24.12.2013 Views

Chromonic mesophases

Chromonic mesophases

Chromonic mesophases

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

<strong>Chromonic</strong> <strong>mesophases</strong><br />

John Lydon<br />

University of Leeds, The School of Biochemistry and Molecular Biology, Leeds LS2 9JT, UK<br />

Abstract<br />

Over the last 10 years, there has been a growing acceptance of the concept of chromonic phases and a wider recognition that<br />

they form a well-defined family of lyotropic liquid crystalline phases, with a package of properties distinct in almost every aspect,<br />

from those of conventional amphiphiles. New chromonogenic compounds have appeared and new technological uses for chromonic<br />

systems are being actively explored. Recent promising investigations include the synthesis of a chromonic dye, C.I. Direct Blue<br />

67, which has an N phase of high order parameter and which can be dried down to give well-oriented films of solid. When dried<br />

down on a ‘command surface’ of photoaligned substrate this can produce a highly patterned film. The use of chromonic phases<br />

in the construction of compensating plates for improving the viewing characteristics of twisted nematic displays has been explored.<br />

Although this technology may not be suitable for commercially exploitation in its present form, the success of the devices is<br />

significant. It is suggested that current studies of the way in which the temperature range of thermotropic discotic <strong>mesophases</strong> is<br />

enhanced in 1:1 CPI mixtures may well lead to improved formulations for chromonic dyes. It is predicted that the marriage of<br />

chromonic phase technology with current biochemical analytical techniques will give rise to a new generation of medical<br />

diagnostic tests.<br />

2004 Elsevier Ltd. All rights reserved.<br />

1. Introduction<br />

About two decades ago it was becoming clear that<br />

there is an extensive and well-defined family of lyotropic<br />

mesogens with properties distinct from those of conventional<br />

amphiphiles. This family consists of various<br />

drugs, dyes, nucleic acids, antibiotics, carcinogens and<br />

anticancer agents. In almost every respect the properties<br />

are different to those of ordinary lyotropic mesogens of<br />

the soapydeterergentyphosphilipid type. The molecules<br />

have aromatic rather than aliphatic structures. They are<br />

rigid rather than flexible and planar disc-like or planklike,<br />

rather than rod-like. The hydrophilic solublising<br />

groups are disposed around the periphery of the molecules<br />

rather than at one end. The molecules aggregate<br />

in solution, not into micelles, but into columns. They<br />

have distinctive optical textures (involving characteristic<br />

types of paramorphosis). Thermodynamic measurements<br />

indicate that the driving force for the aggregation is<br />

enthalpic rather than entropic. Their phase diagrams<br />

tend to be of the peritectic rather than eutectic type.<br />

They do not show cmc’s or Krafft points. These are the<br />

chromonic <strong>mesophases</strong><br />

E-mail address: j.e.lydon@leeds.ac.uk (J. Lydon).<br />

This review is intended to be read as an update of<br />

●●<br />

two previous papers. The first w1 x, in The Handbook<br />

of Liquid Crystals, gives details of the prehistory of<br />

chromonic phases, the patterns of molecular aggregation<br />

in N an M phases and NyM phase diagrams and includes<br />

a description and analysis of characteristic optical textures.<br />

The second w2 x, in this journal 5 years ago,<br />

●●<br />

described the newer brickwall and chimney types of<br />

aggregation encountered in some chromonic dyeywater<br />

systems and discussed the extension of the Israelachvili<br />

approach from lyotropic amphiphile systems to chromonics<br />

(Figs. 1 and 2).<br />

The recent developments discussed in this review fall<br />

into three categories. The first concerns the use of an<br />

optical alignment technique to produce detailed patterns<br />

of alignment in dried down films of dyes. The second<br />

concerns the experimental use of chromonic phases in<br />

compensating plates to improve the performance of the<br />

standard twisted nematic display devices. The third<br />

concerns CPI mixtures—a theme of current interest in<br />

thermotropic discotic systems, which I predict will<br />

become of importance in the future technological exploitation<br />

of chromonic systems.<br />

1359-0294/04/$ - see front matter 2004 Elsevier Ltd. All rights reserved.<br />

doi:10.1016/j.cocis.2004.01.006


J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

481<br />

Fig. 1. The Structures of <strong>Chromonic</strong> N and M Phases. <strong>Chromonic</strong> mesogens can be regarded as being insoluble in one dimension. They tend to<br />

aggregate face-to-face, producing a variety of stacked structures. The classic chromonic phases are the nematic, N phase and the hexagonal, M<br />

phase. In both of these, the molecules are stacked in columns. In the N phase, these lie in a nematic array (i.e. the columns are more or less<br />

parallel, but there is no positional order and there is no orientational order of the columns about their long axes). In the M phase, the columns<br />

lie on a lattice with statistical hexagonal symmetry and have long-range order. Other chromonic structures have been identified by Tiddy et al.<br />

and Harrison et al. where the molecules are aggregated into brickwork patterns or cylindrical chimneys w2,18,19x.<br />

