You are on page 1of 17

Chromism

1. Introduction
Chromism is a reversible change in a substance's colour resulting from a process caused by some form of stimulus. Many materials are chromic, including inorganic and organic compounds and conducting polymers, and the property can result from many different mechanisms. There are also several types of chromism, which are detailed below.

2. Thermochromism
Thermochromism is the reversible colour change of a substance induced by temperature change. A large variety of substances, organic, inorganic, organometallic, supramolecular and polymeric systems exhibit this phenomenon. Examples of these include bianthrones, cobalt hexacyanoferrate, the zirconocene complex of 1,4-diphenyl-1,3-butadiene and poly(3-alkylthiophene). The organic 9,9'-bixanthenylidene is colourless at 90 K, yellow-green at 298 K and dark-blue when melted at 592 K. Heating conducting polymers can cause them to change colour. This is achieved by causing conformational changes to the polymer backbone, resulting in a change in the band gap of the polymer. It has been reported that regioregular P3HT reversibly changes colour upon heating to 220C due to temperature-dependent conformation changes. Thermally crosslinked polymer undergoes the same colour change, but it is much less reversible [1]. Other forms of thermochromism may be commercially important, e.g. to give a visual indication of temperature changes.

3. Photochromism
Photochromism is the reversible colour change as a result of exposure to electromagnetic radiation. Photochromism was first recorded by ter Meer who showed that the potassium salt of dinitromethane reversibly changed colour when exposed to exciting radiation [2]. Research into this phenomenon proceeded slowly until the 1940's and later when research picked up in both quantity and quality [3]. Photochromism can be defined as a reversible change of a single chemical species between two states having distinguishable absorption spectra, such change being induced in at least one direction by the action of electromagnetic radiation [3]. This can be represented by the following equation:

A(1)

B(2)

The inducing radiation is normally in the ultraviolet or visible regions, and the changes in their spectra are in the visible or near infrared. A may be a molecule or ion whereas B may be one or more ions or molecules. Commonly, state B is thermodynamically less stable than A and is more coloured, though some exceptions exist. The reverse reaction often occurs thermally, though it can sometimes occur photochemically. Most photochromes change to state B upon radiation and slowly revert thermally back to state A when removed from the radiation. A wide variety of chemicals exhibit photochromism, including both organic and inorganic compounds [3]. Organic compounds include anils, disulfoxides, hydrazones, osazones, semicarbazones, stilbene derivatives, succinic anhydrides, camphor derivatives, certain dye derivatives, o-nitrobenzyl derivatives, syndrones and spiro compounds. Examples of inorganic

photochromes include metal oxides, alkaline earth metal sulfides, titanates, mercury compounds, copper compounds, certain minerals and transition metal compounds such as carbonyls. Most organic electrochromes are activated by ultraviolet light in the range of 200 to 400 nm, though some are activated by blue light with wavelengths up to 430 nm. Organic molecules change by a number of processes. These include cis-trans isomerism, tautomerism, heterolytic cleavage and homolytic cleavage. Heterolytic cleavage is the term applied if a single bond is cleaved, leaving charged species that exists as individual ions or as moieties still connected by other bonds. They are rarely stable and quickly revert back to their original state. Spiropyrans, triarylmethanes and polymethines are examples of compounds that undergo heterolytic cleavage. Homolytic cleavage is very similar, except that a pair of radicals is produced. These are rarely stable, with fast reversion reaction rates. Examples of compounds showing this mechanism include pyrroles, hydrazines, sydnones, disulfides and nitroso dimers. The coloured forms in these systems fatigue rapidly, limiting their usefulness in many practical devices. Cis-trans isomerism is the process used by living systems that are photochromic. Examples of compounds undergoing this mechanism include urocanic acid, stilbenes, indigoid dyes, azo compounds, polyenes and cyanine dyes. The trans isomer is normally the thermodynamically more stable form and the reverse reaction is normally a thermal process. Depending on the exact nature of the two states and their conditions, temperature, solvent, pH and presence or absence of catalyst, the rate of the reverse reaction can be as fast as a millisecond half-life, or as slow as weeks or even months. The photochromic processes, however, are often insensitive to these factors. These materials can have very high quantum yields, as high as 0.5 to 0.6, which makes some of them very attractive for practical uses. Photochromic tautomerism is the shifting of the equilibrium of two tautomeric forms by the action of electromagnetic radiation. Examples include hydrogen transfer, such as in keto-enol tautomerism, and valence tautomerism where bonds are rearranged, causing an alteration in valences of one or more atoms. An example of this is the isomerism between cis-1,3,5-hexatriene and 1,3cyclohexadiene. Because the lifetimes and reversibility of these systems are altered by such factors as acidity, photochemical stability and oxidation-reduction potentials, their practical applications are limited. A large number of inorganic compounds exhibit photochromism. These solids often have large band gaps of the order of 3 - 12 eV, and excitation of these solids leads to the formation of metastable centres that absorb visible light, giving rise to their colour. They can return to their ground state by heating or by optical excitation within the colour-centre band. In most cases, the photochromism is a structure sensitive phenomenon involving localised defect, impurities or dislocations. Some of these inorganic compounds have the potential for a number of different uses. Photochromic compounds have a number of useful applications. These can be divided according to the most important property that is being used (table 2.4) [3].

Applications Depending Upon Sensitivity to Radiation Self-developing photography Dosimetry and actinometry Protective materials Photododging Reversibility Chemical switches for computers Data displays Optical signal processing Reusable data storage media Photochromic microimages Information encoding and steganography Control of light intensity Specific Colour Changes Camouflage Decoration Thermal, Chemical or Physical Properties Temperature indicators Heat-sensitive recording media Photomasking and photoresist technology Anaytical reagents Photopolymerisation Photocontractile polymers and the photoviscosity effect Q-switches Pyroelectric photochromic materials
Table 2.4: Applications of Photochromic materials.

