Sei sulla pagina 1di 26

Theoretical and Experimental Chemistry, Vol. 41, No.

3, 2005

THE SOLITONIC NATURE OF THE ELECTRONIC STRUCTURE OF THE IONS OF LINEAR CONJUGATED SYSTEMS

A. D. Kachkovskii

UDC 541.61

Features of the electron density distribution in the ground and excited states, the position of the energy levels, and the equilibrium molecular geometry of the ions of linear p-electron systems, including conjugated polymers and oligomers, cationic and anionic polymethine dyes, and the radical-ions of a,w-substituted polyenes, are reviewed. By using the soliton concept it was possible to explain many unusual properties of this type of organic compound: The quasimetallic conductivity, the marked change in the spectral characteristics of ionic dyes absorbing and emitting light in the near IR region of the spectrum, etc. It was shown on the basis of semiempirical and nonempirical calculations that the charges in collective p-electronic systems form soliton waves, the width of which does not depend on the length of the molecule. In the excited state the size of the solitons increases significantly. The introduction of donor and acceptor terminal groups can destroy the symmetry of the charge distribution and molecular geometry at a conjugation chain of critical length exceeding the dimensions of the soliton. It was established that the injection of electrons/holes not only leads to the appearance of a soliton level at the center of the energy gap but is also accompanied by significant displacement of the top of the valence band and the bottom of the conduction band. Key words: conjugated systems, ions, electronic structure, soliton, excited states, quantum-chemical calculations.

INTRODUCTION Among the greatest discoveries of the last quarter of the twentieth century it is necessary to include the unexpected discovery of quasimetallic electrical conductivity in organic conjugated polymers doped with oxidizing or reducing agents. In 2000 this discovery was the subject of a Nobel prize in the chemistry section, and this represents an outstanding measure of both the experimental investigations and the theoretical developments in the field of one specific type of organic compound, i.e., p-electronic molecules. These works formed the basis of contemporary ideas about the electronic and spectral characteristics of conjugated systems. They provide a landmark in molecular design and the practical creation of new prospective materials for molecular electronics and nonlinear optics (see, for example, the review [1] and the references therein). This discovery brought about a reexamination of the theoretical concepts based on classical views about the behavior of p electrons in the ions of conjugated molecules, particularly in anionic and cationic polymethine dyes, the radical cations/anions of polyenes, polyacetylenes, and conjugated systems based on heterocycles, and thereby stimulated the development of a qualitatively new theoretical model taking account of the characteristics of charge distribution in the form of finite waves and the appearance of impurity levels in the forbidden band, brought about by the injection of electrons or holes [1-10]. ___________________________________________________________________________________________________ Institute of Organic Chemistry, National Academy of Sciences of Ukraine, 5 Murmanska Vul., 02094 Kyiv, Ukraine. E-mail: iochkiev@ukrpack.net. Translated from Teoreticheskaya i ksperimentalnaya Khimiya, Vol. 41, No. 3, pp. 133-155, May-June, 2005. Original article submitted June 7, 2005. 0040-5760/05/4103-0139 2005 Springer Science+Business Media, Inc. 139

An important feature of p-electronic molecules as semiconducting systems with a comparatively small forbidden band is the strong absorption of light in the visible region of the spectrum. This formed the basis of the use of such compounds in materials for the conversion of light energy; polyacetylenes and their analogs (see, for example, [11]), polymethine dyes and cyanines [12-14], substituted polyenes and stilbenes [15, 16], and donoracceptor molecules and merocyanines [17, 18] can be considered classical subjects. Recently there has been a significant increase in the attention paid by researchers not only to one-photon absorption by conjugated molecules but also to the two- or three-photon absorption that is realized during excitation, for example, by several accurately focused laser beams [18-20]. As a result it has become important to construct adequate quantum-chemical models that correctly describe the electronic structure of charged conjugated systems in excited states with due regard to the rules of prohibition of electronic transitions induced by one- and multiphoton excitation. The new experimental discoveries and new theoretical concepts support the main theory: The unique properties of conjugated molecules and polymeric systems arise from the specific characteristics of their chemical structure and the electronic state of the extended p-electronic systems and also from the appearance of positive or negative charge in the system. The present work is devoted to examination of such charged conjugated molecules.

FEATURES OF THE CHEMICAL STRUCTURE AND ELECTRONIC STATE In the general form the linear conjugated systems examined here can be represented by the formula [R1(CH)m R2]z, I z = 0, 1,

where R1 and R2 are terminal groups conjugated with the polymethine chain. The carbon atom of the methine group in the conjugation chain is in the sp2-hybrid state; a single generalized system of p electrons is formed from the 2pz electrons. Both individual atoms (O, S, Se, B) or simple groups of atoms (NH2, OH, SH, CN) and complex carbocyclic or heterocyclic groups with their own branched p-electronic systems can be used as terminal groups. Accordingly, even the central part can be constructed as a branched p system, for example, in polyphenylvinylenes and their derivatives
H3C R1
n

R2 II

CH3

or polythiophenes and their hetero analogs


X
n

H3C

CH3

X = S, O, NCH3 III

In all cases, however, the main element is an extended chain of p centers represented in the simplest case by CH groups. Linear conjugated compounds with symmetrical structure (R1 = R2) are traditionally divided, depending the evenness of the number of methine groups in the chain (m), into two types: Polymethine dyes with an odd number of p centers (m = 2n + 1) and a,w-disubstituted polyenes containing an even number of methine groups (m = 2n) in the conjugated chain. On the condition of a closed electron shell the symmetrical polymethine dyes are ions, i.e., cations with donor groups R, for example, cyanine dyes 140

Y N
+

Y
n

CH3 CH3 Y = CH CH, C(CH3)2, NCH3, O, S, Se IV

or anions, for example, oxonols O(=)n=O V Polyenes with a stable closed electron shell can exist in the form of both neutral molecules and diions. Such dications or dianions as compounds VI and VII, for example, are stable, depending on the nature of the terminal groups. H2N(CH=CH)nNH2 H2N+=CH(CH=CH)n1CH=+NH2, VIn VIc O=CH(CH=CH)n1CH=O O (CH=CH)n1O. VIIn VIIa Polymethine dyes and substituted polyenes differ substantially in many electronic, chemical, and spectral characteristics. As a rule, therefore, they have been investigated separately, particularly as concerns spectral investigations [12-14, 21-27]. Thus, it was established on the basis of calculation and spectral data (13C NMR) that in polymethine dyes significant positive and negative charges alternate at neighboring carbon atoms while the lengths of the carboncarbon bonds are approximately equal and amount to approximately 1.4 [14, 28-32]. The bond lengths change little in the transition to the first excited state, whereas the charges at the atoms change sign [12-14]. In polyenes, on the other hand, the charges are practically absent in the ground state, but the lengths of the CC bonds have maximum alternation, and the amplitude of alternation in the long molecules amounts to 0.08 (see, for example, the review [33] and the references therein). Excitation of the polyenes with light quanta leads to significant change in the bond lengths, and the sequence of the longer and shorter bonds is reversed; the electron density distribution, however, remains practically unchanged [12]. The specific characteristics of the electronic structure of conjugated molecules in the ground and excited states were examined in detail by Daehne. He developed the triad theory [34-38], in which the p systems were classified and it was proposed to separate them into three main groups: Aromatic molecules, polyenes, and polymethines. Unlike linear conjugated systems, aromatic molecules have cyclic structure and are characterized by equalization both of the electron density at the atoms and of the bond lengths, and the transition to the excited state does not destroy the equalization of the populations at the atoms and bonds [34-38]. In order to take account of the effect of the topology of the terminal groups on the electronic characteristics of linear conjugated systems, primarily on the position of the boundary levels and the size of the energy gap, Dyadyusha and his students developed a model of long polymethines in the Hckel approximation [39-46]. It was proposed to evaluate the effect of the terminal groups through their topological indices F0 and L. The parameter F0, called the electron-donating capacity, corresponds to displacement of the nodes of the frontier MOs from their positions in the unsubstituted conjugated molecules and, consequently, characterizes the transmission of electron density from the terminal groups to the polymethine chain. In the topological Hckel approximation the index F0 is related to the relative position of the highest occupied MO (xg) and the lowest unoccupied MO (xe): F0 = 90[xe/(xe xg)], and for systems with a stable shell it lies in the region of 0 < F0 < 90. The second topological parameter L characterizes the effective elongation of the conjugation chain of the inherent p-electronic system of the terminal groups and is expressed quantitatively by the equivalent number of methine groups. Both indices, correctly 141

Fig. 1. The absorption spectra of thiacyanines IV, X = S, in alcohol; the number of the curve corresponds to the number of vinylene groups in the chain n (dyes with n > 3 contain trimethylene bridging groups in the chain [51, 52]).