2. Outline of chromonic phase structure and<br />

properties<br />

The name chromonic was derived from the bischromone<br />

structure of the widely marketed anti-asthmatic<br />

known in the UK as INTAL and in the US as<br />

Chromolyn w3–10x—by no means the first mesogen of<br />

this type to be reported—but one of the most extensively<br />

studied. It was considered (by its creator at least) to be<br />

a particularly good name because of the fortuitous<br />

combination of connotations of the word, with both<br />

colour (with reference to dyestuffs) and with chromosomes<br />

(with reference to nucleic acids).<br />

<strong>Chromonic</strong> <strong>mesophases</strong> are the lyotropic counterparts<br />

of the discotic <strong>mesophases</strong>-and there are some parallels<br />

in their history. In both cases, the definition of the<br />

concept came far later than one would have expected.<br />

The so-called ‘carbonaceous phases’ had been known<br />

and characterised by the coking industry for decades<br />

and there were predictions of ‘negative nematic’ liquid<br />

crystalline phases years before Chandrasekhar’s classic<br />

work. Similarly, references to the aggregation of dye<br />

molecules, stacking like piles of pennies or packs of<br />

cards were scattered in the dye chemistry literature for<br />

almost a century. Terms such as H- and J-aggregates<br />

were widely used in the industry—but there was scarcely<br />

ever a mention of liquid crystalline properties.<br />

In many aspects, chromonic systems are closer to<br />

thermotropic systems than to conventional amphiphiles.<br />

In both cases, the driving force causing liquid crystalline<br />

phase formation is the face-to-face aggregation of molecules<br />

forming columns—and the geometrical aspects<br />

of the packing of these columns are more or less the<br />

same in the two cases—the difference being that in one<br />

case the columns lie in a sea of alkyl chains and in the<br />

other, they lie in a sea of water. I had expected that by<br />

now, chromonic counterparts of all of the thermotropic<br />

discotic phases would have been found. Bearing in mind<br />

the extensive literature of the dye industry concerning<br />

tilted stacks of dye molecules (J-aggregates), I find it<br />

surprising that there are, as yet, no well-authenticated<br />

tilted chromonic systems.<br />

In general, there is a strong tendency for chromonic<br />

molecules to aggregate into columns, even in very dilute<br />

solution—just as conventional lyotropic mesogens form<br />

micelles before a mesophase is formed w11,12x.<br />

Although there may be a threshold concentration before<br />

significant aggregation begins to occur, there is no<br />

specific optimum column length and, therefore no analogue<br />

of the critical micelle concentration (cmc) of<br />

Fig. 2. A snapshot plan view of the chromonic M phase. The array<br />

shown in this sketch has orthorhombic symmetry – but the rotational<br />

disorder of the columns between the three possible orientations (as<br />

indicated for the central column) leads to the phase having overall<br />

hexagonal symmetry. The hexagonal lattice spacing is approximately<br />

half the length of the molecule plus half the width of the molecule<br />

plus the thickness of the interlying water w1,4,10,29x. For the dye<br />

molecule shown in Fig. 3, this is approximately 24 A. ˚


482 J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

conventional amphiphiles. The term ‘isodesmic’ (first<br />

used in the study of the aggregation of nucleic acids in<br />

solution) has been applied to the steady build up of<br />

chromonic aggregates where the addition or removal of<br />

one molecule to a stack is always associated with the<br />

same increment of free energy w13,14x. This is in direct<br />

contrast with the situation for conventional amphiphile<br />

association, where the micelle represents a free energy<br />

minimum—and there is a cost to the system in having<br />

either larger or smaller units.<br />

A further fundamental distinction between conventional<br />

amphiphile and chromonic systems concerns what<br />

happens at the lower temperature limit of mesophase<br />

formation. In conventional amphilphiles, there are two<br />

micro-phase regions; the aqueous and the hydrophobic,<br />

aliphatic parts. As the temperature is lowered, the alkyl<br />

chain motion in the hydrophobic region usually freezes<br />

out, forming a gel phase. The system becomes too brittle<br />

to be able to pack into the micelles required for<br />

mesophase formation. This gives a lower temperature<br />

limit characterised by its Krafft temperature. In chromonic<br />

systems the opposite happens. Because of the<br />

absence of alkyl chains in chromonic systems (or at<br />

least the absence of significant lengths of alkyl chains),<br />

chromonic systems do not show a Krafft point and the<br />

lower temperature is limit is marked by the appearance<br />

of ice—usually a few degrees below 0 8C. Note that<br />

there is evidence that one can access monotropic chromonic<br />

phases at sub-zero temperatures by adding an<br />

antifreeze to the system w7,9x.<br />

Misciblity is a feature of liquid crystalline systems –<br />

and in general, an understanding of the rules which<br />

govern miscibility is crucial to our understanding of the<br />

factors which determine the dynamics of each type of<br />

mesophase. The chromonic analogue of miscibility is<br />

intercalation where guest molecules can be accepted<br />

randomly into chromonic stacks. Although there have<br />

been no extended systematic studies of chromonic miscibility<br />

it appears that this is as widespread as miscibility<br />

in other mesophase systems w15x.<br />

There is another related pattern of behaviour of<br />

chromonic mixtures (which is discussed in more detail<br />

below). This occurs where there is a strong preference<br />

for an ABAB alternating arrangement of the two components<br />

in every stack w47–53x. This occurs to such a<br />

pronounced extent that the alternating column must be<br />

regarded as the structural unit of the phase. Since at<br />

least some of these CPI ‘compounds’ give <strong>mesophases</strong><br />

with enhanced stability (and since the aromatic cores of<br />

the compounds in both families of mesophase are<br />