4. Chirochromism and Diastereoselective Photochromism


Chirochromism is the reversible change in rotation of a plane of polarised light between two different chiral diastereomers of a photochromic compound. Diastereoselective photochromism is the photoinduced reversible change of absorption spectra between two diastereomers. The required diastereomer can be selected by exposing the material to light at particular wavelengths depending on which isomer is required. For example, if diastereomers A and B have absorption maxima at 1 and 2 respectively, then A can be converted to B by using light of wavelength at or near 1, and B into A at 2. Often, the same reaction can occur as a result of thermal energy (diastereothermochromism).

5. Halochromism
Halochromism is the reversible colour change due to a change in pH of a solution. Halochromic compounds include phenolphthalein and titanium dioxide. The compounds themselves are weak acids or bases and become involved in acid-base reactions. A change in the pH causes a change in the ratio of ionised and non-ionised states, and, since these two states have different colours, the colour of the solution changes. This colour change can be used in acid-base titrations where the colour change of the halochrome corresponds to the end-point of the reaction.

6. Piezochromism and Tribochromism


Piezochromism is the reversible colour changed caused by mechanical grinding. This induced colour change reverts back to its original colour if the material is left in the dark over time or dissolved in an organic solvent. A chemical that displays this phenomenon is diphenylflavelene. Tribochromism is the same as piezochromism, except that the compound does not revert back to its original colour in time if left in the dark or dissolved in an organic solvent.

7. Solvatochromism
Solvatochromism is the reversible colour changed induced by solvents. This often derives from changes in polarity of various solvents. This affects charge transfer mechanisms in solvatochromic compounds, causing colour changes. Poly(3-alkylthiophenes) are known to be solvatochromic. The effect on the polymers' colour of increasing the solvent polarity is very similar to decreasing temperature of the polymer i.e. the absorption maxima is red-shifted. This is due to the solvent being less able to intermingle with the alkyl chains and so the polymer becomes more planar, causing a decrease in the band gap. The colour of poly(3-hexylthiophene) solution was changed by the addition of a poor solvent. There was a crystallisation of the chains into a folded conformation [4-5]. For P3ATs, solvatochromism is often paralleled by a change in phase as it precipitates out of solution. However, insoluble crosslinked P3HT also changes from red to orange when chloroform is added due to chloroform's ability to penetrate the polymer matrix, causing it to swell.

8. Halosolvatochromism
Halosolvatochromism is the reversible change in colour brought on by a change in ionic strength without a change in the structure of the chromophore. An example of a halosolvatochromic compound is the dye, 2,6-diphenyl-4-(2,4,6-triphenyl-1-pyridino)phenoxide. Upon addition of salts such as KI, Ca(SCN)2, or Mg(ClO4)2 to solutions containing the dye to undergo a shift in the electronic absorption spectrum that increases with a change in cation density.

9. Ionochromism and Acidichromism


Ionochromism is the reversible colour change caused by the addition of ions. It can occur in addition to photochromism and can trigger an alteration in conductivity. When ionochromism does occur along side photochromism, both the relaxed and excited states can undergo ionochromism, i.e., in the photochromic reaction AB, A can react to form AM+ and B can react to form BM+. All four of these compounds can have different colours, producing a four-colour system. A variation to ionochromism is acidichromism, where M = H. Some phenols and aromatic amines are display this phenomenon. One possible use for acidichromism is in non-destructive readout systems where one form is used for readout and the others for writing and erasing [6].

10. Electrochromism
This is potentially the most commercially useful form of chromism. Electrochromic materials have been known since 1968 [7]. Three classes of electrochromic materials are known. These are metal oxide films, molecular dyes and conducting polymers. While some of these systems are all-solution, and some are solution-to-solid electrochromes, most electrochromes are all-solid systems. They have a number of potential uses, including displays, smart mirrors and windows, active optical filters and computer data storage [8]. The key properties of electrochromic materials are the switching times, the contrast ratios, coloration efficiency, electrochromic memory and long term stability. The switching time may be defined as the time for the colour change to become 75% of the ultimate change in % transmittance. The contrast ratio is the difference in transmittance in the visible spectrum. The electrochromic memory is the ability of the material to hold its colour with no current. The coloration efficiency (CE) is the change in optical density (OD) per unit area of the electrode (Qd) for a given wavelength. For amorphous tungsten trioxide films fabricated by thermal evaporation, a CE of 115

