8
1 Aleksandr Ivanovich Ponyaev Dr. Sci. (Chem), professor Chemical Technology of Organic Dyes and Phototropic Compounds Department, e-mail: po- [email protected] A.I. Ponyaev 1 INVESTIGATION OF SPECTRAL–KINETIC PARAMETERS OF PHOTOINDUCED FORMS OF NAPHTHOTHIAZINE, ACRIDINE AND PHENANTHRIDINE SPIROPYRANS BY FLASH PHOTOLYSIS St. Petersburg State Institute of Technology (Technical Uni- versity), Moskovskii pr. 26, St. Petersburg, 190013 Russia Abstract—General relations have been revealed in the reversible photo- chemical coloration–decoloration of naphthothiazine and acridine spiropy- rans that are characterized by favorable (from the viewpoint of possible practical applications) combination of photostability, color depth, and pho- tocoloration quantum yield. Photoinduced merocyanine forms of compounds of the naphtho[1,8-de]thiazine series exhibit positive solvatochromism. The lifetime of the photoinduced form shorters in the series indoline > phenan- thridine > naphthothiazine > acridine derivatives. Photocoloration of acri- dine spiropyrans involves the singlet excited state. Naphthothiazine spiropy- rans in alcohol solution are converted into the colored form through the singlet state, and in toluene, partly through the triplet state. The absorption spectra of photoinduced forms of acridine spiropyrans are displaced toward longer wavelengths as compared to analogous indoline derivatives. Intro- duction of a fluorine atom into the 3-position of phenanthridine spiropyrans increases the rate constant for the reverse decoloration reaction by 2 to 7 orders of magnitude. Factors favoring enhanced fatigue resistance of spiro- pyrans have been determined, in particular the absence of a nitro group in molecule, singlet path of photocoloration, and quinoid structure and short lifetime of the photoinduced form. Key word: spiropyran, photochemistry, photochrome, flash photolysis, naphthothiazine, acridine, singlet, excited state, rate constant, fatigue, quinoid structure, fluorine atom, photoinduced form The Dye Department is one of the oldest departments of the Institute of Technology. Since 1968 it is called Depart- ment of Chemical Technology of Organic Dyes and Photo- tropic Compounds. Dyes have long been known as substances endowing various materials with different colors, whereas phototropic compounds, i.e., those capable of reversibly change their properties under irradiation, have come into practice relatively recently. This wonderful ability of substanc- es to change their properties in reversible manner by the action of absorbed light is widely used in light-controlled sys- tems. Among phototropic substances, the most interesting are organic compounds that undergo light-induced structural rearrangement and thus change their optical properties. Compounds capable of reversibly change their color by the action of absorbed light are called photochromic. Photo- chromic phenomena occur in organic, inorganic, complex, and biological systems, in gaseous, liquid, and solid phases, in polymeric matrices, glasses, gels, melts, and biological mem- branes. Photochromism forms the basis of natural photosyn- thesis and visual process. Photochromic transformations of dyes-photocatalysts mediate photoinduced decomposition of water to hydrogen in solar energy accumulation systems [1– 6]. Photochromism implies that molecules A on exposure to light with a definite spectral composition are converted into state B whose absorption spectrum differs from that of A. The reverse transformation is induced by irradiation with a different spectral composition, or it occurs spontaneously (dark reaction), and molecules revert to the initial state: The above scheme of photochromic process may be regarded as ideal; in a real case it is much more complex. Reversible reactions can be accompanied by irreversible photochemical and thermal transformations of both initial and photoinduced species, which may be illustrated by the following scheme: Many research teams in different countries display sus- tained interest in various aspects of photochromism. Much attention to photochromic materials is given due to wide pro-

1 INVESTIGATION OF SPECTRAL–KINETIC PARAMETERS OF ...science.spb.ru/files/BulletinIT/2013/1/Articles/19/... · polymeric matrices, glasses, gels, melts, and biological mem-branes

  • Upload
    others

  • View
    2

  • Download
    0

Embed Size (px)

Citation preview

  • 1 Aleksandr Ivanovich Ponyaev Dr. Sci. (Chem), professor Chemical Technology of Organic Dyes and Phototropic Compounds Department, e-mail: [email protected]

    A.I. Ponyaev1

    INVESTIGATION OF SPECTRAL–KINETIC PARAMETERS OF PHOTOINDUCED FORMS OF NAPHTHOTHIAZINE, ACRIDINE AND PHENANTHRIDINE SPIROPYRANS BY FLASH PHOTOLYSIS St. Petersburg State Institute of Technology (Technical Uni-versity), Moskovskii pr. 26, St. Petersburg, 190013 Russia Abstract—General relations have been revealed in the reversible photo-chemical coloration–decoloration of naphthothiazine and acridine spiropy-rans that are characterized by favorable (from the viewpoint of possible practical applications) combination of photostability, color depth, and pho-tocoloration quantum yield. Photoinduced merocyanine forms of compounds of the naphtho[1,8-de]thiazine series exhibit positive solvatochromism. The lifetime of the photoinduced form shorters in the series indoline > phenan-thridine > naphthothiazine > acridine derivatives. Photocoloration of acri-dine spiropyrans involves the singlet excited state. Naphthothiazine spiropy-rans in alcohol solution are converted into the colored form through the singlet state, and in toluene, partly through the triplet state. The absorption spectra of photoinduced forms of acridine spiropyrans are displaced toward longer wavelengths as compared to analogous indoline derivatives. Intro-duction of a fluorine atom into the 3′-position of phenanthridine spiropyrans increases the rate constant for the reverse decoloration reaction by 2 to 7 orders of magnitude. Factors favoring enhanced fatigue resistance of spiro-pyrans have been determined, in particular the absence of a nitro group in molecule, singlet path of photocoloration, and quinoid structure and short lifetime of the photoinduced form. Key word: spiropyran, photochemistry, photochrome, flash photolysis, naphthothiazine, acridine, singlet, excited state, rate constant, fatigue, quinoid structure, fluorine atom, photoinduced form

    The Dye Department is one of the oldest departments of

    the Institute of Technology. Since 1968 it is called Depart-ment of Chemical Technology of Organic Dyes and Photo-tropic Compounds. Dyes have long been known as substances endowing various materials with different colors, whereas phototropic compounds, i.e., those capable of reversibly change their properties under irradiation, have come into practice relatively recently. This wonderful ability of substanc-es to change their properties in reversible manner by the action of absorbed light is widely used in light-controlled sys-tems. Among phototropic substances, the most interesting are organic compounds that undergo light-induced structural rearrangement and thus change their optical properties. Compounds capable of reversibly change their color by the action of absorbed light are called photochromic. Photo-chromic phenomena occur in organic, inorganic, complex, and biological systems, in gaseous, liquid, and solid phases, in polymeric matrices, glasses, gels, melts, and biological mem-branes. Photochromism forms the basis of natural photosyn-thesis and visual process. Photochromic transformations of dyes-photocatalysts mediate photoinduced decomposition of water to hydrogen in solar energy accumulation systems [1–6].

