Next Article in Journal / Special Issue
Comparison of the In Vitro Photodynamic Activity of the C1α and C1β Anomers of a Glucosylated Boron Dipyrromethene
Previous Article in Journal
pH-Dependent Coloring of Combination Effect Pigments with Anthocyanins from Brassica oleracea var. capitata F. rubra
Previous Article in Special Issue
Nanoencapsulation of Red Bell Pepper Carotenoids: Comparison of Encapsulating Agents in an Emulsion Based System
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The First Century of Successful Applications of Pinacyanol: Some Noteworthy Results

by
Serghey A. Shapovalov
Research Institute of Chemistry, V. N. Karazin Kharkiv National University, Svobody sq. 4, 61022 Kharkiv, Ukraine
Colorants 2022, 1(2), 165-192; https://doi.org/10.3390/colorants1020011
Submission received: 10 February 2022 / Revised: 7 March 2022 / Accepted: 7 April 2022 / Published: 17 April 2022
(This article belongs to the Special Issue Colorants: Ancient and Modern)

Abstract

:
The most important results obtained during the first century of pinacyanol (PIN) application in fundamental and applied research are summarized. The main usages of PIN during the century are described. The structural, protolytic, and thermodynamic properties of the dye in different systems (aqueous and aqueous–organic solutions, premicellar and micellar surfactant systems, and salt-containing solutions) have been analyzed. Quantitative characteristics of intermolecular interactions are considered. Self-association and dissimilar association of PIN with the participation of other dyes, organic ions and molecules, and metal complex compounds under 5 × 10−7–8 × 10−5 mol/L concentrations of PIN are discussed. The analyzed data indicate that at present, as before, the scientific interest in this dye is very high.

1. Introduction

The era of laboratory synthesis of dyes began in the middle of the 19th century. The first step on this path was the production of aniline as an individual substance. In 1826, the German scientist Otto Unverdorben was one of the first to try to separate a coloring substance from natural indigo. By distilling indigo with lime, an oily substance known as “crystalline” was obtained [1]. In 1843, August Wilhelm von Hofmann summarized the experiments of O. Underdorben and his followers: Friedlieb F. Runge [2], Carl J. Fritzsche [3], and Nikolay N. Zinin [4]. He set that four substances, namely “cristallin” [1], “kyanol” [2], “aniline” [3], and “benzidam” [4], are one substance, i.e., phenylamine, traditionally also called “aniline”. Aniline has become the basis for the synthesis of many dyes [5].
Perhaps the most significant year was 1856. In that year, three dyes were synthesized for the first time at once. The Polish chemist Jakub Natanson discovered that, when technical aniline is heated with dichloroethane, a bright red product is formed (“aniline red”, later called fuchsine) [6]. Eighteen-year-old English chemist William H. Perkin, on the advice of his teacher August W. Hofmann, accidently synthesized purple dye whilst trying to synthesize quinine, an anti-malarial drug [7,8]. Experimenting with coal tar, he unexpectedly synthesized a purple pigment. This pigment was later called “mauveine” (from the French word “mauve”), because its color was similar to the forest flower mallow. Finally, Charles G. Williams (University of Glasgow) succeeded in obtaining the first “ethane” (more precisely, monomethine) dye, “cyanine blue” (1,1′-diisoamyl-4,4′-cyanine iodide) [9]. The chemical structure of monomethine was not established immediately, but only half a century later, thanks to the development of chemical bonding theory and instrumental methods of investigation.
It turned out that, from a chemical point of view, “mauveine” and other synthesized dyes were not individual substances, but mixtures of homologues. However, this was not an obstacle to the industrial production of synthetic dyes, which started already in the middle of the 19th century. Polymethines have become a priority in scientific research. Initial interest in polymethines was related to the desire to expand the range of dyes with new light-sensitive substances. Over time, the need arose for a targeted selection of such compounds in which the spectral-luminescent, protolytic, and aggregate properties are successfully combined.
A common structural feature of polymethines is the presence of two heteroatom (often nitrogen) centers in the molecule [10]. One of them has a charge and is linked to the other heteroatom by a conjugated polyene chain. This chain is made up of an odd number of carbon atoms. The most prominent is pinacyanol (PIN), or quinaldine blue (Figure 1).
The IUPAC name of PIN is 1-ethyl-2-[3-(1-ethylquinolin-1-ium-2-yl)prop-2-enylidene]quinoline. More concisely, although not nomenclaturally exact, PIN is referred to in the scientific literature [11] and on the commercial market [12,13] as 1,1′-diethyl-2,2′-carbocyanine (“carbo” indicates the methyl chain of 3 carbon atoms). Less commonly, the names 1,1′-diethyl-2,2′-quinocarbocyanine or 2,2′-trimethinequinocyanine are used. Taking into account the salt-like nature, the anion (chloride, bromide, iodide, etc.) should be indicated in the name of the molecular compound. For example, “pinacyanol iodide” C25H25IN2 (2E)-1-ethyl-2-[(E)-3-(1-ethylquinolin-1-ium-2-yl)prop-2-enylidene]quinoline iodide (CAS number 605-91-4) (all IUPAC names correspond to PubChem release 2021.05.07 [12]).

2. The Beginning of PIN Research

In contrast to the date of the beginning of the era of the synthesis of dyes, it is difficult to establish the exact date of the first mention of PIN. Despite the variety of chemical and commercial names for this dye, the generic semantic name “pinacyanol” was the first. It has been in use up until the present day since its inception and is ideographically dominant in the synonymic series. The first mentions of PIN date back to the very beginning of the twentieth century.
Researchers drew attention to the usefulness of PIN in modifying photographic plates to improve color rendering and light resolution. The use of “pinacyanol” in the construction of color photographic plates is known, dating from 1902 to 1904 [14]. Around 1902, Arthur Traube and Adolf Miethe proved the applicability of isocyanine dyes for creating panchromatic emulsions, which are sensitive to the entire range of visible light. Three years later, German researcher (of Austrian nationality) Benno Homolka synthesized PIN and modified the emulsions with dye additives. [15,16]. This improved the sensitivity of the long-wave (“red”) part of the visible spectrum. However, the exact chemical structure of pinacyanol could not be set [16] (p. 972). Nevertheless, B. Gomolka’s improvements made full color separation and the creation of color photographic and cinematographic films possible. At the same time, the first color (panchromatic) films appeared. However, this technique remained inaccessible to amateur filmmakers until the 1930s. A publication by Charles Edward Kenneth Mees [17] mentions the synthesis of pinacyanol as early as 1904. Moreover, W.H. Mills and F.M. Hamer note the following [18]: “One of <…> sensitizers is a substance patented in 1905 by the Farbwerke vorm. Meister, Lucius & Brüning (Brit. Pat. 16227, 1905; D.R.-P. 172118), and sold under the name of pinacyanol”.
In 1908, the “Lowell Observatory Bulletin” published a scientific report by the American spectroscopic astronomer Vesto Melvin Slipher on the spectral study of the planets of the solar system [19] (p. 443). The report mentions experiments with “new sensitizing dyes” (pinacyanol, pinachrom, orthochrome T, pinaverdol, blue pinachrom, homokol, and dicyanin). W.M. Slifer draws attention to the use of pinacyanol as early as 1907 or even 1906 as a suitable dye in mixtures with other dyes (pinaverdol, dicyanine), but does not refer to any publications [20]. This message has not gone unheeded. Nature journal published an abstract by Percival Lowell (founder of the Astronomical Observatory) with the same title, “Spectra of the Major Planets”, in the same year [21] (p. 42). P. Lowell briefly summarizes the results of V.M. Slifer’s observations and notes the following: “Mr. V.M. Slipher succeeded in 1907 in rendering some plates (Seed, 23) sensitive far into the red. This was performed <…> in a solution of pinacyanol, pinaverdol, dicyanin, alcohol, and water. With them he photographed the spectra of all the major planets”. According to biographer William G. Hoyt, V.M. Slipher “<…> made more fundamental discoveries than any other observational astronomer of the twentieth century” [19] (p. 411). His most famous discovery was the unusual radial velocity of spiral nebulae, made in 1913/1914, based on the analysis of “red shifts” in the line spectra of emissions from spherical star clusters. In turn, this discovery was a precursor to Edwin P. Hubble’s theory of an expanding universe, proposed somewhat later in the 1930s. V.M. Slifer set the existence of gas and dust in interstellar space and proved that some nebulae only glow with reflected light from nearby stars. The famous researcher of the spectra of stars and planets was born on 11 November 1875 and did not live to see his 94th birthday, which took place three days after his death. Symbolically, more than any other astronomer, he was at the forefront of the development of instrumental spectroscopy of the starry sky at the turn of two centuries and witnessed the beginning of human exploration of space, including the first landing of compatriots on the Moon.
Perhaps the reason for using pinacyanol was V.M. Slifer’s acquaintance with the studies of L. Blythswood, W.A. Scoble, and R.J. Wallace [22,23,24] and an interesting publication by P.P. Lazareff in the Journal of the Russian Physical and Chemical Society (physical section) [25]. An English-language abstract of this article was published in the Journal of the Chemical Society, Abstracts in 1908 [26]. The Russian researcher found a correlation between the decoloration process and the optical properties of the pigments in the visible spectrum, and also set a relationship between the amount of decomposed matter and the wavelength or energy of incident light. Experiments were made with cyanin, lepidinecyanin, pinacyanol, pinaverdol, quinaldinecyanin, and pinachrome in the form of thin layers of collodion impregnated with the dye that were decolorized in the region of the absorption bands, “but in no other part of the spectrum” [26] (p. 4). P. Lazareff also proposed and described the construction of a device for photometric examination of various parts of a colored plate under the influence of rays causing discoloration of the dye. It is worth noting that the use of light-sensitive plates containing PIN has contributed to astronomical discoveries in the study of Mars [27] and comets [28,29,30].
The organic synthesis methods known at the time resulted in the production of PIN mixed with alternative products. Louis E. Wise et al. first tried to avoid impurities in the dye product [31]. They described a detailed procedure for the synthesis of fairly pure PIN, using ethiodides (i.e., compounds of the R2C(C2H5)I type) of quinaldine and quinoline as initial substances. The authors argued that, based on “<…> absorption data and the results of photosensitization experiments, further purification of the dye appears to be unnecessary”.
The suitability of PIN as a photographic (and later “cinematographic”) color sensitizer prompted an increase in production volumes and a switch from laboratory synthesis to industrial-scale production of the dye. Industrial production started in Great Britain in 1916, and in Germany somewhat earlier [31]. At the same time, PIN received the first commercial name “sensitol red” (actually “sensitive to red”; similarly, “pinaverdol” was called “sensitol green”, i.e., “sensitive to green”).
The ability to produce individual forms of dyes and, therefore, to adjust the chemical composition of composite bicolor and multicolor sensitizers has stimulated a number of original inventions, as well as basic and applied scientific research. In addition, scientific laboratories and trading companies were already drawing attention to the need for standardization and the development of quality standards for light-sensitive materials by the 1920s [32,33]. In this regard, not only the devices and designs (light plates, light filters, etc.), but also the chemical composition of the gelatin media, solvents, and developer agents are being modernized [34,35,36,37,38,39,40,41,42,43,44,45,46]. Since the late 1910s and early 1920s, PIN has been proposed as one of the “standards” of photographic materials [31,32,33,47]. It is actively used as an ingredient in panchromatic light-sensitive substrates [17,40], along with naphthacyanol and acetaminocyanol [48,49]. PIN is also being tested as a highly sensitive component of silver halide media, photographic plates, and photographic and cinematographic films [50,51,52,53,54]. Thus, conditions have developed for a more favorable and meaningful practical application of PIN among other cyanine dyes.

3. Further Progress on the Dye

Since then, hundreds of sensitizing dyes have been synthesized and studied [10,14]. Quinoline dyes are gradually giving way to other, more effective sensitizers, due to their lack of light fastness and the short life of photographic materials. However, interest in PIN has not waned.
The properties of PIN are investigated in media of a different nature: in organic solvents [55,56], in solutions with added salts [57,58,59,60,61], in premicellar and micellar systems [62,63,64,65,66,67,68,69,70,71,72,73,74], and in emulsions [75,76]. Temperature effects [55,64] and kinetics of spectral changes [69,70], thermal conductivity [77], viscosity [78], and flocculation [79,80] of the dye have been studied. Pinacynol has been proposed not only as a color indicator (marker) [81] but also as an organic reagent in methods for the qualitative and quantitative determination of mercury and silver [82,83], thorium (IV) [84], tungsten (VI) [85], and rhenium (VII) [86]. The actual Rayleigh scattering of PIN in solution is up to four times higher than the intensity of the theoretically expected values in the red part of the light absorption spectrum [87]. It was found that the adsorption of dye molecules on negatively charged colloids SiO2 or SnO2 induces the formation of H-aggregates [88]. Excitation by a 532 nm laser pulse produces a short-lived (less than 30 ps) singlet excited H-aggregate state on the SiO2 surface. In contrast, long-lived cation radicals are observed on the surface of SnO2 [88].
PIN has been systematically studied as a component of laser fluid media [89,90,91,92,93,94,95,96,97,98]. The dye increases the stability of femtosecond pulses and enhances the efficiency of frequency filtering [99]. Electron paramagnetic resonance spectroscopy revealed the unusual property of PIN as a photodynamic inducer in the processes of mitochondrial and nuclear mutations in yeast cells [100]. A solution of pinacyanol chloride is used to determine the singlet oxygen released during photo-irradiation in this case.
The properties of PIN in alcohol systems (methanol, propanol, isopropanol, and glycerol) [101,102,103], as well as in premicellar and micellar surfactant media [104,105] (including binary ones [105]), are investigated. The dye has proven to be an acceptable color marker for studying the properties of DNA-like structures [106,107,108,109,110,111], biological cells and tissues [112,113,114,115], enzymes [116], and antibiotics and medications [117,118].
Similar to other carbocyanines, PIN forms associated structures, both in solutions [119] and at heterophase interfaces [88,120]. The aggregation of pinacyanol acetate leads to the formation of chromic liquid crystalline phases. They are used as templates to produce mesoporous silica nanofibers, using a template sol–gel reaction [121]. By spectral and other instrumental methods, it has been set that PIN forms molecular films upon interaction with polymers [122] and even acquires optical activity by binding to DNA [123]. Dye interaction with organized structures is accompanied by relief spectral effects [124,125,126,127], so PIN has been successfully used to determine the critical micelle concentration (CMC) of anionic surfactants [128,129].
PIN is applied in the study of polysaccharide properties [130,131,132,133,134,135,136,137], dendrimers [138,139,140], metal-sensitive holographic sensors [141], hydrophobically modified polymers [142], oligomeric cationic surfactants [143], complex lipids and lipopolysaccharides [144,145,146,147], electron transfer mechanisms [148], complexation reactions [149,150], and supramolecular processes [151,152,153]. For example, UV–Vis spectroscopy is used to monitor the phase transition from micelles to vesicles facilitated by the gradual addition of lipid (cholesterol) to micellar solutions of cetyltrimethylammonium bromide. Pinacyanol chloride acts as a solvatochrome indicator that is sensitive to changes in medium polarity [153].
The interest in PIN has further increased with the development of semi-empirical, non-empirical, and molecular dynamics methods [154,155,156,157,158,159]. The energetics and structure of pinacyanol acetate–based supramolecular polymers are studied. It was shown that, when assembled in water, the dye molecules are capable of forming polyaminated supramolecular fibers [157].
Pinacyanol is one of the leading spectroscopic studies involving polymethines. Interactions with organic counterions [160,161,162,163,164], anionic polyelectrolytes [165,166,167], and the formation of host–guest complexes [168,169,170] are characteristic for PIN, as well as other polymethines. Such interactions significantly change the spectral properties of the systems due to the formation of ionic associates, aggregates, and molecular ensembles in solution under 5 × 10−7–8 × 10−5 mol/L of PIN. For such dye contents, electron spectroscopy is an indispensable instrumental method of investigation.
Of particular interest is the association of PIN with dyes of various classes or large molecules in aqueous solutions and in systems with the addition of organic solvents or surfactants. Water is a medium in which interacting protolytic and aggregative forms of dyes exhibit unique spectral–luminescent properties [171]. Under certain conditions, a self-associate or dissimilar associate of two dyes can function as an analytical reagent, a color marker, and a bioprobe. This feature is used in solving a number of topical problems. Examples are as follows: the possibility of controlling the formation of water-soluble bilayers and vesicles [172]; incorporation of photosensitive agents into materials containing cyclodextrins [173]; determination of the mechanism and kinetics of electrically charged phospholipid membranes [174], liposomes [175], and β-amyloid peptides [176,177,178,179,180]; search for medical and pharmaceutical drugs, including antibiotics and antitumor drugs [181,182,183]; and improvement of biological cell imaging technologies in biology and medicine [184]. For these reasons, PIN is very intensively studied in the processes of self- and dissimilar association [185,186,187,188,189,190].