cm2 C-1 (633 nm) can be achieved [9]. The long term stability is the ability of the material to retain its electrochromic properties over a large number of switching cycles. All-solution systems are systems where the electrochemical transfer occurs at the solid-liquid interface. The electrochrome undergoes its redox reaction at the surface of an electrode and then moves back into the solution bulk, its colour changed. An example of an all-solution electrochrome is 1,1'dimethyl-4,4'bipyridilium dication in water [10]. In its initial +2 state it is a clear colourless solution, or faint yellow in the presence of some anions. Upon reduction it turns a bright blue colour as it forms +1 a radical cation. The rate limiting step of the process is the rate at which the dication diffuses to the electrode from the bulk solution. For this system the current observed follows the Cottrell time dependence equation (section 1.6.6), i 1/t [11]. Since this is governed by the rate of diffusion, the absorbance A t, providing the current is wholly faradiac, i.e. each electron causes the reduction of one dication. Other all-solution electrochromes include transition metal ions and their complexes such as iron (III) thiocyanate in water and hexacyanoferrates and quinones in tetracyanoquinodimethane or tetrathiafulvalene in acetonitrile solution. Solution to solid electrochromes are initially clear colourless soluble compounds, which, upon oxidation or reduction, form coloured solid films upon the electrode. An example of this type of electrochrome is 1,1'diheptyl-4,4'bipyridilium dication in water. The dibromide of this compound is a pale yellow colour in solution and forms a deeply crimson coloured deposit on the cathode upon electrochemical reduction, which is initially amorphous but becomes more crystalline soon after formation. The solubility, morphology and colour of these salts depends upon the accompanying counter-ion [12]. Initially the Cottrell current-time relationship is observed; however, the formation of the insoluble salt complicates things as the salt in an insulator. This slows further salt formation, which must proceed via slow electron transport through the solid deposit. Reversing the current causes the solid film to re-dissolve as the dication. Electron mediators such as ferrocyanide are used to facilitate electron transfer, which speeds up the colour change [13-14]. Other examples of this group include benzyl paraquat [15-16], aqueous Cyanophenyl paraquat [17] and polyethoxyfluorene species in acetonitrile solution [18]. The deposition of metal onto an other electrode is an example of this type of system, though this is only rarely used for electrochromism. All-solid systems are the most common systems used for electrochromic displays. They include all conducting polymer systems, metal oxides, Prussian blue and its analogues and rare earth phthalocyanines. For metal oxides and Prussian blue the colour change derives from a change in the oxidation state of one or more metal ions. It is these ions that change colour. The colour change for the rare-earth phthalocyanines comes from the redox behaviour of the ligand; charge transfer to the metal can cause decomposition of the complex [19]. Conducting polymers change colour by the creation and destruction of polarons and bipolarons. These colour changes involve the movement of counter ions into and out of the electrochrome matrix, which is normally the rate limiting step of the colour change. The charged species enters the polymer matrix and migrates slowly through the film. However, it does not react at the electrode, indeed, if it did, an undesirable side reaction would occur, producing either irreversibly reduced or oxidised polymer. Diffusion occurs within the electrochrome as the ions diffuse in from the polymer/electrolyte interface and electron diffusion from the polymer/electrode interface. The slower of these two processes governs the rate of colour change. Table 1.3, page 23, shows some typical diffusion coefficients.

10.1. Design and Construction of Electrochromic Devices


Electrochromic displays have two modes of operation, reflective and transmissive. In reflective displays the electrochromic material is coated onto a reflective electrode, or has a reflective counter electrode. The most common design has a thin film of the primary electrochrome on ITO glass which is placed upon a reflective counter electrode that has a secondary electrochromic layer coated with an electrolytic gel on it. A schematic diagram of such a system is shown in figure 2.21.

Figure 2.21: Schematic diagram of a reflective electrochromic device.

This device uses the absorption of light which is reflected by a material such as platinum [20] or rhodium alloy [21]. The main advantage of the system is that, since the light passes through the layers twice, the electrochromic material needs only to be half the thickness that it does in transmissive electrochromic systems. This results in a faster electrochromic system. In transmissive electrochromic systems the mirror is replaced with another glass substrate. They use light passing straight through to get their effect. Electrochromic devices of this kind often need a backlight to work. Figure 2.22 shows a schematic diagram of a transmittance electrochromic device. Note, both reflective and transmissive electrochromic devices can be made without a secondary electrochrome.

Figure 2.22: Schematic diagram of a transmissive electrochromic device.

The electrolytes used are either solids or high viscosity liquids [10]. They function both as ionic conductors and as a supply/sink for ions moving into and out of the electrochromes. Electrolytes must also be compatible with both the electrochromes used. Three types of electrolytes are most common, viz. polymer electrolytes, polyelectrolytes and solid inorganic electrolytes. Polymer electrolytes are neutral polymers such as poly(ethylene oxide) or poly(propylene glycol) with salts such as LiClO4, CF3SO3H or H3PO4 dissolved in them. The polymers are often made with an intermediate molecular weight as these are highly viscous liquids and are termed 'semi-solid solutions' [22]. Polyelectrolytes are polymers that contain ion-labile groups. Examples include poly(2-acrylamido-2-methylpropanesulfonic acid), poly(ethylene sulfonic acid), poly(methyl methacrylate), poly(styrene sulfonic acid) and poly(vinyl alcohol) [10]. Solid inorganic salt electrolytes are often clear colourless inorganic solids and one or more small ions. Examples include hydrogen uranyl phosphate, Li3N, Ta2O5, water glass and ZrO2 [10]. The advantage of the polymer based electrolytes is that they are easier to use in device fabrication, particularly polymer electrolytes. They are also flexible and can easily handle the stresses caused by the small expansion and contraction the solid electrochromes undergo during their redox processes. However, they are vulnerable to photolytic degradation, and so strict control of internal moisture level is often required. This is not a problem for solid inorganic electrolytes, which do not have this problem. However, they are brittle, and so cracks can form due to the expansion and contraction of the electrochromes. They also tend to have a high electrical resistance which results in slower electrochromic switching. Several techniques are available for preparing thin electrochromic films. These include electrodeposition, vacuum deposition, electron-beam sputtering, chemical vapour deposition, sol-gel coating and dip-coating [10]. Electrodeposition is done by placing the electrode in an aqueous solution of the desired metal through which an electric current is passed to produce a thin film of metal, which is then oxidised to form the desired compound. This technique can also be used to prepare thin conducting polymer films by electrochemical synthesis. Vacuum deposition is done by subliming the electrochromic material in a vacuum. The electrode is placed in the vacuum chamber in a cooled holder. The sublimed material then solidifies upon the electrode, forming a thin film. Sputtering is achieved by bombarding the electrode with an electron beam passing through ionised electrochrome. The electrochrome is ionised by placing it on an inductively heated ceramic metal substrate, and ionised with a tungsten filament, typically at a potential of 5 kV. Chemical vapour