    Photochromism implies that molecules A on exposure to light with a definite spectral composition are converted into state B whose absorption spectrum differs from that of A. The reverse transformation is induced by irradiation with a different spectral composition, or it occurs spontaneously (dark reaction), and molecules revert to the initial state:

    The above scheme of photochromic process may be regarded as ideal; in a real case it is much more complex. Reversible reactions can be accompanied by irreversible photochemical and thermal transformations of both initial and photoinduced species, which may be illustrated by the following scheme:

    Many research teams in different countries display sus-

    tained interest in various aspects of photochromism. Much attention to photochromic materials is given due to wide pro-

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    spects in practical usage of light-controlled system. There is a view that the XXIst century will be the century of photonic technologies [7]. First of all, these are data recording and processing systems [8], nonlinear optical materials [9], mo-lecular machines [10], photo computers [11], variable density filters [12], optical switches [11], sensors (including those for biological targets) [13], Photo hem for photodynamic therapy of cancer [14], etc. Particular application imposes specific requirements on photochromic materials. For example, eye protection requires strong absorption in the visible and UV regions [15], whereas data processing systems need materi-als absorbing in the range of irradiation of semiconductor lasers [11, 16], i.e., generally in the IR region. Compatibility of photochromic materials with polymeric matrix is necessary for the design of optical computer components [17]; light-switchable systems utilize photochromes possessing a high thermal stability [18]; rapidly relaxing photochromes are used in data processing devices [19], whereas data storage re-quires those capable of retaining photoinduced color for many years [20]. Photochromic sensors should be selectively sensi-tive to particular parameters, such as pH value, the presence of metals or charged species, etc. [21]. Solar energy storage systems based on strain energy of metastable but kinetically stable structures require materials that maximally utilize the solar radiation spectrum, while the photoinduced form should be colorless [3]. By contrast, the initial state of photochromic compounds for eyeglasses and antidazzle glasses must be colorless, and they should darken under oncoming traffic light.

    Photochromism is a phenomenon combining a number of fundamental problems of excited state physics and chemistry and “dark” organic chemistry into a whole entity. Population of excited states determines the physical photochromism of organic compounds, whereas chemical bond breakage, isom-erization, redox processes, tautomerism, dimerization, and other processes underlie the chemical photochromism.

    The efficiency of new technologies and their further de-velopment are closely related to understanding of fundamen-tal mechanisms of processes occurring in photochromic com-pounds. Of radical importance is to elucidate how structural modification of a molecule affects its photochromic proper-ties; in particular, the relation between the structure of initial compound and spectral and kinetic parameters of its photoin-duced form in the series of complex heteroaromatic com-pounds is a quite significant problem. Incident light often only initiates a process, while further transformations include com-plex chemical rearrangements resulting from two or more types of elementary reactions.

    Among all known photochromic compounds, valence isomerization of spiropyran derivatives has been studied best. Nevertheless, even that class of photochromic substances is not free from some significant gaps primarily related to fac-tors determining their fatigue stability and synthesis of com-pounds absorbing at maximally long wavelengths.

    Spiropyrans constitute the most important and extensively studied class of organic photochromes. Research work on these very interesting compounds at the Chemical Technology of Organic Dyes and Phototropic Compounds Department was initiated by Prof. L.S. Efros and Dr. E.R. Zakhs. Spiropyrans undergo reversible color change upon exposure to light in amorphous [22] and crystalline states [23], aluminosilicate gels [24], clay interlayers [25], porous glasses [26], liquid solutions [27–30], liquid crystalline state [31], undercooled melts [32], and polymeric matrix [33–36] as a result of trans-formation of closed (spiro) structure A into open (merocya-nine) structure B and backward. There are experimental proofs indicating that this transformation is mediated by cis-cisoid isomer C [37–39].

    O O

    OO

    Z

    R

    Z

    R_

    +

    Z+ _

    R

    Z

    R

    hν, Δ

    hν', Δ'

    А С3 4

    5

    6

    78

    3

    4

    5 6

    7

    8B

    Many spiropyrans also exhibit thermo- and electrochrom-

    ism [40–42]. Variation of the molecular fragments linked to-gether through a spiro carbon atom and introduction of vari-ous substituent groups makes it possible to vary the spectral range of photochromic transitions and lifetime of photoin-duced form over a very broad range [1–5]. Examples of re-verse photochromism were also reported [43, 44]. In this case, light induces transformation of initial colored form into colorless. For instance, spiropyrans of the coumarin series become colored on heating and lose color under irradiation [45].

    The structure–property relations in the spiropyran series were extensively studied using both experimental and theo-retical methods [1–5, 46–52]. The nature of the heteroring affects the relative stability of the Spirocyclic and merocya-nine forms and spectral–kinetic parameters of photoinduced forms [53–55]. The stability of merocyanine isomer and its spectral parameters are determined by the degree of planari-ty of the chromophore conjugation chain, intramolecular hy-drogen bonding, and evenness of π-bond orders [56].

    Only three representatives of acridine spiropyrans (1–3) have been described previously as model compounds for studying the effect of heteroring nature on thermochromic properties of spiro compounds [57]. Twenty years later, Fish-er [58] reported the results of low-temperature flash photoly-sis study on photochromic properties of one of these com-pounds (3). Deeper color of merocyanine isomers prompted us to examine a larger series of acridine spiropyrans (1–9), the more so that they have never been studied in liquid solu-tion at room temperature.

    Photochromic properties of acridine spiropyrans (1–9) were studied by flash photolysis at room temperature.

    Table 1.Spectral and kinetic parameters of photoinduced forms of spiropy-

    rans in toluene (20ºC, c0 = 1.5·10–5 M, E = 125 J, d = 20 cm) Comp.

    No R λmax, nm (D)а kt, s-1

    1 H 470 (0.03), 620 (0.29) 220 2 8'-OCH3

    530 (0.07), 620 (0.10) 360

    3 5',6'-benzo 585 (1.15) 1400 4 6'-ОСН3 610 (0.37) 570 5 6'-Br 660 (0.05), 700 (0.07) 74 6 6',8'-Br2

    550 (0.008), 725 (0.02) 17

    7 6'-Cl-8'-OCH3 655 (0.14) 320 8 6'-NO2 >750 (1.2)а 45 9 6'-NO2-8'-OCH3 >750 (1.3)а 50

    10 Ind. 6'-NO2б 595 (1.90) 0.062 11 Ind. 5',6'-benzoб 550 (0.15) 22

    а - D is the optical density at the absorption maximum (λ 750 nm for compounds 8 and 9). б - Indoline heteroring

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    O

    NMe R

    RNMe

    O

    1

    2 3

    4

    5

    67

    8

    3 4

    5

    6

    78

    '

    ' ''

    ''

    3

    4

    '

    '

    hν , Δ hν ', Δ '

    Pulsed photoexcitation of colorless spiropyrans 1–9 in po-

    lar and nonpolar solvents gave rise to short-lived merocyanine isomers B which absorbed in the visible region (Fig. 1). The absorption spectrum of the photoinduced form depended on the substituent in the chromene fragment. As might be ex-pected, acridine merocyanines absorb at longer wavelengths than the corresponding indoline analogs (Table 1). The ab-sorption region of merocyanines generated from nitro deriva-tives 8 and 9 extends beyond 750 nm (Fig. 1). We failed to determine the exact positions of their absorption maxima because of limitations of the flash photolysis setup in use. Deep color is generally typical of acridine dyes belonging to different classes. For example, the absorption maximum of 9-(p-dimethylaminostyryl)acridinium in alcohol is located at λ 614 nm [59, 60], and acridine monomethine cyanine dye absorbs at λmax 671 nm [61].