similar), it is an obvious suggestion that the search for<br />

similar effects in chromonic systems would be<br />

worthwhile.<br />

<strong>Chromonic</strong> systems can be doped with small soluble<br />

chiral compounds to give a chiral N phase (in an<br />

Fig. 3. The molecular structure of C.I. Direct Blue 67. This new dye<br />

forms a chromonic N phase of high order parameter and which can<br />

be dried down to give an aligned solid film. Matsunaga et al. w29x<br />

have shown that a highly patterned film can be produced by drying<br />

down the N phase on a photaligned command substrate as sketched<br />

in Fig. 5.<br />

analogous way to the chiral doping of thermotropic<br />

nematics). Since the twist produced is proportional to<br />

the concentration of the dopant, this property has been<br />

proposed as a practical assay for chiral compounds<br />

w7,9x. A new exploitation of the chirally-doped N phase<br />

is in compensating devices for TN cells as described<br />

below w38–40x.<br />

3. New chromonic materials<br />

The range of compounds which form chromonic<br />

phases now includes xanthoses (used as antiasthmatic<br />

drugs, azo dyes, cyanine dyes, nucleic acids (guanosine<br />

derivatives) and perylenes w16–26x.<br />

Over the last few years, reports of new chromonic<br />

<strong>mesophases</strong> have grown from a trickle to a stream.<br />

These include a report of a new metallo-mesogen and<br />

excitingly, a non-aqueous chromonic system (where the<br />

solvent is dimethyl formamide) w27,28x (Fig. 3).<br />

Significant amongst new chromonic materials is the<br />

●<br />

azo dye, C.I Direct 67 w29 x. This shows typical NyM<br />

phase behaviour promises to become a useful new<br />

chromonic compound. The chromonic phases formed by<br />

aqueous solutions of this dye were investigated by means<br />

of temperature-controlled X-ray diffraction, polarised<br />

light microscopy and UV–visible spectroscopy. The dye<br />

molecules form columnar aggregates, even at low concentrations.<br />

The structural study gave results very similar<br />

to those of earlier work on the INTALywater system<br />

w5,6,10x. As one would expect, the stacking repeat was<br />

3.4 A ˚ (and was found to be independent of concentration<br />

and temperature). The diameter of the column was<br />

found to be comparable with the length of the molecule—suggesting<br />

that the column is a unimolecular<br />

stack. One interesting feature of this investigation was<br />

the observation that the addition of a small amount of<br />

anionic surfactant (0.01% by wt.) was found to enhance<br />

the stability of the nematic phase (at the expense of the


J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

483<br />

Fig. 4. The production of an aligned dye coated polymer film using<br />

a ‘command plate’ produced by the Weigert effect. A spin-coated film<br />

of the azo compound is photaligned with a beam of plane polarised<br />

light to produce a ‘command plate’.<br />

M phase). Clearly at this concentration there is no<br />

suggestion of the added amphiphile having a direct<br />

structural effect on the mesophase such as coating the<br />

columns. The effect must be more subtle. Perhaps in<br />

some way it stabilises column ends or simply alters the<br />

chemical potential of the counter ions.<br />

Over the last 10 years, a photoalignment approach<br />

has been developed which appears to have the potential<br />

for achieving this w30–37x. This technique utilises the<br />

Weigert effect. This is the sensitivity of some photohemical<br />

reactions to the orientation of the plane of polarised<br />

light striking the molecule. Reactions for which the<br />

Weigert effect has been observed include photobleaching,<br />

photodimerization and photoisomerism. In the photoalignment<br />

of photoisomerisable molecules the final<br />

state of the sample has the molecular director aligned<br />

normal to the electric vector of the incident light (Figs.<br />

4 and 5).<br />

Azobenzene has been a favourite compound for producing<br />

aligned films using the Weigert effect. Unfortunately,<br />

it is only weakly absorbing in the visible<br />

wavelengths range and it is, therefore by no means<br />

ideal. There is a way round this problem, however. It<br />

has been found that a film of phoaligned azobenzene<br />

molecules is able to epitaxially align liquid crystalline<br />

phases. The photoaligned substrate can, therefore be<br />

used as a ‘command surface’, which is in turn able to<br />

direct the alignment if the mesophase. Photoinduced<br />

alignment of this kind was first demonstrated with<br />

thermotropic <strong>mesophases</strong> (using films of azodoped polymer,<br />

polymers with azobenzene side groups or polymers<br />

with cinnamic acid side groups). However, it also works<br />

for lyotropic phases and can be used to align the<br />

chromonic N phase.<br />

●<br />

In their recent paper, Matsunga et al. w30 x describe<br />

the production of a patterned film of dye using this<br />

approach. They used a striped template command surface<br />

prepared from the photoinduced alignment of a<br />

polyamide with azo side-groups. The chromonic N phase<br />

of the dye, C.I. Direct Blue 67 is aligned by this surface<br />

and then dried down to give a patterned film. They<br />

4. The production of patterned dye films<br />

In the production of optical elements such as polarisers,<br />

retarders and optical compensators, it is necessary<br />

for us to be able to control the alignment of birefringent<br />

material embedded in (or deposited on) a film. A variety<br />

of manufacturing processes have been tried, including<br />

classical mechanical methods such as stretching the<br />

film, shearing, rubbing the surface and newer approaches<br />

such as electric field poling. The disadvantage of all of<br />

these approaches is that they are only able to produce<br />

surface films with same alignment over the entire area<br />

of film being treated. What would be much more<br />

desirable is a technique which could align small areas<br />

of film in different orientations, ultimately giving us the<br />

ability to align individual pixels in any required<br />

orientation.<br />

Fig. 5. The epitaxial alignment of a chromonic N phase by a photoaligned<br />

command plate. (redrawn from Ref. w30x).