deposition can be the same as vacuum deposition except that the starting material is one that can be oxidised by a trace amount of oxygen that is deliberately bled into the vacuum chamber. Sol-gel coating is used to coat large area electrochromic displays. A gelatinous material containing the electrochrome is applied to the electrode, which is then subjected to thermal curing that drives off solvent and/or chemically alters the component of the gel. Dip coating or the similar technique of spin coating is the simplest method for fabrication of electrochromic devices. The substrate is dipped into solutions containing ether solvated or colloidal electrochrome that is then allowed to dry, sometimes with heating. Multiple layers can be prepared this way by successive dipping. Spin coating is the same except that the substrate is spun dry; this can produce very even films.

10.2. Metal Ion Electrochromism


Many transition metal oxides are capable of redox reactions that result in colour change. Metal oxide films are commonly prepared as thin layers of either tungsten, nickel, molybdenum or other metal compounds by a number of techniques. These include sol-gel electrochemical, by d.c. or r.f. reactive sputtering techniques, electron-beam evaporation, by anodic or cathodic electrodeposition or by solution dipping of the electrochromic metal compounds (or compounds that can be changed into these metal compounds) onto optically transparent electrodes (OTE) [10, 23-28]. Their electrochromism is derived from the colour change associated with a change in the oxidation state of the metal anion. The behaviour of these materials is dependent upon pH, moisture and exposure to the atmosphere [29]. Generally, the switching rates of these films is somewhat slow, with typical switching times of about 15 - 60 seconds to achieve 100% conversion to either coloured or bleached state [23-28, 30-34]. An Example of this includes nickel oxide, which changes from transparent (pale green) to brown/black, taking about 30 seconds to do so [35]. Other examples include [(NH4) 5+ and [(NH4)5Ru]2(4,4'-bipyridine)5+ whose electrochromism is significantly 5Ru]2(pyrazine) different due to the effect of the ligand [36]. Table 2.5 below gives some examples of metal oxide films with electrochromic properties [10]. Metal Oxide Reaction Colour Change Reference Cobalt Oxide Indium Tin Oxide Iridium Oxide Molybdenum Trioxide Nickel Oxide Tungsten Trioxide Vanadium Pentoxide Cerium Oxide Manganese Oxide Niobium Pentoxide Ruthenium Dioxide 3CoO + 2OH Co3O4 + H2O + 2eIn2O3 + 2x(Li+ + e-) Li2xInIII(1-x)InIxO3 Ir(OH)3 IrO2H2O + H+ + eMoO3 + x(Li+ + e-) LixMoVI(1-x)MoVxO3 NiOxHy [NiII(1-z)NiIIIz]OxH(y-z) + zH+ + zeWO3 + x(Li+ + e-) LixWVI(1-x)WVxO3 LixV2O5 V2O5 + x(Li+ + e-) CeO2 + x(Li+ + e-) LixCeO2 MnO2 + ze- + zH+ MnO(2-z)(OH) Nb2O5 + x(Li+ + e-) LixNb2O5 RuO22H2O + H2O + e- (Ru2O35H2O) + OHgreen brown colourless pale blue colourless blue/grey colourless blue colourless brown/black very pale blue blue very pale blue (brown/yellow) yellow very pale yellow brown colourless pale blue (blue/brown) black 37 38 39 7 40 41 42 43 44 45 46

Table 2.5: Some examples of electrochromic metal oxides.

These metal oxides can be mixed with small amounts of metal oxide to produce doped metal oxides. Very small quantities of alien metal oxide can have a large impact on the spectroscopic characteristics of the material, its conductivity and the potential window available for electrochromic operation [10]. Of the electrochromic metal systems, the most durable known is Prussian blue [47-49]. This can be reduced in the presence of alkali metal ions to yield white Everitt's salt. KFe3+Fe2+(CN)6 + K+ + e- K2Fe2+Fe2+(CN)6 Prussian blue Everitt's salt Prussian blue can also be oxidised to a green colour, known as Prussian green by oxidation of some of the Fe2+ to Fe3+. Further oxidation causes the electrochrome to turn golden-brown (Prussian brown). Reduction proceeds at -200 mV and oxidation +1000 mV for complete oxidation [10]. A closely related electrochromic material is vanadium hexacyanoferrate which changes from yellow to green upon oxidation [50]. However, the greatest electrochromic change occurs in the UV region. Electrochemical and XPS data indicates that its electrochromism originated from the iron centres in the film. Other analogues of Prussian blue include ruthenium, osmium, nickel and copper hexacyanoferrate.