4. Spectral, Protolytic, and Aggregative Properties of the Dye

The widespread use of PIN in various studies is due to its unique spectral and protolytic properties. The dye is stable in organic and mixed solutions, in contrast to aqueous solutions. The absorption spectrum of PIN iodide in methanol is shown in the Figure 2 (see References [191,192]; the databases are open access), and the properties of the bands are shown in Table 1.
The dimension of molar absorption coefficient ε is L × mol−1 × cm−1 hereinafter. The quantum yield of fluorescence is 0.001 (λmax = 622 nm, methanol, excitation wavelength is 550 nm) [193]. The λmax and εmax values are traditionally assigned to the more energetically favorable cis-isomer (PIN isomerism will be discussed later). However, relaxation energy studies (excitation of molecules by 565 nm radiation) in ethanol medium also show sub-bands of absorption apparently inherent to the trans-isomer [148]. The rather developed PIN vibrational structure is more clearly visible in the absorption spectra (three bands, Figure 2) at low temperatures [193,195].
The maxima in the absorption band correspond to transitions from the lowest vibrational level of the ground state to different vibrational levels of the excited state (S0S1 for the α-band). For the coordinates Iabsorb/ν (absorption) and Ilumines4 (luminescence), a mirror symmetry of the absorption and fluorescence spectra of the PIN alcohol solutions at 77 K is observed [195,196]. The long-wave absorption band is due to electronic transitions in the conjugated bond system between two nitrogen atoms (Figure 3).
Excitation of the PIN changes the length of the polymethine fragment [195,197]. If the intensity of the α-band maximum is taken as 1, then the second (β-band) and third (γ-band) maximums have intensities of 0.305 and 0.052 (methanol; calculated values are 0.305 and 0.047) [195].
In aqueous solutions, the single-charged PIN cation (Ct+) dominates over a fairly wide pH range of 4.5 to 9.5 (the pKa value of the double-charged cation is 3.5 [198] 3.48 ± 0.03 [199]). PIN is protonized at a pH that is lower than 2.5 [82] and hydrolyzes markedly at a pH that is higher than 10. These protolytic processes are accompanied by the disruption of the chromophore system and consequently lead to discoloration of the aqueous solution. Reducing the dielectric constant of the medium markedly reduces the pKa value (down to −1.39 when the dielectric constant varies from 78.36 (water) to 9.53 (82 % 1,4-dioxane) [199]).
In contrast to alcohol and other organic systems, the light absorption of PIN in water depends significantly on the concentration. The deviation of the optical density from the basic law of light absorption becomes noticeable even at relatively low concentrations of PIN (1 × 10−6 mol/L) [200].
Numerous experiments [187,190,194,199,200,201,202,203] show that increasing the concentration causes an increase in light absorption throughout the spectral range (Figure 4). As can be seen from curves 1–4, the intensity of the α-band with νmax = 16,670 cm−1max = 600 nm) is higher than the β-band with νmax = 18,180 cm−1max = 550 nm), while the situation is reversed on curves 5 and 6. At concentrations greater than 1 × 10−4 mol/L, the γ-band νmax = 19,610 cm−1max = 510 nm), which is characteristic of H-type self-aggregates, becomes noticeable [201]. Spectral changes in coordinates A × C−1 × l−1 vs. ν (Figure 5; normalized spectra of PIN are presented) indicate a redistribution of the intensities of the α- and β-bands [190,194,202,203]. This phenomenon has been interpreted as the conversion of the PIN monomer into a dimer (Ct22+) [190,201]. Similar band redistribution is also observed at elevated temperatures [202], and not only in water but also in media with small additions of organic solvents (e.g., in mixtures of ethanol (7.5%) with water [194] or 5% mixtures of ethylene glycol, 2-methoxyethanol, dioxane, tetrahydrofuran, ethylamine, diethylamine, triethylamine, ethanolamine, diethanolamine, and triethanolamine with water [203]). On the contrary, lowering the temperature of the solution induces the dimerization of PIN. This fact was first brought to the attention of S. Sheppard and H. Brigham back in 1944. By using mixtures of ethanol with methanol, diethyl ether, and hexane, they were able to lower the temperature of the solutions from 298 to 202 K, and even to 108 K. Under these conditions, a relative increase in the intensity of the β-band and the manifestation of the γ-band as the temperature decreases has been observed [55].
The appreciable self-association of PIN in water is due not only to the high value of the dielectric constant but also to the manifestation of hydrophobic interactions [203]. The latter are particularly characteristic of planar structures, including cyanines [160,161,162,163,164,171,185,186,187,188,189,190,195,204]. As for non-aqueous or organic solvents mixed with water, the basic law of light absorption is observed in a much wider concentration range; the absorption intensity of the α-band is greater than that of the β-band at higher concentrations of the dye. The maxima of the α- and β-bands of PIN monomer in monatomic n-alcohols, DMSO, acetone, and chloroform are bathochromically shifted with respect to the aqueous spectrum. The α-band shift increases with hydrocarbon radical lengthening and is 4.61 nm (methanol), 13.33 nm (1-octanol), and 17.52 nm (benzyl alcohol) [205]. The value of the bathochromic shift of the α-band is linearly dependent on the dielectric constant of the solvent for a number of the pure alcohols (from methanol to 1-octanol) [199,206].
The normalized absorption spectrum of the monomeric form of PIN in water has four overlapping bands (the deconvolution model is an equally flat Gaussian; the search for the positions of the maxima was carried out by using the second derivative) [190] (Figure 6).
The very narrow corridor of the optical density difference curve (not more than ± 0.018 units, C = AB) indicates the coincidence of the experimental (A) and total (B) light absorption curves and, therefore, the eligibility of the light absorption curve contour as a sum of four bands (Table 2). Spectrum B (Figure 6) has the following features: νmax, cm−1, 17,920 (558.0 nm) and 16,680 (599.5 nm); and the surface area is 1762.2 cm−1. Deconvolution calculation procedures may vary and the number of bands may be different. For example, the authors of Reference [202] point out that five bands are possible, but conclude that only three of them are significant. However, it must be considered whether there is a statistical difference between the parameters of each band and the error in the experimental determination of the light absorption intensity values (as a rule, not more than 3%), as well as the error in the calculation of the band area [207,208]. Furthermore, it makes sense to relate deconvolution results to the existence of one or another particle type (monomer, dimer, or self-aggregate) at specific dye concentrations. The consideration of these circumstances leads to the conclusion that four bands [190] are valid in the aqueous spectrum of PIN monomer. The ratio of the intensities of the maximums for the bands is as follows: IM:IIM:IIIM:IVM = 1:0.53:0.078:0.058. Meanwhile, the IVM band (νmax = 19,614 cm−1, λmax = 509.7 nm) indicates the existence of more aggregated forms of PIN, that is, H-aggregates (with νmax = 19,610 cm−1, λmax = 510 nm, and γ-band [201]). Note that the closely spaced bands IIIM and IVM have similar intensities. When the temperature rises from room temperature to 325 K [202], these bands practically merge into one, and the IM and IIM bands change insignificantly. A similar situation is observed for somewhat higher concentrations of PIN, even at normal temperatures (e.g., for 4.32 × 10−6 mol/L according to [176]); their ratio of band intensities was found to be 1:0.49:0.21. In the methanol medium [193], three bands of the general absorption spectrum contour were also identified. The ratios obtained from the experimentally measured spectrum of 1:0.305:0.052 (calculated as 1:0.305:0.047) are known. The normal-logarithmic deconvolution of the PIN spectrum (5 × 10−5 mol/L) also demonstrates three bands in the micelles of n-tetradecyltrimethylammonium bromide C14H38NBr [205].
An experimental study of the self-association of nCt22+ → Ctnn+ requires the creation of higher concentrations of the dye. However, this is hindered by the limited solubility of PIN in water. This can create an additional error in the measurement of the light absorption of concentrated PIN solutions.
The equilibrium dimerization constant 2Ct+ ⇆ Ct22+ is determined on a combination of the basic law of light absorption, the law of acting masses, the conditions of material balance, and the principle of additivity of optical density [202,203,209]:
A ( λ )   = ε D ( λ )   ×   ( C 2     1 ± 1 + 8 K D × C 8 K D ) + ε M ( λ )   × ( C 2     1 ± 1 + 8 K D × C 4 K D ) ,  
where A is the optical density at a wavelength of λ, εM(λ) and εD(λ) are the coefficients of molar light absorption of the monomer and dimer at a wavelength of λ, C is the initial molar concentration of PIN, and KD is the equilibrium dimerization constant. This ratio ignores the presence of higher-order self-associates. The nonlinear approximation of the least squares method gave the values KD = 35,002 L/mol (294.16 K; the calculated spectra of the monomer and dimer were used, and the presence of particles of higher orders of dye aggregation in the solution was taken into account) and KD = (350 ± 5) × 102 L/mol [202,209]. Spectrophotometrically, the value log β20 = 4.59 ± 0.06 has been obtained for the thermodynamic (ionic strength I→0) dimerization constant β20 = aD × aM−2, β20 = (Aλ − εM × C × l) × (εD × l − 2εM × l) × (C × εD × l − 2Aλ)−2, where aM and aD are monomer and dimer activities, and l is the thickness of the absorbing layer [190]. An increase in temperature to 325.16 K lowers KD by about five times, and the dimerization process is exothermic (the standard Gibbs energy is −25.59 kJ/mol at 294.16 K [202]).
The formation of higher-order aggregated particles under certain conditions can be observed in non-aqueous or mixed systems. One example is a mixture of “ethanol (7.5%)–water (92.5%)” at a PIN concentration of 7.5 × 10−5 mol/L. Deconvolution of the two-band spectrum (601 and 546 nm), using the PeakFit program, reveals seven constituent bands with maxima at 601, 572, 557, 546, 522, 507, and 486 nm (the deconvolution model uses the fourth derivative of the light absorption curve) [194]. The calculated values of the molar absorption coefficient are 1.92 × 105, 1.56 × 105, 2.82 × 105, and 3.765 × 105 for the maxima of the monomer (601 nm), dimer (546 nm), trimer (522 nm), and tetramer (507 nm) bands respectively. It is noteworthy that the data obtained are in satisfactory agreement with the value εmax = 1.42 × 105 for the PIN dimer located in the cavity of γ-cyclodextrin [210].
The properties of PIN in mixed aqueous–organic solutions (with 5% of organic component under 303.15 K) [203] demonstrate the importance of hydrophobic interactions as a driving force of self-association. Dye aggregates are better stabilized in a more hydrophobic environment (alkylamines), despite the lower dielectric constant of mixed solvents with respect to water. In the ethylamine–diethylamine–triethylamine series, the dielectric constant changes insignificantly (76.51…76.37…76.31), but the dimerization constant of PIN increases appreciably (1.873 × 104…2.089 × 104…2.424 × 104) [203].
The values given for the dimerization constant of PIN in water are quite similar, despite the different approaches in the calculations used. However, it should be noted that different methods often produce statistically different results. For example, it has been noted [190] that, for rhodamine B, the log β20 value takes from 2.98 ± 0.03 [211] to 3.24 ± 0.02 [212,213]. PIN is inferior to only a few carbocyanines (e.g., 3,3’-diethylthiatricarbocyanine with a constant of 2.4 × 105 [214]) in dimerization ability, surpassing safranin T (the log β20 value is 2.41 ± 0.01), methylene violet (2.66 ± 0.01), rhodamine 6G (3.35 ± 0.01), and methylene blue (3.70 ± 0.01) [215]).
The normalized absorption spectrum of the PIN dimer in water is interpreted as consisting of three bands (Figure 7 and Table 3) [190,202].
Spectrum A (Figure 7) has the following features: νmax, cm−1 = 18,380 (544.1 nm) and 16,840 (593.8 nm); and the surface area is 1712.0 cm−1. A comparison of the data in Table 2 and Table 3 shows that the α-band intensity in dimer (ID) is significantly lower than in monomer (IM), and the course of spectral changes in Figure 4 is explained by the increasing intensity of band IID and decreasing intensity of band ID. An analysis of the spectra of the monomer and dimer leads to an unobvious conclusion: even in the “monomer” PIN, the bands inherent to the dimer and H-aggregates (γ-band) are recognized [190]. This is indicated by the almost identical arrangement of the IIIM and IIID absorption bands, as well as the IVM band and the H-aggregate spectrum given in S. Sheppard’s publication [201]. The two-band spectrum of the dimer can be explained by Davydov’s splitting effect [216]. This is observed when two neighboring molecules are arranged at some (less than 90°) angle to each other [164,217]. The first of the bands is hypsochromic with respect to the monomer and is related to H-type self-aggregation (coplanar arrangement of chromophores, “one above the other”), while the bathochromic band is related to J-type self-aggregation (arrangement of chromophores in one plane, “adjacent to each other”).

