261.86 (m, 120H; CH2), 3.97, 4.02 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.97 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.03 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.14 (d, 3J = 16.4 Hz, 3H; olefin. H), 7.20 (d, 3J = 16.4 Hz, 2H; olefin. H),
This journal is The Royal Society of Chemistry 2006

7.21 (d, 3J = 16.4 Hz, 1H; olefin. H), 7.28 (m, 1H; aromat. H), 7.38 (m, 2H; aromat. H), 7.50, 7.53 (AA9BB9, 8H; aromat. H), 7.55 (s, 3H; 2H, 4H, 6H), 7.55 (m, 2H; aromat. H); 13C NMR (100 MHz, CDCl3): d = 14.0 (CH3), 22.731.9 (CH2), 69.3, 73.5 (OCH2), 105.5, 123.9, 126.6, 126.7, 126.9 (aromat. CH), 127.2 (olefin. CH), 127.7 (aromat. CH), 128.1, 128.4, 129.2 (olefin. CH), 128.7 (aromat. CH), 128.8, 129.0 (olefin. CH), 132.5, 136.4, 137.0, 137.3, 138.1, 138.6 (Cq, CqO), 153.3 (CqO); FD MS: m/z (%): 1693.6 (72, M+N), 1694.5 (100), 1695.5 (56), 1696.7 (44); elemental analysis: calcd for C118H180O6: C 83.63, H 10.71; found C 83.63, H 10.73. (E,E,E,E,E)-1,3-Bis{2-[4-(2-{3,4,5-tridodecyloxyphenyl}ethenyl)phenyl]ethenyl}-5-[2-(4-bromophenyl)ethenyl]benzene (3d) NaH (0.30 g, 12 mmol) was given to 60 ml THF under argon and cooled with an ice bath to 0 uC. A mixture of aldehyde 2 (1.52 g, 0.93 mmol) and diethyl 4-bromobenzylphosphonate 4d21 (0.32 g, 1.05 mmol) in 40 ml THF were added dropwise. The reaction mixture was then stirred at RT for 24 h, poured on 100 g ice and 50 ml HCl (2 N) were added. The precipitate was collected and recrystallised from CHCl3ethanol = 1 : 1. 1.01 g of the yellow solid (1.50 g) was then further purified by column chromatography (silica, CH2Cl2hexane = 4 : 6) Yield 0.42 g (37%) of light yellow solid, Tcl = 65 uC; 1H NMR (400 MHz, CDCl3): d = 0.87 (m, 18H; CH3), 1.251.85 (m, 120H; CH2), 3.96, 4.01 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.96 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.03 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.12 (AB, 2H; olefin. H), 7.13 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.19 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.40 (AA9BB9, 2H; aromat. H), 7.50 (m, 12H; aromat. H), 7.56 (s, 1H; 2H); 13C NMR (100 MHz, CDCl3): d = 14.1 (CH3), 22.731.9 (CH2), 69.3, 73.6 (OCH2), 105.5 (aromat. CH), 121.5 (Cq), 124.0, 124.2, 126.8, 126.9 (aromat. CH), 127.2 (olefin. CH), 128.0, 129.2 (olefin. and aromat. CH partially superimposed), 129.1 (olefin. CH), 131.8 (aromat. CH), 132.5, 136.2, 136.3, 137.1, 137.7, 138.2, 138.7 (Cq, CqO), 153.4 (CqO); FD MS: m/z (%): 1771.7 (76, M+?), 1772.9 (91), 1773.8 (100), 1774.9 (86), 1776.1 (69); elemental analysis: calcd for C118H179BrO6: C 79.91, H 10.17; found C 79.72, H 10.20. (E,E,E,E,E)-1,3-Bis{2-[4-(2-{3,4,5-tridodecyloxyphenyl}ethenyl)phenyl]ethenyl}-5-[2-(4-cyanophenyl)ethenyl]benzene (3e) Preparation analogous to 3d using diethyl 4-cyanobenzylphosphonate 4e.23 The crude product was purified by column chromatography on silica (toluene) and subsequent recrystallisation from CHCl3ethanol = 1 : 1, which afforded 496 mg (31%) of a yellow solid, Tcl = 92 uC; 1H NMR (400 MHz, CDCl3): d = 0.87 (m, 18H; CH3), 1.261.85 (m, 120H; CH2), 3.97, 4.02 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.96 (d, 3 J = 16.4 Hz, 2H; olefin. H), 7.03 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.13 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.16, 7.24 (2d, 3J = 16.2 Hz, 2H; olefin. H), 7.19 (d, 3J = 16.4 Hz, 2H, olefin. H), 7.50, 7.53 (AA9BB9, 8H, aromat. H), 7.54 (s, 2H; 4H, 6H), 7.59 (s, 1H; 2H), 7.60, 7.65 (AA9BB9, 4H; aromat. H); 13C NMR (100 MHz, CDCl3): d = 14.0 (CH3), 22.631.9 (CH2),
J. Mater. Chem., 2006, 16, 441451 | 449