calculated mathematically by means of a topological matrix (or a contiguity matrix), proved extremely useful for estimating the donor characteristics of the terminal groups and their effect on the shift of the long-wave absorption band of polymethine dyes [12, 39-46]. In the triad theory the concept of ideal electronic states was introduced, i.e., the aromatic state with equalized populations at the atoms and bonds, the polyene state with maximum alternation of the bond lengths and no charges at the atoms, and the polymethine state with maximum alternation of positive and negative charges at the carbon atoms and completely equalized carboncarbon bond lengths [34-38]. By using the long polymethine model and the parameters L and F0 it was possible to determine the conditions for the realization of these ideal polymethine and polyene states in linear conjugated compounds with arbitrary terminal groups [12]. However, neither triad theory nor the long polymethine model explains the reasons for the alternation or equalization of the populations at the atoms and bonds. Complications also arise during the classification of conjugated systems with an open electron shell. Thus, in the radicals of polyenes the bond lengths must be equalized while retaining the equalization of electron density at the atoms [47, 48]. In the polymethines, however, the appearance of the unpaired electron leads to equalization of the electron density at the atoms, whereas the bond lengths do not undergo significant changes [47-49]. Thus, the difference in the p-electronic structure of the two types of linear conjugated compounds disappears. A series of problems arose during the experimental investigation of vinylogous series of linear conjugated compounds, which cannot be interpreted correctly in terms of triad theory and the long polymethine model. The most characteristic, for example, are the marked broadening of the absorption bands of ionic polymethine dyes and the stepwise decrease in the quantum yields of fluorescence with increase in the length of the fluorophore. Thus, according to the concept of the ideal polymethine state the introduction of vinylene groups into the conjugation chain of cyanine dyes IV regularly shifts the maximum of the long-wave absorption band each time by approximately 100 nm (the so-called vinylene shift, detected experimentally [50]); in the limit the energy of the transition tends toward zero, while the intensity increases continually. It is seen from Fig. 1 that these relationships are actually fulfilled for dyes with a comparatively short chain (n = 1, 2, 3), but the transition to the higher homologs is accompanied by a decreases in the intensity of the spectral bands, while an additional band very sensitive to polarity appears in the short-wave region of the spectrum [51, 52]. A similar change in the form of the spectral bands is observed for other dyes that absorb in the near IR region [53-56]. Attempts to explain this effect by the solvent effect can hardly be considered satisfactory. Certain effects in the fluorescence spectra also remained unexplained. For example, it was established that at a critical polymethine chain length in the cyanines IV (n = 3) the subsequent introduction of a vinylene group is accompanied by a 142

TABLE 1. The Quantum Yield of Fluorescence j (%) for Dyes IV in Ethanol [57]
n Y = C(CH3)2 Y=S

1 2 3 4 5

3 25 28 0.3 0.1

5 43 34 0.1 0.06

decrease of more than two orders of magnitude in the quantum yield of fluorescence, whereas the transition from dicarbocyanines (n = 2) to tricarbocyanines (n = 3) has comparatively little effect on the effectiveness, as seen from the spectral data presented in Table 1 [57]. investigations of the fluorescence spectra of thiacyanines IV, Y = S, showed that two components differing significantly in their life times exist in the excited state [58, 59]. These effects cannot be explained satisfactorily in terms of traditional theories about the nature of the electronic transitions in polymethine dyes. It has been suggested in a series of papers that relaxation of the excited state can take place in two ways, leading to symmetrical and unsymmetrical isomers [60-63]. However, the main problem during investigation of the electronic structure of linear conjugated compounds remained the unusually high electric conductivity or, in other words, the current carrier problem. Its continued importance is demonstrated by the fact that in 2004 a separate issue (No. 11) of the leading chemical journal Chemical Reviews (see, for example, the introductory paper [64] and the references therein) was devoted to advances in the investigation of molecular conductivity. All this prompted the development of new models, based on quantitatively new ideas about the electronic structure of linear conjugated systems.

THE SOLITON CONCEPT Many of the effects observed for the ions of linear p-electronic compounds can be explained by using the concept of charge waves or solitons. Such physical models have been used by a series of authors [2-10, 65-75]. The idea was based mainly on the fact that linear conjugated molecules, both comparatively short polymethine dyes and polyenes and oligomers and conjugated polymers, can be regarded as a one-dimensional collective system formed as a result of the strong interaction of p electrons (see, for example, the review [76]). In contrast to the three-dimensional case of crystals, excitations in a one-dimensional system are only collective in nature. The injection of electrons or holes can provide the excitation, while the injected charge is not delocalized over the whole system, as is usually assumed by chemists [12-14, 21, 22, 34-38], but are self-localized in the form of a solitary wave, i.e., a kink or soliton [2, 76]. When the length of the polymethine chain exceeds the dimensions of the soliton wave, the soliton becomes mobile, in so far as it can be localized in any segment of the conjugated chain without change in the total energy of the system. Such movements of charge in the form of a soliton wave can serve to explain the unusually high quasimetallic conductivity of conjugated molecules (polyacetylenes, polythiophenes, etc. [3]). As shown by investigations (for example, [65-75]), the appearance of charge in the chain of conjugated bonds leads to equalization of the carboncarbon bond lengths in the segment of the chain where the soliton is localized, whereas in the remaining part of the conjugated molecule the adjacent CC bonds alternate strongly in length. It can be considered that the charge generates a defect in the molecular geometry, which in solid-state physics is usually called a supersonic soliton [2]. Physicists most often regard the interaction of charge and defect in a one-dimensional lattice as capture of an electron or a hole by the soliton. 143

Fig. 2. Diagram of the arrangement of the conduction band (CB), valence band (VB), and soliton level in a neutral p-electronic molecule (a), cation (b), radical-ion (c), and anion (d).

According to Davydov [2], a system consisting of an electron (hole) and its surrounding deformation, moving at a constant rate (~1 km/s) and described by a hyperbolic function, is customarily called a self-localized electron (hole) or an electronic (hole) soliton. The self-localization of the electron (hole) is due to far-reaching interaction between the charge and the vector polarization field arising under the influence of this same charge [2]. Such a self-localized state is also called a polaron. The last name (and the name bipolaron for two-charge excitation) is also used by other authors, while the term soliton remains for waves in neutral systems [3, 4, 76]. It was subsequently proposed to treat the electron density wave (electronic or hole soliton) and the wave with variation of the bond lengths, called topological or geometric soliton, separately [65, 67-69]. In fact we will also use such a distinction between the charge and geometric solitons in the present paper during analysis of the nature of the soliton waves in the ions of linear conjugated systems. It should be noted in particular that for molecules present at a global and stable minimum the localizations of the charge and geometric solitons must coincide. However, such an assertion is not so obvious for molecules at the moment of injection of charges, during the transfer of charge from one polymeric chain to another, or during relaxation of the excited electronic state, when the movement of the geometric soliton (as the wave of the vibrations of the carbon atoms) can fall behind the more mobile charge wave. The injection of multiple charges must lead to the appearance of the corresponding number of solitons [3, 10]. In the present work, however, we will restrict ourselves to examination of only the monoions of conjugated molecules in order to analyze the behavior of the charge waves and the geometric defects in the ground and excited states in relation to the molecular topology for the simplest example. In terms of band theory [2, 3] neutral conjugated molecules (even-numbered polyenes and the radicals of odd-numbered polymethines) can be regarded as systems in which each cell contains one electron (so-called Mott insulators [77]); they form the valence band of the filled levels and the conduction band, separated by a forbidden band (about 3 eV) (Fig. 2a). The appearance of the soliton during the injection of charges (as the excited state of the Mott system) is then interpreted as the appearance of an impurity level in the forbidden band, as shown diagrammatically in Fig. 2b-d. It is usually assumed that the length of the conjugation chain is infinite, the soliton level lies in the middle of the energy gap, and its position does not depend on its occupation with electrons, i.e., the system can be a radical in the case of the ions of polyenes (Fig. 2c) and a cation (Fig. 2b) or an anion (Fig. 2d) for polymethines with a closed electron shell. Increase in the number of solitons (multiple ions or highly doped p systems) naturally leads to the appearance of the corresponding number of soliton levels in the energy gap, which by interacting with each other split and form soliton (impurity) bands. The width of such a band for polypyrrole, for example, amounts to 0.4 eV, whereas the width of the most forbidden band increases from 3.2 to 3.6 eV [2, 3]. Since the soliton levels in doped conjugated polymers are close to the Fermi level, the transfer of electron to this level (reduction) or the removal of an electron from it (oxidation) takes place almost without expenditure of energy. Processes with transfer of an electron (or hole) between the chains in films of conjugated polymers are, supposedly, responsible for the appearance of conductivity [2, 3, 76, 78-82], although the solution of the charge transfer problem is undoubtedly more complicated. In the present work, however, we will restrict ourselves to examination of features of the electronic structure in the ions of conjugated molecules and the possible application of ideas that arise in terms of the soliton concept to p-electronic compounds with a comparatively short polymethine chain, such as polymethine dyes and disubstituted polyenes.

144

QUANTUM-CHEMICAL MODELLING On account of the simplicity of their chemical structure and the ease of variation of the molecular topology linear conjugated compounds have traditionally been used as a polygon for the development of new quantum-chemical models (see, for example, [1, 5, 12, 35, 83, 84]). In all approximations, from the simplest Hckel method [85] up to multiconfiguration theories [86, 87], it has invariably been assumed that the system of p electrons in linear conjugated molecules is collective in nature. Here we restrict ourselves to only those approaches that are based on a one-electron approximation (the Hartree Fock method). Among these approaches the most popular has been the MO-LCAO approximation, in which the ni molecular orbitals (MOs) are constructed as a linear combination (LC) of basis atomic orbitals (AOs) cm: ji = C i , where Cim are the

coefficients of the MO normalized in such a way that C i = 1 [85, 88]. It was then possible in all the approximations to model correctly features of the electronic structure in a one-dimensional system of p electrons. In the simplest, so-called metallic, model or the free electron model, proposed by Kuhn [89-92], Bayliss [93], Basu [94], Simpson [95], Ruedenberg [96], and Olszewski [97], it was suggested that the p electrons of the conjugated molecule are in the field of the s core and form an electron gas, as in a metal. Thus, the molecule can be represented as an ideal conductor of length L, and the movement of the free electron in it is described by standing electronic waves of sinusoidal form: ji(x) = L1/2 sin [(pix/L)], i = 1, 2, N,