    Fig. 1. Electronic absorption spectra of photoinduced forms of (1–3) acri-

    dine and (4) indoline spiropyrans in benzene at 22°C, c0 = 1.5·10–5 M, E = 125 J; (1) 10-methylspiro[acridine-9,2′-[2H]chromene] (1), (2) 10-

    methylspiro[acridine-9,3′-[3H]benzo[f]chromene] (3), (3) 10-methyl-6′-nitrospiro[acridine-9,2′-[2H]chromene] (8), (4) 1′,3′,3′-trimethyl-6-nitro-

    1′H,3′H-spiro[chromene-2,2′-indole] (10)

    The decay of the colored form of all the examined spiro-pyrans in nonpolar (benzene, toluene) and polar (ethanol) solvents follows first-order kinetics. Electron-withdrawing substituents stabilize the open isomer, thus slowing down the ring closure. By contrast, electron-donating groups shorten the lifetime of the colored form, which is consistent with the generally observed substituent effects on the rate of bleach-ing of spiropyrans of other classes [51]. Table 1 also contains photochromic parameters of model 1′,3′,3′-trimethyl-6-nitro-1′H,3′H-spiro[chromene-2,2′-indole] (10) determined under analogous conditions. It is seen that the bleaching rate con-

    stant of 10 is lower by three orders of magnitude than that found for nitro-substituted acridine spiropyran 8. This differ-ence may be rationalized by change of the heteroring nature and steric hindrances intrinsic to acridine spiropyrans due to hydrogen atoms in positions 1 and 8. The merocyanine form of acridine–benzochromene derivative 3 turned colorless at a rate exceeding the rate of decoloration of its indoline analog 11 by two orders of magnitude. Thus acridine spiropyrans are characterized by deeper color and faster decoloration of their merocyanine isomers as compared to extensively studied indoline analogs.

    Photochromic parameters of benzo derivative 3 were measured in ethanol and benzene. The absorption maximum of the merocyanine isomer shifts from λ 580 to 640 nm in going from benzene to ethanol, indicating positive solvato-chromism and hence dominant contribution of the quinoid structure to the photoinduced form. The bleaching rate con-stant in alcohol was lower than in benzene only by 25%, which suggests the absence of hydrogen bonding between the merocyanine isomer and solvent, in contrast to indoline spiropyrans [62].

    An important aspect of photoinduced transformations of spiropyrans is the nature of excited state that undergoes ring opening. The relative quantum yield for photocoloration of acidine spiropyrans decreases in going from unsubstituted compound 1 to bromo and dibromo derivatives 5 and 6 (1, 0.3, 0.1). Solutions of these compounds with equal initial concentrations were subjected to flash photolysis under simi-lar conditions. The relative quantum yield was calculated as the ratio of the optical densities at the absorption maxima of the photoinduced forms assuming equal molar absorption coefficients at the respective wavelengths. This assumption seems to be reasonable, for the molar absorption factors of the photoinduced forms of phenanthridine spirochromenes with the same substituents differ by a factor of no more than 1.7 [63]. Moreover, the absorbance of the dibromo derivative is even higher than that of the mono bromo derivative.

    Addition of bromobenzene to a benzene solution of spiro-pyran 2 (c0 = 1.96·10–5 M) did not change the absorption spectrum of the colored form, but the optical density at λ 610 nm decreased almost twofold as the concentration of bromo-benzene increased from 0 to 1.6 M. To avoid internal filter effect, the jacket of the spectrophotometric cell and the beakers encompassing the flash lamps were filled with pure bromobenzene so that the overall bromobenzene layer thick-ness was about 1 cm. Therefore, bromobenzene present in solution at a low concentration could not absorb exciting light and act as internal filter. This means that the observed drop of the optical density is the result of competition between the photochemical ring opening reaction in the singlet excited state and external heavy atom-facilitated intersystem crossing to the triplet state. Atmospheric oxygen almost does not af-fect the bleaching rate, and the optical density of the colored form in air-saturated solution is lower by 6–15% than in oxy-gen-free solution.

    The above experimental results, namely reduction of the photoisomerization quantum yield in the series 1 > 5 > 6 and effects of external heavy atom and atmospheric oxygen, sug-gest singlet nature of the reactive state of acridine spiropy-rans having no nitro group. The effect of oxygen on the be-havior of nitro-substituted spirochromenes 8 and 9 was more appreciable: the optical density of their oxygen-free solutions was higher by 35–40% as compared to air-saturated ones. However, the kinetics of the bleaching process almost did not change. This pattern implies some contribution of triplet states to the photocoloration in direct photoexcitation of ni-tro-substituted spiropyrans.

    The photocoloration process can be sensitized by appro-priate triplet sensitizers. With benzophenone as triplet donor (Et = 69 kcal/mol), ring opening in acridine spiropyrans 3 and

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    9 was observed in deoxygenated benzene. In this case, the absorption spectrum and back reaction kinetics were identical to those observed on direct photoexcitation. The only differ-ence was a sharp reduction of the light fatigue resistance upon triplet sensitization. Even the most photostable spiropy-ran 3 decomposed to an appreciable extent after a few flash-es under flash photolysis in benzene solution in the presence of benzophenone. When the flash photolysis was carried out under simultaneous continuous irradiation with a DRSh-1000 lamp, signal from the photoinduced form disappeared in 15 min. In the absence of benzophenone, the parameters of the photoinduced form did not change over a period of more than 2 h. To prevent direct absorption of the exciting light by spi-ropyran, the flash lamps were immersed in a solution of the same spiropyran with a tenfold concentration. Therefore, we concluded that the photodecomposition is mediated by the triplet excited state which is populated via triplet–triplet trans-fer from benzophenone. Addition of benzophenone to an air-saturated benzene solution where triplet–triplet transfer is ruled out by triplet quenching with oxygen did not lead to an appreciable decomposition (Fig. 2).

    Fig. 2. Relative variation of the photoinduced optical density at λmax of

    spiropyrans (at λ 750 nm for compound 9) versus the number of flashes (N); benzene, c = 1.5·10–5 M, E = 125 J; (1–3) on exposure to air; (4, 5)

    oxygen-free solution; (1, 4) 10-methylspiro[acridine-9,3′-[3H]benzo[f]chromene] (3), (2, 3, 5) 8′-methoxy-10-methyl-6′-

    nitrospiro[acridine-9,2′-[2H]chromene) (9); (1, 2) in the absence of benzo-phenone; (3–5) in the presence of benzophenone.