484 J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

Fig. 6. Conventional (uncompensated) twisted nematic cells in the on<br />

and off states. There are two forms of the twisted nematic cell: normally<br />

white (NW) and normally black (NB) cells – the difference<br />

lies only in the orientation of the upper polarising layer plate.<br />

The twisted nematic (TN) device continues to dominate<br />

the flat panel liquid crystal display market. However,<br />

it is not without its shortcomings. The<br />

characteristics of an ideal device are high contrast, a<br />

wide viewing angle, good colour rendition and a completely<br />

achromatic dark state. The standard (uncompensated)<br />

TN device fails to live up to every one of these<br />

ideals. In particular, the phenomenon known as ‘greyscale<br />

inversion’ can be severe. This arises where the<br />

contrast between ‘on’ and ‘off’ areas falls to zero and<br />

is then reversed, as the viewing angle is changed.<br />

The defects of the standard TN cell can, in some<br />

measure, be corrected by the addition of a compensator.<br />

This is an optical system, which matches the optical<br />

characteristics of the sample and reduces the contrast<br />

loss as the viewing angle is increased. An ideal compensating<br />

device would ‘correct’ the optics of both the<br />

‘on’ and ‘off’ states, but this is an over-ambitious target<br />

at the present time. Current compensators are passive<br />

devices, which are designed to correct only the more<br />

critical of the two states of the device. In an uncompensated<br />

TN cell there is both leakage of light through the<br />

dark state and a decrease of intensity of the light state.<br />

However, the effects of these deficiencies are not symmetrical<br />

and the leakage of light through the dark state<br />

is the more serious. Because of this, compensators are<br />

specifically designed to improve only the dark state<br />

optics.<br />

The structures of the two variants of the standard (i.e.<br />

non-compensated) TN cell are shown in Fig. 6. In a<br />

cell in the ‘off’ state, the twisting molecular alignment<br />

rotates the plane of polarisation of the light through 908.<br />

In the ‘on’ state, the molecules are realigned to give a<br />

homeotropic state where the molecular axes lie perpendicular<br />

to the cell surface. This arrangement does not<br />

change the polarisation state of the light. There are two<br />

geometries that are used in these devices—with the two<br />

polarisers lying either parallel or perpendicular. The<br />

parallel arrangement gives a ‘normally black’ (NB)<br />

device in the absence of a field and the perpendicular<br />

arrangement gives a ‘normally white’ (NW) device.<br />

Compensators have, therefore been devised to improve<br />

the optics of the ‘off’ state of the NB device and the<br />

‘on’ state of the NW device.<br />

were able to produce a solid dye film consisting of<br />

alternating 300-mm wide rows of orthogonally aligned<br />

dye material with impressively sharp-edged resolution.<br />

5. The use of planar and twisted lyotropic chromonic<br />

liquid crystal cells as optical compensators for twisted<br />

nematic cells<br />

Fig. 7. A more accurate view of the mesophase alignment of the NW<br />

cell in the ‘on’ mode. To a first approximation, the applied field aligns<br />

the mesophase in a homeotropic pattern as sketched in Fig. 6. However,<br />

molecules near to the surfaces tend to retain a parallel alignment<br />

and the phase adopts a complex splay pattern shown.


J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

485<br />

7. Compensating a normally white cell<br />

For a NW cell, the situation is more complex. In the<br />

dark ‘on’ state the director field is not perfectly homeotropic<br />

and the arrangement sketched in Fig. 6 is only<br />

an approximation to reality. In practice, the situation is<br />

more like that sketched in Fig. 7 where the molecules<br />

near to the upper and lower substrate surfaces retain a<br />

parallel alignment and, as the director field curves<br />

towards the homeotropic region in the centre, there is<br />

appreciable splay distortion. A plate compensating specifically<br />

for this splay has been developed by Fuji.<br />

In addition to the splay distortion, the optics require<br />

correction for birefringence and Lavetrovitch et al.<br />

●<br />

w40 x have explored the use of chromonic cells to<br />

compensate for the birefringence and twist. The layout<br />

of their complete device is shown in Fig. 9.<br />

Fig. 8. The design of a compensated NB cell (Lavrentrovich et al.<br />

w40x). Compensators have been added to improve the optics of the<br />

dark ‘off’ state. The twisted compensating cell is filled with a chirally<br />

doped N* phase, the compensating A-plate contains a parallel-aligned<br />

chromonic mesophase. (redrawn from Ref. w40x).<br />

6. Compensating a normally black TN cell<br />

The root cause of the optical problems of a TN cell<br />

is the large positive birefringence of the cell. One can<br />

improve the optical performance of a NB display by<br />

compensating this with a passive retarder. The compensating<br />

plate should have negative birefringence and a<br />

twisted structure that mirrors that of the cell in the ‘off’<br />

mode. An early attempt at constructing a compensator<br />

with these properties consisted of several superposed<br />

polymer films with negative birefringence. These were<br />

stacked in a twisted arranged to match the twist director<br />

twist of the cell in the ‘off’ state. More sophisticated<br />

compensators have been constructed by Laventrovitch<br />

●<br />

et al. using liquid crystalline phases w38–49 x. To<br />

counteract the positive birefringence of the nematic<br />

phase, phases with negative optical anisotropy are<br />

required. This suggests the use of thermotropic discotic<br />

and lyotropic chromonic phases—and both have been<br />

tried.<br />

The design of an experimental compensated NB TN<br />

●<br />

cell devised by Laventovitch et al. w40 x is shown in<br />

Fig. 8. Note that this employs a combination of a planar<br />

N-phase chromonic cell and a twisted N* chromonic<br />

cell to match the uncompensated optics.<br />

Fig. 9. The design for a compensated NW cell (Lavrentrovich et al.<br />

w40x). Compensators have been added to improve the optics of the<br />

dark, ‘on’ state. The Fuji plates compensate for the splay shown in<br />

Fig. 7. The compensating A-plates contain a parallel-aligned chromonic<br />

mesophase. (redrawn from Ref. w40x).