(1)

where N is the number of p electrons, x is the coordinate along the conductor, and i is the number of the wave that can be compared with the molecular orbital ji. The electron density distribution along the conductor is then represented by the square of function 2 (x), which also has the form of a sinusoidal wave. i The energies of the electronic states ei are described in the following way: ei = h2i2/8mL (2)

where h is Plancks constant and m is the mass of the electron. If the electronic levels are filled, no dependence of their positions on the number of electrons is assumed, i.e., they are not sensitive to the charge. Therefore, in terms of the metallic model it is not possible to describe correctly either the formation of the bands of levels for the neutral conjugated molecules or the appearance of a specific level in the forbidden band during the injection of electrons or holes. Some increase of the energy gap in the even polyenes can be achieved if account is taken of the alternation of the bond lengths as the alternation of the barriers at the bottom of the potential box [91, 97]. However, calculations in terms of higher approximations that take account of electronic interaction predict a significant value for the energy gap in the neutral radicals of the odd polymethines, in spite of the equality of the CC bond lengths in the chain [48, 49]. The reason for the appearance of the gap is the fact that the neutral molecule corresponds to the conditions of the Mott isolator [77]. In the Hckel approximation, which is still called the topological approximation, the interaction of the electrons is also not taken into account in explicit form [85, 88]. In the case of an unsubstituted conjugated chain the solutions of the Hamiltonian (the coefficients Cim and the energies of the orbitals ei) are written as normal trigonometrical functions of the total number of p centers N, the number of the orbital i, and the number of the atom in the chain m and do not depend on the populations of the levels [85, 88]: ei = a + 2 cos [ip/(N + 1)]b, Cim = [2/(N + 1)]1/2 sin [imp/(N + 1)], (3) (4)

where a and b are empirical parameters (coulombic and resonance integrals). The analytic expressions for ei and Cim were also obtained for polymethine compounds (I) with the simplest terminal groups R (on the condition that aR = a b) [99-102] and also for conjugated systems with alternating single and double CC 145

Fig. 3. The charges at the carbon atoms in the polymethine cation (VIII), n = 8, and polymethine anion (IX); the ab initio 6-31G** approximation.

bonds [103-105]. Unlike the metallic model or the free electron method, the Hckel approximation takes account of the topology of the conjugated molecule mathematically correctly, since the Hckel determinant corresponds to the contiguity matrix in graph theory, and the set of molecular levels coincides with the spectrum of the graph for the conjugated system of the molecule [106]. Although the Hckel approximation is fairly rough, it was nevertheless used in the development of the first soliton model the so-called SSH soliton theory (Su, Schrieffer, Heeger) [65, 66]. It was shown that the doping of polyacetylenes with alternating CC bond lengths leads to the appearance of a segment of chain where the difference in the lengths of two adjacent bonds, Drj = rj rj+1, is described not by the usual trigonometric function but by the soliton hyperbolic function: Drj = Drtanh [(j jc)/l] (5)

where j is the number of the current bond, jc is the number of the bond at the center of the soliton, l is the half-width of the soliton wave, and Dr is the limiting value of the alternation of bond lengths for an infinitely long conjugation chain. Subsequent calculations showed that Dr 0.1 , while the parameter l is equivalent to approximately seven CC bond lengths, i.e., the soliton is located on a section of chain containing approximately 15-20 CH groups [65-69]. The charge at the atoms in the SSH approximation is also described by a hyperbolic function as the soliton wave: qm = l1sech2[(m mc)/l] cos2[(m mc)p/2] (6)

where m is the number of the atom in the chain. The charges reach their maximum value at the center of the soliton (mc), where the bond lengths are equalized. By refinement of the quantum-chemical approximations that take proper account of the interelectronic interaction, ranging from the p-electron version of the Parr Pariser Pople (PPP) method and the semiempirical CNDO/2, MINDO/3, AM1, PM3, and other methods including valence electrons to the nonempirical ab initio method with ever-increasing basis sets it was possible to investigate the distribution of charge in the ions of conjugated systems and features of their molecular geometry more reliably and to make a correlation with experimental data and, primarily, with the chemical shifts in the NMR spectra. The charges at the carbon atoms in unsubstituted polymethine cations (VIII) and polymethine anions (IX) with n = 8, calculated in the ab initio approximation, are given as an example in Fig. 3. H2C+(CH=CH)nCH=CH2, VIII H2C(CH=CH)nCH=CH2. IX

146

Fig. 4. The alternation of charge Dqm in the polymethine cation (VIII), n = 8, calculated in various approximations; m is the number of the bond.

Fig. 5. The charges q (a), alternation of charge Dq (b), and bond lengths l (c); the alternation of bond lengths Dl (d) and Dl (e) in the polymethine cation (VIII), n = 40; AM1 approximation. 147

As seen, the values of the charges alternate strongly, and the alternation of the charges in the cations and anions is opposite in phase. This agrees with the alternation of the 13C chemical shifts of neighboring atoms in both cationic and anionic polymethine dyes: 90-120 ppm for the positions with excess electron density and 150-170 ppm for the positions with an electron density deficiency [29, 107-109]. At the same time the difference in the value of the chemical shifts in neutral polyenes amounts to 1-2 ppm, while the values themselves are close to 130 ppm [110-112]. The ab initio calculations predict only negative charges at the atoms both in the anions and in the cations (Fig. 3), and this is due to significant polarization of the CH bond; thus, the charges at the hydrogen atoms are approximately +0.23. The calculations in the other approximations give somewhat smaller values: +0.13 (AM1), +0.01 to +0.04 (CNDO/2, ZINDO/S). Calculations in the last two approximations show that charges differing in sign alternate at the carbon atoms in the chain [48]. It is also seen from Fig. 3 that the amplitude of alternation decreases from the center of the chain to its edges, so that the charges form a wave or wave packet that is described, as shown in the fundamental papers [67-70], by a hyperbolic function. For analysis of the dependence of the charges on the molecular topology it is more convenient to use the difference between the values of the electron density at neighboring atoms Dqm = (1)m(qm qm+1) or simply the absolute value of the amplitude Dqm = qm qm+1. (8) (7)

Figure 4 shows the Dqm = f(m) graphs for the polymethine cation (VIII) (n = 8), calculated in the most widely used approximations. Although the semiempirical AM1 and CNDO/2 methods somewhat overestimate the degree of alternation of electron density compared with the ab initio method while the spectroscopic ZINDO/S method gives somewhat lower values for the amplitude Dqm, all the approximations nevertheless predict the soliton character of the distribution of the injected charge along the conjugation chain. If the conjugated system is long enough, the charge wave breaks away from the edges of the chain and now appears as a typical solitary wave of finite dimensions, i.e., as a group soliton (Fig. 5a, b). True, the calculations never lead to a zero value for the amplitude Dqm at the edges of the chain but give a constant value (0.012 in the AM1 approximation), which is probably due to nonequivalent polarization of the CH bonds at the even and odd positions of the chain. Figure 5c shows the lengths of the CC bonds calculated for polymethine cations (VIII) with 40 vinylene groups: It is seen that they alternate strongly at the edges of the chain, so that the difference in the lengths of neighboring bonds moves asymptotically toward 0.099 , whereas the calculations predict equalization of the bond lengths at the location of the charge wave. The form of the geometric soliton can be determined from graphs for the functions Dln (Fig. 5d) and Dln (Fig. 5e), which like functions (7) and (8) characterize the lengths of the CC bonds along the conjugation chain [67-70]: Dln = (-1)n(ln ln+1), Dln = ln ln+1, (9) (10)

where n is the number of the bond. After optimization of the molecular geometry and the attainment of a global energy minimum the centers of the geometric and charge solitons naturally coincide, although in the dynamic vibrational relaxation process, for example after excitation by a light quantum or after the injection of an electron/hole, the localization of both waves can differ. In contrast to the distribution of charges at the atoms in the cations (VIII) and anions (IX), which alternate in opposite phase (Fig. 3), the alternation of the bond lengths in both ions, on the other hand, coincides in phase. Moreover, the bond lengths themselves are very close, as can be seen from the data in Table 2. According to the definition of a soliton as a solitary wave, its dimensions must be preserved for any length of conjugated system. Figure 6 shows the graphs for the Dqm function for polymethine cations (VIII) with a variable number of 148

TABLE 2. The Bond Lengths ln () in Polymethine Cations (VIII) and Anions (IX) (n = 8), Calculated in ab initio (6-31G**) and AM1 Approximations
Ab initio Bond number ln(VIII) ln(IX) D ln(VIII) AM1 ln(IX) D

1 2 3 4 5 6 7 8 9 10 11 12 13 14

1.4565 1.3412 1.4368 1.3552 1.4179 1.3736 1.3960 1.3960 1.3736 1.4179 1.3552 1.4368 1.3412 1.4565

1.4467 1.3417 1.4264 1.3538 1.4075 1.3696 1.3874 1.3874 1.3696 1.4075 1.3538 1.4264 1.3417 1.4467

0.0098 0.0005 0.0104 0.0014 0.0104 0.0040 0.0086 0.0086 0.004 0.0104 0.0014 0.0104 0.0005 0.0098

1.4428 1.3557 1.4290 1.3671 1.4143 1.3813 1.3978 1.3978 1.3813 1.4143 1.3671 1.4290 1.3557 1.4428

1.4401 1.3537 1.4267 1.3646 1.4116 1.3781 1.3946 1.3946 1.3781 1.4116 1.3646 1.4267 1.3537 1.4401

0.0027 0.0020 0.0023 0.0025 0.0027 0.0032 0.0032 0.0032 0.0032 0.0027 0.0025 0.0023 0.0020 0.0027

Note. D = ln(IX) ln(VIII).