10.3. Organic Dyes


A number of types of electrochromic organic dyes exist. Their electrochromic behaviour derives from the effect of changing the charge on large conjugated systems. They include phthalocyanine complexes, bipyridilium systems, carbazoles, methoxybiphenyl, quinones, diphenylamine and pyrazolines. A number of electrochromic heavy rare-earth metal diphthalocyanine complexes are known. These heavy rare-earth metal diphthalocyanine (Pc2) complexes (Pc2M, M = Lu, Yb, Er, Ho, Dy, Gd, ect.) have been extensively studied [51-57]. The colour change of these species results from the reduction of the diphthalocyanine ring. However, these compounds are difficult to process, as they are only slightly soluble in organic solvents such as chloroform and benzene. As a result, the films produced tend to be inhomogeneous, resulting in a slow response time and bad adhesion to the electrode surface. A more soluble system is R14R24Pc2M (where R1 = propoxy, R2 = tert-butyl, M = Er, Lu) [58]. This exhibits good colour changes with four distinct colours, being green, brown-red, blue and purple. Another type of electrochromic organic material is molecular dyes. These dyes can be deposited as electrode-confined films and they can can exhibit fast response times. Dyes dispersed throughout an electrolyte are much slower. These materials have been used to construct electrochromic devices (ECDs) that are fully transparent with no applied potential and opaque dark green or blue at a potential of 1 V [59]. The colour change of electrochromic systems based on bipyridilium derives from the reduction of the bipyridilium dication to form an intensely coloured radical cation (figure 2.23). The intense colour of the radical cation results from an internal photo-effected electronic excitation. The radical cation can be further reduced to a neutral molecule that tends to be weakly coloured.

R'

N R'' Colourless

+e-

R'

N R''

Intensely Coloured +e-

R'

N R''

Pale colour
Figure 2.23: Electrochromism of bipyridilium.

The precise colour of the bipyridilium depends upon the substituents on the two nitrogen atoms. Alkyl substituted bipyridilium radical cations are a blue/violet colour, changing to crimson with longer chains, whereas aryl bipyridilium cations have a green hue [10]. For soluble systems, the solvent used can affect their colour. The solubility of bipyridilium compounds decreases as the length of the substituents increases. Since the radical cation is less soluble than the dication, thin films of the radical cation can form upon reduction for some of the bipyridilium derivatives. Some examples of some symmetrical bipyridiliums are shown in table 2.6 [10]. Substituent Methyl Ethyl Propyl Butyl Pentyl Hexyl Heptyl Octyl Iso-pentyl Benzyl Solid Bromide Salt Film on Platinum No No No No Yes Yes Yes Yes Yes Yes Colour Blue Blue Blue Blue Purple Purple Mauve Crimson Purple Mauve

Table 2.6: Symmetrical bipyridiliums: The effect of varying the substituents on the colour of the radical cation.

The counter anion can affect the performance of the electrochromic device by changing their solubility and stability [60]. Studies with 1,1'diheptyl-4,4'bipyridilium showed that the best anion with water was dihydrogen phosphate. sulfate, fluoride, formate, acetate, bromide, chloride, tetrafluoroborate were found to be satisfactory [12]. A number of other organic chemicals can also form radical cations in the same fashion. These include carbazole, methoxybiphenyl, quinones, diphenylamine and pyrazolines and all of their derivatives. Table 2.7 gives the colours obtainable by these systems [10].

Electrochromic Compound Carbazoles Carbazole N-ethylcarbazole N-phenylcarbazole N-carbazycarbazole Methoxybiphenyl Compounds 2,7-dimethoxyfluorene 2,3,4,5,6,7-hexamethylfluorene 2,7-dimethoxy-9,10-dihydrophenanthrene 2,3,6,7-tetramethoxy-9,10-dihydrophenanthrene Quinones o-3,4,5,6-tetrachlorobenzoquinone o-3,4,5,6-tetrabromobenzoquinone 1,4-benzoquinone p-3,4,5,6-tetrafluorobenzoquinone p-3,4,5,6-tetrachlorobenzoquinone p-2,3-dicyano,5,6-dichlorobenzoquinone 5-aminonaphthoquinone 1,5-diaminonathraquinone

Colour of Radial Cation Dark Green Green Iridescent Yellow Brown Blue Blue Green Green Intense Blue Blue Light Blue Yellow Yellow Pink Purple/blue Purple
Table 2.7: Some examples of electrochromic dyes.

10.4. Conducting Polymers


The third class of electrochromic material is conducting polymers. Their electrochromic properties result from the oxidation or reduction of the polymer backbone, resulting in the formation of bipolaron bands. This often has large effects on the electronic absorption spectra, producing high contrast colour changes. The speed of the colour change is dependent on the speed at which the dopant ions can migrate in and out of the polymer matrix. Because some ions are relatively slowmoving, electrochromism can be a slow process, with the fastest systems (2-methoxy and 2-ethoxy substituted polyaniline with 1M hydrochloric acid electrolyte) taking from about 2.5 ms for a colour change to occur fully [61]. However, most systems take about 100 ms to change colour. The use of ITO glass can have a slowing effect too. The thickness of the ITO layer is typically 0.3m [10] and it is a semiconductor with a conductivity of about 0.08 Sm-1 [62]. The low conductivity can restrict the current enough to have an appreciable effect on the response time. Ideally, the polymer should change from colourless to an intense colour. However, in practice, the colour change is often from a pale colour to a different intense colour. For example, poly[4,4'-bis(2-(3,4-ethylenedioxythiophene)) biphenyl] switches from an intense deep blue oxidised state to a pale yellow-orange reduced state [63]. The colour of the polymer is dependent on the energy gap between the valence and conduction bands when the polymer is in its neutral state, and the gap between the bipolaron and the conduction bands when it is in its oxidised or reduced state. These, in turn, are dependent on the conjugation of the polymer, the electrochemical nature of any side groups and their effects on the polymer backbone. As the response time is dependent on the movement of charge compensating counterions,