5. Structural and Thermodynamic Features of the Dye

The PIN cation has a planar structure; only the ethyl groups of are not in the plane of the heterocyclic system due to the sp3-hybridization of the carbon atoms of the alkyl groups. Assuming that the aromaticity is distributed over all π–electron fragments of the cation, the rotation angle in relation to axis 1 (see Figure 8) ranges from 4.6° to 6.7° (vacuum, semi-empirical methods AM1, PM3, PM7; MOPAC, GAMESS packages [190]).
A slightly larger angle of rotation along the polymethine chain is given by calculations for the boundary resonance positions (alternating single and double bond) and calculations using non-empirical methods. For example, the rotation angle is 23.4° (the GIAO-HF/DFT method using the B3LYP/6-31G (d) basis [218]), or 8° (DFT/B3LYB, 6-31G basis [219]) and even ~36° (MM2 method and molecular orbitals calculation [202]). The transition of the cation to an excited state also entails disruption of the planar conformation [95,98,220].
The discrepancy in the semi-empirical calculations of the standard enthalpy of formation of PIN (ΔfH°) is not more than 11%. The AM1 method gives values of 1070.9 kJ/mol (MOPAC) and 1110.6 kJ/mol (GAMESS), the PM3 method gives 965.9 and 974.2 kJ/mol respectively, and the MNDO method gives 1121.7 and 1125.2 kJ/mol [190]. Taking into account the systematic error of semi-empirical calculations, we see that these ΔfH° values are very close to each other. However, there is no unambiguity with regard to the electron density distribution on heteroatoms. The AM1, MNDO, and CNDO methods on nitrogen atoms give −0.13…−0.25 (see “−0.131” in Figure 8, AM1 method), the Hartree–Fock method (HF, 6-31G basis) gives −1.13, whereas the PM3 method gives +0.31 (MOPAC, GAUSSIAN, GAMESS) [190]. Obviously, significant differences in electron densities are due to the particular parameterization of one or the other method. One must also take into account the fact that the electron density at the atom is not a measurable quantity, in contrast to, for example, the dipole moment of a molecule [218]. In addition, differences in the geometry of the molecular system may be the cause of the ambiguity of the electron density on the atoms.
The isomerization of cis- and trans-positions is possible for the PIN cation. A uniform distribution of the electron density is observed at the lowest angle of rotation of the polyene hydrocarbon chain. The electron density symmetry is broken when the heterocyclic fragments are mutually rotated. The calculation of the electron oscillation forces for a number of different conformers leads to the conclusion [190] that the cis-conformation of the PIN is more energetically stable. However, under certain conditions (changes in the dielectric constant of the medium) a trans-isomer is also possible (e.g., in ethanol [148]).
Isomerization is also possible for a dimer. In a cis-dimer, the ethyl radicals of the two cations point in the same direction. A trans-dimer is characterized by an opposite arrangement of ethyl radicals. Its cation planes are nearly coplanar, while for the cis-dimer, no predominant arrangement can be calculated. In the latter case, the total energy values of the two cations are found to be approximately the same for several different mutual arrangements of the two cations. For this reason, the trans-dimer is more energetically advantageous. The self-association of the PIN trans-dimer (Figure 9; stereo image; distance between the planes of the structure is equal to 3.7…3.8 Å; the energy is 79.1 kcal/mol, and MM+ method) leads to the experimentally observed H-aggregates [201,221].

6. Intermolecular Interactions Involving PIN

The processes of interactions involving PIN are accompanied by a rather complex combination of different forces, including non-Coulomb, π-electronic, and solvophobic effects. Below we focus on some of them, including multi-atomic organic molecules (dyes), metal complex particles, and surfactants.
A spectral feature of cation–anion interactions in solutions involving dyes is the significant change in the intensity of the PIN absorption bands. A typical example is the aqueous system of “PIN + bromocresol purple (BCP)”. The light absorption bands of HAn and An2− forms of BCP have λmax (HAn) = 430 nm and λmax (An2−) = 588 nm with εmax (HAn) = 24,900 and εmax (An2−) = 67,200. For this reason, the absorption bands of BCP and Ct+ are well spectrally spaced.
As the initial concentration of BCP increases (Figure 10, curves 27), the intensity of the short-wave part of the spectrum increases, which is related to the light absorption of the single-charged sulphonephthalein anion (hereafter, arrows indicate the direction of the spectral shift). At the same time, a decrease in the intensity of the light absorption of the α-band of PIN is noticeable. The absorption bands of BCP2− and Ct+ are mutually superimposed. However, also in this case, a decrease of the α- and β-bands intensity of PIN for the mixture “Ct+ + BCP2−” is detected if the absorption spectra of the system are measured not against water but against a solution with the same concentration of sulphonephthalein (Figure 11).
Similar effects have been observed for systems involving not only PIN [171,186,187,188,190,200] but also other polymethine dyes [162,171,185]. In all cases, the analysis of the electronic absorption spectra of mixtures of dyes reveals a violation of the additivity principle of light absorption. The optical density of the mixture is systematically lower than the total absorbance of the individual ions. This phenomenon is observed over a fairly wide range of initial concentrations of both PIN (4.9 × 10−7–4.8 × 10−5 mol/L) and other dyes. Similar cases are observed in the interaction of PIN with phenolsulfonephthalein [190,222,223], its alkyl [223,224,225] and halogen derivatives [187,226,227], oxyxanthenes [187,228], alizarines [229,230], xylenol orange [231], arsenazo II [232], and anionic rhodamine 200C [233], as well as with colorless bulky organic anions (tetraphenylborate [200]), polyelectrolytes (humic acids [232,234]). Spectral shifts in the absence of new or splitting of existing bands indicate a solvate-separated type of associative structure (in accordance with the general features of spectral shifts for associative dye systems [161,163,171]). The results of numerous studies give us reason to consider PIN as a kind of “colored” cation standard in the spectral study of supramolecular cation–anion interactions.
Both the Ct+ monomer and the Ct22+ dimer interact with organic bulk anions. However, the association of Ct+ under the influence of excess anion is not accompanied by dimerization of the Ct+ (in contrast, for example, to some cyanines, bisindocarbothiocyanines, which induce the formation of dimeric particles upon interaction with borate anions [235]) or H-aggregation (in contrast to the interaction of cyanines with DNA duplexes [236]). PIN interactions have also been experimentally detected with dissimilar associates [171,187,200]. The PIN cation is involved in the formation of more complex particles: (Ct+)j·Anj + kAnj− ⇆ (Ct+)j·(Anj)k+1. However, the nature of these interactions cannot be studied in more detail due to the aggregative instability of such systems. Nevertheless, a similar mechanism has been established by 1H NMR spectroscopy in the systems “daunomycin + ethidium bromide” [237] and “acridine orange + caffeine” [238]. Inorganic ions at the same low concentrations do not cause noticeable spectral changes.
The association of the PIN with various counterions is characterized by common spectral signatures. This includes the presence of an isosbestic point (similar to Figure 10) and areas of “saturated” optical density. The interactions of PIN or other cationic polymethines with anions are considered according to the equilibrium model: jCt+ + Anj ⇆ (Ct+)j·Anj [171,185,187,190,230,232,239]. The equilibrium association constant (Kas) can be calculated on the basis of the mass action law:
Kas = [(Ct+)j·Anj] × (CAnj − [(Ct+)j·Anj])−1 × (CCt+j × [(Ct+)j·Anj])j,
where [(Ct+)j·Anj−] = (εCt+ × CCt+ × lA) × (j × εCt+ − εas)−1 × l−1, and the equilibrium molar concentration of particles are in brackets; CAnj− is the initial molar concentration of the anion; CCt+ is the initial molar concentration of PIN; l is the thickness of the absorbing layer, cm; A is the optical density of the solution; and εCt+ and εas are the molar extinction coefficients of cation and an associate, respectively.
When the ionic strength of the solutions is low, the concentration value of Kas practically does not differ from the thermodynamic value (Table 4; the data are given according to References [187,190,222,223,224,225,226,227,228,230]). Note that the ratio of stoichiometric coefficients in dissimilar associates is determined not only by the electrostatic factor, but also by the effect of π-electron fragments of counterions. An example is the associate of the monomeric form of the PIN with the single-charged spherical tetraphenylborate-anion (TFB). Interaction occurs according to the scheme: 2Ct+ + TFB ⇆ (Ct+)2·TFB with log Kas = 13.4 ± 0.1 [200]. An indirect confirmation of this association is the fact that the calculations result in meaningless negative Kas values for 1:1 stoichiometry. Other examples are the interactions of PIN with multiply charged anions of humic acids. The log Kas values are from 1.74 ± 0.12 to 3.09 ± 0.03 depending on the composition of the polyelectrolyte [234]. Moreover, the stoichiometry is not directly related to the charges of anionic particles in PIN associates with arsenazo II [232].
The Kas values for the dye associates (Table 4) are comparable with the known values of similar values for the complexation process [161,163,171].
The rather high Kas values may indicate the formation of contact ionic associates in the solution. The dye association process is characterized by some deformation of the flattened π-electron system of the PIN (Figure 12; AM1 method; stereo image; the distance between neighboring marked in green ions is 4.6…5.0 Å, and the inner central angle with the vertex on the central carbon atom of the PIN polymethine chain equals 144°).
The equilibrium nature of dissimilar associates is experimentally confirmed by the fact of their destruction under the influence of ionic surfactant additives: in micellar systems, the absorption band contours of each dye are practically restored [171,186,233,240].
The interaction of the PIN cation, not only with anions but also with cationic dyes, was found experimentally. The effect of flattened cations of neutral red (NR) and astrafloxin (AF) on PIN is not obvious [241]. AF and NC form singly charged cationic forms in aqueous solution (λmax are 540 and 535 nm, respectively). AF predominates in the cationic form at pH 3.0–9.5, and NR at pH 1–5 (according to various estimates pKa is 6.5–7.4 of the NR cationic form; see the discussion of these values in Reference [241]). Our analysis of the spectral changes leads to the conclusion that, in addition to the disturbance of the additivity of light absorption, hyper or hypochromic effects are observed depending on the concentration of PIN. Increasing the concentration of the indocarbocyanine dye AF ambiguously alters the initial absorption of PIN. At 3.0 × 10−6 mol/L, and by increasing the AF concentration from 3.0 × 10−6 to 9.0 × 10−6 mol/L, the light absorption of PIN increases throughout the spectral range, while the position of the α- and β-bands of PIN does not change. The situation is different if the initial concentration of PIN is about 10 times higher. Increasing the AF concentration from 1.0 × 10−5 to 7.0 × 10−5 mol/L first increases the light absorption and then (at more than equimolar concentration) significantly weakens it. The dimeric form of PIN, interacting with AF, tends to form an additional absorption band with a maximum of 519 nm (19,250 cm−1).
Thus, the spectral changes observed during “cation + cation” association depend on the structure of the dyes. The appearance of a new absorption band could be due to the rather close arrangement of the flattened chromophore systems in the “contact” associate of PIN with AF. The electrostatic repulsion of the cations of the two dyes is compensated by the affinity of their developed π-electronic systems. The stronger the induced electrostatic field and the more similar the chromophores are in the nature of π-electronic systems (e.g., PIN and AF have the same polymethine fragments), the stronger the resonance interactions appear; they lead to the appearance of new absorption bands. The formation of new bands has been observed in the interaction of identically charged dye ions in non-aqueous and mixed media (e.g., for cyanine dyes in non-polar solvents [162]) and even for oppositely charged ions (interaction of rhodamine 6G with bromothymol blue [242]).
This conclusion is supported by the study of the interactions of PIN with the multi-charged anions (Figure 13) of organic reagents (arsenazo II), with polyelectrolytes [232], and with the metal complex ions [Fe(CN)6]3− and [Fe(CN)6]4−, which have a high negative charge [243,244].
It was experimentally found that, as increasing amounts of anion are added to the unchanged content of pinacyanol in aqueous solution, the α- and β-bands of PIN decrease in light absorption, similar to the cation–anion dye interactions discussed above. However, in addition, a new short-wavelength band appears (Figure 13) with λmax = 475 nm (21,050 cm−1). Remarkably, the spectral changes occur at low initial concentrations of the dye (1.0 × 10−6–4.0 × 10−6 mol/L), at which its monomeric form dominates. Changes in optical density occur at anion concentrations that are even lower than the PIN concentration.
The manifestation of the short-wave band is induced not only by interaction with multi-charged inorganic ions, but also by an increase in the ionic strength of the aqueous solution. In this case, high concentrations of electrolyte salts favor a change in the hydrate environment of the PIN cation. As the ionic strength (I) of the solution increases by increasing the concentration of potassium chloride, the intensity of the α- and β-bands decreases markedly. A new short-wave absorption band appears, as evidenced by the deconvolution of the light absorption curve (Figure 14; the deconvolution model is similar to Figure 6, I = 0.2). Table 5 details the characteristics of the normalized spectrum components.
The experimental spectrum (curve A) is characterized by absorption bands with maxima of 18,140 cm−1 (551.21 nm), I = 0.6910; and 16,660 cm−1 (600.24 nm), I = 1.000 un. The total (spectrum B of five bands) has 18,140 cm−1 (551.27 nm), I = 0.6924 un.; and 16,680 cm−1 (599.62 nm), I = 0.9863 un. There is practically no difference between the experimentally measured light absorption (A) and the calculated sum (B) of the components of the five bands, 1–5. Bands 1–4 belong to the PIN cation (as indicated above), while the new band (5) indicates the tendency of the cation to self-association due to decreasing in the distance between neighboring dye cations, and some enhancement of the resonance interactions.

7. PIN in Surfactant-Containing Systems

“Dye + surfactant” systems occupy a special place among supramolecular interactions. In fact, there is no class of dyes that has not been studied in the context of interactions with surfactants. In practical terms, these systems are widely used for the quantitative determination of pharmaceutical components, in the chemical analysis of surfactants, in the creation of polyintercalated and aggregated structures, in medical and biological research (study and modification of DNA and polypeptide structure), and in other areas. Pinacyanol is no exception in this aspect [186,199,205,230,240,245,246].
Surfactants have a significant influence on the spectral and protolytic properties of dyes in the premicellar and micellar concentration ranges. In the premicellar concentration range, compounds (cation–anionic associates) with different stoichiometry are formed between dyes and surfactants. In the micellar region, surfactants act as a microenvironment in which the dielectric constant is lower than that of water.
For electrostatic reasons, the interaction between the Ct+ and the cationic surfactant is not sufficiently appreciable. Indeed, premicellar and even micellar contents of cetylpyridinium bromide (CPB) do not change the shape of the absorption bands of PIN over a wide range of own concentrations from 1 × 10−6 to 5 × 10−5 mol/L (Figure 15); some bathochromic shift of the absorption bands [186] relative to pure water is noticeable in the micelles.
Only high concentrations (0.9 mol/L) of branched bromide n-tetradecyltrimethylammonium C17H38NBr restore the absorption band contour characteristic of PIN monomer in water [205]. Micelles of other cationic surfactants also cause bathochromic band shifts. The shift value increases with hydrocarbon radical lengthening: 12.4 nm for C15H34NBr, 13.4 nm for C17H38NBr, and 14.6 nm for C19H42NBr [199].
The effect of anionic surfactants on PIN is quite different (Figure 16). When a cation interacts with sodium dodecyl sulfate (SDS), the formation of compounds between Ct+ and the SDS anion is observed [245], and even the formation of salts (Ct+m·SDSn, where m > n [129]) is detected. If the dye content is unchanged and the concentration of SDS increases, the light absorption of the dye decreases. For a dye concentration of 2 × 10−6 mol/L, this is already noticed at an equimolar concentration of SDS. At higher surfactant concentrations (1.1 × 10−3–3.2 × 10−3 mol/L), even H-aggregates of PIN are induced, absorbing at 510 nm. However, in micellar systems with an anionic surfactant (8.5 × 10−3 mol/L or more), all compounds and aggregates are destroyed by solubilization of the dye on the surface of the homomycells. The spectrum of micellar solutions is similar to that of the PIN cation in water, but the absorption α- and β-bands are slightly batochromically shifted (λmax = 610.5 nm (16,380 cm−1), α-band; and λmax = 669–570 nm (17,560 cm−1), β-band [246]).
In micellar systems of cationic surfactants, the apparent dissociation constants of PIN were determined (pK values: 3.09 ± 0.01 (n-dodecyltrimethylammonium bromide C15H34NBr), 3.11 ± 0.01 (n-tetradecyltrimethylammonium bromide C17H38NBr), 3.17 ± 0.01 (n-hexadecyltrimethylammonium bromide C19H42NBr), 1.82 ± 0.01 (Triton TX-100)) [199]. These values of constants are somewhat lower compared to the aqueous environment. PIN interacts most strongly with anionic surfactants (the constant of association with SDS under standard conditions is 1117.45 × 103 L/mol), and associations with cationic ones are much weaker (for example, 10.77 × 103 L/mol with cetyltrimethylammonium bromide); the non-ionic surfactant Triton TX-100 naturally occupies an intermediate position (with a constant of 418.99 × 103 L/mol) [205]. Such spectral effects are not only characteristic of PIN; they are manifested in micellar [247,248,249,250,251], premicellar [252,253], and mixed systems of ionic and nonionic surfactants [254,255,256] and often have a kinetic character [71,126,240,249,252].
Similar spectral shifts are manifested in the interaction of PIN with heparin [257], calixarenes [258], and cyclodextrins [210]; in “dye + alginates” systems [259]; and in dye binding by nucleic acids [260]. These and several other publications show that PIN is used in spectroscopic studies not only in homogeneous, microheterogeneous, and ultramicroheterogeneous systems, but also in studies of supramolecular schemes and mechanisms, or as a marker. For example, the interaction of PIN with nucleic acids [260] is characterized by noticeable hypochromic effects and frequency shifts of the dye absorption bands. The analysis of spectra, DNA melting point, and electrophoresis indicated that the binding of PIN to nucleic acids occurs via an external stack in which nucleic acids serve as matrices. The use of PIN has made it possible to define a non-intercalating mechanism of dye binding to DNA and also to determine that the interaction between PIN and DNA is not only electrostatic.