Fig. 6. The alternation of charge Dqm in the polymethine cation (VIII), n = 4k; AM1 approximation.

vinylene groups in the chain: n = 4k, where k = 1-5. It is easy to see that the graphs practically coincide with the exception of the shortest molecules, where the effect of the ends of the chain, as a potential barrier of infinite height, must have the greatest influence. Nevertheless, the soliton character of the charge wave shows up completely even with a short chain length: It does not fit into the dimensions of the conjugated system, and the remaining part breaks away without substantial distortion of the displayed section. In principle, it is possible mentally to construct the remaining section of the soliton. This is a very important and fundamental result, since it makes it possible to apply the soliton concept, i.e., the distribution of charges in the 149

Fig. 7. The positions of the energy levels in neutral polymethine radicals (X) (a, Pm-Cr) and in the polymethine cations (VIII) (b, Pm-C+); AM1 approximation.

form of a solitary wave, to oligomers and even to such comparatively short conjugated ions as polymethine dyes. By such an approach it is possible to explain correctly the unusual experimental effects that appear, for example, in the electronic absorption spectra. We note also that standard quantum-chemical calculations give stationary solutions for the Hamiltonian, whereas mobility of the charge wave (and accordingly the geometric kink) is usually postulated in soliton theory [2]. In view of the fact that the velocity of the soliton, as a geometric defect, is determined by vibrational relaxation of the atoms and is equal to 1 km/s, in the conjugated chain it can travel approximately 5-6 CC bonds in 1 ps. The movement of a self-localized charge (electron/hole) together with its surrounding molecular geometry defect is customarily regarded a the reason for the high quasimetallic conductivity in films of conjugated polymers [1, 4, 65, 66, 76, 83]. The transfer of charge from one molecule to another, the change here in the geometry, and the energy of reorganization of individual molecules and molecular nanostructures are usually also taken into account during more detailed modelling of electric conductivity processes [113-123]. We will now examine the position of the soliton level in the energy gap. The simplest Hckel p-electron approximation gives a continuous spectrum of electronic levels, and the energy gap (the distance between the highest occupied MO and the lowest unoccupied orbital De = eLUMO eHOMO) decreases regularly with increase of the conjugation chain. For unsubstituted polymethines De can be written in the analytic form: De (N + 1)1/2 sin [p/2(N + 1)], where N is the number of p centers; in the limit lim De 0. Since the Hckel MO method does not take account of electronic interaction, neutral radicals and the anions and cations of odd polymethines correspond to one and the same levels; in exactly the same way the same set of levels corresponds to neutral polyenes and their radical-ions. Calculations that properly take account of electronic interaction give a qualitatively different distribution of the p levels [98]. As an example Fig. 7 shows the series of highest occupied and lowest unoccupied levels for the vinylogous series of neutral radicals of unsubstituted polymethines H2C (CH=CH)nCH=CH2 X and the corresponding cations (VIII). It is easy to see that the levels of the neutral p system (Fig. 7a) form energy bands, i.e., a valence band and a conduction band, widely separated by the energy gap. With increase in the length of the conjugation chain the density of the levels near the bottom of the conduction band and the ceiling of the valence band increases, but the size of the forbidden band decreases 150

Fig. 8. The positions of the electronic levels in the polymethine cations (VIII) (Pm-C+) and the polyene radical-cations (XII) (Pe-Cr+) (a), in neutral polymethine radicals (X) (Pm-Cr) and polyenes (XI) (Pe-C) (b), and in the anion polymethines (IX) (Pm-C) and radical-ion polyenes (XIII) (Pe-Cr) (c); AM1 approximation.

slightly, asymptotically approaching some limiting value. Calculations in the AM1 approximation give De = 8.11 eV for n = 10; for neutral polyenes H2=(=)nCH=CH2 XI this value is somewhat lower: 6.66 eV (n = 10), but the centers of the gaps practically coincide for both series of vinylogs of the radical polymethines (X) and polyenes (XI): 4.44 eV and 4.47 eV respectively [98]. The transition from neutral p systems to the cations leads to the appearance of an unoccupied level clearly separated from the conduction band the soliton level, which in the higher vinylogs is hardly sensitive at all to the length of the chain (Fig. 7b). By comparing Fig. 7 with the scheme in Fig. 2 it is possible to see that the appearance of an impurity (soliton) level in the gap during the transition from the neutral molecule (X) to the charged polymethine (VIII) is accompanied by a significant shift of the conduction band and the valence band, as a result of which the density of the levels in the valence band increases, whereas close to the bottom of the conduction band it decreases. On the other hand in the transition to the anions the calculations predict an upward shift of the levels with almost mirror symmetry compared with the cations, as seen from Fig. 8. It is interesting to note that in the cations the soliton level is 1.03 eV below the center of the gap in the neutral p systems (the AM1 approximation, n = 10), and in the anions it is shifted by exactly the same amount [98]. A similar shift of the energy levels occurs with the appearance of the soliton level in the even polyenes during the transition from the neutral molecules to their ions (XII) and (XIII) (Fig. 8). H2C (CH=CH)n +CH2, XII H2C (CH=CH)n CH2. XIII

151

From Figs. 7 and 8 it is also seen that the gap itself is significantly reduced on account of the appearance of the soliton level. The decrease in the distance between the boundary levels, one of which is the soliton level, is also confirmed qualitatively by the spectral data; in the ions of the conjugated systems the energy of the first electronic transition is significantly lower than in the corresponding neutral molecules, and they are as a rule characterized by a deep color, primarily in cationic and anionic polymethine dyes, in contrast to the polyenes and their a,w-disubstituted derivatives (see, for example, [12-14, 24-27, 53]). Quantum-chemical investigations also showed that the introduction of the terminal groups has a significant effect on the position of the soliton level and the energy bands, and their effect is substantially stronger in the charged systems. In the cations the soliton level and the top of the valence band are displaced toward the levels of the donor substituents, while in the anions the corresponding displacement of the energy levels makes them more sensitive to the close unoccupied levels of the acceptor terminal groups [98].

DESTRUCTION OF THE SYMMETRY OF THE SOLITON WAVE Quantum-chemical investigations showed that the values of the charges at the carbon atoms in the conjugated chain of ionic polymethine dyes and the radical-ions of polyenes are sensitive to the topology of the terminal groups and primarily to their donoracceptor characteristics [12, 14, 47, 48]. Thus, it was shown that increase in the electron-donating character of the terminal groups (the topological index F0) leads regularly to an increase in the electron density at the a position of the polymethine chain both in symmetrical and in unsymmetrical dyes [12, 124, 125]. This agrees with the experimental data, e.g., with the dependence of the protonation constants of the dyes on the basicity of the terminal groups [126]. It was also found that the orders of the p bonds pmn are sensitive to the nature of the terminal groups so that for dyes with highly basic heterocyclic groups (pyridinium, quinolinium) the order of the first CC bond in the chain is reduced, whereas the introduction of low-basicity groups (thia- or selenopyrylium) leads to an increase in the order of this bond [12, 127]. On the basis of the fact that the order of the p bond correlates with the length of this bond, according to the known relation lmn () = 1.54 0.14pmn [85], it was shown that the alternation of the CC bond lengths depends on the parameter F0 both in the ionic conjugated systems and in the neutral substituted polyenes [12, 47, 48, 128, 129]. As a rule the introduction of electron-donating and electron-accepting groups leads to decrease in the amplitude of alternation. As example Fig. 9 shows graphs for the dependence of Dln on the length of the chain in unsubstituted polymethine cation (VIII) and in streptocyanines (H3C)2N+=CH(CH=CH)nN(CH3)2 XIV with dimethylamino groups as terminal groups. As seen, with identical chain length (with n = 2-6) the amplitude Dln in the substituted cyanine dyes (XIV) is significantly smaller than in the corresponding unsubstituted polymethine cations (VIII), i.e., the introduction of substituents leads to equalization of the CC bond lengths, whereas in the unsubstituted compound (VIII) the degree of alternation increases regularly, asymptotically approaching a limiting value. As seen from comparison of Figs. 9, a and b, this value depends on the method of calculation, and the limiting values themselves are close to the values of Dln for the neutral polyenes [33, 48]. True, it should be noted that the experimental value for the alternation of the bond lengths in the polyenes is somewhat smaller (0.08 ) than the calculated value: 0.118 (ab initio, STO 6-31G**) and 0.091 (AM1) [33]. The main result predicted by the quantum-chemical calculations presented in Fig. 9 is that the symmetry of the molecular geometry is destroyed at a certain critical number of vinylene groups ncr, and the center of the topological soliton is gradually shifted from the center of the conjugated chain to one of it ends. The localizations of the soliton obtained in the ab initio and AM1 approximations do not always coincide, and this is probably due to the different values of the amplitude of alternation for molecules with one and the same length.

152

Fig. 9. The alternation of the bond lengths Dln in polymethine cations (VIII) (a, b) and streptocyanines (XIV) (c, d) calculated in the AM1 and ab initio (6-31G**) approximations; n is the number of the bond in the chain.

Fig. 10. The alternation of the charges Dqm in polymethine anions (IX) (a, b) and their thia analogs (XV) (c, d); AM1 and ab initio (6-31G**); m is the number of the atom in the chain. The center of the charge wave is shifted similarly. As example Fig. 10 shows graphs for the function Dqm for the highest vinylogs of the series of unsubstituted polymethines (IX) and their thia analogs S(CH=CH)nCH=S. XV As seen, if n = ncr = 7 the center of the charge soliton in the anionic dyes (XV) is shifted to one of the terminal groups, whereas in the unsubstituted anions (IX) the symmetrical distribution of the charge is retained irrespective of the length of the conjugation chain. The conclusion that destruction of the symmetry does not arise in unsubstituted conjugated systems is likewise confirmed by Figs. 5 and 6, where analogous graphs for comparatively long cation polymethines (VIII) are presented.

153

Fig. 11. The charges at the carbon atoms in the chain: a) polymethine cation (VIII), n = 20; b) the aza analog of the dye (XVI) (the terminal carbon atoms are substituted by nitrogen atoms with the same molecular geometry); c) the unsymmetrical form of the cation (XVI) after optimization; m is the number of the atom in the chain.