open polymer morphology often results in a reduced response time [63]. The smaller the band gap, the longer the wavelength of light absorbed. As polarons and bipolarons are formed, the absorbance of the main inter-band peak decreases and a new peak lower in energy is formed. The ideal electrochromic polymer will have a large contrast between its colours, have a very short switching time, maintain its colour after the current has been switched off and retain its electrochromism over a large number of switches. Unfortunately, bipolarons may be relatively high energy species that can take part in side reactions, damaging the polymer's conjugated system and causing a loss in electrochromic activity. There are many examples of electrochromic conducting polymers. They include a number of functionalised polypyrroles, polythiophenes and polyanilines. For example, poly(3methylthiophene) changes from red to blue upon oxidation. Another example is poly(3,4ethylenedioxythiophene) (PEDOT) which is a low band gap polymer, and changes from dark blue to very light blue [64]. These polymers are prepared or formed into thin films on OTE such as ITO glass and placed in an electrolyte solution or gel with another OTE. Insoluble polymers normally need to be formed by electrochemical polymerisation of a monomer in an electrolyte solution. Soluble polymers, however, can be sprayed, printed, spun or cast onto an OTE. Thin films of polyaniline can be prepared electrochemically by oxidation of the monomer, aniline. Typical conditions used are 1 M aniline in 2 M HCl solution and at a current density of 1 A m-2 [65]. Aniline has 4 redox states, which, starting from the fully-reduced polymers, are called leucoemeraldine, emeraldine salt, emeraldine base and pernigraniline. This is shown in figure 2.24 [10].

H X-

H X*

H N N H

H N
X+ N

N * -2 HX pKa 0 H N H +2 HX H N

n
*

X+ N H

H -2e* N H N

H N
+

XX*

N * H Leucoemeraldine

+2e*

N H

N H

-2 HX pKa = 3 -4 H N * N

Emeraldine salt (conductor) +2 HX

- (2 HX+e-)

N * N

N * Pernigraniline

+ (2 HX+e-) *

N H

Emeraldine base

Figure 2.24: Structure of the various redox state of polyaniline.

Each of the states shown in figure 2.24 has its own distinct colour, shown in table 2.8 [10].

State Leucoemeraldine Emeraldine salt Emeraldine base Pernigraniline

Colour Yellow Green Blue Black

max / nm 305 740 420 740 < < 840

Potential Range /mV < -200 0 - 600 600 - 1000 > 1000

Comments Undoped partially doped partially doped Fully doped

Reference [65] [65] [66] [66]

Table 2.8: The various states of polyaniline.

Polyaniline has been used in conjunction with Prussian blue to produce a complementary electrochromic device that exhibits a deep blue to light green colour change [67]. The device works by reducing the highly coloured oxidised polyaniline to produce the much less coloured emeraldine, and at the same time, oxidising the strongly coloured Prussian blue to the less coloured, Prussian green. This produces a device with a very stable coloured state, but a less stable bleached state that becomes re-coloured over a six-hour period [68]. Bleached Coloured 2+ +2 3+ 2K2Fe Fe (CN)6 + Pani 2KFe Fe2+(CN)6 + K+ + Pani
2+

Polypyrrole was first prepared electrochemically in 1965 by MacNeill et. al. [69]. The polymer, as formed, has an intense blue/violet colour which turns to a yellow colour upon dedoping [70]. Polypyrrole is not the most stable of electrochromes, with a typical life span of 104 cycles [71]. Derivatives of polypyrrole have been prepared have have been reported to have enhanced life spans [72] and a changed colour. Table 2.9 gives some examples of this in tetra-n-butylammonium tetrafluoroborate in acetonitrile, where R1 and R2 are substituents on the 3 and 4 positions respectively and X is in the N (1) position [10]. Polymer Colour X H H H CH3 R1 H H CH3 H R2 H COCH3 CH3 H Oxidised Brown Purple Brown/red Reduced Yellow Green Orange/yellow Potential Range/ mV 0 to 700 0 to 1100 -500 to 500 0 to 800 max / nm Oxidised 500 580 440 500

Yellow/brown Brown/yellow

Table 2.9: Electrochromic properties of substituted polypyrroles.

Polythiophene and its derivatives can also be prepared electrochemically, though this is done in non-aqueous medium such as propylene carbonate or acetonitrile. The resulting films have a relative low cycle life with falling contrast ratios and increasing resistance to charge transfer. A possible mechanism that has been proposed is that nucleophilic attack at the ring causes chain cleavage and eventual decomposition of the polymer [73]. The nucleophiles originate from the solvent or the electrolyte. Table 2.10 gives examples of the electrochromic properties of polythiophene and its derivatives [70, 74] and figure 2.25 shows UV-visible scans of poly(3-hexylthiophene) over a range of potentials vrs. Ag/AgCl.

Polymer Polythiophene Poly(3-methylthiophene) Poly(3-methylthiophene) Poly(3-methylthiophene) Poly(3,4-dimethylthiophene) Poly(2,2'-bithiophene) Poly(2,2'-bithiophene) Poly(3-phenylthiophene) Poly(3,4-diphenylthiophene)

Doping Anion ClO4ClO4 BF4CF3SO3CF3SO3BF4 BF4


-

Polymer Colour Oxidised Blue Deep blue Blue/green Blue Dark-blue Blue/grey Blue/grey Green/blue Blue/grey Reduced Red Red Red Red Pale brown Red/orange Red Yellow Yellow

Picryl

BF4-

Table 2.10: Electrochromic properties of polythiophene and its derivatives.


1.4

1.2

1 P3HT at 0 mV P3HT at 100 mV P3HT at 200 mV P3HT at 400 mV P3HT at 500 mV P3HT at 600 mV P3HT at 700 mV P3HT at 800 mV P3HT at 900 mV P3HT at 1000 mV P3HT at 1100 mV

Absorption

0.8

0.6

0.4

0.2

0 300 400 500 600 700 800 900

Wavelength (nm)
Figure 2.25: Electrochromism of poly(3-hexylthiophene).