    Thus the main path of singlet state deactivation of acri-

    dine spiropyrans is photochemical pyran ring opening. The contribution of intersystem crossing to the triplet state is small, but it increases upon introduction of a nitro group, which is reflected in increased induced absorbance of oxygen-free solutions.

    Rise in the concentration of photoinduced merocyanine and just the possibility for triplet sensitization suggest that the photochemical ring opening could involve the triplet state. On the one hand, participation of the triplet state in the pho-tocoloration process increases the overall yield of the colored form, and on the other, sharply reduces the light fatigue re-sistance of spiropyrans since irreversible photodegradation (as follows from the obtained data; see Fig. 2) occurs mainly through triplet states.

    Photochromic parameters of spiropyrans are determined by the nature of both heterocyclic fragments linked by a spiro atom. The relative stabilities of colored merocyanine isomer B and colorless spiropyran A strongly depend on the stereoelec-tronic parameters of the constituent fragments and substitu-ents therein [64–66]. For example, the presence of bulky substituents in positions neighboring to the spiro carbon atom

    is essential for enhanced stability of the closed isomer of ben-zothiazoline and azine spiropyrans [67, 68].

    ONR

    R,

    R,,

    X

    X = C(Alk)2, S, O, Se; R' = Alk, Ar, OAlk, OAr, SMe, SPh.

    It is also known that spiropyrans of the perimidine series under ambient conditions are more stable than isomeric mer-ocyanines [69], whereas their azole analogs, benzimidazole derivatives, in most cases are stable only in the merocyanine form [66, 70].

    O

    N

    NMe

    Me

    RNCH3

    +

    N

    OCH3-

    We presumed that the presence of a 1,3-thiazine frag-

    ment in the peri-naphthothiazine system should favor spiro cyclization as well.

    O

    S

    NMe

    S

    N

    OMe

    R +

    _

    β

    12А-15А 12B-15B 12 R=H, 13 R=5',6'-benzo, 14 R=8'-OMe, 15 R=7'-NEt2

    Fig. 3. Electronic absorption spectra of (1) 3′-methyl-2′,3′-dihydrospiro[2H-

    chromene-2,2′-naphtho[1,8-de][1,3]thiazine] (12) in ethanol and (2, 3) N,N-diethyl-3′-methyl-2′,3′-dihydrospiro[2H-chromene-2,2′-naphtho[1,8-

    de][1,3]thiazin]-7-amine (15) in (2) heptane and (3) ethanol.

    Crystalline spiropyrans 12A–15A are colorless, and their solutions in common organic solvents are also colorless. Only compound 15A having a diethylamino group undergoes par-tial isomerization upon dissolution in ethanol; its solution ac-quires an intense pink color, and the long-wave absorption band features two maxima with almost equal intensities at λ 526 and 547 nm (apparent molar absorption coefficients ε =

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    1200 and 1250 cm2/mmol; Fig. 3). To evaluate spectral pa-rameters of isomer 15B we used styryl dye SD1 as model; the absorption maximum of the latter almost coincides with that of colored isomer 15B [71, 72]. Assuming that the molar absorption coefficients of 15B and SD1 are similar, the equi-librium concentration of 15B in alcohol was estimated at about 2.5%.

    S

    N NMe2CH3

    +

    ClO4_

    SD1 Strong stabilization of colored structure B is achieved by

    introduction into the 7′-position of a strong electron-donating substituent which, after opening of the pyran ring, appears in direct polar conjugation with the cationic heteroatom. The thermal equilibrium of diethylamino-substituted peri-naphthothiazine derivative 15A at room temperature is dis-placed toward the spirocyclic structure to a considerably greater extent as compared to indoline and phenanthridine analogs 16 and 17. This is consistent with the lower acidity of the peri-naphthothiazine system relative to indoline and phenanthridine, which was estimated by the deviations of λmax of the corresponding styryl dyes. The minimal deviation was found for naphthothiazine dye SD1 (16%), a larger deviation, for indoline derivative SD2 (24%), and the maximal, for phe-nanthridine analog SD3 (~65%).

    To calculate the relative deviations we used λmax for styryl dyes SD1 and SD3 and previously reported values for indo-line derivative SD2 (λmax 545 nm) [74], symmetric indoline trimethinecyanine (λmax 548 nm) [75], Michler’s hydrol blue (λmax 607.5 nm) [76], and symmetric phenanthridine (λmax 610 nm) [77] and peri-naphthothiazine trimethinecyanines (λmax 530 nm) [71].

    ON

    MeMe

    MeNEt2

    NMe

    MeMe

    NMe2+

    I_

    16 SD2

    λmax (EtOH) nm, (lg ε): 545, (4.92)

    O

    NMe

    NEt2

    NMe

    NMe2

    +

    MeSO4_

    17 SD3

    λmax (EtOH) nm, (lg ε): 513, (4.42)

    Fig. 4. Electronic absorption spectra of the colored form of N,N-diethyl-3′-methyl-2′,3′-dihydrospiro[2H-chromene-2,2′-naphtho[1,8-de][1,3]thiazin]-7-amine (15) in deoxygenated solutions in (1) toluene, c = 4.1·10–5 M, 100 μs after flash start; (2) ethanol, c = 6.8·10–6 M, 100 μs after flash start;

    and (3) toluene, c = 1.7·10–6 M, 50 ms after flash start. Flash energy 125 J.

    We were the first to reveal photochromic properties of

    spiropyrans 12–17 in liquid solution [78]. Pulsed photoexcita-tion of solutions of 12A–15A in toluene or ethanol at room temperature induced absorption in visible region, and the absorption pattern did not change to an appreciable extent if air was present in the system (Fig. 4, Table 2). The absorp-tion maximum of the colored form in more polar solvent ap-peared at longer wavelengths.

    Unlike indoline [1, 47] and phenanthridine spiropyrans [79] studied previously, colored isomers 13B–15B exhibit positive rather than negative solvatochromism, indicating their predominantly quinoid structure. The absorption maxima of the colored form of unsubstituted derivative (12B) in etha-nol and toluene are almost similar.

    Table 2. Thermal bleaching rate constants of the photoinduced forms of

    spiropyrans 12–15

    Comp. No

    R

    k, s-1 Published data for analogs of 12–15 with different heterocycles (toluene)

    Tolu-ene

    96 % Etha-nol

    Phe-nanthri

    thri-dine

    Indo-line

    1,3-Dithi-ole

    12 H 91 25 29 (0.96) —13 5',6'-benzo 2800 700 100 17 (3.3)a14 8'-OCH3 70 16 8,70 1,9 —15 7'-N(C2H5)2

    130 34 220b 1.1b,c (0.8)a a In dioxane. b Determined in this work. c In heptane

    5,6-Benzo annulation of the chromene fragment and in-

    troduction of a diethylamino group into its 7-position consid-erably increase the optical density at λmax of the colored forms (B) of compounds 13 and 15 in toluene and ethanol relative to that observed for unsubstituted derivative 12, other condi-tions (initial concentration, flash energy, spectral composition of the exciting radiation, and excitation geometry) being equal. Provided that the molar absorption coefficients of mer-ocyanines 12B–15B are fairly similar, as was determined for phenanthridine derivatives [63], the observed difference in the photoinduced optical densities should be attributed to considerably higher photocoloration quantum yield of 13A and 15A.