69.3, 73.5 (OCH2), 105.5 (aromat. CH), 110.8, 118.9 (Cq), 124.2, 124.7, 126.7, 126.93, 129.90 (aromat. CH), 127.1, 127.3, 132.1, 127.8, 129.1, 129.2 (olefin. CH), 132.4 (aromat. CH and Cq superimposed), 136.2, 137.1, 137.2, 138.3, 138.7, 141.7 (Cq, CqO), 153.4 (CqO). FD MS: m/z (%): 1719.0 (86, M+?), 1720.0 (100), 1721.0 (70), 1722.0 (34), 1723.0 (12); elemental analysis: calcd for C119H179NO6: C 83.11, H 10.49, N 0.81; found C 83.04, H 10.34, N 0.89. (E,E,E,E,E)-4-{2-[3,5-Bis(2-{4-[2-(3,4,5-tridodecyloxyphenyl)ethenyl]phenyl}ethenyl)phenyl]ethenyl}benzoic acid amide (3f) Preparation analogous to 3a using diethyl 4-cyanobenzylphosphonate 4e.23 The precipitated product was purified by column chromatography (MPLC) on silica (petrol ether (40 70 uC)acetone = 3 : 2), which yielded 430 mg (27%) of a waxy yellow solid, Tcl = 189 uC; 1H NMR (400 MHz, CDCl3): d = 0.86 (m, 18H; CH3), 1.251.85 (m, 120H; CH2), 3.96, 4.01 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.96 (d, 3J = 16.1 Hz, 2H; olefin. H), 7.03 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.13 (d, 3 J = 16.2 Hz, 2H; olefin. H), 7.19 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.19, 7.24 (2d, 3J = 16.4 Hz, 2H; olefin. H), 7.49, 7.53 (AA9BB9, 8H; aromat. H), 7.55 (s, 2H; aromat. H), 7.57 (s, 1H; aromat. H), 7.60, 7.82 (AA9BB9, 4H; 2H, 3H, 5H, 6H); 13C NMR (100 MHz, CDCl3): d = 14.0 (CH3), 22.731.9 (CH2), 69.3, 73.6 (OCH2), 105.5, 124.1, 124.4, 126.6, 126.8, 127.0 (aromat. CH), 127.2 (olefin. CH), 127.9 (aromat. CH), 128.0, 128.1, 132.2, 129.1 (olefin. CH), 130.7, 132.5, 136.3, 137.1, 137.5, 138.2, 138.7, 140.9 (Cq, CqO), 153.4 (CqO), 168.8 (1Cq, amide); FD MS: m/z (%): 1737.2 (100, M+?), 1738.2 (77), 1739.2 (53), 1740.3 (29); elemental analysis: calcd for C119H181NO7: C 82.25, H 10.50, N 0.81; found C 82.05, H 10.62, N 0.76. (E,E,E,E,E)-4-{2-[3,5-Bis(2-{4-[2-(3,4,5-tridodecyloxyphenyl)ethenyl]phenyl}ethenyl)phenyl]ethenyl}benzaldehyde (3g) Preparation analogous to 3d using diethyl 4-diethoxymethylbenzylphosphonate 4f.22 After 24 h, the reaction mixture was poured on 100 g ice, 50 ml HCl (18%) and 100 ml CHCl3 were added. The mixture was stirred until the cleavage of the acetal was completed (approx. 3 h). The crude product was purified by column chromatography on silica (petrol ether (4070 uC) toluene = 1 : 4). Yield 501 mg (35%) of a yellow solid, mp = 100 uC; 1H NMR (400 MHz, CDCl3): d = 0.88 (m, 18H; CH3), 1.261.85 (m, 120H; CH2), 3.97, 4.02 (2t, 12H; OCH2), 6.71 (s, 4H; aromat. H), 6.96 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.03 (d, 3 J = 16.2 Hz, 2H; olefin. H), 7.13 (d, 3J = 16.2 Hz, 2H; olefin. H), 7.18 (d, 3J = 16.4 Hz, 2H; olefin. H), 7.20, 7.28 (2d, 3J = 16.4 Hz, 2H; olefin. H), 7.49, 7.53 (AA9BB9, 8H; aromat. H), 7.55 (s, 2H; 4H, 6H), 7.57 (s, 1H; 2H), 7.67, 7.88 (AA9BB9, 4H; aromat. H), 9.98 (s, 1H, CHO); 13C NMR (100 MHz, CDCl3): d = 14.1 (CH3), 22.731.9 (CH2), 69.4, 73.6 (OCH2), 105.7, 124.2, 124.7, 126.8, 127.0, 127.0, 130.3 (aromat. CH), 127.2, 127.9, 128.0, 132.0, 129.2, 129.3 (olefin. CH), 132.5, 135.6, 136.3, 137.2, 137.5, 138.4, 138.9, 143.4 (Cq, CqO), 153.4 (CqO), 191.4 (CHO); FD MS: m/z (%): 1721.6 (67, M+?), 1722.6 (100), 1723.5 (55), 1724.6 (10); elemental analysis: calcd for C119H180O7: C 82.97, H 10.53; found C 82.67, H 10.37.
450 | J. Mater. Chem., 2006, 16, 441451