A range of polythiophene derivatives has been prepared from a monomer with two thiophene units surrounding an aromatic group. If the aromatic group is benzene then the polymer was found to have a pale yellow to green electrochromism. Another group of polythiophene derivatives are with fused ring systems. One example of this is poly (3,2-b;2,3-d)thiophene) which is semi-transparent when doped and opaque when undoped [75]. Other fused rings with electrochromic properties are ones with aromatic ring system attached to the thiophene ring. An example of this is poly(benzo[c]thiophene) which is black when reduced and is yellow in colour upon oxidation [76]. A number of different electrochromic systems exist that have a wide range of colours. Often, complementary systems can be devised using different electrochromes that simultaneously change

colour. Although it is unlikely that extremely rapid changing displays such as monitors and televisions will be made using electrochromes, for other applications where high speed is not so vital, when power consumption is of vital importance, or when large displays are needed, then these needs can be met using an electrochromic system. The aim of this project was to develop novel electrochromic conducting polythiophene derivatives with a tethered metal containing moiety. It was hoped that the metals would alter the properties of the conducting polymer to produce a faster and more stable electrochromic system.

11. REFERENCES
1 K.A. Murray, A.B. Holmes, S.C. Moratti, G. Rumbles, J. Mater. Chem., 9, 2109, 1999 2 E. ter Meer, Ann. Chem., 181, 1, 1876 3 G.H. Brown, Photochromism, John Wiley & Sons Inc., 1971 4 K. Ihn, J. Moulton, P. Smith, J. Polym. Sci. Part B, Polym Phys. 31, 735, 1993 5 J. Mardelen, E. Samuelsen, A. Pendersen, Sythn. Met., 55, 378, 1993 6 Y. Yokoyama, T. Yamane, Y. Kurita, J. Chem. Soc., Chem. Commun., 1722, 1991 7 S.K. Deb, J.A. Chopoorian, J. Appl. Phys., 37, 4818, 1968 8 W.R. Salaneck, D.T. Clark, E.J. Samuelsen, Science and Aplications of Conducting Polymers, IOP Publishing Limited, 1991 9 B.W. Faughnan, R.S. Crandall, P.M. Heyman, RCA Rev., 36, 177, 1975 10 P.M.S. Monk, R.J. Mortimer, D.R. Rosseinsky, Electrochromism: Fundamentals and Applications, VCH Inc., Weinheim, 1995 11 F.G. Cottrell, Z. Physik. Chem., 42, 385, 1902 12 R.J. Jasinski, J. Electrochem. Soc., 124, 637, 1977 13 G. Beni, C.E. Rice, in P. Vashishta, J.N. Mundy, G.K. Shenoy (eds.), Fast Ion Transport in Solids: Electrodes and Electrolytes, Elsevier, 1979 14 J.-P. Randin, J. Electrochem. Soc., 129, 1215, 1982 15 B. Scharifker, C. Wehrmann, J. Electrochem. Soc., 185, 93, 1985 16 D.R. Rosseinsky, J.D. Slocombe, A.M. Soutar, P.M.S. Monk, A. Glidle, J. Electroanal. Chem., 258, 233, 1989 17 D.R. Rosseinsky, P.M.S. Monk, R.A. Hann, Electrochim. Acta, 35, 1113, 1990 18 B. Grant, N. Clecak, M. Oxsen, A. Jaffe, G. Kellar, J. Org. Chem., 45, 702, 1980 19 J.T.S. Irvine, B.R. Eggins, J. Grimshaw, J. Electroanal. Chem., 271, 161, 271 20 F.G.K. Baucke, Rivista della Staz. Sper. Vetro, 6, 119, 1986 21 W. Wagner, F. Rauch, T. Bange, Nuc. Inst. Meth. Phy. Res., B50, 27, 1990 22 P. Hagenmuller, W. Van Gool, Solid Eletrolytes: General Principles, Characterisation, Materials, Applications, Academic Press, 1978 23 B.W. Faughnan, R.S. Crandall, P.M. Heyman, RCA Rev., 36, 177, 1975 24 S.F. Cogan, T.D. Plante, M.A. Parker, R.D. Rauh, J. Appl. Phys., 60, 2735, 1986 25 S. Passerini, B.J. Scrosati, Electrochem. Soc., 141, 889, 1994 26 P.M.S. Monk, S.L. Chester, Electrochim. Acta, 38, 1521, 1993