8. Conclusions

As follows from what was discussed above, the use of PIN is effective both in respect to the identification of spectral, equilibrium, thermodynamic regularities, taking into account specific or non-specific interactions in homogeneous or heterogeneous systems, and in connection with the promising practical use of these systems. The unique spectral, thermodynamic, aggregative properties of polymethine dyes are fully attributed to PIN. That is why the use of this dye in scientific and technical research has received continuous attention for many years.
PIN was initially the standard in a range of sensitizing materials, but now the results of numerous studies give us reason to consider PIN as a kind of “colored” cation standard in the spectral study of supramolecular interactions.
For more than a century, only in sources indexed by leading services (WebSci, Science Direct, PubMed, Google Scholar, eLIBRARY, and Index Copernicus, Scopus), more than 1000 scientific works have been published on “pinacyanol” (or “quinaldine blue”), including 93.3% of original articles, 2.9% of reviews, 2.0% of monographs, brief reviews, messages, and abstracts. Furthermore, more than 900 patents for inventions have also been issued. It is noteworthy that a leading number of studies are in the fields of physical, colloidal, supramolecular chemistry, and molecular spectroscopy. The increasing percentage of publications in recent years is related to biology, pharmacy, medicine, and biochemical engineering. All of this suggests that pinacyanol is successfully entering the second century of fundamental research and applications.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Unverdorben, O. Ueber das Verhalten der organischen Körper in höheren Temperaturen. Ann. Phys. Chem. 1826, 84, 397–410. [Google Scholar] [CrossRef] [Green Version]
  2. Runge, F.F. LXXV. Ueber einige Produkte der Steinkohlendestillation. Ann. Phys. Chem. 1834, 31, 513–524. [Google Scholar] [CrossRef]
  3. Fritzsche, J. Ueber das Anilin, ein neues Zersetzungsproduct des Indigo. Justus Liebigs Ann. Chem. 1840, 36, 84–90. [Google Scholar] [CrossRef] [Green Version]
  4. Zinin, N. Beschreibung einiger neuer organischer Basen, dargestellt durch die Einwirkung des Schwefelwasserstoffes auf Verbindungen der Kohlenwasserstoffe mit Untersalpetersäure. J. Prakt. Chem. 1842, 27, 140–153. [Google Scholar] [CrossRef] [Green Version]
  5. Hofmann, A. Chemische Untersuchung der organischen Basen im Steinkohlen-Theeröl. Ann. Chem. Pharm. 1843, 47, 37–87. [Google Scholar] [CrossRef]
  6. Natanson, J. Ueber Acetylamin und seine Derivate. J. Prakt. Chem. 1856, 67, 242–247. [Google Scholar] [CrossRef] [Green Version]
  7. Hübner, K. 150 Jahre Mauvein. Chem. Zeit 2006, 40, 274–275. [Google Scholar] [CrossRef]
  8. Travis, A.S. Perkin’s mauve: Ancestor of the organic chemical industry. Technol. Cult. 1990, 31, 51–82. [Google Scholar] [CrossRef]
  9. Williams, C. XXVI.—Researches on chinoline and its homologues. Trans. R. Soc. Edinb. 1856, 21, 377–401. [Google Scholar] [CrossRef] [Green Version]
  10. Zollinger, H. Color Chemistry: Syntheses, Properties, and Applications of Organic Dyes and Pigments; John Wiley & Sons: Zürich, Switzerland, 2003; 637p, ISBN 978-3-906-39023-9. [Google Scholar]
  11. Mishra, A.; Behera, R.; Behera, P.; Mishra, B.; Behera, G. Cyanines during the 1990s: A review. Chem. Rev. 2000, 100, 1973–2012. [Google Scholar] [CrossRef]
  12. PubChem Open Chemistry Database. Available online: https://pubchem.ncbi.nlm.nih.gov/#query=Pinacyanol&tab=compound (accessed on 27 January 2022).
  13. Sigma Trade Catalog. Available online: http://www.sigmaaldrich.com/catalog/search/ (accessed on 27 January 2022).
  14. Joule, J.; Mills, K. Heterocyclic Chemistry; Wiley-Blackwell: New York, NY, USA, 2010; 720p, ISBN 978-1-405-13300-5. [Google Scholar]
  15. Mitchell, E. Photographic Science; Wiley: New York, NY, USA, 1984; 404p, ISBN 978-0-471-09046-5. [Google Scholar]
  16. Mees, C.E.K. The Theory of Photographic Process; The Macmillan Company: New York, NY, USA, 1942; 1124p, Available online: https://openlibrary.org/works/OL1156030W/The_theory_of_the_photographic_process (accessed on 13 February 2020).
  17. Mees, C.E.K. The color sensitivity of photographic materials. J. Frankl. Inst. 1926, 201, 525–551. [Google Scholar] [CrossRef]
  18. Mills, W.; Hamer, F. CLXXI.—The cyanine dyes. Part III. The constitution of pinacyanol. J. Chem. Soc. Trans. 1920, 117, 1550–1562. [Google Scholar] [CrossRef] [Green Version]
  19. Hoyt, W. Biographical Memoirs (Vesto Melvin Slipher); National Academy of Sciences: New York, NY, USA, 1980; 495p, ISBN 978-0-309-03099-1. [Google Scholar]
  20. Slipher, V. The spectra of the major planets. Lowell Observ. Bull. 1908, 42, 231–238. [Google Scholar]
  21. Lowell, P. The spectra of the major planets. Nature 1908, 79, 42. [Google Scholar] [CrossRef] [Green Version]
  22. Blythswood, L.; Scoble, W. Photographs of certain ARC spectra from Å 5800 to Å 8300. Astrophys. J. 1906, 24, 125–127. [Google Scholar] [CrossRef]
  23. Wallace, R. Color filters for astronomical photography with reflecting telescopes. Astrophys. J. 1906, 24, 268–277. [Google Scholar] [CrossRef]
  24. Wallace, R. Orthochromatism by bathing. Astrophys. J. 1907, 26, 299–325. [Google Scholar] [CrossRef]
  25. Lazareff, P. Decoloration of pigments in the visible spectrum. J. Russ. Phys. Chem. Soc. Phys. Part 1907, 39, 236–246. [Google Scholar]
  26. General and physical chemistry. J. Chem. Soc. Abstr. 1908, 94, B1–B31. [CrossRef] [Green Version]
  27. Slipher, V. The spectrum of Mars. Astrophys. J. 1908, 28, 397–410. [Google Scholar] [CrossRef]
  28. De la Baume Pluvinel, A.; Baldet, F. Spectrum of comet morehouse (1908 c). Astrophys. J. 1911, 24, 89–111. [Google Scholar] [CrossRef]
  29. Wallace, R. The function of a color-filter and “isochromatic” plate in astronomical photography. Astrophys. J. 1908, 27, 106–124. [Google Scholar] [CrossRef]
  30. Ives, H. Experimental study of the Lipmann color photograph. Astrophys. J. 1908, 27, 325–352. [Google Scholar] [CrossRef]
  31. Wise, L.; Adams, E.; Stewart, J.; Lund, C. Synthesis of photosensitizing dyes: Pinaverdol and pinacyanol. J. Ind. Eng. Chem. 1919, 11, 460–463. [Google Scholar] [CrossRef]
  32. Hitchins, A. Maintaining photographic standards. J. Frankl. Inst. 1917, 184, 179–203. [Google Scholar] [CrossRef] [Green Version]
  33. Hitchins, A. Modern applications of photography. J. Frankl. Inst. 1919, 187, 129–146. [Google Scholar] [CrossRef]
  34. Sheppard, S.; Elliott, F. The reticulation of gelatine. J. Ind. Eng. Chem. 1918, 10, 727–732. [Google Scholar] [CrossRef]
  35. Sheppard, S.; Meyer, G. Chemical induction in photographic development. I. Induction and the Watkins factor. J. Am. Chem. Soc. 1920, 42, 689–700. [Google Scholar] [CrossRef]
  36. Sheppard, S.; Elliott, F. Photometric methods and apparatus for the study of colloids. J. Am. Chem. Soc. 1921, 43, 531–539. [Google Scholar] [CrossRef] [Green Version]
  37. Sheppard, S.; Sweet, S. The setting and melting points of gelatins. J. Ind. Eng. Chem. 1921, 13, 423–426. [Google Scholar] [CrossRef]
  38. Wightman, E.; Sheppard, S. The size-frequency distribution of particles of silver halide in photographic emulsions and its relation to sensitometric characteristics. J. Phys. Chem. 1921, 25, 181–195. [Google Scholar] [CrossRef] [Green Version]
  39. Sheppard, S.; Elliott, F. The drying and swelling of gelatin. Preliminary note. J. Am. Chem. Soc. 1922, 44, 373–379. [Google Scholar] [CrossRef] [Green Version]
  40. Sheppard, S. Recent contributions of chemistry to photography. J. Ind. Eng. Chem. 1922, 14, 820–822. [Google Scholar] [CrossRef] [Green Version]
  41. Wightman, E.; Trivelli, A.; Sheppard, S. The distribution of sensitivity and size of grain in photographic emulsions. J. Frankl. Inst. 1922, 194, 485–520. [Google Scholar] [CrossRef] [Green Version]
  42. Sheppard, S. Plasticity in relation to gelatin. J. Frankl. Inst. 1925, 200, 815. [Google Scholar] [CrossRef]
  43. Sheppard, S. Photographic sensitivity: A colloid chemical problem. J. Frankl. Inst. 1926, 201, 503–504. [Google Scholar] [CrossRef]
  44. Sheppard, S. Behavior of gelatin in the processing of motion picture film. J. Frankl. Inst. 1928, 205, 571–572. [Google Scholar] [CrossRef]
  45. Sheppard, S.; Trivelli, A. Comparison of some developers for sensitometric standardization. J. Frankl. Inst. 1929, 208, 434. [Google Scholar] [CrossRef]
  46. Sheppard, S. The function of gelatin in photographic emulsions. J. Frankl. Inst. 1929, 208, 559. [Google Scholar] [CrossRef]
  47. Wise, L.; Adams, E. Photographic sensitizing dyes: Their synthesis and absorption spectra. J. Ind. Eng. Chem. 1918, 10, 801–802. [Google Scholar] [CrossRef] [Green Version]
  48. Mees, C.; Gutekunst, G. Some new sensitizers for the deep red. J. Frankl. Inst. 1922, 194, 549. [Google Scholar] [CrossRef]
  49. Mees, C.; Gutekunst, G. Some new sensitizers for the deep red. J. Ind. Eng. Chem. 1922, 14, 1060. [Google Scholar] [CrossRef]
  50. Sheppard, S. Antagonism of radiations in photochemical and photographic reactions. Chem. Rev. 1928, 4, 319–353. [Google Scholar] [CrossRef]
  51. Sheppard, S.; Lambert, R.; Keenan, R. The adsorption of organic materials to the silver halides. J. Phys. Chem. 1932, 36, 174–184. [Google Scholar] [CrossRef]
  52. Bancroft, W.; Ackerman, J.; Gallagher, C. Optical sensitization in photography. J. Phys. Chem. 1932, 36, 154–173. [Google Scholar] [CrossRef]
  53. Doja, M. The cyanine dyes. Chem. Rev. 1932, 11, 273–321. [Google Scholar] [CrossRef]
  54. Kornfeld, G. The action of optical sensitizers on the photographic plate. J. Phys. Chem. 1938, 42, 795–813. [Google Scholar] [CrossRef]
  55. Sheppard, S.; Brigham, H. Some effects of solvents upon the absorption spectra of dyes. III. Temperature and organic solutions of cyanine dyes. J. Am. Chem. Soc. 1944, 66, 380–384. [Google Scholar] [CrossRef]
  56. Sheppard, S.; Geddes, A. Effect of solvents upon the absorption spectra of dyes. IV. Water as solvent: A common pattern. J. Am. Chem. Soc. 1944, 66, 1995–2002. [Google Scholar] [CrossRef]
  57. Klotz, I. The effects of salts and proteins on the spectra of some dyes and indicators. Chem. Rev. 1947, 41, 373–399. [Google Scholar] [CrossRef]
  58. Corrin, M.; Harkins, W. The effect of salts on the critical concentration for the formation of micelles in colloidal electrolytes. J. Am. Chem. Soc. 1947, 69, 683–688. [Google Scholar] [CrossRef] [PubMed]
  59. Merrill, R.; Spencer, R.; Getty, R. The effect of sodium silicates on the absorption spectra of some dyes. J. Am. Chem. Soc. 1948, 70, 2460–2464. [Google Scholar] [CrossRef]
  60. Merrill, R.; Spencer, R. Spectral changes of pinacyanol chloride in sodium silicate–salt solutions. J. Am. Chem. Soc. 1950, 72, 2894–2899. [Google Scholar] [CrossRef]
  61. Herzfeld, S. Effect of salt on the critical concentrations of potassium n-alkanecarboxylates as determined by the change in color of pinacyanole. J. Phys. Chem. 1952, 56, 959–963. [Google Scholar] [CrossRef]
  62. Corrin, M.; Harkins, W. Critical concentrations for micelle formation in mixtures of anionic soaps. J. Coll. Sci. 1946, 1, 469–472. [Google Scholar] [CrossRef]
  63. Corrin, M.; Harkins, W. Determination of the critical concentration for micelle formation in solutions of colloidal electrolytes by the spectral change of a dye. J. Am. Chem. Soc. 1947, 69, 679–683. [Google Scholar] [CrossRef]
  64. Klevens, H. The effects of temperature upon the critical concentrations of anionic and cationic detergents. J. Phys. Coll. Chem. 1947, 51, 1143–1154. [Google Scholar] [CrossRef]
  65. Harkins, W.; Mittelmann, R.; Corrin, M. Types of solubilization in solutions of long-chain colloidal electrolytes. J. Phys. Coll. Chem. 1949, 53, 1350–1361. [Google Scholar] [CrossRef]
  66. Klevens, H. The effect of added hydrocarbons upon critical concentrations of soap and detergent solutions. J. Phys. Coll. Chem. 1950, 54, 1012–1016. [Google Scholar] [CrossRef]
  67. Mukerjee, P. Use of ionic dyes in analysis of ionic surfactants and other ionic organic compounds. Anal. Chem. 1956, 28, 870–873. [Google Scholar] [CrossRef]