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    As noted above, compound 15 in alcoholic solution exists

    mainly as spiropyran 15A, while the fraction of merocyanine isomer 15B is as small as ~2.5%. Pulsed UV radiation leads to strong increase of absorbance in the visible region due to rise in the concentration of 15B as a result of A → B pho-toisomerization. In all cases, the photoinduced optical density is strictly linear in the flash energy at the initial part, which implies a single-quantum photocoloration process. When the initial concentration of 15A in alcohol is low (c0 = (10–6 M), the induced optical density tends to saturation as the flash energy increases, though it does not reach a limiting value even at a flash energy of 400 J. The gain in the optical densi-ty per energy unit merely decreases at high flash energies (200–400 J). Therefore, the limiting optical density may be estimated only by extrapolation. The limiting value corre-sponds to the maximum possible (under the given excitation conditions) displacement of the A↔B equilibrium toward merocyanine isomer. The minimal molar absorption coefficient of the photoinduced isomer can be evaluated assuming that there are no processes reducing the concentration of isomer B (e.g., back photoreaction B→A) and that the concentration of merocyanine 15B in the saturation region is equal to the initial concentration of 15 (i.e., 100% conversion A→B). Its values should be no less than 1.1·104 cm2/mmol at the ab-sorption maximum (λ 550 nm). Correspondingly, the fraction of merocyanine 15B in alcohol solution at room temperature should not exceed 10%. The difference from the value de-termined with the use of model styryl dye SD1 (2.5%) is likely to reflect processes neglected in the simplified ap-proach, which reduce the real concentration of 15B.

    As follows from the data in Table 2, the bleaching rate constant for all the examined compounds decreases in going from toluene to 96% ethanol. Such relation between the bleaching rate constant and solvent polarity is generally typi-cal for negative solvatochromism of the colored isomer (e.g., for indoline [42] and phenanthridine spiropyrans [80]) rather than positive as in our case. Despite predominating quinoid structure of the merocyanine isomer, it remains more polar than the initial spiropyran, and lower bleaching rate in more polar solvent is determined by better stabilization of more polar merocyanine B compared to less polar spiropyran A [81]. Nevertheless, in the series of heterocyclic spiropyrans there are examples of the reverse bleaching rate constant–solvent polarity relation. For instance, naphthofuran [82] and 2-oxaindan spiropyrans [83] display increase in the bleaching rate constant with rise in solvent polarity. These compounds, as well as naphthothiazine spiropyrans 13–15, are character-ized by positive solvatochromism. The above data indicate an important role of the heteroring nature in the bleaching pro-cess. The cyclization to colorless spiropyran with almost or-thogonal orientation of molecular fragments is preceded by isomerization and charge transfer from the heteroatom to the carbon atom in position 2. This charge transfer is more facile for less basic heterocycle [84], while the ease of isomerization in solution is determined mainly by the orders of bonds in-volved in that isomerization. Therefore, provided that the cyclization is the rate-limiting step of bleaching of spiro-chromenes, the nature of the other heteroring becomes cru-cial, for it determines the magnitude of local charges on the oxygen and carbon atoms. As already noted, naphthothiazine ring is less basic than indoline (according to Brooker [73]); in addition, quinoid structure of merocyanine isomer implies a lower order of the C3′–C4′ bond; these factors should facilitate the isomerization. Thus, the main reasons favoring the cy-clization of merocyanines 12B–15B to the corresponding spiropyrans apply better to naphthothiazine spiropyrans as compared to indoline and phenanthridine derivatives. From the kinetic viewpoints, this is reflected in shortening of the lifetime of the colored merocyanine isomer in the series indo-line > phenanthridine > naphthothiazine derivatives (Table

    2). The lifetime of the colored form of differently substituted naphthothiazine spiropyrans, determined under unimolecular bleaching conditions, changes in approximately the same order of substituents as that observed for indoline, phenan-thridine, and dithiole spiropyrans. Benzo derivative 13 turned out to be most kinetically labile at room temperature, though, according to [85], 5,6-benzo annulation of the chromene ring should enhance thermodynamic stability of the colored (open) form compared to spirocyclic structure. Presumably, the acti-vation barrier to the back cyclization of 13B is lower than those for compounds with different substituents, which is responsible for the shorter lifetime of the merocyanine iso-mer.

    Thus, a new class of photochromic spiropyrans belonging to the naphthotiazine series has been revealed and character-ized by spectral–kinetic method.

    We tried to control the rate of bleaching in the series of phenanthridine spiropyrans via introduction of a substituent into the 3-position of the chromene moiety. Substituted 3-fluoro-5-methyl-5′,6′-dihydrospiro[2H-chromene-2,6′-phenanthridines] 18–22 were subjected to flash photolysis in benzene at room temperature, and parameters characterizing their photochromic behavior were compared with those for fluorine-free analogs.

    O

    NM e

    F

    R R = H (18), 8'-ОСН3 (19),

    6'-Cl-8'-ОСН3 (20), 7'-N(C2H5)2 (21), 6'- NO2 (22)

    Unlike fluorine-free analogs, the colored forms of fluorine-

    containing spiropyrans 18–22 displayed more uniform ab-sorption pattern in a broad range (λ 420–750 nm). The ther-mal bleaching reaction conformed to first-order kinetic equa-tion at least within the flash energy range from 20–500 J and in the flash duration range from 50 μs to complete decolora-tion. By contrast, the initial part of the kinetic curves for fluo-rine-free spiropyrans indicated formation of a short-lived iso-mer in a low concentration. The rate constants for bleaching of the colored isomers of 18–20 were higher by two orders of magnitude than those found for the corresponding fluo-rine-free analogs, and the bleaching rate constant for nitro derivative 22 was higher even by seven orders of magnitude (Table 3).

    Table 3. Thermal bleaching rate constants for colored forms of fluorine-

    containing (kF) and fluorine-free (k) phenanthridine spiropyrans in toluene Comp. No R k, s-1 (20º С) kF, s-1

    18 Н 3·10 5,5.10319 8'-ОСН3 9 1,6.10320 6'-Cl-8'-ОСН3 5,1 1,5.10321 7'-N(C2H5)2 2,2.102 b 2,9.10222 6'- NO2 3,8.10-4 a 1,4.103

    a In dioxane. b In heptane

    The reason is that the presence of a fluorine atom in the α-

    position with respect to the cyclization reaction center increases the positive charge on the latter, thus favoring the reverse reac-tion. Furthermore, larger size of fluorine atom compared to hy-drogen causes the open photoinduced form to deviate from pla-nar structure, which also reduces its stability.