486 J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

I have given prominence to the use of chromonic<br />

compensators in this review. This was not done, because<br />

I believe that devices of the kind described using the<br />

chromonic solutions are immediately feasible for largescale<br />

commercial production. The narrow temperature<br />

range of the chromonic phase used would alone make<br />

this impracticable. The importance lies in the way it<br />

stresses the versatility of this family of <strong>mesophases</strong>.<br />

8. Complementary polytopic interaction (C.P.I.) in<br />

discotic systems<br />

The so-called p–p interactions which hold chromonic<br />

molecules face to face have been discussed by Hunter<br />

et al. w41–43x and Bates and Luckhurst w44x. They argue<br />

that these interactions are in fact a combination of van<br />

der Waals forces and electrostatic interactions and that<br />

no specific properties of p systems need be invoked. In<br />

recent studies of thermotropic columnar <strong>mesophases</strong>, in<br />

the search for enhanced electronic conductors, a phenomenon<br />

has been encountered which suggests that in<br />

certain cases, two different molecules can form 1:1<br />

‘compounds’ with enhanced mesophase-forming properties<br />

w45–53x. In these cases, the structural unit is<br />

thought to be the (AB) n column. The reason why these<br />

alternating columns are ‘better’ mesogenic units than<br />

those of either of the two separate species is not<br />

immediately evident. Some form of ‘bonding’ must be<br />

occurring between the two types of molecules, which is<br />

not describable in classical chemical terms. The interaction<br />

appears to be more subtle than a simple covalent,<br />

hydrogen bonding or electrostatic effect. There is no<br />

evidence that the electronic structure of either molecule<br />

is significantly disturbed and there is no detectable<br />

charge-transfer interaction. The explanation for the complementary<br />

nature of the two types of molecule appears<br />

to lie in the sum total of a large number of atom-byatom<br />

van der Waals type interactions. The term complementary<br />

polytopic interaction (C.P.I.) has been coined.<br />

Computer modelling of the interactions between the two<br />

types of molecule in such systems has been carried out<br />

using the XED program and the results appears to<br />

confirm this view.<br />

The two component phase diagram of two compounds<br />

which interact in this fashion, recently reported by<br />

●<br />

Boden, Bushby and Lozman w53 x is shown in Fig. 10.<br />

Note that one of these compounds is mesogenic on its<br />

own, and the other, although having a similar overall<br />

structure (with a polyaryl core and a halo of alky<br />

chains) is not. The thin vertical region at the centre of<br />

the diagram represents the 1:1 C.P.I. compound. The<br />

extreme narrowness of this single phase area implies<br />

that there is a very positive interaction between the two<br />

types of molecule and that we are dealing with a system<br />

qualitatively different to one based simply on the mutual<br />

solubility of the two component. Note in particular, that<br />

the discotic columnar temperature range is enhanced in<br />

both directions, with the mesophase extending to higher<br />

and lower temperatures than those of the pure mesogenic<br />

component. The result of computer modelling the docking<br />

of an ABA unit of the column using the XED<br />

program is shown in Fig. 11.<br />

The reason for highlighting a property of thermotropic<br />

mixtures in this review is that I can see no reason why<br />

such effects should not arise in chromonic systems—<br />

especially since in some cases, the core polyaryl parts<br />

of the molecules are the same. In the past, the suggestion<br />

that it ought be possible to convert intransigent, insoluble<br />

dyes into conveniently soluble <strong>mesophases</strong> by the<br />

addition of some magic colourless agent was greeted<br />

with dismissive scepticism by the (British) dye industry.<br />

Perhaps it now looks one increment more plausible.<br />

9. The future<br />

The development of chromonic mesohase studies has<br />

not proceeded in the way I had expected. Bearing in<br />

mind the large industry concerned with the production<br />

and use of dyestuffs for printing fabric and paper and<br />

the apparently widespread occurrence of chromonic<br />

phases amongst dyes, I had expected that over the last<br />

decade there would have emerged a large literature<br />

concerned with enhancing the solubility properties (i.e.<br />

the mesophase-forming properties) of dyes, the investigation<br />

of mesophase enhancing mixtures, the fine tuning<br />

of colour by the addition of non-dye chromonic additives,<br />

the specalist use of the alignment of <strong>mesophases</strong><br />

on prepared surfaces and with electric and magnetic<br />

fields for printed material—perhaps for high grade<br />

security printing such as banknotes, credit cards and<br />

identification. As far as I am aware, little of this has<br />

occurred. However, the production of aligned dye films<br />

has materialised (albeit requiring a command surface).<br />

In the previous review, I referred to two grails. The<br />

first was of these is the production of cheap electrically<br />

conducting organic material, the second, a viable lightharvesting<br />

device. With the development of techniques<br />

to control the alignment of chromonic phases to produce<br />

multilayer stacks of aligned phases w54x, and to produce<br />

highly-aligned films by drying down chromonic phases<br />

(enabling circuits to be produced by ink jet printer),<br />

both appear to have been brought a step closer.<br />

The use of chromonic phases as compensating plates<br />

for TN devices came as a surprise—but since an<br />

optically negative nematic phase was required, there<br />

were not a lot of alternative systems to choose from.<br />

Polymer films and discotic and chromonic mesphases<br />

are the obvious three—and all have been investigated.<br />

The major distinction between chromonic phases and<br />

thermotropic phases is of course the fact that they are


J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

487<br />

Fig. 10. The production of ABAB stacks by complementary polytopic interactions (CPI). In some mixtures of thermotropic discotic and ‘near<br />