  68. Mukerjee, P. The nature of the association equilibria and hydrophobic bonding in aqueous solutions of association colloids. Adv. Coll. Interface Sci. 1967, 1, 242–275. [Google Scholar] [CrossRef]
  69. Yasunaga, T.; Takeda, K.; Harada, S. Kinetic study of sodium dodecyl sulfate micelle dissociation by a stopped-flow method. J. Coll. Interface Sci. 1973, 42, 457–463. [Google Scholar] [CrossRef]
  70. Takeda, K.; Tatsumoto, N.; Yasunaga, T. Kinetic study on solubilization of pinacyanol chloride into the micelle of sodium dodecyl sulfate by a stopped-flow technique. J. Coll. Interface Sci. 1974, 47, 128–133. [Google Scholar] [CrossRef]
  71. Reeves, R. Nature of mixed micelles from anionic dyes and cationic surfactants. Kinetic study. J. Am. Chem. Soc. 1975, 97, 6019–6024. [Google Scholar] [CrossRef]
  72. Timbers, A.; Lingafelter, E. Irreversible decolorization of pinacyanol chloride in the presence of paraffin-chain salts. J. Am. Chem. Soc. 1949, 71, 4155–4156. [Google Scholar] [CrossRef]
  73. Wu, J.; Harwell, J.; O’Rear, E. Two-dimensional reaction solvents: Surfactant bilayers in the formation of ultrathin films. Langmuir 1987, 3, 531–537. [Google Scholar] [CrossRef]
  74. Grieser, F.; Drummond, C. The physicochemical properties of self-assembled surfactant aggregates as determined by some molecular spectroscopic probe techniques. J. Phys. Chem. 1988, 92, 5580–5593. [Google Scholar] [CrossRef]
  75. West, W.; Carroll, B.; Whitcomb, D. The adsorption of sensitizing dyes in photographic emulsions. J. Phys. Chem. 1952, 56, 1054–1067. [Google Scholar] [CrossRef]
  76. Herz, A. Aggregation of sensitizing dyes in solution and their adsorption onto silver halides. Adv. Coll. Interface Sci. 1977, 8, 237–298. [Google Scholar] [CrossRef]
  77. Lutsenko, V.; Lutsenko, E.; Rosenshtein, L. Thermostimulated conductivity in organic dyes (pinacyanol). Phys. Status Sol. A 1970, 2, K165–K168. [Google Scholar] [CrossRef]
  78. Eske, A.; Razi Naqvi, K. Viscosity dependence of the fluorescence lifetimes of cryptocyanine, pinacyanol and DDI. Chem. Phys. Lett. 1979, 63, 128–132. [Google Scholar] [CrossRef]
  79. McGarry, M.; Hazel, J. Flocculation of pinacyanol chloride in alkali silicate solutions. J. Coll. Sci. 1964, 19, 294–301. [Google Scholar] [CrossRef]
  80. Kainz, G. Mit der spektralphotometrischen Bestimmung der Farbstoffe Pinacyanol und Pinakryptolgrün. Fresenius’ Z. Anal. Chem. 1962, 187, 225. [Google Scholar] [CrossRef]
  81. McQuillen, K.I. Effect of cetyl-trimethyl-ammonium bromide on the electrophoretic mobility of certain gram-positive bacteria. Biochim. Biophys. Acta 1950, 5, 463–471. [Google Scholar] [CrossRef]
  82. Babko, A.; Vasilieva, E. Complex compounds of pinacyanol with mercury. (For use in the analysis of complexes of cyanine dyes with metals). J. Anal. Chem. 1947, 2, 159–166. (In Russian) [Google Scholar]
  83. Vasilieva, E. Qualitative discovery of silver and mercury ions and their quantitative (colorimetric) determination. J. Anal. Chem. 1947, 2, 167–172. (In Russian) [Google Scholar]
  84. Babko, A.; Turov, D.; Turov, P. Polymolybdate of pinacyanol. Ukr. Chem. J. 1969, 35, 953–957. (In Russian) [Google Scholar]
  85. Turov, P.; Turov, D. Microtitrimetric determination of tungsten with pinacyanol. J. Anal. Chem. 1970, 25, 934–939. (In Russian) [Google Scholar]
  86. Guseinov, I.; Rustamov, N.; Bagbagly, S. Pinacyanol as a reagent for the extraction-photometric determination of rhenium (VII). Azerb. Chem. J. 1976, 4, 138–142. (In Russian) [Google Scholar]
  87. Anglister, J.; Steinberg, I. Resonance Rayleigh scattering of cyanine dyes in solution. J. Chem. Phys. 1983, 78, 5358–5368. [Google Scholar] [CrossRef]
  88. Barazzouk, S.; Lee, H.; Hotchandani, S.; Kamat, P. Photosensitization aspects of pinacyanol H-aggregates. Charge injection from singlet and triplet excited states into SnO2 nanocrystallites. J. Phys. Chem. B 2000, 104, 3616–3623. [Google Scholar] [CrossRef]
  89. Sundstroem, V.; Gillbro, T. Viscosity-dependent isomerization yields of some cyanine dyes. A picosecond laser spectroscopy study. J. Phys. Chem. 1982, 86, 1788–1794. [Google Scholar] [CrossRef]
  90. Gillbro, T.; Sundstrom, V. Picosecond spectroscopy using continuous analyzing light. Electronic relaxation in pinacyanol. Chem. Phys. Lett. 1980, 74, 188–192. [Google Scholar] [CrossRef]
  91. Arvis, M.; Mialocq, J. Flash photolysis of cyanine dyes. Pinacyanol chloride (1,1′-diethyl-2,2′-carbocyanine chloride). J. Chem. Soc. Faraday Trans. 2 Mol. Chem. Phys. 1979, 75, 415–421. [Google Scholar] [CrossRef]
  92. Fassler, D.; Feller, K.-H. Picosecond spectroscopy of polymethine dyes. J. Mol. Struct. 1988, 173, 377–387. [Google Scholar] [CrossRef]
  93. Werncke, W.; Tschol, T.; Weigmann, H.; Pfeiffer, M.; Lau, A.; Rentsch, S.; Graness, A. Photoisomerization studies of pinacyanol using nanosecond time-resolved resonance cars. J. Raman Spectr. 1987, 18, 323–326. [Google Scholar] [CrossRef]
  94. Stiel, H.; Teuchner, K.; Dähne, S. Photophysikalische Parameter des Pinacyanols aus nichtlinearen Absorptionsmessungen mit Nanosekundenimpulsen. Z. Chem. 2010, 25, 264–265. [Google Scholar] [CrossRef]
  95. Kaschke, M.; Kleinschmidt, J.; Wilhelmi, B. Calculation of the viscosity dependence of the excited-state lifetime in pinacyanol solution. Chem. Phys. Lett. 1984, 106, 428–430. [Google Scholar] [CrossRef]
  96. Sundstrom, V.; Gillbro, T. Transient absorption spectra of pinacyanol and cyanine photoisomers obtained with a sync-pumped picosecond dye laser and independently tunable probe light. Chem. Phys. Lett. 1983, 94, 580–584. [Google Scholar] [CrossRef]
  97. Thistlethwaite, P.; Griesser, H. The temperature dependence of the fluorescence lifetime of pinacyanol iodide. Chem. Phys. Lett. 1982, 91, 58–62. [Google Scholar] [CrossRef]
  98. Rentsch, S.; Danielius, R.; Gadonas, R. Ultrafast photophysical processes in pinacyanol after selective excitation. Chem. Phys. 1981, 59, 119–123. [Google Scholar] [CrossRef]
  99. Hébert, P.; Marguet, S.; Gustavsson, T.; Mialocq, J. Pinacyanol chloride as a saturable absorber in rhodamine B and 6G hybridly mode-locked dye lasers. Opt. Commun. 1992, 90, 85–88. [Google Scholar] [CrossRef]
  100. Iwamoto, Y.; Murakami, K.; Danno, M.; Tsuchiya, M.; Masuzawa, T.; Shimizu, T.; Morita, T.; Yanagihara, Y. Singlet oxygen production and photobiological effects of pinacyanol chloride on yeast Saccharomyces cerevisiae. J. Pharm.-Dynam. 1990, 13, 316–320. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  101. Mialocq, J.; Jaraudias, J.; Goujon, P. Picosecond spectroscopy of pinacyanol (1,1′-diethyl-2,2′-monocarbocyanine chloride). Chem. Phys. Lett. 1977, 47, 123–126. [Google Scholar] [CrossRef]
  102. Mialocq, J. Picosecond study of pinacyanol photophysics. Chem. Phys. 1982, 73, 107–115. [Google Scholar] [CrossRef]
  103. Heisel, F.; Miehé, J.; Rachidi, J. The dynamics of the intramolecular and orientational motions of pinacyanol. Chem. Phys. Lett. 1987, 134, 379–384. [Google Scholar] [CrossRef]
  104. Sato, H.; Kawasaki, M.; Kasatani, K. Fluorescence and energy transfer of dye-detergent systems in the premicellar region. J. Photochem. 1981, 17, 243–248. [Google Scholar] [CrossRef]
  105. Manna, K.; Panda, A. Spectral studies on the interaction of pinacyanol chloride with binary surfactants in aqueous medium. Spectrochim. Acta A Mol. Biomol. Spectr. 2009, 74, 1268–1274. [Google Scholar] [CrossRef]
  106. Kodama, M.; Tagashira, Y.; Nagata, C. The interaction of pinacyanol with nucleic acids. Biochim. Biophys. Acta 1966, 129, 638–640. [Google Scholar] [CrossRef]
  107. Kodama, M.; Tagashira, Y.; Nagata, C. The interaction of pinacyanol and quinoline-1-oxides with nucleic acids: The base specificity of binding. J. Biochem. 1968, 64, 167–170. [Google Scholar] [CrossRef]
  108. Klinger, H.; Hammond, D. Rapid chromosome and sex-chromatin staining with pinacyanol. Stain Technol. 1971, 46, 43–47. [Google Scholar] [CrossRef] [PubMed]
  109. Kamimura, A. Colorimetric determination of long-chain n-acylglutamic acids with pinacyanol. Agric. Biol. Chem. 1973, 37, 457–464. [Google Scholar] [CrossRef]
  110. Pal, M.; Mandel, M. Differentiation between the α-helical and coillike conformations of poly(L-glutamic acid) in aqueous solution through binding of pinacyanol. Biopolymers 1979, 18, 2267–2277. [Google Scholar] [CrossRef]
  111. Narayan, R. A technique for C-banding chromosomes with pinacyanol chloride. Stain Technol. 1980, 55, 9–11. [Google Scholar] [CrossRef]
  112. Watanabe, M.; Williams, C. Mitochondria in the flight muscles of insects. J. Gen. Physiol. 1951, 34, 675–689. [Google Scholar] [CrossRef] [Green Version]
  113. Bensley, S. Pinacyanol erythrosinate as a stain for mast cells. Stain Technol. 1952, 27, 269–273. [Google Scholar] [CrossRef]
  114. Tanaka, Y. Effect of the ruby laser microbeam on mitochondria of KB cells supravitally stained by pinacyanol. J. Cell Biol. 1969, 41, 424–430. [Google Scholar] [CrossRef]
  115. Hellberg, H.; Bjerkås, I.; Vågnes, Ø.; Noga, E. Mast cells in common wolffish Anarhichas lupus L.: Ontogeny, distribution and association with lymphatic vessels. Fish Shellfish Immunol. 2013, 35, 1769–1778. [Google Scholar] [CrossRef]
  116. Bucholtz, M.; Light, R. Acetylation of 13-sophorosyloxydocosanoic acid by an acetyltransferase purified from Candida bogoriensis. J. Biol. Chem. 1976, 251, 424–430. [Google Scholar] [CrossRef]
  117. Dufrenoy, J. Emploi du pinacyanol pour l’étude du mécanisme d’action d’antibiotiques. Ann. Inst. Pasteur 1953, 85, 396–397. [Google Scholar]
  118. Sugimura, T.; Okabe, K.; Kodama, M. Induction of respiration-deficient mutant of Saccharomyces cerevisiae by pinacyanol. J. Bacteriol. 1969, 97, 964–965. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  119. Shklyarevskiy, I.; Boamfa, M.; Christianen, P.; Touhari, F.; van Kempen, H.; Deroover, G.; Callant, P.; Maan, J. Magnetic field induced alignment of cyanine dye J-aggregates. J. Chem. Phys. 2002, 116, 8407–8410. [Google Scholar] [CrossRef]
  120. Dan, P.; Willner, I.; Dixit, N.; Mackay, R. Effects of colloidal SiO2 and cyclodextrins on the aggregation of dyes in aqueous solutions. J. Chem. Soc. Perkin Trans. 1984, 2, 455–459. [Google Scholar] [CrossRef]
  121. Rodríguez-Abreu, C.; Torres, C.; Tiddy, G. Chromonic liquid crystalline phases of pinacyanol acetate: Characterization and use as templates for the preparation of mesoporous silica nanofibers. Langmuir 2011, 27, 3067–3073. [Google Scholar] [CrossRef] [PubMed]
  122. Kobayashi, M.; Yuan, X.; Sakai, H.; Hoshi, T.; Okubo, J.; Tanizaki, Y. Preliminary analysis of the electronic absorption spectrum of a solute in a biaxially-stretched polymer film. Bull. Chem. Soc. Jpn. 1992, 65, 463–470. [Google Scholar] [CrossRef]
  123. Permogorov, V.; Lazurkin, Y. Study of anomalous dispersion of optical activity of complexes of desoxyribonucleic acid of phage T2 with pinacyanol. Polym. Sci. 1966, 8, 1238–1247. [Google Scholar] [CrossRef]
  124. Nunn, C.; Schechter, R.; Wade, W. Visual evidence regarding the nature of hemimicelles through surface solubilization of pinacyanol chloride. J. Phys. Chem. 1982, 86, 3271–3272. [Google Scholar] [CrossRef]
  125. Sato, H.; Kawasaki, M.; Kasatani, K. Energy transfer between rhodamine 6G and pinacyanol enhanced with sodium dodecyl sulfate in the premicellar region. Formation of dye-rich induced micelles. J. Phys. Chem. 1983, 87, 3759–3769. [Google Scholar] [CrossRef]
  126. Takeda, K.; Ohtawa, T.; Hachiya, K.; Aoki, K. Kinetics of the solubilization of pinacyanol chloride into a complex between bovine serum albumin and sodium dodecyl sulfate. Bull. Chem. Soc. Jpn. 1987, 60, 1174–1176. [Google Scholar] [CrossRef] [Green Version]
  127. Pal, M.; Pal, P. Dichroic probe of different modes of aggregation of pinacyanol by bis(2-ethylhexyl)sulfosuccinate below and above its critical micelle concentration. J. Phys. Chem. 1990, 94, 2557–2559. [Google Scholar] [CrossRef]
  128. Corrin, M.; Klevens, H.; Harkins, W. The determination of critical concentrations for the formation of soap micelles by the spectral behavior of pinacyanol chloride. J. Chem. Phys. 1946, 14, 480–486. [Google Scholar] [CrossRef]