As shown by the calculations, the symmetrical distribution of charge and the symmetrical optimized geometry are not destroyed in the vinylogous series of polymethine ions containing phenyl groups as terminal groups [48, 127]. The authors [67-70, 73] name the pseudo-Jahn Teller effect, i.e., stabilization of the molecule on account of mixing of the ground and excited states, as the reason for destruction of the symmetry of the charge and geometric solitons; this is a special case of Peierls theorem about geometric violations in the radical-ions of linear hydrocarbons. The destruction of the symmetry of the soliton waves can be interpreted more simply and more explicitly in the following way. The introduction of donor or acceptor substituents additionally induces alternation of electron density at the atoms at the ends of the chain, and the form of the charge wave is therefore distorted compared with the ions of unsubstituted conjugated systems. Figure 11 shows the distribution of charges at the atoms of the unsubstituted polymethine cation (IX) with 20 vinylene groups (Fig. 11a) and the variation of the charges with replacement of the terminal carbon atoms by nitrogen atoms [the so-called Brooker ions (XVI)] with the same molecular geometry (Fig. 11b). H2N+=CH(CH=CH)nNH2 XVI

154

It is clearly seen that such variation of the chemical structure hardly changes the charges at the center of the chain at all (with accuracy up to 0.001), but at the ends of the chain close to the terminal groups the amplitude of alternation of the charges is significantly increased. A further geometry optimization procedure leads to displacement of the center of the group soliton wave to the end of the chain (Fig. 11c). Consequently, in order to retain its dimensions (as a solitary charge wave of finite dimensions, according to the definition of a soliton), the mobile soliton moves to the end of the chain, where the parameter Dqm even exceeds that at the center of the initial symmetrical molecule (Fig. 11a, b). It is seen from Fig. 9b that with destruction of the symmetry only half the charge wave (although nevertheless somewhat distorted on account of the effect of the terminal group) is actually observed in the molecule, while the second half must be projected mentally onto an imaginary extension of the molecule. It must also be emphasized that simultaneously with the induction of the charges the introduction of the terminal groups will lead, as mentioned above, to equalization of the CC bond lengths at the ends of the chain, and the degree of variation correlates unambiguously with the electron-donating power F0. This also favors movement of the mobile geometric soliton (as a section of equalized bonds) together with the charge soliton to the end of a conjugated chain with already equalized CC bonds, i.e., such movement must be energetically favorable, and the optimization procedure will therefore lead to unsymmetrical localization of the solitons (Fig. 11c). In [12, 127, 129] it was shown that the terminal groups with minimum values for the topological index F0 have practically no effect on the order and, consequently, on the length of the CC bonds; conversely, the maximum equalization of the bonds is realized in the so-called ideal polymethine state, where F0 = 45. For phenyl groups, like any alternant rings, F0 = 0, and the introduction of such terminal groups does not therefore lead to destruction of the symmetry even with a conjugated system of comparatively large length [130]. The calculations showed that with the introduction of more complex heterocyclic groups the destruction of the symmetry can occur at a significantly smaller chain length than in the streptocyanines (XIV) or anionic dyes (XVI). Some examples of series of polymethine dyes for which the ncr values were calculated in the ab initio approximation (the STO 3-21G** basis set), are given below:
(C2H5)2N N(C2H5)2

H3C N+
n

CH3 N

O+
n

ncr = 5 XVII

(C2H5)2N

ncr = 2 XVIII

N(C2H5)2

H3C
+

CH3

N CH3

N CH3

ncr = 4 XIX

ncr = 4 XX

On the basis of an analysis of the results from quantum-chemical calculations and spectral data Przhonska and the author of the present paper put forward a hypothesis about the simultaneous existence of two forms of ionic polymethine dyes with a long polymethine chain, i.e., with symmetrical and unsymmetrical localization of the soliton waves [51]. This made it possible to interpret correctly the marked change in the form of the spectral bands of the dyes absorbing in the near IR region: Thiacyanines (IV) (Y = S, ncr = 5) (see Fig. 1) [51, 52], pyridopentacarbocyanines (XVII) (ncr = 5) [73], thiapyrylotetracarbocyanines [52], pyroninedicarbocyanines (XX) with the smallest value ncr = 2 of all known polymethine dyes [128]. The long-wave band corresponds to the symmetrical form of the dye, while the appearance of the broad short-wave band is due to the existence of the unsymmetrical form of the dye with a displaced charge wave and with a significant dipole

155

Fig. 12. The bond lengths in polymethines, n = 7 (a): in the anion (XV) (Pm-S-7), in the neutral radical (XXI) (Pm-S-7R); in polyenes, n = 8 (b): in the neutral molecule (XXII) (Pe-S-8), in the radical-anion (XXII) (Pe-S-8R); the ab initio approximation (6-31G**).

moment, as a result of which the band is very sensitive to the polarity of the solvent. Calculations predict that in the dyes that are derivatives of benz[c,d]indole (XVI) or azaazulene (XVII) destruction of the symmetry and the above-mentioned spectral effects must be expected for the higher vinylogs: ncr = 4. Quantum-chemical calculations give analogous destruction of the symmetry of soliton waves for the radical-ion polyenes that are derivatives of compounds (XII) and (XIII) containing donor or acceptor terminal groups [47, 48]. It should be noted particularly that the destruction of the symmetry does not arise in neutral p-electronic systems by definition on account of the absence of a soliton charge wave. As an illustration Fig. 12 shows the bond lengths of anionic sulfur-containing dyes (XV) and their neutral radicals S (CH=CH)7CH=S, XXI and also the corresponding neutral polyenes S=CH(CH=CH)8CH=S XXII and their radical-anions
& S =CH(CH=CH)8CH=S. XXIII

It is seen that the calculations only predict destruction of the symmetry of the molecular geometry for the ions (XV) and (XXIII). Destruction of the symmetry was also established during calculations on the ions of polyphenylvinylenes (II) (R1 = R2 = H) [132]. However, as shown by more detailed analysis, it is different in nature and is not due to the effect of the terminal 156

groups [133]. On the other hand the calculations showed that the two soliton waves in the dications of a,w-diaminopolyenes are arranged symmetrically, i.e., destruction of the symmetry does not arise [134].

SOLITON WAVES IN THE EXCITED STATE As known, the excitation of the ions of conjugated systems by a quantum of light is accompanied by a substantial redistribution of electron density at the atoms and by a change in the population of p bonds (see, for example, the reviews and monographs [12-14, 56, 135] and the references therein). Thus, a distinguishing feature of polymethine dyes is the transfer of electron density during the transition to the first excited state from the p centers at the odd positions in the chain to the atoms at the even positions, whereas in the neutral polyenes there are no changes in the charges during excitation, and only the orders of the p bonds change [12, 135-138]. Change in the charge distribution affects the intensity of the electronic transitions, while change in the bond orders and, accordingly, their lengths during subsequent vibrational relaxation in the excited state determines the width of the absorption bands [139-141], the magnitude of the Stokes shifts in the fluorescence spectra [142-145], and the paths for nonradiative degradation of the excited state [142, 146]. A number of authors have postulated that the main reason for the decrease in the quantum yields of fluorescence is transcis conformational transformations both in the polymethine dyes [147-155] and in polyenes, stilbenes, and their substituted derivatives and hetero analogs [156, 157]. By pulse spectroscopy (time-resolved spectroscopy) it was shown that two components with life times differing by 1-2 orders of magnitude coexist in the excited state of polymethine dyes [58, 60, 158-160]. To explain this fact it was suggested that two vibrational relaxation paths are possible: With retention of the symmetry of the molecular geometry and with destruction of the symmetrical structure. This was confirmed by appropriate quantum-chemical modelling [58, 158]. The possible destruction of the symmetry and the simultaneous existence of two components, symmetrical and unsymmetrical, are also indicated by the splitting of the bands in the reabsorption spectra, corresponding to electronic transitions from the excited state (transient absorption spectra) [161-164], the destruction of the perpendicularity of polarization of the highest electronic transitions during measurement of the anisotropy in the fluorescence excitation spectra [165], or the generation of a second harmonic by solutions of symmetrical thiacyanines [158]. Information on the electronic structure of conjugated molecules in the excited state and its dependence on the molecular topology (chain length, evenness of the number of p centers, the nature of the terminal groups) is necessary during investigation of the generation characteristics of conjugated compounds and, primarily, of ionic polymethine dyes [140, 141, 166]. Only the Frank Condon excited state, corresponding to an electronic transition without change in the equilibrium molecular geometry and appearing in the absorption spectra, will be examined in the present paper; examination of vibrational relaxation processes and conformational transformations is a subject of a separate investigation. Since two states, symmetrical and unsymmetrical, are possible for the ions of substituted conjugated systems with a long chain in the ground state, we will examine them separately. All the configurations corresponding to electronic transitions from the four highest occupied MOs to the four unoccupied MOs with the lowest energy were used in the calculations of the excited states. Figure 13 shows the charges qm and the alternation of the charges Dqm, calculated according to Eq. (7), for the ground and excited states respectively of the unsubstituted polymethine cation (VIII) and its diamino-substituted analog (XVI) with an identical number of vinylene groups n = 7, where the symmetry has not yet been destroyed. It is seen that the amplitude of alternation of the charges at the center of the chain changes sign during the transition to the excited state, so that the polarization of the CC bonds (the ionicity of the bonds) becomes opposite compared with the ground state. At the ends of the chain in unsubstituted polymethines, however, the directions of the polarization vectors coincide (Fig. 13c); in this section the Dqm values also coincide for both states. A more detailed analysis of the graphs for the Dqm function for the excited state for the ions of polymethine dyes containing both the simplest groups and complex heterocyclic pyridinium, benzothiazolium, and thiapyrylium groups showed that there is a segment of chain where the order of the atoms with an excess and a deficiency of electron density is reversed compared with the electron density distribution in the ground state; it was proposed to treat the charge wave on this segment as a soliton in the excited state [51, 52, 166]. At the ends of this segment the amplitude of alternation of the charges (the Dqm parameter) in the excited state reaches its own minima, as seen from Fig. 14, which gives 157