Acknowledgements
We are grateful to Prof. Jochen Gutmann and Michael Bach for their strong support during the X-ray measurements, to Dr Volker Abetz for measurements with the Kratky camera and to the Deutsche Forschungsgemeinschaft and the Fonds der Chemischen Industrie for financial support.

References
1 H. Meier, Angew. Chem., Int. Ed. Engl., 1992, 31, 1399. 2 J. H. Burroughes, D. D. C. Bradeley, A. R. Brown, R. N. Marks, K. Mackay, R. H. Friend, P. L. Burn and A. B. Holmes, Nature, 1990, 347, 539. 3 T. Piok, C. Brands, P. J. Neyman, A. Erlacher, C. Soman, M. A. Murray, R. Schroeder, W. Graupner, J. R. Heflin, D. Marciu, A. Drake, M. B. Miller, H. Wang, H. Gibson, H. C. Dorn, G. Leising, M. Guzy and R. M. Davis, Synth. Met., 2001, 116, 343. 4 H. Meier and M. Lehmann, Angew. Chem., 1998, 110, 666 (Angew. Chem., Int. Ed., 1998, 37, 643). 5 H. Meier and M. Lehmann, Encyclopedia of Nanoscience and Nanotechnology, vol. 10, ed. Nalwa, H. S., p. 95, American Scientific Publishers, Stevensons Ranch, 2005. 6 R. Beavington, M. J. Frampton, J. M. Lupton, P. L. Burn and I. D. W. Samuel, Adv. Funct. Mater., 2003, 13, 211. 7 (a) A. Adam, P. Schuhmacher, J. Simmerer, L. Ha usling, K. Siemmensmeyer, K. H. Etzbach, H. Ringsdorf and D. Haarer, Nature, 1994, 371, 141; (b) A. M. van de Craats, J. M. Warman, M. P. de Haas, D. Adam, J. Simmerer, D. Haarer and P. Schuhmacher, Adv. Mater., 1996, 8, 823. 8 M. Lehmann, I. Fischbach, H. W. Spiess and H. Meier, J. Am. Chem. Soc., 2004, 126, 772. 9 H. Meier, M. Lehmann and U. Kolb, Chem. Eur. J., 2000, 6, 2462. 10 C. Tschierske, J. Mater. Chem., 2001, 11, 2647. 11 A. J. Attias, C. Cavelli, B. Donnio, D. Guillon, P. Hapiot and J. Malthete, Chem. Mater., 2002, 14, 375. 12 H. C. Holst, T. Pakula and H. Meier, Tetrahedron, 2004, 60, 6765. 13 (a) V. Percec, W.-D. Cho, G. Ungar and D. J. P. Yeardley, Angew. Chem., 2000, 112, 1661(Angew. Chem., Int. Ed., 2000, 39, 1597); (b) G. Johansson, V. Percec, G. Ungar and J. P. Zhou, Macromolecules, 1996, 29, 646. 14 C. Tschierske, Annu. Rep. Prog. Chem., Sect. C, 2001, 97, 191. 15 V. Percec, M. Glodde, T. K. Bera, Y. Miura, I. Shiyanovskaya, K. D. Singer, V. S. K. Balagurusamy, P. A. Heiney, I. Schnell, A. Rapp, H. W. Spie, S. D. Hudson and H. Duan, Nature, 2002, 417, 384. 16 M. Lehmann, Ph.D. Dissertation, University of Mainz, 1999. 17 M. Lehmann, B. Schartel, M. Hennecke and H. Meier, Tetrahedron, 1999, 55, 13377. 18 F. Albrecht, F. Frickel, R. Schlecker and P. C. Thieme, Synthesis, 1979, 712. 19 F. Boberg, A. Jackiewicz and A. Garming, Phosphorus, Sulfur Silicon Relat. Elem., 1992, 72, 1. 20 S. Gronowitz, K. Stenhammar and L. Svensson, Heterocycles, 1981, 15, 947. 21 G. Kennedy and A. D. Perboni, Tetrahedron Lett., 1996, 37, 7611. 22 E. Sugiono, T. Metzroth and H. Detert, Adv. Synth. Catal., 2001, 343, 351. 23 F. Kagan, R. D. Birkenmeyer and R. E. Strube, J. Am. Chem. Soc., 1959, 81, 3026. 24 1H NMR data, UV-Vis absorption spectra and FT-IR spectrum of a thin film of 3f are available in the electronic supplementary information. 25 B. Krasovitskii and B. M. Bolotin, Organic Luminescent Materials, VCH Weinheim, 1988, p. 46. 26 Y. Bouligand, J. Phys., 1980, 41, 1307. 27 The correlation length was determined using the Scheerer formula and the half width and reflection maximum obtained from the fit function. (a) R. Jenkins and R. L. Snyder, Introduction to X-ray Powder Diffractometry, Chemical Analysis, vol. 138, ed. J.D. Winefordner, Wiley, New York, 1996; (b) P. Scheerer, Nachr.