    Spiropyrans having no substituent in the 3-position of the chromene moiety conform to an empirical rule according to which electron-withdrawing substituents in the chromene frag-

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    ment reduce the rate of thermal bleaching while electron-donating groups increase it. Although the dependence of the bleaching rate constant upon Hammett substituent constant qualitatively reflects this relation, it is far from being linear. Espe-cially strong deviation is observed for 7-diethylamino derivative 21. Deviations from linearity may be rationalized by steric and conjugation effects in addition to electronic. In particular, a strong electron-donating substituent in the 7-position appears, after pyran ring opening, to be conjugated with the cationic het-eroatom via direct polar effect, which should strongly stabilize the colored form (B) and slow down the isomerization B→A.

    On the other hand, electron-donor effect of the diethylamino

    group should accelerate the same process due to increase of the negative charge on the oxygen atom. Just these opposite effects cause the Hammett dependence to deviate from linearity.

    The relative light resistance of spiropyrans 12–22 belonging to different classes was estimated in benzene using a combina-tion of continuous irradiation with a DRSh-1000 lamp (λ 313 nm) and flash photolysis. The optical density at the absorption maxi-mum of the colored isomers (at λ 750 nm for acridine nitro deriv-atives) generated by identical flashes was measured every 15 min under continuous irradiation at λ 313 nm. Under these condi-tions, 5,6-benzo derivative 3 of the acridine series turned out to be the most stable (it was superior to indoline analog 11), while 6-nitro indoline spiropyran 10 was the least stable. Nitro substi-tution on the chromene fragment in the series of acridine spiro-pyrans reduces the light resistance as well, but even least stable (among acridine spiropyrans) 6′-nitro derivative 8 displayed bet-ter stability than known indoline spiropyran 10 (Fig. 5). Thus the light resistance of spiropyrans increases in the following heteror-ing series: indoline < phenanthridine < phenanthridine (3-fluorochromene) < naphthothiazine < acridine

    Fig. 5. Relative variation of the optical density at λmax of the colored forms

    of spiropyrans 3, 7, and 9–11 in benzene solution upon continuous irradia-tion with a DRSh-1000 lamp: (1) 10-methyl-10H-spiro[acridine-9,3′-[3H]benzo[f]chromene] (3), (2) 1′,3′,3′-trimethyl-1′H,3′H-spiro[3H-

    benzo[f]chromene-3,2′-indole] (11), (3) 10-methyl-6′-chloro-8′-methoxy-10H-spiro[acridine-9,2′-[2H]chromene] (7), (4) 8′-methoxy-10-methyl-6′-nitro-10H-spiro[acridine-9,2′-[2H]chromene] (9) (λ 750 nm), (5) 1′,3′,3′-

    trimethyl-6-nitro-1′H,3′H-spiro[2H-chromene-2,2'-indole] (10).

    To conclude, we have revealed experimentally the follow-ing factors responsible for light resistance of spiropyrans: predominately singlet path of the photocoloration process,

    quinoid structure and short lifetime of the colored photoin-duced form, and the absence of nitro substitution. Obviously, spiropyrans of the acridine series are characterized by favora-ble combination of light resistance, color depth, and photo-coloration quantum yield, which makes them promising for practical application.

    The results presented in this article were obtained at the Chemical Technology of Organic Dyes and Phototropic Com-pounds Department in collaboration with E.R. Zakhs, V.P. Matrynova, R.P. Polyakova, N.G. Leshenyuk, and L.A. Zvenigorodskaya.

    The study was performed under financial support by the Russian Foundation for Basic Research (project no. 13-08-1425-а).

    References

    1. Photochromism, Brown, G.H., Ed., New York: Wiley, 1971. 853 pp.

    2. Barachevskii, V.A., Lashkov, G.I., and Tsekhomskii, V.A., Fotokhromizm i ego primenenie (Photochromism and Its Applica-tion), Moscow: Khimiya, 1977. 280 pp.

    3. V.L. Ermolaev, E.B. Sveshnikova, L.Yu. Mironov. Nano-technologies in Russia. 2012. V. 7. No 3-4. рр. 110–116.

    4. Organic Photochromes, Eltsov, A.V., Ed., New York: Ple-num, 1990. 269 pp.

    5. Photochromism. Molecules and Systems, Dürr, H. and Bouas-Laurent H., Eds., Amsterdam: Elsevier, 1990. 1068 pp.

    6. Organic Photochromic and Thermochromic Compounds, Crano, J.C. and Guglielmetti, R.J., Eds., New York: Plenum, 1999, vols. 1, 2. 820 pp.

    7. CRC Handbook of Organic Photochemistry and Photobiolo-gy, Horspool, W.M. and Lenci, F., Eds., Boca Raton: CRC, 2004, 2nd ed. 173 pp.

    8. Barachevskii, V.A., Kozenkov, V.M., and Gerulaitis, Yu.N., Zh. Nauch. Prikl. Fotogr. Kinematogr., 1974, vol. 19, no. 3, pp. 161–174.

    9. Barachevskii, V.A., Oliveira-Kampos, A.M., Stebunova, L.V., et al., Zh. Nauch. Prikl. Fotogr., 2002, vol. 47, no. 2, pp. 4–8.

    10. Balzani, V., Gomes, L.M., and Stoddart, J.-F., Acc. Chem. Res., 1998, vol. 31, no. 7, pp. 405–414.

    11. Minkin, V.I., Ros. Khim. Zh., 2000, vol. 44, no. 6, pp. 3–13.

    12. Barachevskii, V.A., Zh. Vses. Khim. Ob–va., 1974, vol. 19, no. 4, pp. 423–433.

    13. Pfeifer, U., Fukumura, H., Misawa, H., et al., J. Am. Chem. Soc., 1992, vol. 114, no. 11, pp. 4417–4418.

    14. Bonett, R., Chem. Soc. Rev., 1995., vol. 24, no. 1, p. 19–33.

    15. Barachevskii, V.A., Gal’bershtamm, M.A., Gerasimenko, Yu.E., and Gerulaitis, Yu.N., Zh. Vses. Khim. Ob–va., 1974, vol. 19, no. 1, pp. 85–92.

    16. Heller, H.G., Kon, K., Kose, M., and Rowles, N., Mol. Cryst. Liq. Cryst., 1997, vol. 297, pp. 73–80.

    17. Vinogradov, A.M., Kuznetsov, E.I., Marevtsev, V.S., et al., Biol. Membr., 1990, vol. 7. no. 10, pp. 1097–1104.

    18. Gritsan, N.P. and Klimenko, L.S., J. Photochem. Photobiol. A: Chem., 1993, vol. 70, no. 2, pp. 103–117.

    19. Kaymo, F.M. and Tomasulo, M., J. Phys. Chem. A, 2005, vol. 109, no. 3, pp. 7343–7352.

    20. Barachevskii, V.A., Besserebryanye i neobychnye fotograficheskie protsessy (Non-silver and Unusual Photographic Processes), Vil’nyus: SSO Lit. SSR, 1980, pp. 31–33.