discotic’ compounds, it appears that the 1:1 mixture gives a columnar phase with a wider temperature range (and enhanced conductivity and<br />

photoconductivity) than that of either of the two separate compounds. It is presumed that the two molecules complement each other in some way<br />

and that the enhanced stability of the mesophase is due to the enhance stability of the ABAB columns. This example of the CPI effect is for a<br />

binary system of a mesogen, hexaalyltryphenylene(HAT6) – based derivative and the non-mesogenic compound, hexakis(4-nonylphenyl)dipyrazino<br />

w2,3-f:2.3.hxquoinoxalene( PDQ9). The nematic phase of the CPI 1:1 mixture occurs as a narrow vertical band at the centre of the<br />

phase diagram. The shaded area at the bottom of this band indicates where the chromonic N phase of the CPI compound is in a glassy state.<br />

(redrawn from Ref. w53x).


488 J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

proposed). However, there are still books being produced<br />

which purport to give a broad overview of ‘soft<br />

matter’, which omit all mention of chromonics. A single<br />

large-scale commercial application of chromonics will<br />

of course change this picture overnight. The widespread<br />

technological use of chromonic systems has not yet<br />

materialised, but the continuing discovery of unique<br />

properties and versatility of these systems promises<br />

much. I would hazard the guess that wherever nanotechnology<br />

takes us, the liquid crystalline state will never<br />

be far away—and chromonic systems will have something<br />

vital to offer.<br />

Acknowledgments<br />

I am indebted to Richard Bushby, Owen Lozman and<br />

to Gordon Tiddy for their continuing encouragement.<br />

References and recommended reading<br />

● of special interest<br />

●● of outstanding interest<br />

Fig. 11. Orthogonal views of the optimum stacked columnar structure<br />

of a sequence of HAT1–PDQ1–HAT1 molecules as predicted by the<br />

XED program. The material in this figure is taken from Ref. w53x.<br />

The XED program is described in Refs. w41–43x and Vinter JC, Saunders<br />

MR: Ciba F Symp 1991, 158:249–265, Vinter JC: J Comp-Aid<br />

Mol Des 1994, 8:653–668, Vinter JC: J Comp-Aid Mol Des 1996,<br />

10:417–426.<br />

‘water based’. This should makes possible a marriage<br />

between established display technology and established<br />

biochemical techniques. I foresee the diagnostic biomedical<br />

tools of the next century operating via a combination<br />

of liquid crystal display technology and biochemical<br />

recognition.<br />

10. Conclusion<br />

The recognition of chromonic <strong>mesophases</strong> as an<br />

important and distinct class of lyotropic <strong>mesophases</strong> is<br />

still patchy. Papers are now appearing where the term is<br />

used without the authors feeling the need to define it<br />

(and to refer to the literature where the word was first<br />

w1x<br />

●●<br />

Lydon JE. <strong>Chromonic</strong>s. In: Demus D, Goodby J, Gray GW,<br />

Speiss H-W, Vill V, editors. Weinheim: Wiley VCH, 1998.<br />

p. 981 –1007.<br />

w2x Lydon J. <strong>Chromonic</strong> liquid crystal phases. Curr Opin Colloid<br />

●● Interface 1998;3:458 –66.<br />

w3x Altounyan REC. Pharmacology of disodium cromoglycate.<br />

Schweiz Med Wochenschr 1980;110:179 –81.<br />

w4x Pepys J, Frankland AW. Disodium cromoglycate in allergic<br />

airways disease. London: Butterworths, 1970.<br />

w5x Hartshorne NH, Woodward GD. Mesomorphism in the system<br />

disodium chromoglycate-water. Mol Cryst Liq Cryst<br />

1973;23:343 –68.<br />

w6x Lydon JE. New models for the <strong>mesophases</strong> of disodium<br />

chromoglycate. Mol Cryst Liq Cryst Lett 1980;64:19 –24.<br />

w7x Hiu YW, Labes MM. Structure and order parameter of a<br />

nematic lyotropic liquid crystal studied by FTIR spectroscopy.<br />

J Phys Chem 1986;90:4064 –7.<br />

w8x Goldfarb D, Luz Z, Spielberg N, Zimmermann H. Structural<br />

and orientational characteristics of the disodiumychromoglycate-water<br />

<strong>mesophases</strong> by deuterium NMR and X-ray diffraction.<br />

Mol Cryst Liq Cryst 1984;124:225 –46.<br />

w9x Lee H, Labes MM. Lyotropic cholesteric and nematic phases<br />

of disodium chromoglycate in magnetic fields. Mol Cryst<br />

Liq Cryst 1982;84:137 –57.<br />

w10x Attwood TK, Lydon JE. Lyotropic mesophase formation by<br />

anti-asthmatic drugs. Mol Cryst Liq Cryst 1984;108:349 –57.<br />

w11x Israelachvili JN, Mitchell DJ, Ninham BW. Theory of selfassembly<br />

of hydrocarbon amphiphiles into micelles and<br />

bilayers. J Chem Soc Faraday Trans II 1976;72:1525 –68.<br />

w12x Israelachvili JN. Intermolecular and surface forces. London:<br />

Academic Press, 1991. p. 349.<br />

w13x Edwards RG, Henderson JR, Pinning RL. Simulation of selfassembly<br />

and lyotropic liquid crystal phases in model discotic<br />

solutions. Mol Phys 1995;6:567 –93.<br />

w14x Maiti PK, Lansac Y, Glaser MA, Clark NA. Isodesmic selfassembly<br />

in lyotropic chromonic systems. Liq Cryst<br />

2002;29(5):619 –26.