27 K.-C. Ho, T.G. Rukavina, C.B. Greenberg, Proc. Symp. Electrochromic Mater., 94-2, 195, 1994 28 H. Inaba, M. Iwaku, K. Nakase, H. Yasukawa, I. Seo, N. Oyama, Electrochim. Acta, 40, 227, 1995 29 S.A. Sapp, G.A. Sotzing, J.R. Reynolds, Chem. Mater.,10, 2101, 1998 30 S.K. Deb, Solar Energy Mater. Solar cells, 25, 327, 1992 31 M.S. Habib, S.P. Maheswari, Solar Energy Mater. Solar cells, 25, 195, 1992 32 C. Arbizzani, M. Mastragostino, L. MeneghelloM. Morselli, A.J.Zanelli, J. Appl. Electrochem., 26, 121, 1996 33 Q. Pei, G. Yu, C. Zhang, Y. Yang, A.J. Heeger, J. Science, 269, 1086, 1995 34 M. Granstom, O. Inganas, Adv. Mater., 7, 1012, 1995 35 J. Scarminio, A. Urbano, B.J. Gardes,J. Of Mater. Sci. Lett., 11, 562, 1992 36 D.H. Oh, S.G. Boxer, J. Am Chem. Soc., 112, 8161, 1990 37 P.M.S. Monk, S.L Chester, D.S. Higham, Proc. Electrochem. Soc., 94-2, 100, 1994 38 J.S. Golden, B.C.H. Steele, Solid State Ionics, 28-30, 1755,1988 39 S. Gottesfield, J.D.E. McIntyre, G.Beni, J.L. Shay, Appl. Phys. Lett., 33, 208, 1978 40 H. Suiyang, C. Fengbo, Z. Jicai, in B.V.R. Chowdari, S. Radhakrishna (eds.), Proceeding of the International Seminar on Solid State Ionic Devices, World Publishing, 521, 1988 41 S.K. Deb, Appl. Optics, Supp., 3, 192, 1969 42 A.I. Gavrilyuk, F.A. Chudnovski, Sov. Tech. Phys. Lett., 3, 69, 1977 43 P. Baudry, A.C.M Rodriguez, M.A. Aegerter, L.O. Bulhes, J. Non-Cryst. Solids, 121, 319, 1990 44 L.D. Burke, O.J. Murphy, J. Electroanal. Chem., 109, 373, 1980 45 S.F. Cogan, T.D. Plante, E.J. Anderson, R.D. Rauh, Proc. S.P.I.F., 562, 23, 1985 46 L.D. Burke, D.P. Whelan, J. Electroanal. Chem., 103, 179, 1979 47 K. Itaya, K. Shibayama, H. Akahoshi, S. Toshima, J. Appl. Phys., 53, 804, 1982 48 K. Itaya, T. Ataka, S. Toshima, J. Am. Chem. Soc., 104, 4767, 1982 49 K. Itaya, I. Uchida, Inorg. Chem., 27, 389, 1986 50 M.K. Carpenter, R.S. Conell, S.J. Simko, Inorg. Chem., 29, 845, 1990 51 M.M. Nicholson, F.A. Pizzarello, J. Electrochem. Soc., 128, 1740, 1981 52 G.S.E. Collins, D.J. Schiffrin, J. Electroanal. Chem., 139, 335, 1982 53 I.S. Kirin, P.N. Moskalev, Y.A. Makashev,Russ. J. Inorg. Chem., 10, 1065, 1965 54 M.T. Riou, M. Auregan, C.J. Clarisse, J. Electroanal. Chem., 187, 349, 1965 55 L.G. Tomilova, E.V. Chernykh, V.I. Gavrilov, I.V. Shelepin, V.M. Derkacheva, E.A. Luk'yenets, Zh. Obshch. Khim., 52, 2606, 1982 56 L.G. Tomilova, E.V. Chernykh, E.V. Ioffe, E.A. Luk'yenets, Zh. Obshch. Khim., 53, 2594, 1983 57 F. Castaneda, C. Piechocki, V. Plichon, S. Simon, Vaxiviere, J. Electrochim. Acta, 31, 131, 1986

58 Y. Liu, K. Shigehara, M. Hara, A. Yamada, J. Am. Chem. Soc., 113, 440, 1991 59 H.J. Byker, Proc. Symp. Electrochromic Mater., 94-2, 3, 1994 60 R.G. Compton, P.M.S. Monk, D.R. Rosseinsky, A.M. Waller, J.C.S. Faraday Trans., 86, 2583, 1990 61 P.J.S. Foot, R. Simon, J. Phys. D: Appl. Phys., 22, 1598, 1989 62 B. Scrosati, Applications of Electroactive Polymers, Chapman and Hall, 1993 63 B.C. Thompson, P. Schottland, G. Sonmez, J.R. Reynolds, Synth. Met., 119, 333, 2001 64 T. Kobayashi, H. Yoneyama, T. Tamura, J. Electroanal. Chem., 161, 419, 1984 65 M. Kobayashi, J. Chen, T.C. Moraes, A.J. Heeger, F. Wudl, Syth. Met., 9, 77, 1984 66 A. Watanabe, K. Mori, Y. Iwaski, Y. Nakamura, S. Niizuma, Macromolecules, 20, 1793, 1987 67 E.A.R. Duek, M.-A. De Paoli, M Mastragostino, Adv. Mater., 4, 287, 1992 68 E.A.R. Duek, M.-A. De Paoli, M Mastragostino, Adv. Mater., 5, 650, 1993 69 R. MacNeill, D.E. Weiss, D. Willist, Aust. J. Chem., 18, 477, 1965 70 J.E. Dubois, F. Garnier, G. Tourillon, M. Gazard. J. Electroanal. Chem., 148, 299, 1983 71 M.A. De Paoli, S. Panero, P. Prosperi, B. Scrosati, Electrochim. Acta, 35, 1145, 1990 72 R.B. Bjorkund, I. Lundstrm, J. Electron. Mater., 14, 39, 1985 73 J. Wang, Electrochim. Acta, 39, 417, 1994 74 M. Gazard, in T.A. Skotheim, R.L. Elsenbaumer, J.R. Reynolds (ed), Handbook of Conducting Polymers, Vol. 1, Marcel Dekker, 1986 75 M. Mastrogostino, A.M. Marinangeli, A. Corradini, C. Arbizzani, Electrochim. Acta, 32, 1589, 1987 76 H. Yashima, M. Kobayashi, K.-B. Lee, D. Chung, A.J. Heeger, F. Wudl, J. Electrochem. Soc., 134, 46, 1987

You might also like