    21. Bren’, V.A., Usp. Khim., 2001, vol. 70, no. 12, pp. 1152–1174.

    22. Sychev, I.Yu., Khamchukov, Yu.D., Marevtsev, V.S., et al., Zh. Nauchn. Prikl. Fotogr., 1998, vol. 43. no. 5, pp. 54–58.

    23. Asaki, T., Suzuki, M., and Masuhara, H., J. Phys. Chem. A, 2002, vol. 106, no. 10, pp. 2335–2340.

    24. Preston, D., J. Phys. Chem., 1990, vol. 94, no. 10, pp. 4167–4172.

  • I. CHEMISTRY AND CHEMICAL TECHNOLOGY • TECHNOLOGY OF ORGANIC SUBSTANCES

    25. Takagi, K., Kurematsu, T., and Sawaki, Y., J. Chem. Soc.,

    Perkin Trans. 2, 1991, no. 9, pp. 1517–1521. 26. Veselova, T.V., Obyknovennaya, I.E., and Cherkasov,

    A.S., Opt. Spektrosk., 1990, vol. 69, no. 6, pp. 1300–1305. 27. Kalisky, Y., Orlowski, T.E., and Williams, D.J., J. Phys.

    Chem., 1983, vol. 87, no. 26, pp. 5333–5338. 28. Lashkov, G.I. and Shablya, A.V., Izv. Akad. Nauk SSSR,

    Ser. Fiz., 1968, vol. 32, no. 9, pp. 1569–1574. 29. Heiligman-Rim, R., Hirshberg, Y., and Fischer, E., J. Phys.

    Chem., 1962, vol. 66, no. 12, pp. 2465–2470. 30. Appriou, P., Brelivet, J., Trebaul, C., and Guglielmetti, R.,

    J. Photochem., 1976–1977, vol. 6, no. 1, pp. 47–54. 31. Song, H.H., Yoo, J.-H., Doe, J.-H., et al., Mol. Cryst. Liq.

    Cryst. Sci. Technol., Sect. A, 2000, vol. 349, pp. 267–270. 32. Zelichenok, A., Bucholtz, F., Ratner, J., et al., J.

    Photochem. Photobiol. A, 1994, vol. 77, nos. 2–3, pp. 201–206. 33. Semyakina, G.N., Merzabekova, T.G., Popova, N.I., and

    Zakhs, E.R., Izv. Akad. Nauk SSSR, Ser. Khim., 1983, no. 6, pp. 1277–1282.

    34. Applied Photochromic Polymer Systems, McArdle, C.B., Ed., Glasgow: Blackie, 1992

    35. Hama, H., Miyashita, A., Yamaoka, K., and Nakahara, H., Chem. Lett., 1996, vol. 25, no. 4, pp. 313–314.

    36. Malkin, Ya.N., Popova, N.I., Shashkov, A.S., et al., Dokl. Akad. Nauk SSSR, 1990, vol. 246, no. 2, pp. 384–388.

    37. Krysanov, S.A. and Alfimov, M.V., Dokl. Akad. Nauk SSSR, 1983, vol. 272. no. 2, pp. 406–410.

    38. Petrukhin, A.N Antipin, S.A., Gostev, F.E., et al., Khim. Fiz., 2000, vol. 19, no. 2, pp. 90–97.

    39. Malkin, Ya.N., Lysak, N.A., Tikhomirov, S.A., et al., Izv. Akad. Nauk SSSR, Ser. Khim., 1987, no. 9, pp. 2129–2132.

    40. Minkin, V.I., Chem. Rev., 2004, vol. 104, no. 5, pp. 2751–2776.

    41. Kozlovskaya, I.E., Sychev, I.Yu., Luchina, V.G., and Marevtsev, V.S., Izv. Akad. Nauk, Ser. Khim., 1994, no. 5, pp. 836–838.

    42. Marevtsev, V.S., Zaichenko, N.L., Beshenko, S.I., et al., Izv. Akad. Nauk SSSR, Ser. Khim., 1980, no. 10, pp. 2272–2278.

    43. Anisimov, V.N., Ataev, A.S., Mordaleishvili, I.R., et al., Izv. Akad. Nauk SSSR, Ser. Khim., 1991, no. 11, pp. 2643–2644.

    44. Liu, S.-H. and Wu, C.T., Synth. Commun., 2000, vol. 30, no. 5, pp. 895–902.

    45. Metelitsa, A.V., Knyazhansky, M.I., Ivanitsky, V.V., et al., Mol. Cryst. Liq. Cryst., 1994, vol. 246, no. 1, pp. 37–40.

    46. Tyer, N.-W. and Becker, R.S., J. Am. Chem. Soc., 1970, vol. 92, no. 5, pp. 1289–1294.

    47. Dzhaparidze, K.G., Spirokhromeny (Spirochromenes), Tbilisi: Metsniereba, 1979, 112 pp.

    48. Simkin, B.Ya., Minkin, V.I., and Nivorozhkin, L.E., Khim. Geterotsikl. Soedin., 1974, no. 1, pp. 76–83.

    49. Minkin V.I, Chemical Reviews, 2004, vol. 104, no 5, pp. 2751-2776.

    50. Kellmann, A., Tfibel, F., Dubest, R., et al., J. Photochem. Photobiol. A, 1989, vol. 49, no. 1, pp. 63–73.

    51. Berman, E., Fox, R.E., and Thomson, F.D., J. Am. Chem. Soc., 1959, vol. 81, no. 10, pp. 5605–5609.

    52. Simkin, B.Ya., Minkin, V.I., and Nivorozhkin, L.E., Khim. Geterotsikl. Soedin., 1978, no. 9, pp. 1180–1192.

    53. Malkin, Ya.N., Cand. Sci. (Chem.) Dissertation, Moscow, 1976.

    54. Dvornikov, A.S., Malkin, Ya.N., and Kuz’min, M.G., Izv. Akad. Nauk SSSR, Ser. Khim., 1982, no. 7, pp. 1520–1526.

    55. Lenoble, C. and Becker, R.S., J. Phys. Chem., 1986, vol. 90, no. 1, pp. 62–65.

    56. Aldoshin, S.M., Atovmyan, L.O., D’yachenko, O.A., and Gal’bershtam, M.A., Izv. Akad. Nauk SSSR, Ser. Khim., 1981, no. 12, pp. 2720–2725.