J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

489<br />

w15x Mundy K, Sleep JC, Lydon JE. The intercalation of ethidium<br />

bromide in the chromonic lyotropic phases of drugs and<br />

nucleic acids. Liq Cryst 1995;19:107 –12.<br />

w16x Attwood TK, Lydon JE, Jones F. The chromonic phases of<br />

dyes. Liq Cryst 1986;1:499 –507.<br />

w17x Turner JE, Lydon JE: <strong>Chromonic</strong> mesomorphism: the range<br />

of lyotropic discotic phases. Mol Cryst Liq Cryst Lett, 5:93–<br />

99.<br />

w18x Tiddy GJT, Mateer DL, Ormerod AP, Harrison WJ, Edwards<br />

DJ. Highly ordered aggregates in dilute water–dye systems.<br />

Langmuir 1995;11:390 –3.<br />

w19x Harrison WJ, Mateer DL, Tiddy GJT. Liquid crystalline J-<br />

aggregates formed by aqueous ionic cyanine dyes. J Phys<br />

Chem 1996;100:2310 –21.<br />

w20x (a) Spada GP, Carcuro A, Colonna FP, Garbesi A, Gottarelli<br />

G. Liq Cryst 1988;3:651<br />

(b) Mariani P, Mazabard C, Garbesi A, Spada GP. A study<br />

of the structure of the lyo<strong>mesophases</strong> formed by the dinucleoside<br />

phosphate d(GpG). An approach by X-ray diffraction<br />

and optical microscopy. J. Am Chem Soc<br />

1989;111:6369.<br />

w21x Usol’tseva N, Espinet P, Buey J, Serrano JL. Liquid-crystalline<br />

behaviour of di- and mono-palladium organyls: two<br />

ways of lyomesophase formation. J Mater Chem 1997;7:215.<br />

w22x Harrison WJ, Mateer DL, Tiddy GJT. J-aggregates and liquid<br />

crystal structures of cyanine dyes. Faraday Discuss<br />

1996;104:139 –54.<br />

w23x Yevdokimov YM, Skuridin SG, Salanov VI. The liquid<br />

crystalline phases of double-stranded nucleic acids in vitro<br />

and in vivo. Liq Cryst 1988;3:1443 –59.<br />

w24x Livolant F, Levelut AM, Doucet J, Benoit JP. The highly<br />

concentrated liquid crystalline phase of DNA is columnar<br />

liquid crystal. Nature 1989;339:724 –6.<br />

w25x Bonazzi S, De Morais M, Garbesi A, Gottarelli G, Mariani<br />

P, Spada GP. <strong>Chromonic</strong> lyo<strong>mesophases</strong> formed by the selfassembly<br />