  129. Mukerjee, P.; Mysels, K. A re-evaluation of the spectral change method of determining critical micelle concentration. J. Am. Chem. Soc. 1955, 77, 2937–2943. [Google Scholar] [CrossRef]
  130. Chakrabarti, A.; Nath, R.; Chakraborty, A. Induction of metachromasy in cationic dye by bacterial cell surface polysaccharide from Klebsiella K10. Spectrochim. Acta Part A Mol. Spectr. 1989, 45, 981–982. [Google Scholar] [CrossRef]
  131. Chakrabarti, A.; Nath, R.; Chakraborty, A. Chromotropic character of bacterial acidic polysaccharides: Part II–Induction of metachromasy in cationic dye pinacyanol chloride by Klebsiella K10 capsular polysaccharide. Ind. J. Biochem. Biophys. 1989, 26, 74–79. [Google Scholar]
  132. Nath, R.; Singh, T.; Dasgupta, S.; Mitra, A.; Panda, A. Interaction of cationic dye/surfactants with Klebsiella K18 capsular polysaccharides: Physico-chemical studies. Mater. Sci. Eng. C 2010, 30, 549–554. [Google Scholar] [CrossRef]
  133. Panda, A.; Chakraborty, A. Studies on the interaction of bacterial lipopolysaccharide with cationic dyes by absorbance and fluorescence spectroscopy. J. Photochem. Photobiol. A Chem. 1997, 111, 157–162. [Google Scholar] [CrossRef]
  134. Mitra, A.; Nath, R.; Chakraborty, A. Studies on induction of metachromasy in cationic dye pinacyanol chloride by Klebsiella K7 capsular polysaccharide. Ind. J. Biochem. Biophys. 1992, 29, 411–414. [Google Scholar]
  135. Mitra, A.; Chakraborty, A. Absorption and fluorescence studies on interaction between cationic dyes and Klebsiella K7 capsular polysaccharide. Ind. J. Biochem. Biophys. 1992, 29, 291–295. [Google Scholar]
  136. Mitra, A.; Chakrabarti, A.; Nath, R.; Chakraborty, A. Chromotropic character of bacterial acidic polysaccharides: Part III-Interaction of cationic dye pinacyanol chloride with Klebsiella K15 capsular polysaccharide. Ind. J. Biochem. Biophys. 1990, 27, 291–294. [Google Scholar]
  137. Mitra, A.; Chakraborty, A. Spectrophotometric and spectrofluorometric studies on interaction of cationic dyes with bacterial capsular polysaccharide. Ind. J. Biochem. Biophys. 1998, 35, 241–246. [Google Scholar]
  138. Teobaldi, G.; Zerbetto, F. Molecular dynamics of a dendrimer-dye guest-host system. J. Am. Chem. Soc. 2003, 125, 7388–7393. [Google Scholar] [CrossRef] [PubMed]
  139. Gopidas, K.; Whitesell, J.; Fox, M. Metal-core-organic shell dendrimers as unimolecular micelles. J. Am. Chem. Soc. 2003, 125, 14168–14180. [Google Scholar] [CrossRef] [PubMed]
  140. Aulenta, F.; Hayes, W.; Rannard, S. Dendrimers: A new class of nanoscopic containers and delivery devices. Eur. Polym. J. 2003, 39, 1741–1771. [Google Scholar] [CrossRef]
  141. Mayes, A.; Blyth, J.; Millington, R.; Lowe, C. Metal ion-sensitive holographic sensors. Anal. Chem. 2002, 74, 3649–3657. [Google Scholar] [CrossRef] [PubMed]
  142. Hait, S.; Majhi, P.; Blume, A.; Moulik, S. A critical assessment of micellization of sodium dodecyl benzene sulfonate (SDBS) and its interaction with poly(vinyl pyrrolidone) and hydrophobically modified polymers, JR 400 and LM 200. J. Phys. Chem. B 2003, 107, 3650–3658. [Google Scholar] [CrossRef]
  143. Laschewsky, A.; Wattebled, L.; Arotçaréna, M.; Habib-Jiwan, J.; Rakotoaly, R. Synthesis and properties of cationic oligomeric surfactants. Langmuir 2005, 21, 7170–7179. [Google Scholar] [CrossRef]
  144. Jones, G.; Jiang, H. Detection of lipopolysaccharide and lipid a employing a spermine−pyrene conjugate. Bioconjug. Chem. 2005, 16, 621–625. [Google Scholar] [CrossRef]
  145. Jana-Sur, P.; Chakraborty, A. Studies on the interaction of bacterial capsular polysaccharide-Klebsiella K16 with cationic dyes. J. Photochem. Photobiol. A Chem. 2005, 173, 64–69. [Google Scholar] [CrossRef]
  146. Mitra, A.; Nath, R.; Biswas, S.; Chakraborty, A.; Panda, A. A comparative study on the physico-chemical properties of bacterial capsular polysaccharides from different serotypes of Klebsiella. J. Photochem. Photobiol. A Chem. 2006, 178, 98–105. [Google Scholar] [CrossRef]
  147. Dasgupta, S.; Nath, R.; Biswas, S.; Mitra, A.; Panda, A. Interactions of bacterial polysaccharides with cationic dyes: Physicochemical studies. Ind. J. Biochem. Biophys. 2009, 46, 192–197. [Google Scholar]
  148. Voigt, G.; Nowak, F.; Ehlert, J.; Beenken, W.; Leupold, D.; Sandner, W. Substructures and different energy relaxation time within the first electronic transition of pinacyanol. Chem. Phys. Lett. 1997, 278, 380–390. [Google Scholar] [CrossRef]
  149. Mitchell, D.; Ryabov, A.; Kundu, S.; Chanda, A.; Collins, T. Oxidation of pinacyanol chloride by H2O2 catalyzed by FeIII complexed to tetraamidomacrocyclic ligand: Unusual kinetics and product identification. J. Coord. Chem. 2010, 63, 2605–2618. [Google Scholar] [CrossRef]
  150. Tang, L.; Gunderson, W.; Weitz, A.; Hendrich, M.; Ryabov, A.; Collins, T. Activation of dioxygen by a TAML activator in reverse micelles: Characterization of an FeIIIFeIV dimer and associated catalytic chemistry. J. Am. Chem. Soc. 2015, 137, 9704–9715. [Google Scholar] [CrossRef] [Green Version]
  151. Cano-Sarabia, M.; Angelova, A.; Ventosa, N.; Lesieur, S.; Veciana, J. Cholesterol induced CTAB micelle-to-vesicle phase transitions. J. Coll. Interface Sci. 2010, 350, 10–15. [Google Scholar] [CrossRef]
  152. Berlepsch, H.; Ludwig, K.; Böttcher, C. Pinacyanol chloride forms mesoscopic H- and J-aggregates in aqueous solution—A spectroscopic and cryo-transmission electron microscopy study. Phys. Chem. Chem. Phys. 2014, 16, 10659–10668. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  153. Grygorova, G.; Klochkov, V.; Yefimova, S.; Malyukin, Y. Effects of orthovanadate-based nanoparticles of various sizes on the aggregation behavior of polymethine dyes in aqueous solutions. Chem. Phys. Lett. 2015, 621, 46–51. [Google Scholar] [CrossRef]
  154. Christensson, N.; Milota, F.; Hauer, J.; Sperling, J.; Bixner, O.; Nemeth, A.; Kauffmann, H. High frequency vibrational modulations in two-dimensional electronic spectra and their resemblance to electronic coherence signatures. J. Phys. Chem. B 2011, 115, 5383–5391. [Google Scholar] [CrossRef]
  155. Delgado, J.; Selsby, R. Density functional theory calculations on rhodamine B and pinacyanol chloride. Optimized ground state, dipole moment, vertical ionization potential, adiabatic electron affinity and lowest excited triplet state. Photochem. Photobiol. 2012, 89, 51–60. [Google Scholar] [CrossRef]
  156. Galestian Pour, A.; Lincoln, C.; Perlík, V.; Šanda, F.; Hauer, J. Anharmonic vibrational effects in linear and two-dimensional electronic spectra. Phys. Chem. Chem. Phys. 2017, 19, 24752–24760. [Google Scholar] [CrossRef]
  157. Magana, J.; Gumí-Audenis, B.; Tas, R.; Gascoigne, L.; Atkins, D.; Voets, I. Bioinspired scaffolding by supramolecular amines allows the formation of one- and two-dimensional silica superstructures. Chem. Eur. J. 2020, 26, 15330–15336. [Google Scholar] [CrossRef]
  158. Mewes, L.; Ingle, R.; Al Haddad, A.; Chergui, M. Broadband visible two-dimensional spectroscopy of molecular dyes. J. Chem. Phys. 2021, 155, 034201. [Google Scholar] [CrossRef] [PubMed]
  159. Yu, G.; Walker, M.; Wilson, M. Atomistic simulation studies of ionic cyanine dyes: Self-assembly and aggregate formation in aqueous solution. Phys. Chem. Chem. Phys. 2021, 23, 6408–6421. [Google Scholar] [CrossRef] [PubMed]
  160. Kiprianov, A. Absorption spectra of organic dyes containing two chromophores. Russ. Chem. Rev. 1971, 40, 594–607. [Google Scholar] [CrossRef]
  161. Yuzhakov, V. Association of dye molecules and its spectroscopic manifestation. Russ. Chem. Rev. 1979, 48, 1076–1091. [Google Scholar] [CrossRef]
  162. Ishchenko, A. Structure and spectral-luminescent properties of polymethine dyes. Russ. Chem. Rev. 1991, 60, 865–884. [Google Scholar] [CrossRef]
  163. Yuzhakov, V. Aggregation of dye molecules and its influence on the spectral luminescent properties of solutions. Russ. Chem. Rev. 1992, 61, 613–628. [Google Scholar] [CrossRef]
  164. Shapiro, B. Molecular assemblies of polymethine dyes. Russ. Chem. Rev. 2006, 75, 433–456. [Google Scholar] [CrossRef]
  165. Nandini, R.; Vishalakshi, B. A study of interaction of cationic dyes with anionic polyelectrolytes. Spectrochim. Acta Part A Mol. Biomol. Spectr. 2010, 75, 14–20. [Google Scholar] [CrossRef]
  166. Nandini, R.; Vishalakshi, B. A comparitive study of polyelectrolyte–dye interactions. Spectrochim. Acta Part A Mol. Biomol. Spectr. 2009, 74, 1025–1030. [Google Scholar] [CrossRef]
  167. Nandini, R.; Vishalakshi, B. Spectophotometric study of interaction between sodium carrageenate and cationic dyes. J. Chem. 2011, 8, S253–S265. [Google Scholar] [CrossRef]
  168. Gadde, S.; Batchelor, E.; Weiss, J.; Ling, Y.; Kaifer, A. Control of H- and J-aggregate formation via host−guest complexation using cucurbituril hosts. J. Am. Chem. Soc. 2008, 130, 17114–17119. [Google Scholar] [CrossRef] [PubMed]
  169. Gadde, S.; Batchelor, E.; Kaifer, A. Controlling the formation of cyanine dye H- and J-aggregates with cucurbituril hosts in the presence of anionic polyelectrolytes. Chem. Eur. J. 2009, 15, 6025–6031. [Google Scholar] [CrossRef] [PubMed]
  170. Batchelor, E.; Gadde, S.; Kaifer, A. Host–guest control on the formation of pinacyanol chloride H-aggregates in anionic polyelectrolyte solutions. Supramol. Chem. 2010, 22, 40–45. [Google Scholar] [CrossRef]
  171. Ishchenko, A.; Shapovalov, S. Heterogeneous association of the ions of dyes in solutions (review). J. Appl. Spectr. 2004, 71, 605–629. [Google Scholar] [CrossRef]
  172. Namani, T.; Ishikawa, T.; Morigaki, K.; Walde, P. Vesicles from docosahexaenoic acid. Coll. Surf. B BioInterface 2007, 54, 118–123. [Google Scholar] [CrossRef]
  173. Scalia, S.; Tursilli, R.; Bianchi, A.; Nostro, P.; Bocci, E.; Ridi, F.; Baglioni, P. Incorporation of the sunscreen agent, octyl methoxycinnamate in a cellulosic fabric grafted with β-cyclodextrin. Int. J. Pharm. 2006, 308, 155–159. [Google Scholar] [CrossRef]
  174. Solier, J.; Galera-Cortés, E.; Sabaté, R.; Estelrich, J. Translational diffusion in charged phospholipid bilayer membranes. Coll. Surf. A Physicochem. Eng. Asp. 2005, 270–271, 88–92. [Google Scholar] [CrossRef]
  175. Haro-Pérez, C.; Quesada-Pérez, M.; Callejas-Fernández, J.; Sabaté, R.; Estelrich, J.; Hidalgo-Álvarez, R. Probing the jellium model with colloidal dispersions of charged liposomes. Coll. Surf. A Physicochem. Eng. Asp. 2005, 270–271, 352–356. [Google Scholar] [CrossRef]
  176. Sabaté, R.; Estelrich, J. Pinacyanol as effective probe of fibrillar β-amyloid peptide: Comparative study with Congo red. Biopolymers 2003, 72, 455–463. [Google Scholar] [CrossRef]
  177. Sabaté, R.; Gallardo, M.; Estelrich, J. An autocatalytic reaction as a model for the kinetics of the aggregation of β-amyloid. Biopolymers 2003, 71, 190–195. [Google Scholar] [CrossRef]
  178. Sabaté, R.; Gallardo, M.; Estelrich, J. Temperature dependence of the nucleation constant rate in β amyloid fibrillogenesis. Int. J. Biol. Macromol. 2005, 35, 9–13. [Google Scholar] [CrossRef] [PubMed]
  179. Sabaté, R.; Estelrich, J. Stimulatory and inhibitory effects of alkyl bromide surfactants on β-amyloid fibrillogenesis. Langmuir 2005, 21, 6944–6949. [Google Scholar] [CrossRef] [PubMed]
  180. Sabaté, R.; Estelrich, J. Evidence of the existence of micelles in the fibrillogenesis of β-amyloid peptide. J. Phys. Chem. B 2005, 109, 11027–11032. [Google Scholar] [CrossRef] [PubMed]
  181. Poltev, V.; Grokhlina, T.; González, E.; Deriabina, A.; Cruz, A.; Gorb, L.; Leszczynski, J.; Djimant, L.; Veselkov, A. The study of three-dimensional structure of caffeine associates using computational and experimental methods. J. Mol. Struct. THEOCHEM 2004, 709, 123–128. [Google Scholar] [CrossRef]
  182. Veselkov, A.; Evstigneev, M.; Rozvadovskaya, A.; Hernandez Santiago, A.; Zubchenok, O.; Djimant, L.; Davies, D. 1H NMR structural and thermodynamical analysis of the hetero-association of daunomycin and novatrone in aqueous solution. J. Mol. Struct. 2004, 701, 31–37. [Google Scholar] [CrossRef]
  183. Davies, D.; Evstigneev, M.; Veselkov, D.; Veselkov, A. Hetero-association of anticancer antibiotics in aqueous solution: NMR and molecular mechanics analysis. Biophys. Chem. 2005, 117, 111–118. [Google Scholar] [CrossRef]