Fig. 13. The charges q (a, b) and the alternation of the charges Dq (c, d) in the ground (solid line) and excited (dashed line) states in the polymethine cation (VIII) (n = 7) (a, c) and the diamino-substituted analog (XVI) (b, d); AM1 approximation.

corresponding graphs for the vinylogous series of unsubstituted cation polymethines (VIII) and their diamino-substituted analogs (XVI). The distance between the minima does not depend on the length of the conjugated system and can serve as the length of the soliton wave. It is seen from Fig. 14 that the calculated dimensions of the excited soliton (determined as proposed above) are significantly less than the dimensions of the charge wave in the ground state: 15-17 carbon atoms (cf. Figs. 5 and 6). In contrast to the unsubstituted polymethine (VIII), as seen from Fig. 14d, the increase in the length of the chain in the diamino analog (XVI) gives rise to some decrease in Dqm* close to both minima, as a result of which the boundaries of the soliton become blurred. The calculations predict similar behavior for polymethine dyes, pyridinium derivatives, pyrylium derivatives, etc. The dimensions of the charge wave in the excited state are increased somewhat with increase in the basic characteristics of the hetero groups (the F0 index) [167]. The significant change in the form of the soliton in the excited state and the appearance of two alternation minima (which must correspond to the maxima in the alternation of the p-bond order) can lead to instability in the symmetrical form. It is logical to suppose that, like the ground state, in the excited state two forms exist after relaxation, i.e., with symmetrical and unsymmetrical localization of the charge and topological solitons. Quantum-chemical calculations of the ions of the conjugated molecules in the excited state predict destruction of the symmetry of the optimized geometry, and with a significantly shorter length for the polymethine chain [130]. Also, in contrast to the ground state, destruction of the symmetry even arises in a,w-diphenylpolymethines (F0) [130]. Movement of the topological soliton toward one of the terminal groups in the relaxed unsymmetrical form must naturally lead to elongation of the CC bond near the other terminal group and, accordingly, to a decrease of the barrier to conformational transformations along it. Probably, it is the unsymmetrical form that is the short-lived component in the excited state, whereas the barriers to rotation about the equalized bonds in the symmetrical form are significantly higher, which corresponds to an increase in the life time of such a component. During measurement of the fluorescence spectra of thiacyanine dyes by pulse spectroscopy at low temperatures, when the conformational processes are impossible, it was possible to detect both components with different life times [168]. It is also seen from Fig. 14b that in the excited state of substituted polymethines (XVI) increase in the length of the chain will lead to decrease in the amplitude of alternation of charges at the points corresponding to the minimum of Dqm* which must naturally be accompanied by a significant increase in the alternation of the CC bond orders and, after vibrational relaxation, their lengths. As a result the barriers of the conformational transformations can be substantially reduced, which may be the reason for the marked decrease in the quantum yields of fluorescence in the series of cyanine dyes (IV) during the transition to the higher vinylogs (Table 1), although it is of course necessary to take proper account of other possible factors [169, 170].

158

Fig. 14. The alternation of the charges Dqm in the ground (S0) and excited (S1) states of the polymethine cations (VIII) (a, b) and diamino-substituted dyes (XVI) (c, d); AM1 approximation.

Fig. 15. The alternation of charge Dqm in the ground (S0) and excited (S1) states of the unsymmetrical form of cationic cyanines (XVI) (a, b) and anionic malocyanines (XXIV) (c, d); AM1 approximation.

Now we will examine the form of the charge waves in the excited states for the unsymmetrical form of the ions of conjugated systems with a long chain. Figure 15 show the graphs for the Dqm functions in the ground and excited states of the Brooker ions (XVI) and malocyanines (NC)2C(CH=CH)nCH=C(CN)2 XXIV for n > ncr.

159

Fig. 16. Diagrams of the redistribution of electron density at the atoms during excitation in the symmetrical (a) and unsymmetrical (b) forms of thiapentacarbocyanine (IV), Y = S, n = 5; the dark circles indicate increase of density, light circles indicate decrease.

As can be seen, the form of the charge waves in the excited state of both series of dyes is similar to the form of the solitons in the symmetrical form of the polymethine cations (VIII) (cf. Fig. 14), although somewhat distorted, except that the center of the soliton does not coincide with the center of the chain but is displaced. By comparison with the form of the corresponding charge waves in the ground state it is possible to conclude that excitation of the unsymmetrical form will lead to substantial symmetrization of the electron density distribution in the conjugation chain. Increase in the length of the molecule leads to gradual movement of the excited soliton from the center of the chain, and its form becomes irregular, in contrast to the comparatively smooth curves in the graphs of Dqm for the symmetrical form (cf. Figs. 14 and 15). It can be supposed that vibrational relaxation will lead to smoothing of the form of the charge waves compared with their form in the Frank Condon excited state. Similar symmetrization of charge waves was obtained for dyes with complex groups [51, 52]. Movement of the center of the charge wave in the unsymmetrical form leads to substantial change in the nature of the electronic transition in relation to the symmetrical form. Diagrams of the redistribution of electron density at the atoms during 0 excitation dqm = q* q proved useful for analysis of the character of the electronic transitions [12, 171-172]. Such diagrams for the two forms of thiapentacarbocyanine (IV) (Y = S, n = 5) are shown in Fig. 16; it is seen that excitation in the symmetrical form gives rise to transfer of electron density to the neighboring atoms (a typical high-intensity polymethine transition) (Fig. 16a), while destruction of the symmetry (Fig. 16b) leads to the result that the transition becomes a typical charge-transfer transition [173], accompanied by a significant change in the stationary dipole moment of the molecule, as a result of which it is extremely sensitive to the polarity of the solvent.

CONCLUSION Thus, numerous theoretical and experimental investigations have shown that the ions of linear conjugated molecules containing an extended collective system of p electrons have a series of unique properties. In the energy gap there are soliton levels that significantly reduce the distance between the frontier MOs and, consequently, the energy of the first electronic transition. This gives a deep and highly intense color, while the charges at the atoms (p centers) are not delocalized uniformly along the conjugation chain but form waves of the soliton type with finite dimensions, which become mobile with increase in the length of the conjugation chain and can move easily along the molecule. Self-localization of the charge in the form of the soliton is accompanied by change in the equilibrium molecular geometry. A wave a topological (geometric) soliton is also formed as a result of the change in the lengths of the carboncarbon bonds. In the excited state the dimensions of the soliton decrease in comparison with the ground state. The introduction of donor or acceptor terminal groups leads to distortion in the form of the soliton waves and increase in their mobility; with a chain length exceeding the dimensions of the soliton it can move to one end of the conjugated chain, and the symmetrical structure of the molecule is destroyed, as observed experimentally in the electronic spectra. 160

The use of modern physical concepts on the electronic structure of molecules and the soliton concept, in particular, proved very fruitful during investigation of the specific characteristics of conjugated systems both in polymers or oligomers and in the comparatively short molecules of polymethine dyes and a,w-disubstituted polyenes and has opened up new paths for their widespread application.