This journal is The Royal Society of Chemistry 2006

28

29

30 31 32

Ges. Wiss. Goettingen, Math. Phys. Kl., Fachgruppe 1, 1918, 2, 96100. The preliminary apparent model of antiparallel stacked mesogens agrees well with the columnar diameter, however, the assembly of four molecules in a columnar slice of only 4.05 A in height (hcol) is not yet evident. The experimental values (hcol and z) may be explained, if two antiparallel aligned mesogens are inclined versus the columnar axis (see also ref. 29a) and a helical arrangement is assumed. (a) B. Donnio, B. Heinrich, H. Allouchi, J. Kain, S. Diele, D. Guillon and D. W. Bruce, J. Am. Chem. Soc., 2004, 126, 15258; (b) M. Marcos, R. Gimenez, J.-L. Serrano, B. Donnio, B. Heinrich and D. Guillon, Chem. Eur. J., 2001, 7, 1006. The stilbenoid scaffold can not back-fold. It can only rotate about the single bonds, which will not change the molecular extension. (a) Beilstein E IV 9/2, 892; (b) Beilstein E IV, 7/2, 506; (c) Beilstein E IV, 9/2, 725. The dipoles were also calculated from AM1 minimised models (Chem3D Ultra 8.0): Dipoles along the bisect of 4.08 D for the

33

34 35 36 37 38

cyano derivative and 3.35 D for the aldehyde derivative are very close to the measured dipoles of cyanobenzene and benzaldehyde (see ref. 30). (a) G. R. Desiraju in Comprehensive Supramolecular Chemistry, vol. 6, ed. J. L. Atwood, J. E. D. Davies, D. D. Macnicol, F. Vogtle and J.-M. Lehn, pp. 122, Elsevier Science Ltd., Oxford, 1996; (b) D. A. Dougherty in Comprehensive Supramolecular Chemistry, Vol. 2, ed. J. L. Atwood, J. E. D. Davies, D. D. Macnicol, F. Vogtle, J.-M. Lehn, pp. 195209, Elsevier Science Ltd., Oxford, 1996. U. Beginn and G. Lattermann, Mol. Cryst. Liq. Cryst., 1994, 241, 215. R. I. Gearba, A. Bondar, M. Lehmann, B. Goderis, W. Bras, M. H. J. Koch and D. A. Ivanov, Adv. Mater., 2005, 17, 671. M. Kasha, H. R. Rawls and M. Ashraf El-Bayoumi, Pure Appl. Chem., 1965, 11, 371. R. Petermann, C. Schnorpfeil, M. Lehmann, M. Fetten and H. Meier, J. Inf. Rec., 2000, 25, 259. S. S. Soomro, R. Benmouna, R. Berger and H. Meier, Eur. J. Org. Chem., 2005, 16, 3586.

This journal is The Royal Society of Chemistry 2006

J. Mater. Chem., 2006, 16, 441451 | 451

You might also like