  184. Grigorev, I.; Korzhevskii, D. Modern imaging technologies of mast cells for biology and medicine (review). Mod. Technol. Med. 2021, 13, 93–109. [Google Scholar] [CrossRef]
  185. Shapovalov, S.; Koval, V.; Mchedlov-Petrossyan, N.; Kleschevnikova, V.; Derevyanko, N.; Ishchenko, A. Influence of the polymethine chain length of indopolycarbocyanine cations on their association with anionic dyes in water. Rep. NAS Ukr. 1999, 2, 156–161. (In Russian) [Google Scholar]
  186. Shapovalov, S.; Svishcheva, Y. Effect of cationic surfactants on association of pinacyanol with organic counterions. Russ. J. Appl. Chem. 2002, 75, 452–456. [Google Scholar] [CrossRef]
  187. Shapovalov, S.; Svishcheva, Y.; Larin, V. Association of pinacyanol with sulfonephthalein and oxyxanthene anions in aqueous solutions. Ukr. Chem. J. 2002, 68, 107–110. (In Russian) [Google Scholar]
  188. Shapovalov, S. Interaction of cationic forms of pinacyanol with organic anions in aqueous solutions: Estimation of structural factor. Ukr. Chem. J. 2004, 70, 25–29. (In Russian) [Google Scholar]
  189. Shapovalov, S. Effect of temperature on the stability and stoichiometric composition of ion associates of the pinacyanol cation. Russ. J. Phys. Chem. A 2005, 79, 478–482. Available online: https://elibrary.ru/item.asp?id=27770914 (accessed on 27 January 2022).
  190. Shapovalov, S.; Samoilov, E. Regularities of homo- and heteroassociation of the pinacyanol cation in aqueous solution. Russ. Chem. Bull. 2008, 57, 1405–1415. [Google Scholar] [CrossRef]
  191. Taniguchi, M.; Du, H.; Lindsey, J. Photochemcad 3: Diverse modules for photophysical calculations with multiple spectral databases. Photochem. Photobiol. 2018, 94, 277–289. [Google Scholar] [CrossRef] [PubMed]
  192. Taniguchi, M.; Lindsey, J. Database of absorption and fluorescence spectra of >300 common compounds for use in Photochem CAD. Photochem. Photobiol. 2018, 94, 290–327. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  193. Rentsch, S.; Danielius, R.; Gadonas, R. Bestimmung von Lebensdauern und Transientenabsorptionsspektren von Polymethinfarbstoffen aus pikosekunden-spektroskopischen Messungen. J. Signalaufz.-Mater. 1984, 12, 319–328. [Google Scholar]
  194. Khouri, S.; Buss, V. UV/Vis spectral study of the self-aggregation of pinacyanol chloride in ethanol–water solutions. J. Sol. Chem. 2009, 39, 121–130. [Google Scholar] [CrossRef]
  195. Permogorov, V.; Serdyukova, L.; Frank-Kamenetsky, M. On the nature of the long-wavelength absorption band and luminescence of cyanine dyes. Opt. Spectrosc. 1967, 22, 979–981. (In Russian) [Google Scholar]
  196. Higgins, D.; Reid, P.; Barbara, P. Structure and exciton dynamics in J-aggregates studied by polarization-dependent near-field scanning optical microscopy. J. Phys. Chem. 1996, 100, 1174–1180. [Google Scholar] [CrossRef]
  197. Shindy, H. Fundamentals in the chemistry of cyanine dyes: A review. Dye. Pigment. 2017, 145, 505–513. [Google Scholar] [CrossRef]
  198. Herz, A. Protonation equilibria of cyanines of solution and at AgBr surfaces. Photogr. Sci. Eng. 1974, 18, 207–221. [Google Scholar]
  199. Sabaté, R.; Gallardo, M.; Estelrich, J. Location of pinacyanol in micellar solutions of n-alkyl trimethylammonium bromide surfactants. J. Coll. Interface Sci. 2001, 233, 205–210. [Google Scholar] [CrossRef] [PubMed]
  200. Svisheva, Y.; Shapovalov, S. Peculiarities of the association of cationic forms of pinacyanol with counterions in aqueous solution. Kharkiv Univ. Bull. Chem. Ser. 2000, 477, 112–116. (In Russian) [Google Scholar]
  201. Sheppard, S. The effects of environment and aggregation on the absorption spectra of dyes. Rev. Mod. Phys. 1942, 14, 303–340. [Google Scholar] [CrossRef]
  202. Sabaté, R.; Estelrich, J. Determination of the dimerization constant of pinacyanol: Role of the thermochromic effect. Spectrochim. Acta Part A Mol. Biomol. Spectr. 2008, 70, 471–476. [Google Scholar] [CrossRef]
  203. Naorem, H.; Devi, S. Determination of the dimerization constant of pinacyanol dye in mixed aqueous–organic solvent media with different hydrophobic character. J. Mol. Liq. 2012, 173, 119–123. [Google Scholar] [CrossRef]
  204. Yao, H.; Morita, Y.; Kimura, K. Effect Of organic solvents on J aggregation of pseudoisocyanine dye at mica/water interfaces: Morphological transition from three-dimension to two-dimension. J. Coll. Interface Sci. 2008, 318, 116–123. [Google Scholar] [CrossRef]
  205. Sabaté, R.; Estelrich, J. Determination of micellar microenvironment of pinacyanol by visible spectroscopy. J. Phys. Chem. B 2003, 107, 4137–4142. [Google Scholar] [CrossRef]
  206. Sabaté, R.; Freire, L.; Estelrich, J. Influence of dielectric constant on the spectral behavior of pinacyanol. A spectrophotometric experiment for physical chemistry. J. Chem. Educ. 2001, 78, 243. [Google Scholar] [CrossRef]
  207. Perram, J. Interpretation of spectra. J. Chem. Phys. 1968, 49, 4245–4246. [Google Scholar] [CrossRef]
  208. Antonov, L. UV–Vis spectroscopic and chemometric study on the aggregation of ionic dyes in water. Talanta 1999, 49, 99–106. [Google Scholar] [CrossRef]
  209. Sabaté, R.; Gallardo, M.; de la Maza, A.; Estelrich, J. A spectroscopy study of the interaction of pinacyanol with n-dodecyltrimethylammonium bromide micelles. Langmuir 2001, 17, 6433–6437. [Google Scholar] [CrossRef]
  210. Khouri, S.; Richter, D.; Buss, V. Circular dichroism and theoretical calculations of pinacyanol dimer inclusion in γ-cyclodextrin. J. Incl. Phenom. Macrocycl. Chem. 2009, 65, 287–292. [Google Scholar] [CrossRef]
  211. Ramette, R.; Sandell, E. Rhodamine B equilibria. J. Am. Chem. Soc. 1956, 78, 4872–4878. [Google Scholar] [CrossRef]
  212. Lavorel, J. Influence of concentration on the absorption spectrum and the action spectrum of fluorescence of dye solutions. J. Phys. Chem. 1957, 61, 1600–1605. [Google Scholar] [CrossRef]
  213. Rohatgi, K.; Singhal, G. Nature of bonding in dye aggregates. J. Phys. Chem. 1966, 70, 1695–1701. [Google Scholar] [CrossRef]
  214. West, W.; Pearce, S. The dimeric state of cyanine dyes. J. Phys. Chem. 1965, 69, 1894–1903. [Google Scholar] [CrossRef]
  215. Poluektov, N.; Beltyukova, S.; Meshkova, S. Dimerization constants of some basic dyes used in extraction-photometric methods of analysis. J. Anal. Chem. 1971, 26, 1042–1047. (In Russian) [Google Scholar]
  216. Davydov, A. The theory of molecular excitons. Sov. Phys. Uspekhi 1964, 7, 145–178. [Google Scholar] [CrossRef]
  217. Alfimov, M.; Shtykova, A.; Razumov, V. Photo- and thermoinitiated formation of J- and H-aggregates in amorphous dispersion of a carbocyanine dye. High Energy Chem. 2006, 40, 18–21. [Google Scholar] [CrossRef]
  218. Kim, S.; Kim, D. Determination of the structure of 1,1′-diethyl-2,2′-carbocyanine iodide using NMR spectra and GIAO-HF/DFT calculations. Dye. Pigment. 2008, 79, 1–6. [Google Scholar] [CrossRef]
  219. Adeagbo, W.; Buss, V.; Entel, P. Inclusion complexes of dyes and cyclodextrins: Modeling supermolecules by rigorous quantum mechanics. J. Incl. Phenom. 2002, 44, 203–205. [Google Scholar] [CrossRef]
  220. Ma, F.; Yartsev, A. Ultrafast photoisomerization of pinacyanol: Watching an excited state reaction transiting from barrier to barrierless forms. RSC Adv. 2016, 6, 45210–45218. [Google Scholar] [CrossRef]
  221. Shapovalov, S. Dimerization of Cyanine Dyes: Results of Quantum-Chemical Simulation. In Proceedings of the XXIV Galyna Puchkovska International School-Seminar “Spectroscopy of Molecules and Crystals”, Odessa, Ukraine, 25–30 August 2019; p. 47. Available online: https://isssmc.org.ua/files/books/2019-XXIV-ISSSMC.pdf (accessed on 27 January 2022).
  222. Shapovalov, S. Interaction of sulfophthaleine anions with cationic dyes in aqueous solution. Russ. J. Gen. Chem. 2010, 80, 953–963. [Google Scholar] [CrossRef]
  223. Shapovalov, S. Association of anions of phenolsulfonephthalein and its alkyl-substituted derivatives with single-charged cations of polymethines. Russ. Chem. Bull. 2011, 60, 465–473. [Google Scholar] [CrossRef]
  224. Shapovalov, S.; Kiseliova, Y. Association of thymolsulfonephthalein and cresolsulfonephthalein anions with cationic cyanines in aqueous solution. Chem. Chem. Technol. 2010, 4, 271–276. [Google Scholar] [CrossRef]
  225. Shapovalov, S. Association constants and enthalpies of formation of heteroassociates of anions of cresol red and thymol blue. Russ. J. Phys. Chem. A 2010, 85, 9–15. [Google Scholar] [CrossRef]
  226. Shapovalov, S.; Kiseleva, Y. Heteroassociation of the bromine-containing anions of sulfophthaleins in aqueous solution. Russ. Chem. Bull. 2010, 59, 1317–1326. [Google Scholar] [CrossRef]
  227. Shapovalov, S. Bromocresol purple in the processes of cation-anionic interactions in aqueous solutions: UV–Vis spectroscopy and computer simulation of dissimilar association. Spectrochim. Acta A Mol. Biomol. Spectr. 2022, 269, 120741. [Google Scholar] [CrossRef]
  228. Shapovalov, S. Cation-anionic association of organic dyes in aqueous solutions: Structure and properties of associates. Mod. Org. Chem. Res. 2017, 2, 195–203. [Google Scholar] [CrossRef]
  229. Shapovalov, S. Heteroassociates of singly- and doubly-charged anions of alizarin red S with the pinacyanol cation. Russ. Chem. Bull. 2011, 60, 89–98. [Google Scholar] [CrossRef]
  230. Shapovalov, S. Ionic association with anions of alizarin red S in aqueous solutions with surfactants. Russ. J. Phys. Chem. A 2011, 85, 1369–1375. [Google Scholar] [CrossRef]
  231. Svishchova, Y.; Shapovalov, S. Interactions between xylenol orange and pinacyanol in aqueous solutions. Kharkiv Univ. Bull. Chem. Ser. 2012, 21, 197–199. (In Russian) [Google Scholar]
  232. Shapovalov, S.; Svishchova, Y. Heteroassociates of pinacianol cation in aqueous solutions: Formation and their interaction with organic multiply charged anions. Fr. Ukr. J. Chem. 2018, 6, 21–30. [Google Scholar] [CrossRef]
  233. Shapovalov, S.; Ponomariov, V.; Mosharenkova, O.; Butenko, S. Studies of the association of dyes with surfactants in aqueous solutions: Spectrophotometry and computer simulation. Fr. Ukr. J. Chem. 2019, 7, 9–19. [Google Scholar] [CrossRef]
  234. Svisheva, Y.; Shapovalov, S. Dissimilar association of cationic dyes with humic acids in aqueous solutions. Kharkiv Univ. Bull. Chem. Ser. 2014, 23, 47–53. (In Ukrainian) [Google Scholar]
  235. Armitage, B.; Retterer, J.; O’Brien, D. Dimerization of cyanine dyes in water driven by association with hydrophobic borate anions. J. Am. Chem. Soc. 1993, 115, 10786–10790. [Google Scholar] [CrossRef]
  236. Nicoli, F.; Roos, M.; Hemmig, E.; Di Antonio, M.; de Vivie-Riedle, R.; Liedl, T. Proximity-induced H-aggregation of cyanine dyes on DNA-duplexes. J. Phys. Chem. A 2016, 120, 9941–9947. [Google Scholar] [CrossRef]
  237. Veselkov, D.A.; Evstigneev, M.P.; Kodintsev, V.V.; Djimant, L.N.; Davies, D.; Veselkov, A.N. Investigation of the heteroassociation of daunomycin and ethidium bromide molecules in aqueous solution by 1H NMR spectroscopy. J. Phys. Chem. 2001, 75, 879–884. (In Russian) [Google Scholar]
  238. Veselkov, D.A.; Sigaev, V.A.; Vysotskii, S.A.; Djimant, L.N.; Davies, D.; Veselkov, A.N. Investigation of the heteroassociation of caffeine with acridine orange dye in aqueous solution by 1H NMR spectroscopy. J. Struct. Chem. 2000, 41, 86–96. (In Russian) [Google Scholar] [CrossRef]
  239. Shapovalov, S.; Koval, V.; Chernaya, T.; Pereverzev, A.; Derevyanko, N.; Ishchenko, A.; Mchedlov-Petrossyan, N. Association of indopolymethine cyanine cations with anions of sulfonephthalein and xanthene dyes in water. J. Braz. Chem. Soc. 2005, 16, 232–240. [Google Scholar] [CrossRef] [Green Version]
  240. Shapovalov, S. Kinetics of interaction between cyanine dye ion association species and surfactants in aqueous solution. J. Phys. Chem. 2005, 79, 1586–1591. [Google Scholar]
  241. Shapovalov, S.; Lvovskaya, M. About interaction between pinacyanol and cations of astrafloxine or neutral red in aqueous solution. Kharkiv Univ. Bull. Chem. Ser. 2002, 9, 157–161. (In Russian) [Google Scholar]