    57. Wizinger, R. and Wenning, H., Helv. Chim. Acta, 1940, vol. 23, pp. 247–271.

    58. Heiligman-Rim, R., Hirshberg, Y., and Fischer, E., J. Phys. Chem., 1962, vol. 66, no. 12, pp. 2470–2477.

    59. Izmail’skii, V.A. and Vishnevskii, L.D., Dokl. Akad. Nauk SSSR, 1958, vol. 121, no. 1, pp. 111–114.

    60. Kitrosskii, N.A. and Izmail’skii, V.A., Dokl. Akad. Nauk SSSR, 1960, vol. 132, no. 3, pp. 598–601.

    61. Albert, A., The Acridines, London: Edward Arnold, 1966, 2nd ed., 604 pp.

    62. Murin, V.A., Mandzhikov, V.F., and Barachevskii, V.A., Opt. Spektrosk., 1974, vol. 37, no. 6, pp. 1174–1176.

    63. Malkin, Ya.N., Kuz’min, V.A., and Zakhs, E.R., Izv. Akad. Nauk SSSR, Ser. Khim., 1979, no. 3, pp. 643–644.

    64. Appriou, P., Guglielmetti, R., and Garnier, F., J. Photochem., 1978, vol. 8, no. 2, pp. 145–165.

    65. Komisarov, V.N. and Metelitsa, A.V., Panorama sovremennoi khimii Rossii. Sbornik obzornykh statei (Panorama of Modern Chemistry in Russia. A Collection of Review Articles), Moscow: Khimiya, 2003, pp. 325–340.

    66. Zakhs, E.R., Martynova, V.P., and Efros, L.S., Khim. Geterotsikl. Soedin., 1979, no. 11, pp. 4435–4459.

    67. Kister, J., Pommier, H. P., Metzger, J., and Guglielmetti, R., Can. J. Chem., 1984, vol. 62, no. 11, pp. 2372–2378.

    68. Samat, A., Guglielmetti, R., and Metzger, J., Helv. Chim. Acta, 1972, vol. 55, no. 6, pp. 1782–1801.

    69. Anisimova, A.A., Pozharskii, A.F., Nivorozhkin, L.E., and Minkin, V.I., Khim. Geterotsikl. Soedin., 1978, no. 3, pp. 108–112.

    70. Maguet, M., Le Baccon, M., Poirier, Y., and Guglielmetti, R., Can. J. Chem., 1982, vol. 60, no. 21, pp. 2644–2653.

    71. Hamer, F.M. and Rathbone, R.J., J. Chem. Soc., 1943, pp. 487–491.

    72. Joy, H. van B. and Bogert, M.T., J. Org. Chem., 1936, vol. 01, no. 3, pp. 236–244.

    73. Brooker, L.G.S., Sklar, A.L., Gressman, H.W.J., et al., J. Am. Chem. Soc., 1945, vol. 67, pp. 1875–1889.

    74. Hamer, F.M. The Cyanine Dyes and Related Compounds, New York: Wiley, 1964, p. 438.

    75. Kogan, I.M., Khimiya krasitelei (Chemistry of Dyes), Moscow: Goskhimizdat, 1956, p. 393.

    76. Gordon, P.F. and Gregory, P., Organic Chemistry in Colour, Springer: Berlin, 1983. Translated under the title Organicheskaya khimiya krasitelei, Moscow: Mir, 1987, p. 266.

    77. Jesthi, P.K., Roy, M., and Mohapatra, S.K., J. Indian Chem. Soc., 1978, vol. 55, no. 7, pp. 712–715.

    78. Zakhs, E.R., Martynova, V.P., and Ponyaev, A.I., Russ. J. Gen. Chem., 1996, vol. 66, no. 8, pp. 1349–1361.

    79. Malkin, Ya.N., Kuz’min, V.A., Martynova, V.P., et al., Izv. Akad. Nauk SSSR, Ser. Khim., 1975, no. 10, pp. 2189–2193.

    80. Kuzmin, V.A., Malkin, Y.N., and Zakhs, E.R., Photogr. Sci. Eng., 1979, vol. 23, no. 1, pp. 20–26.

    81. Samat, A. and Metzger, J., J. Phys. Chem., 1972, vol. 76. no. 24, pp. 3554–3558.

    82. Dvornikov, A.S., Malkin, Ya.N., Mezheritskii, V.V., et al., Izv. Akad. Nauk SSSR, Ser. Khim., 1982, no. 9, pp. 2014–2019.

    83. Murin, V.A., Barachevskii, V.A., Voloshin, N.A et al., Opt. Spektrosk., 1979, vol. 46, no. 5, pp. 1027–1029.

    84. Gal’bershtam, M.A., Przhiyalgovskaya, N.M., Khrolova, et al., Khim. Geterotsikl. Soedin., 1977, no. 12, pp. 1640–1644.

    85. Minkin, V.I., Nivorozhkin, L.E., Trofimova, N.S., et al., Zh. Org. Khim., 1975, vol. 11, no. 4, pp. 828–836.

    /ColorImageDict > /JPEG2000ColorACSImageDict > /JPEG2000ColorImageDict > /AntiAliasGrayImages false /CropGrayImages true /GrayImageMinResolution 300 /GrayImageMinResolutionPolicy /OK /DownsampleGrayImages true /GrayImageDownsampleType /Subsample /GrayImageResolution 300 /GrayImageDepth 8 /GrayImageMinDownsampleDepth 2 /GrayImageDownsampleThreshold 1.50000 /EncodeGrayImages true /GrayImageFilter /FlateEncode /AutoFilterGrayImages false /GrayImageAutoFilterStrategy /JPEG /GrayACSImageDict > /GrayImageDict > /JPEG2000GrayACSImageDict > /JPEG2000GrayImageDict > /AntiAliasMonoImages false /CropMonoImages true /MonoImageMinResolution 1200 /MonoImageMinResolutionPolicy /OK /DownsampleMonoImages true /MonoImageDownsampleType /Bicubic /MonoImageResolution 1200 /MonoImageDepth -1 /MonoImageDownsampleThreshold 1.50000 /EncodeMonoImages true /MonoImageFilter /CCITTFaxEncode /MonoImageDict > /AllowPSXObjects false /CheckCompliance [ /None ] /PDFX1aCheck false /PDFX3Check false /PDFXCompliantPDFOnly false /PDFXNoTrimBoxError true /PDFXTrimBoxToMediaBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXSetBleedBoxToMediaBox true /PDFXBleedBoxToTrimBoxOffset [ 0.00000 0.00000 0.00000 0.00000 ] /PDFXOutputIntentProfile () /PDFXOutputConditionIdentifier () /PDFXOutputCondition () /PDFXRegistryName () /PDFXTrapped /False

    /CreateJDFFile false /Description > /Namespace [ (Adobe) (Common) (1.0) ] /OtherNamespaces [ > /FormElements false /GenerateStructure false /IncludeBookmarks false /IncludeHyperlinks false /IncludeInteractive false /IncludeLayers false /IncludeProfiles false /MultimediaHandling /UseObjectSettings /Namespace [ (Adobe) (CreativeSuite) (2.0) ] /PDFXOutputIntentProfileSelector /DocumentCMYK /PreserveEditing true /UntaggedCMYKHandling /LeaveUntagged /UntaggedRGBHandling /UseDocumentProfile /UseDocumentBleed false >> ]>> setdistillerparams> setpagedevice