of the cyclic dinucleotide d(cGpGp). Liq Cryst<br />

1991;10:495 –506.<br />

w26x Mariani P, Demorais MM, Goterelli G, Spada GP, Delacroix<br />

H, Tondelli L. Structural analysis of the lyotropic polymorphism<br />

of four-stranded aggregates of 29-deoxyguanosine<br />

39monoposphate derivatives. Liq Cryst 1993;15:757 –78.<br />

w27x Bykova VV, Usol’tseva NV, Anan’eva GA, Smirnova AI,<br />

Shaposhnikov GP, Maizlish VE. Sulfamoyl-substituted copper<br />

phthalocyanines and their mesomorphic properties. Russ<br />

J Gen Chem 2000;70(1):145 –7.<br />

w28x Donnio B. Lyotropic metallomesogens. Curr Opin Colloid<br />

Interface 2002;7:371 –94.<br />

w29x<br />

●<br />

w30x<br />

●<br />

Ruslim C, Matsunaga D, Hashimoto M, Tamaki T, Ichimura<br />

K. Structural characteristics of the chromonic <strong>mesophases</strong> of<br />

CI Direct Blue 67. Langmuir 2003;19(9):3686 –91.<br />

Matsunaga D, Tamaki T, Ichimura K. Azo-pendant polyamides<br />

which have the potential to photoalign chromonic<br />

lyotropic liquid crystals. J Mater Chem 2003;13:1558 –64.<br />

w31x Ichimura K, Fujiwara T, Momose M, Matsunaga D. Surfaceassisted<br />

photoalignment control of lyotropic liquid crystals.<br />

Part 1. Characterisation and photoalignment of aqueous<br />

solutions of a water-soluble dye as lyotropic liquid crystals.<br />

J Mater Chem, 2002; 12(12):3380–3386.<br />

w32x Matsunaga D, Tamaki T, Akiyama H, Ichimura K. Photofabrication<br />

of micro-patterned polarising elements for steroscopic<br />

displays. Adv Mater 2002;14:1477.<br />

w33x Iverson IK, Casey SM, Seo W, Tam-Chang SW, Pindzola<br />

BA. Controlling molecular orientation in solid films via selforganisation<br />

in the liquid-crystalline phase. Langmuir<br />

2002;18(9):3510 –6.<br />

w34x Remizow S, Krivoschchepov A, Nazarov V, Grodsky A.<br />

Rheology of the lyotropic liquid crystalline material for thin<br />

film polarisers. Mol Mater 2001;14(2):179 –90.<br />

w35x (a) Iverson IK, Tam-Chang S-W. Cascade of molecular order<br />

by sequential self-organization, induced orientation, and<br />

order transfer processes. J Am Chem Soc 1999;121:5801<br />

(b) Iverson IK, Casey SM, Seo W, Tam-Chang S-W. Controlling<br />

molecular orientation in solid films via self-organization<br />

in the liquid-crystalline phase. Langmuir<br />

2002;18:3510.<br />

w36x (a) Schneider T, Lavrentovich OD. Self-assembled monolayers<br />

and multilayered stacks of lyotropic chromonic liquid<br />

crystalline dyes with in-plane orientational order. Langmuir<br />

2000;16:5227<br />

(b) Sergan T, Schneider T, Kelly J, Lavrentovich OD.<br />

Polarising-alignment layers for twisted nematic cells. Liq<br />

Cryst 2000;27:567.<br />

w37x Hahn C, Spring I, Thunig C, Platz G, Wokaun A. Investigation<br />

of the photoinduced optical anisotropy of azo dye<br />

<strong>mesophases</strong>. Langmuir 1998;14(24):6871 –8.<br />

w38x Sergan T, Kelly J. Negative uniaxial films from lyotropic<br />

liquid crystalline material for liquid crystal display applications.<br />

Liq Cryst 2000;27:1481.<br />

w39x Lavrentovich M, Sergan T, Kelly J: In Proceedings of the<br />

19th International Liquid Crystal Conference, Edinburgh<br />

2002, to be published in Mol Cryst Liq Cryst.<br />

w40x<br />

●<br />

Lavrentovich M, Sergan T, Kelly J. Planar and twisted<br />

lyotropic chromonic liquid crystal cells as optical compensators<br />

for twisted nematic displays. Liq Cryst 2003;30:851 –<br />

9.<br />

w41x Hunter A, Sanders JKM. The nature of p–p interactions. J<br />

Am Chem Soc 1990;112:5525 –34.<br />

w42x Hunter CA. Angew Chem Int Edit 1993;32:1584 –6.<br />

w43x Hunter C, Lawson KR, Perkins J, Urch CJ. Aromatic stacking<br />

interactions. J Chem Soc Perkin Trans 2001;2:651 –9.<br />

w44x Bates MA, Luckhurst GR. Computer simulation studies of<br />

anisotropic systems XXIX. Quadrupolar Gay–Berne discs<br />

and chemically induced liquid crystal phases. Liq Cryst<br />

1998;24:229 –41.<br />

w45x Bushby RJ, Evans SD, Lozman OR, McNeill A, Movaghar<br />

B. Enhanced charge conduction in discotic liquid crystals. J<br />

Mater Chem 2001;11:1982 –4.<br />

w46x Kreouzis T, Scott K, Donovan KJ, Boden N, Bushby RJ,<br />

Lozman OR, et al. Enhanced electronic transport properties<br />

in complementary binary discotic liquid crystal systems.<br />

Chem Phys 2000;262:489 –97.<br />

w47x Boden N, Bushby RJ, Liu QY, Lozman OR. CPI (complementary<br />

polytopic interaction) stabilised liquid crystal compounds<br />

formed by esters of<br />

2hydroxy3,6,7,10,11pentakis(hexyloxy)triphenyl. J Mater<br />

Chem 2001;11:1612 –7.<br />

w48x Boden N, Bushby RJ, Lu ZB, Lozman OR. CPI induction<br />

of liquid crystal behaviour in triphenylenes with a mixture<br />

of hydrophobic and hydrophilic side chains. Liq Cryst<br />

2001;28:657 –61.<br />

w49x<br />

●<br />

Boden N, Bushby RJ, Headdock G, Lozman OR, Wood A.<br />

Syntheses of new ‘large core’ discogens based on the<br />

triphenylene, azatriphenylene and hexabenztrinaphthylene<br />

nuclei. Liq Cryst 2001;28:139 –44.<br />

w50x Lozman OR, Bushby RJ, Vinter JG. Complementary polytopic<br />

interactions (CPI) as revealed by molecular modelling<br />

using the XED force field. J Chem Soc Perkin Trans 2<br />

2001;9:1446 –52.<br />

w51x Boden N, Bushby RJ, Lu ZB. A rational synthesis of<br />

polyacrylates with discogenic side groups. Liq Cryst<br />

1998;25:47 –58.


490 J. Lydon / Current Opinion in Colloid and Interface Science 8 (2004) 480–490<br />

w52x Boden N, Bushby RJ, Cooke G, Lozman OR, Lu Z. CPI: A<br />

recipe for improving applicable properties of discotic liquid<br />

crystals. J Am Chem Soc 2001;123:7915 –6.<br />

w53x<br />

●<br />

Boden N, Bushby RJ, Lozman OR Designing better columnar<br />

<strong>mesophases</strong> presented at the Anglo-Japanese meeting in York<br />

2001 to be published in Mol Cryst. Liq Cryst.<br />

w54x Schneider T, Lavrentovich OD. Self-assembled monolayers<br />

and multilayered stacks of lyotropic chromonic liquid crystalline<br />

dyes with in-plane orientational order. Langmuir<br />

2000;16:5227.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!