  242. Hamai, S. Complex formation in cationic dye-organic anion systems in aqueous solution. Bull. Chem. Soc. Jap. 1985, 58, 2099–2106. [Google Scholar] [CrossRef] [Green Version]
  243. Larin, V.; Shapovalov, S. Association between complex ferrumcyanide anions with pinacyanol cation in aqueous solution. Chem. Chem. Technol. 2012, 6, 25–29. [Google Scholar] [CrossRef]
  244. Samoilov, E. Interaction between complex ferrum-cyanide ions and polymethine cations in aqueous solutions. Kharkiv Univ. Bull. Chem. Ser. 2009, 17, 208–212. (In Russian) [Google Scholar]
  245. Ganguly, B.; Nath, R. Spectral studies of pinacyanol chloride in sodium alkyl sulphate. Chem. Mater. Res. 2012, 2, 13–24. Available online: https://www.iiste.org/Journals/index.php/CMR/article/view/2183 (accessed on 27 January 2022).
  246. Shapovalov, S. Effect of surfactants on homo- and heteroassociation of pinacyanol cation in aqueous solution. Russ. J. Appl. Chem. 2008, 81, 1160–1166. [Google Scholar] [CrossRef]
  247. Maity, B.; Chatterjee, A.; Seth, D. Photophysics of lumichrome in anionic and cationic micellar media. RSC Adv. 2015, 5, 3814–3824. [Google Scholar] [CrossRef] [Green Version]
  248. Ali, A.; Bhushan, V.; Malik, N.; Behera, K. Study of mixed micellar aqueous solutions of sodium dodecyl sulfate and amino acids. Coll. J. 2013, 75, 357–365. [Google Scholar] [CrossRef]
  249. Khan, M.; Siddiqui, Z.; Uddin, F. Kinetic and mechanism study of the oxidative decolorization of neutral red by bromate in micellar medium. J. Iran. Chem. Soc. 2009, 6, 533–541. [Google Scholar] [CrossRef]
  250. Zaghbani, N.; Hafiane, A.; Dhahbi, M. Removal of safranin T from wastewater using micellar enhanced ultrafiltration. Desalination 2008, 222, 348–356. [Google Scholar] [CrossRef]
  251. Göktürk, S.; Yeşim Talman, R.; Erdinç, N.; Tunçay, M. Solution behaviour of rivanol in micellar environments. Spectr. Lett. 2006, 39, 357–372. [Google Scholar] [CrossRef]
  252. Barnadas-Rodríguez, R.; Estelrich, J. Photophysical changes of pyranine induced by surfactants: Evidence of premicellar aggregates. J. Phys. Chem. B 2009, 113, 1972–1982. [Google Scholar] [CrossRef] [PubMed]
  253. Hayakawa, K.; Sato, M.; Nagata, I.; Yamato, N.; Murata, H.; Tanokura, M. Temperature-induced chirality reversal of induced circular dichroism of premicellar aggregates of acridine orange derivatives and dodecanoyl-L-threonine in aqueous solution. Coll. Polym. Sci. 2006, 284, 1453–1458. [Google Scholar] [CrossRef]
  254. Tiwari, A.; Sonu; Sowmiya, M.; Saha, S.K. Study on premicellar and micellar aggregates of gemini surfactants with hydroxyl substituted spacers in aqueous solution using a probe showing TICT fluorescence properties. J. Photochem. Photobiol. A Chem. 2011, 223, 6–13. [Google Scholar] [CrossRef]
  255. Sowmiya, M.; Tiwari, A.; Saha, S. Fluorescent probe studies of micropolarity, premicellar and micellar aggregation of non-ionic Brij surfactants. J. Coll. Interface Sci. 2010, 344, 97–104. [Google Scholar] [CrossRef]
  256. Matsumoto, T.; Otsuki, T.; Sueishi, Y.; Yamamoto, S. Effects of anionic surfactants on the hydride-transfer reaction of NADH with methylene blue: In premicellar and micellar regions. Bull. Chem. Soc. Jap. 2008, 81, 291–297. [Google Scholar] [CrossRef]
  257. Nandini, R.; Vishalakshi, B. A spectroscopic study of interaction of cationic dyes with heparin. Orbital 2009, 1, 255–272. [Google Scholar]
  258. Pandey, S.; Kar, J.; Azam, A.; Pandey, S.; Chawla, H. Competitive self and induced aggregation of calix [4]arene ethers and their interaction with pinacyanol chloride and methylene blue in nonaqueous media. J. Sol. Chem. 2010, 39, 107–120. [Google Scholar] [CrossRef]
  259. Khouri, S.; Knierim, R.; Buss, V. Induced circular dichroism of the interaction between pinacyanol and algal alginates. Carbohyd. Res. 2009, 344, 1729–1733. [Google Scholar] [CrossRef] [PubMed]
  260. Wu, H.; Li, W.; Miao, K.; He, X.; Yu, C.; Huang, X.; Liang, H. Studies on the binding mode of pinacyanol chloride to nucleic acids. Chin. J. Chem. 2010, 20, 462–466. [Google Scholar] [CrossRef]
Figure 1. Structure of PIN (X = Cl, Br, I).
Figure 1. Structure of PIN (X = Cl, Br, I).
Colorants 01 00011 g001
Figure 2. Absorption spectrum of PIN iodide in methanol and its constituent bands.
Figure 2. Absorption spectrum of PIN iodide in methanol and its constituent bands.
Colorants 01 00011 g002
Figure 3. Alternation of π-bonds in the electronic system of the cation.
Figure 3. Alternation of π-bonds in the electronic system of the cation.
Colorants 01 00011 g003
Figure 4. Redistribution of absorption band intensities of PIN in water. Concentration, mol/L: 2.1 × 10−6 (1), 4.1 × 10−6 (2), 6.2 × 10−6 (3), 8.2 × 10−6 (4), 1.9 × 10−5 (5), and 3.1 × 10−5 (6). The thickness of the absorbing layer is 1.00 cm; pH = 6.4.
Figure 4. Redistribution of absorption band intensities of PIN in water. Concentration, mol/L: 2.1 × 10−6 (1), 4.1 × 10−6 (2), 6.2 × 10−6 (3), 8.2 × 10−6 (4), 1.9 × 10−5 (5), and 3.1 × 10−5 (6). The thickness of the absorbing layer is 1.00 cm; pH = 6.4.
Colorants 01 00011 g004
Figure 5. Normalized spectra of two concentration of PIN in water. Concentration, mol/L: 2.2 × 10−6 (1) and 8.2 × 10−6 (2); pH = 6.4.
Figure 5. Normalized spectra of two concentration of PIN in water. Concentration, mol/L: 2.2 × 10−6 (1) and 8.2 × 10−6 (2); pH = 6.4.
Colorants 01 00011 g005
Figure 6. Normalized spectrum of PIN monomer (A); the algebraic sum (B) of bands IM (I), IIM (II), IIIM (III), and IVM (IV); and the difference (C) of curves A and B.
Figure 6. Normalized spectrum of PIN monomer (A); the algebraic sum (B) of bands IM (I), IIM (II), IIIM (III), and IVM (IV); and the difference (C) of curves A and B.
Colorants 01 00011 g006
Figure 7. Normalized spectrum of PIN dimer in water (A). Bands: ID (I), IID (II), and IIID (III).
Figure 7. Normalized spectrum of PIN dimer in water (A). Bands: ID (I), IID (II), and IIID (III).
Colorants 01 00011 g007
Figure 8. Electron density in PIN cation (AM1 semi-empirical method; hydrogen is gray, nitrogen is violet, and some carbon atoms show electron density or are marked in blue).
Figure 8. Electron density in PIN cation (AM1 semi-empirical method; hydrogen is gray, nitrogen is violet, and some carbon atoms show electron density or are marked in blue).
Colorants 01 00011 g008
Figure 9. Optimized construction of PIN pentamer (no hydrogen atoms are shown).
Figure 9. Optimized construction of PIN pentamer (no hydrogen atoms are shown).
Colorants 01 00011 g009
Figure 10. Absorption spectra in the system “Ct+ + BCP”. Concentrations, mol/L; PIN, 4.4 × 10−6 (1–7); BCP, 0 (1), 3.1 × 10−6 (2), 5.0 × 10−6 (2), 1.0 × 10−5 (3), 1.5 × 10−5 (4), 2.0 × 10−5 (5), 2.6 × 10−5 (6), and 3.0 × 10−5 (7). The thickness of the absorbing layer is 1.0 cm; pH = 4.7. Comparison solutions are water.
Figure 10. Absorption spectra in the system “Ct+ + BCP”. Concentrations, mol/L; PIN, 4.4 × 10−6 (1–7); BCP, 0 (1), 3.1 × 10−6 (2), 5.0 × 10−6 (2), 1.0 × 10−5 (3), 1.5 × 10−5 (4), 2.0 × 10−5 (5), 2.6 × 10−5 (6), and 3.0 × 10−5 (7). The thickness of the absorbing layer is 1.0 cm; pH = 4.7. Comparison solutions are water.
Colorants 01 00011 g010
Figure 11. Absorption spectra in the system “Ct+ + BCP2−”. Concentrations, mol/L; PIN, 5.4 × 10−7 (1–4); BCP, 0 (1), 1.2 × 10−6 (2), 2.6 × 10−5 (3), and 5 × 10−5 (4). The thickness of the absorbing layer is 5.0 cm; pH = 9.1. Comparison solutions are water (1) and BCP solution at the appropriate concentration (2–4).
Figure 11. Absorption spectra in the system “Ct+ + BCP2−”. Concentrations, mol/L; PIN, 5.4 × 10−7 (1–4); BCP, 0 (1), 1.2 × 10−6 (2), 2.6 × 10−5 (3), and 5 × 10−5 (4). The thickness of the absorbing layer is 5.0 cm; pH = 9.1. Comparison solutions are water (1) and BCP solution at the appropriate concentration (2–4).
Colorants 01 00011 g011
Figure 12. Dissimilar Ct+ associate with BCP2− (anion is located in the middle; no hydrogen atoms are shown).
Figure 12. Dissimilar Ct+ associate with BCP2− (anion is located in the middle; no hydrogen atoms are shown).
Colorants 01 00011 g012
Figure 13. Absorption spectra in the system “Ct+ + K4[Fe(CN)6]”. Concentration, mol/L; PIN, 1.4 × 10−6 (13); K4[Fe(CN)6], 0 (1), 4.1 × 10−6 (2), and 1.7 × 10−5 (3); pH = 6.8. The thickness of the absorbing layer is 1.00 cm. Comparison solutions are water.
Figure 13. Absorption spectra in the system “Ct+ + K4[Fe(CN)6]”. Concentration, mol/L; PIN, 1.4 × 10−6 (13); K4[Fe(CN)6], 0 (1), 4.1 × 10−6 (2), and 1.7 × 10−5 (3); pH = 6.8. The thickness of the absorbing layer is 1.00 cm. Comparison solutions are water.
Colorants 01 00011 g013
Figure 14. Normalized absorption spectrum of Ct+ with potassium chloride at I = 0.2 and its constituent bands.
Figure 14. Normalized absorption spectrum of Ct+ with potassium chloride at I = 0.2 and its constituent bands.
Colorants 01 00011 g014
Figure 15. Light absorption of PIN at different content of CPB. Concentration, mol/L; PIN, 4.7 × 10−5 (1–3); CPB, 0 (1), 4 × 10−6 (2), and 2 × 10−3 (3); pH = 6.9. The thickness of the absorbing layer is 0.20 cm.
Figure 15. Light absorption of PIN at different content of CPB. Concentration, mol/L; PIN, 4.7 × 10−5 (1–3); CPB, 0 (1), 4 × 10−6 (2), and 2 × 10−3 (3); pH = 6.9. The thickness of the absorbing layer is 0.20 cm.
Colorants 01 00011 g015
Figure 16. Light absorption in the “PIN + SDS” system. Concentration, mol/L; PIN, 2 × 10−5 (1–4); SDS, 0 (1), 4 × 10−4 (2), 1 × 10−3 (3), and 2 × 10−2 (4); pH = 6.2. The thickness of the absorbing layer is 0.20 cm.
Figure 16. Light absorption in the “PIN + SDS” system. Concentration, mol/L; PIN, 2 × 10−5 (1–4); SDS, 0 (1), 4 × 10−4 (2), 1 × 10−3 (3), and 2 × 10−2 (4); pH = 6.2. The thickness of the absorbing layer is 0.20 cm.
Colorants 01 00011 g016
Table 1. Spectral properties of PIN bands.
Table 1. Spectral properties of PIN bands.
Mediumλmax, nm (Band)εmax (α-Band)
Water [82]600 (α)1.2 × 105
Methanol [193]603.5 (α)1.28 × 105
Ethanol, 30% [85]608 (α)1.35 × 105
Ethanol, abs. [194]606 (α)
563 (β)
522 (γ)
1.752 × 105
Table 2. Features of normalized PIN monomer bands.
Table 2. Features of normalized PIN monomer bands.
FeatureBand
IMIIMIIIMIVM
νmax, cm−116,66117,93919,13219,614
λmax, nm600.2557.4522.3509.7
I, relative units0.9770.5220.0760.057
Half-width, cm−189212059161543
Surface area, cm−1925.0669.373.894.1
Table 3. Features of normalized PIN dimer bands.
Table 3. Features of normalized PIN dimer bands.
FeatureBand
IDIIDIIID
νmax, cm−116,84018,30019,120
λmax, nm593.8546.5523.0
I, relative units0.1560.7980.446
Half-width, cm−113409501444
Surface area, cm−1223.5800.1688.4
Table 4. Dissimilar associates of Ct+ with dyes.
Table 4. Dissimilar associates of Ct+ with dyes.
A Dyelog Kas
Ct+·An(Ct+)2·An2−
Alizarin red C6.12 ± 0.1010.92 ± 0.10
Bengal Rose B6.91 ± 0.1213.46 ± 0.11
Bromocresol green6.74 ± 0.0412.09 ± 0.09
Bromocresol purple6.67 ± 0.0511.07 ± 0.10
Bromophenol blue6.88 ± 0.0513.73 ± 0.10
Bromothymol blue5.95 ± 0.1110.89 ± 0.09
Cresol red4.59 ± 0.0310.96 ± 0.10
Eosin5.15 ± 0.0911.25 ± 0.10
Erythrosin6.60 ± 0.1211.84 ± 0.10
Ethyleosin5.45 ± 0.09
Fluorescein4.58 ± 0.119.75 ± 0.10
Phenol red5.83 ± 0.1011.81 ± 0.10
Tetrabromophenylfluoron6.38 ± 0.11
Thymol blue5.29 ± 0.0511.0 ± 0.1
Table 5. Absorption bands of the normalized absorption spectrum (Figure 14) of PIN.
Table 5. Absorption bands of the normalized absorption spectrum (Figure 14) of PIN.
FeatureBand
12345
νmax, cm−116,66518,14019,12019,56020,460
λmax, nm600.24551.27523.01511.25488.76
I, relative units0.9660.68870.09470.15870.1382
Half-width, cm−1101112975968041349
Surface area, cm−11039.120947.03360.144135.700198.255
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Shapovalov, S.A. The First Century of Successful Applications of Pinacyanol: Some Noteworthy Results. Colorants 2022, 1, 165-192. https://doi.org/10.3390/colorants1020011

AMA Style

Shapovalov SA. The First Century of Successful Applications of Pinacyanol: Some Noteworthy Results. Colorants. 2022; 1(2):165-192. https://doi.org/10.3390/colorants1020011

Chicago/Turabian Style

Shapovalov, Serghey A. 2022. "The First Century of Successful Applications of Pinacyanol: Some Noteworthy Results" Colorants 1, no. 2: 165-192. https://doi.org/10.3390/colorants1020011

Article Metrics

Back to TopTop