REFERENCES 1. 2. 3. 4. 5. 6. 7. 8. 9. 10. 11. 12. 13. 14. 15. 16. 17. 18. 19. 20. 21. 22. 23. 24. 25. 26. 27. 28. 29. 30. 31. 32. 33. 34. 35. 36. 37. 38. 39. J. L. Bredas, D. Belionne, J. Cornil, et al., Synth. Met., 125, 107-116 (2002). A. S. Davydov, Solitons in Molecular Systems [in Russian], Naukova Dumka, Kiev (1984). J. L. Bredas and G. B. Street, Accounts Chem. Revs., 18, 309-315 (1985). A. O. Patil, A. J. Heeger, and F. Wudl, Chem. Rev., 88, 183-200 (1988). L. M. Tolbert, Accounts Chem. Res., 25, 561-568 (1992). J. R. Reimers, J. S. Craw, G. B. Bacskay, et al., Biosystems, 35, 107-111 (1995). J. G. Gaudiello, G. E. Kellog, S. M. Tetrick, et al., J. Am. Chem. Soc., 111, 5259-5271 (1989). V. A. Kuprievich, Teor. ksp. Khim., 22, No. 4, 385-393 (1986). A. Yu. Kon and I. A. Misurkin, Teor. ksp. Khim., 22, No. 2, 146-153 (1986). A. Yu. Kon and I. A. Misurkin, Teor. ksp. Khim., 25, No. 4, 393-398 (1989). V. D. Pokhodenko and N. F. Guba, Teor. ksp. Khim., 30, No. 5, 241-263 (1994). A. D. Kachkovskii, Usp. Khim., 66, 715-734 (1997). A. A. Ishchenko, Usp. Khim., 60, 865-880 (1991). N. Tyutyulkov, J. Fabian, A. Mehlhorn, et al., Polymethine Dyes. Structure and Properties, St. Kliment Ohridski Univ. Press, Sofia (1991). F. Meyers, S. R. Marder, and J. W. Perry, Chemistry of Advanced Materials, L. V. Interrante, M. J. Hampden-Smith (eds.), Wiley-VCH Inc., New York, etc. (1998), pp. 207-268. P. Cronstrand, Yi. Luo, and H. gren, Chem. Phys. Lett., 352, 262-269 (2002). W. Rettig and M. Dekhtyar, Chem. Phys., 293, 75-90 (2003). D. Beljonne, Z. Shuai, J. L. Bredas, et al., Chem. Phys. Lett., 279, 1-8 (1997). M. Albota, D. Beljonne, J. L. Bredas, et al., Science, 281, 653-656 (1998). V. M. Geskin, M. Yu. Balakina, J. Li, et al., Synth. Met., 116, 263-267 (2001). A. I. Kiprianov, Color and Structure of Cyanine Dyes [in Russian], Naukova Dumka, Kiev (1979). L. G. S. Brooker, Rev. Mod. Phys., 14, 275-293 (1942). S. Huenig and H. Berneth, Top. Curr. Chem., 92, 1-44 (1980). S. Huenig, D. Schewtzow, H. Schlaf, et al., Ann. Chem., 1436-1449 (1974). S. Huenig and F. Linhard, Ann. Chem., 2116-2189 (1975). S. Huenig and F. Linhard, Ann. Chem., 317-335 (1976). S. Huenig and H. C. Steinmetzer, Ann. Chem., 1060-1089 (1976). A. Zedler and S. Kulpe, J. Prakt. Chem., 317, 199-213 (1975). Yu. L. Slominskii, S. V. Popov, I. V. Repyakh, et al., Teor. ksp. Khim., 23, No. 6, 687-692 (1987). R. Radeglia, E. Gey, K.-D. Nolte, et al., J. Prakt. Chem., 315, 586-599 (1973). S. Kulpe, A. Zedler, S. Daehne, et al., J. Prakt. Chem., 315, 865-872 (1973). S. Barlow, L. M. Henling, M. W. Day, et al., Chem. Commun., 1567-1568 (1999). G. Orlandi, F. Zerbetto, and M. Z. Zgierski, Chem. Rev., 91, 867-891 (1991). S. Daehne and R. Radeglia, Wiss. Z. Techn. Univ. (Dresden), 29, 101-107 (1980). S. Daehne, Z. Chem., 168-183 (1970). S. Daehne and R. Radeglia, Tetrahedron, 27, 3673-3693 (1971). S. Daehne and K.-D. Nolte, J. Chem. Soc., Chem. Communs., 1056-1067 (1972). S. Daehne, Science, 199, 1163-1167 (1978). G. G. Dyadyusha and A. D. Kachkovskii, Ukr. Khim. Zh., 41, 1176-1181 (1975). 161

40. 41. 42. 43. 44. 45. 46. 47. 48. 49. 50. 51. 52. 53. 54. 55. 56. 57. 58. 59. 60. 61. 62. 63. 64. 65. 66. 67. 68. 69. 70. 71. 72. 73. 74. 75. 76. 77. 78. 79. 80. 81. 82. 83. 84. 85. 86. 162

G. G. Dyadyusha and A. D. Kachkovskii, Teor. ksp. Khim., 15, No. 2, 152-161 (1979). G. G. Dyadyusha and A. D. Kachkovskii, Usp. Fotogr. Nauk, 22, 55-66 (1984). G. G. Dyadyusha and M. N. Ushomirskii, Teor. ksp. Khim., 21, No. 3, 268-279 (1985). G. G. Dyadyusha and M. N. Ushomirskii, Teor. ksp. Khim., 22, No. 2, 137-145 (1986). G. G. Dyadyusha, A. D. Kachkovskii, and M. L. Dekhtyar, J. Mol. Struct., 217, 195-205 (1990). A. D. Kachkovskii, G. G. Dyadyusha, and M. L. Dekhtyar, Dyes Pigments, 15, 191-202 (1991). M. L. Dekhtyar, Dyes Pigments, 28, 261-274 (1995). A. D. Kachkovskii, D. A. Yushchenko, and D. A. Kachkovskii, Teor. ksp. Khim., 38, No. 6, 341-346 (2002). A. D. Kachkovskii, D. A. Yushchenko, G. A. Kachkovskii, et al., Dyes Pigments, 66, 211-221 (2005). H. Oeling and F. Baer, Org. Magn. Reson., 8, 623-627 (1976). W. Koenig, J. Prakt. Chem., 112, 1-36 (1925). R. S. Lepkowich, O. V. Przhonska, J. M. Hales, et al., Chem. Phys., 305, 259-270 (2004). A. D. Kachkovskii, A. I. Tolmachev, Yu. L. Slominskii, et al., Dyes Pigments, 64, 207-216 (2005). M. Matsuoka (ed.), Near-Infrared Absorbing Dyes, Plenum Press, New York (1990). J. Fabian, H. Nakazumi, and M. Matsuoka, Chem. Rev., 92, 1197-1230 (1992). A. I. Tolmachev, Yu. L. Slominskii, and A. A. Ischchenko, Near-Infrared Dyes for High Technology Applications, S. Daehne, U. Resh-Genge, U. Wolfbeis (eds.), Kluwer Acad. Publ., Berlin (1998), pp. 384-515. A. A. Ishchenko, Structure and Luminescence Spectral Characteristics of Polymethine Dyes [in Russian], Naukova Dumka, Kiev (1994). V. A. Mostovnikov, A. N. Rubinov, and M. A. Alperovich, Zh. Prikl. Spektrosk., 20, 42-47 (1974). L. Martini and G. V. Hartland, Chem. Phys. Lett., 258, 180-186 (1996). T. Zhang, C. Chen, Qi Gong, et al., Chem. Phys. Lett., 298, 236-240 (1998). M. Garavelli, P. Celani, F. Bernardi, et al., J. Am. Chem. Soc., 119, 11487-11494 (1997). J. Quenneville, M. Ben-Nun, and T. J. Martinez, J. Photochem. Photobiol. A, 144, 229-235 (2001). A. Yartsev, J.-L. Alvarez, U. Aberg, et al., Chem. Phys. Lett., 243, 281-289 (1995). W. Wu, D. Danovich, A. Shurki, et al., J. Phys. Chem. A, 104, 8744-8758 (2000). P. Batail, Chem. Rev., 104, 4887-4890 (2004). W. P. Su, J. R. Schrieffer, and A. J. Heeger, Phys. Rev. Lett., 42, 1698-1701 (1979). W. P. Su, J. R. Schrieffer, and A. J. Heeger, Phys. Rev., 22, 2099-2111 (1980). J. S. Craw, J. R. Reimers, G. B. Bacskay, et al., Chem. Phys., 167, 77-99 (1992). J. S. Craw, J. R. Reimers, G. B. Bacskay, et al., Chem. Phys., 167, 101-109 (1992). J. R. Reimers and N. S. Hush, Chem. Phys., 176, 407-420 (1993). J. R. Reimers, J. S. Craw, G. B. Bacskay, et al., Mol. Cryst. Liquid Cryst. A, 234, 51-55 (1993). L. M. Tolbert and M. E. Ogle, J. Am. Chem. Soc., 112, 9519-9527 (1990). L. M. Tolbert and M. E. Ogle, Synth. Met., 51, 391-393 (1992). L. M. Tolbert and X. Zhao, J. Am. Chem. Soc., 119, 3253-3258 (1997). J. V. Caspar, V. Ramamurthy, and D. R. Corbin, J. Am. Chem. Soc., 113, 600-610 (1991). T. Bally, K. Rith, W. Tang, et al., J. Am. Chem. Soc., 114, 2440-2446 (1992). T. Giamarchi, Chem. Rev., 104, 5037-5055 (2004). N. F. Mott, Metal-Insulator Transitions, Taylor and Francis, London (1990). J. R. Reimers and N. S. Hush, Chem. Phys., 134, 323-354 (1989). J. R. Reimers and N. S. Hush, Chem. Phys., 146, 89-103 (1990). D. Comoreto, D. Moses, H. Okumoto, et al., Synth. Met., 84, 539-544 (1997). D. Moses, A. Dogarin, and A. J. Heeger, Synth. Met., 116, 19-22 (2001). H. Seo, C. Hotta, and K. Fukuyama, Chem. Rev., 104, 5005-5036 (2004). G. K. Hutchison, M. A. Ratner, and T. J. Marks, J. Am. Chem. Soc., 127, 2339-2350 (2005). I. A. Misurkin and A. A. Ovchinnikov, Usp. Khim., 46, 1835-1870 (1977). A. Streitwieser, Molecular Orbital Theory for Organic Chemists, Wiley, New York (1961). A. V. Luzanov, Teor. ksp. Khim., 27, No. 4, 413-426 (1991).

87. 88. 89. 90. 91. 92. 93. 94. 95. 96. 97. 98. 99. 100. 101. 102. 103. 104. 105. 106. 107. 108. 109. 110. 111. 112. 113. 114. 115. 116. 117. 118. 119. 120. 121. 122. 123. 124. 125. 126. 127. 128. 129. 130. 131. 132. 133.

R. Carbo and J. M. Riera, A General SCF Theory, Lecture Notes in Chemistry, Springer, Berlin (1998). M. J. S. Dewar, Molecular Orbital Theory for Organic Chemists, McGraw-Hill, New York (1969). H. Kuhn, J. Chem. Phys., 16, 840-841 (1948). H. Kuhn, J. Chem. Phys., 17, 1098-1212 (1949). H. Kuhn, W. Hyber, G. Handschig, et al., J. Chem. Phys., 32, 467-470 (1960). C. Kuhn and H. Kuhn, Synth. Met., 68, 173-181 (1995). N. S. Bayliss, J. Chem. Phys., 16, 287-292 (1948). S. Basu, J. Chem. Phys., 22, 1270-1271 (1954). W. T. Simpson, J. Am. Chem. Soc., 73, 5359-5363 (1951). K. Ruedenberg, J. Chem. Phys., 21, 1565-1581 (1953). S. Olszewski, J. Chem. Phys., 44, 3297-3306 (1966). O. D. Kachkovskii, D. A. Yushchenko, G. O. Kachkovskii, et al., Dyes Pigments, 66, 223-229 (2005). J. Fabian and H. Hartmann, J. Signal AM, 2, 457-460 (1974). N. Tyutyulkov, J. Petkov, O. Polansky, et al., Theor. Chim. Acta, 38, 1-8 (1975). J. Fabian, H. Hartmann, and N. Tyutyulkov, J. Signal AM, 4, 101-114 (1976). H. Hartmann, J. Signal AM, 7, 101-118 (1976). C. A. Coulson and G. S. Rashbrooke, Proc. Cambridge Phil. Soc., 36, 193-200 (1940). H. C. Longuet-Higgins, J. Chem. Phys., 18, 265-287 (1950). C. A. Coulson and H. C. Longuet-Higgins, Proc. Roy. Soc. London, 191, 39-47 (1947). N. Trinaismich, Semiempirical Methods for Calculation of Electronic Structure [Russian translation], J. Sigal (ed.), Vol. 1, Mir, Moscow (1980), pp. 13-46. W. Grahn, Tetrahedron, 32, 1931-1939 (1976). P. M. Henrichs and S. Gross, J. Am. Chem. Soc., 98, 7169-7175 (1976). M. Wahnert, S. Daehne, and R. Radeglia, Adv. Mol. Relax Interact. Process, 11, 263-282 (1977). J. Schaefer and E. O. Steiskal, J. Am. Chem. Soc., 98, 1031-1032 (1976). M. M. Maricq, J. S. Waugh, A. G. MacDiarmid, et al., J. Am. Chem. Soc., 100, 7729-7730 (1978). K. Knoll, S. A. Krouse, and R. R. Schrock, J. Am. Chem. Soc., 110, 4424-4425 (1988). F. Garnier, A. Yassar, R. Hajlaoui, et al., J. Am. Chem. Soc., 115, 8716-8720 (1993). D. M. Adams, J. Kerito, E. J. C. Olson, et al., J. Am. Chem. Soc., 119, 10608-10619 (1997). X.-C. Li, H. Sirringhaus, F. Garnier, et al., J. Am. Chem. Soc., 120, 2206-2207 (1998). S. Tretiak, V. Chernyak, and S. Mukamel, J. Am. Chem. Soc., 119, 11408-11419 (1997). V. H. F. Bunz, Chem. Rev., 100, 1605-1644 (2000). J. A. Levitsky, J. Kim, and T. M. Swager, J. Am. Chem. Soc., 121, 1466-1472 (1999). G. Pourtois, D. Beljonne, J. Cornil, et al., J. Am. Chem. Soc., 124, 4436-4447 (2002). E. Neuteboom, S. C. J. Meskers, P. A. Hal, et al., J. Am. Chem. Soc., 125, 8625-8638 (2003). M. Heeney, C. Bailey, K. Genericius, et al., J. Am. Chem. Soc., 127, 1078-1079 (2005). M.-H. Yoon, S. A. DiBenedetto, A. Facchetti, et al., J. Am. Chem. Soc., 127, 1348-1349 (2005). D. A. Scherlis and N. Marzarl, J. Am. Chem. Soc., 127, 3207-3212 (2005). G. G. Dyadyusha, I. V. Repyakh, and A. D. Kachkovskii, Teor. ksp. Khim., 20, No. 4, 398-406 (1984). G. G. Dyadyusha, I. V. Repyakh, and A. D. Kachkovskii, Teor. ksp. Khim., 22, No. 3, 347-351 (1986). A. I. Kiprianov, S. G. Fridman, L. S. Pupko, and A. I. Kiprianov, Color and Structure of Cyanine Dyes [in Russian], Naukova Dumka, Kiev (1979), pp. 501-521. G. G. Dyadyusha, I. V. Repyakh, and A. D. Kachkovskii, Teor. ksp. Khim., 21, No. 2, 138-146 (1985). A. D. Kachkovskii and M. L. Dekhtyar, Teor. ksp. Khim., 28, 313-319 (1992). A. D. Kachkovskii and M. L. Dekhtyar, Dyes Pigments, 30, 43-54 (1996). A. D. Kachkovskii and O. A. Zhukova, Teor. ksp. Khim., 37, No. 5, 280-284 (2001). M. P. Shandura, Ye. M. Poronic, and Yu. P. Kovtun, Dyes Pigments, 66, 171-177 (2005). L. Zuppiroli, A. Bieber, D. Michoud, et al., Chem. Phys. Lett., 374, 7-12 (2003). V. M. Geskin, J. Cornil, and J.-L. Bredas, Phys. Lett., 403, 228-231 (2005). 163

134. 135. 136. 137. 138. 139. 140. 141. 142. 143. 144. 145. 146. 147. 148. 149. 150. 151. 152. 153. 154. 155. 156. 157. 158. 159. 160. 161. 162. 163. 164. 165. 166. 167. 168. 169. 170. 171. 172. 173.

Yu. N. Bernatskaya and A. D. Kachkovskii, Teor. ksp. Khim., 35, No. 3, 150-154 (1999). A. D. Kachkovskii, Structure and Color of Polymethine Dyes [in Russian], Naukova Dumka, Kiev (1989). G. G. Dyadyusha and I. V. Repyakh, Teor. ksp. Khim., 24, No. 2, 129-138 (1988). J. Fabian and A. Melhorn, J. Mol. Struct. Theochem., 109, 17-26 (1984). V. Buss, M. Schreder, and M. P. Fuelscher, Angew. Chem. Int. Ed., 40, 3189-3190 (2001). A. A. Ishchenko, N. A. Derevyanko, V. M. Zubarovskii, et al., Teor. ksp. Khim., 20, No. 4, 443-451 (1984). A. A. Ishchenko, Zh. Prikl. Spektrosk., 55, 717-724 (1991). A. A. Ishchenko, Quantum Electronics, 24, 471-492 (1994). V. G. Plotnikov, Usp. Khim., 49, 327-361 (1980). E. F. McCoy and I. J. Ross, Aust. J. Chem., 15, 573-590 (1962). V. P. Klochkov, Optika Spektroskopiya, 19, 337-344 (1965). V. I. Permogorov, L. A. Serdyukova, and M. D. Frank-Kamenetskii, Optika Spektroskopiya, 22, 979-981 (1967). B. Genri and M. Kasha, Usp, Fiz. Nauk, 108, 113-141 (1982). M. M. Awad, P. K. McCarthy, and D. Blanchard, J. Phys. Chem., 98, 1454-1458 (1994). K. Schoffel, F. Dietz, and T. Krosser, Chem. Phys. Lett., 172, 187-192 (1990). F. Momicchioli, L. Baraldi, and G. Berthier, Chem. Phys., 123, 103-112 (1988). G. Ponterini and F. Momicchioli, Chem. Phys., 151, 111-126 (1991). I. Baraldi, A. Carnevali, F. Momicchioli, et al., Spectrosc. Acta. A, 49, 471-495 (1993). E. N. Kalitievskaya, T. K. Razumova, and G. M. Rubanova, Optika Spektroskopiya, 60, 272-275 (1986). D. I. Berezin, G. P. Gurinovich, and F. Enikhen, Zh. Prikl. Spektrosk., 51, 932-938 (1989). Yu. L. Lifanov, V. A. Kuzmin, A. V. Karyakin, et al., Izv. Akad. Nauk. SSSR, Ser. Khim., 787-790 (1973). F. Dietz and S. K. Rentsch, Chem. Phys., 96, 145-151 (1985). U. Mazzucato and F. Momicchioli, Chem. Rev., 91, 1679-1719 (1991). B. E. Kohler, Chem. Rev., 93, 41-54 (1993). A. F. Marks, A. K. Noah, and M. R. V. Sahyun, J. Photochem. Photobiol. A, 139, 143-149 (2001). M. I. Demchuk, A. A. Ischchenko, Zh. A. Krasnaya, et al., Chem. Phys. Lett., 167, 774-780 (1990). O. Przhonska, Yu. Slominskii, U. Stahl, et al., J. Luminescence, 69, 105-113 (1996). P. Cronstrand, O. Christian, P. Norman, et al., Phys. Chem. Chem. Phys., 2, 5357-5363 (2000). Y. H. Meyer, M. Pittman, and P. Plaza, J. Photochem. Photobiol. A, 114, 1-21 (1998). R. Negres, O. Przhonska, D. Hagan, et al., IEEE J. on Selected Topics in Quantum Electronics, 849-863 (2001). R. S. Lepkowicz, O. Przhonska, J. M. Hales, et al., Chem. Phys., 286, 277-291 (2003). O. V. Przhonska, D. J. Hagan, E. Novikov, et al., Chem. Phys., 273, 235-248 (2001). A. Mishra, R. K. Behera, P. K. Behera, et al., Chem. Rev., 100, 1973-2011 (2000). A. D. Kachkovskii, M. Yu. Kornilov, and D. N. Shut, Teor. ksp. Khim., 40, No. 2, 83-87 (2004). Yu. P. Piryatinskii, V. G. Nazarenko, R. M. Vasyuta, et al., Teor. ksp. Khim., 39, No. 4, 225-229 (2003). A. D. Kachkovskii and O. A. Zhukova, Teor. ksp. Khim., 36, No. 3, 162-168 (2000). A. D. Kachkovski and O. O. Zhukova, Dyes Pigments, 63, 323-332 (2004). A. D. Kachkovski, Dyes Pigments, 24, 171-183 (1994). D. G. Krotko, K. V. Fedotov, A. D. Kachkovski, et al., Dyes Pigments, 64, 79-84 (2005). Z. R. Grabowski, K. Rotkiewicz, and W. Rettig, Chem. Rev., 103, 3899-4031 (2003).

164

Potrebbero piacerti anche