Next Article in Journal
The Association Between the Flesh Colour and Carotenoid Profile of 25 Cultivars of Mangoes
Previous Article in Journal
Solubility of Metal Precursors in Supercritical CO2: Measurements and Correlations
Previous Article in Special Issue
Beyond Spin Crossover: Optical and Electronic Horizons of 2,6-Bis(pyrazol-1-yl)pyridine Ligands and Complexes
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives: Syntheses, Photophysical Properties and Applications

by
Ekaterina K. Pylova
1,2,3,
Taisiya S. Sukhikh
1,
Alexis Prieto
3,
Florian Jaroschik
3,* and
Sergey N. Konchenko
1,*
1
Nikolaev Institute of Inorganic Chemistry, Siberian Branch of the Russian Academy of Sciences, 3 Lavrentiev Ave., 630090 Novosibirsk, Russia
2
Department of Natural Sciences, Novosibirsk State University, 630090 Novosibirsk, Russia
3
ICGM, Univ Montpellier, CNRS, ENSCM, 34090 Montpellier, France
*
Authors to whom correspondence should be addressed.
Molecules 2025, 30(8), 1659; https://doi.org/10.3390/molecules30081659
Submission received: 13 March 2025 / Revised: 1 April 2025 / Accepted: 3 April 2025 / Published: 8 April 2025
(This article belongs to the Special Issue Advances in Coordination Chemistry, 3rd Edition)

Abstract

:
2-(2′-aminophenyl)benzothiazole is a readily tunable fluorescent core with widespread applications in coordination chemistry, sensing, light-emitting processes, medicinal chemistry, and catalysis. This review provides an overview of the synthetic methodologies to access 2-(2′-aminophenyl)benzothiazole and its organic derivatives, including various phosphorous and silane pincer ligands. The luminescent properties will be discussed, with a special focus on ESIPT and AIE processes. The coordination of transition metals and lanthanides is presented, as well as their influence on biological and light-emitting properties. 2-(2′-aminophenyl)benzothiazole derivatives have also been employed as sensors for a range of cations and anions due to their various binding modes, as well as for bioimaging purposes. Recently, the first application in photocatalysis has emerged, showing one of the many openings for these organic building blocks in the future.

Graphical Abstract

1. Introduction

The chemistry of sulfur–nitrogen heterocyclic compounds is one of the main areas of the chemical industry and is involved in nearly all fields of modern life. The developed compounds have significant contributions in molecular and cellular biology [1,2,3,4], medicine [5,6,7,8,9], optoelectronic technologies [10,11,12,13,14,15,16,17,18,19,20,21,22,23], organic field-effect transistors (OFET) [24,25,26], sensors [27,28,29,30], solar cells [31,32,33,34,35], dyes [36,37], and the chemistry of other functional materials [38,39]. Moreover, such compounds have also found applications as fungicides [40,41] or catalysts [42,43,44]. A particularly notable example is the 2-phenylbenzothiazole (pbt) fragment, which is an important chromophore group, contributing to the utilization of its derivatives in many practical applications. Derivatives of pbt are widely used as building blocks in organic and organometallic chemistry due to their interesting photophysical properties, the ease of their synthesis, and the low cost of the starting reagents. The chemistry of pbt derivatives is a well-studied area that continues to be actively researched: the known organic derivatives include over 120 k compounds, combining amines, amides, alcohols, thiols, phosphines, silicium-, boron-, selenium-containing compounds. The photophysical and electronic properties of these compounds enable their use as ligands in transition metal complexes [45,46,47,48,49,50,51,52], in the biochemical field (bio-probes, antioxidants, etc.) [53,54,55,56,57,58], as light-emitting elements in chemical sensors [59,60], or in luminescent materials (organic light-emitting devices, OLEDs) [61,62,63,64,65] (Figure 1).
The pbt derivatives bearing different functional groups such as -OH, -SH, or -NHR in the 2′-position of the phenyl group have attracted scientific interest as they often undergo intramolecular proton transfer in the excited state (ESIPT) or display Aggregation-Induced Emission (AIE) phenomenon [66,67,68,69]. These processes are related to hydrogen bonding between the acidic hydrogen of these functional groups and the N atom of the heterocycle. In particular, 2-(2′-aminophenyl)benzothiazole (2-NH2-pbt, 1) is a promising core due to the possibility of introducing donor/electron-withdrawing substituents at the amino group. This feature gives ready access to tunable photophysical properties related to ESIPT and/or AIE and also provides additional metal coordination centers [70,71,72,73]. This review will focus on the various synthetic pathways of 1 and its derivatives, as well as on the numerous applications of these emerging organic building blocks in coordination chemistry, sensing, medicinal chemistry, and catalysis.

2. Overview of the Properties and Syntheses of 2-(2′-Aminophenyl)benzothiazole Derivatives

2.1. Structural Characterizations of 2-(2′-Aminophenyl)benzothiazole 1 and Its Derivatives

There are two primary ways of numbering atoms in 2-NH2-pbt 1 found in the literature, namely the IUPAC nomenclature and a commonly used alternative, which will be used in this work, as shown in Figure 2.

2.1.1. XRD Analysis

The only reported crystal structure of compound 1 was published as a 1:1 solvate with DMSO (Figure 2), where each solvent molecule is disordered in two positions (not shown in the figure for simplicity) [74]. The two amine molecules in the unit cell exhibit an almost planar structure, with the angle between the benzothiazole plane and phenyl ring in the range of 4.1–5.4° and an intramolecular hydrogen bond between N2′-N3 (2.68–2.75 Å). Similar planar structures were obtained for closely related 2-(2′-(N-methyl)aminophenyl)benzothiazole (2, 3.1–3.5°) [75] and 2-(2′-amino-5′-methylphenyl)benzothiazole (3, 3.0°) [76], whereas for other derivatives, including amides, imines, aminophosphines, etc., the angle could reach up to 20° [27,48,50,62,72,77,78,79,80,81,82,83,84,85]. This implies some influence of the substituent at the N atom on the geometry of the pbt moiety and, as a consequence, on the strength of the hydrogen bond.

2.1.2. NMR Spectroscopy

Compound 1 is an aromatic system in which all proton signals appear in the range of 6.10–8.10 ppm in the 1H NMR spectrum. In the ¹³C NMR spectrum, carbons C2′, C2, and C3a, which are bonded to nitrogen atoms, have chemical shifts in the low-field region, while other carbons show signals in the aromatic range. An extensive NMR investigation of 43 2-aryl benzothiazole derivatives was performed in 2006, where one- and two-dimensional NMR experiments were recorded in DMSO-d6 to characterize pbt derivatives [86]. Later, Pierens et al. recorded NMR spectra in solvents with different polarity for imidazoles, oxazoles, and thiazoles and tried to predict the NMR spectra via DFT calculations using B3LYP/6-31G(d) [85]. This approach turned out to be reliable for most solvents, except for the prediction of 1H spectra in benzene. Table 1 presents the data for 1 from reference [87]. Venkatachalam et al. performed deuterium exchange experiments by adding D2O to a solution of alkyl derivatives 2-AlkNH-pbt (Alk = Me, Et, nPr, nBu) in CDCl3. As a result, the equilibrium between the deuterated and non-deuterated forms of 2-AlkNH-pbt was observed in the NMR spectra. Based on the chemical shifts for atoms directly involved in intramolecular hydrogen bonding or connected to the amino group in 1H, 13C, 15N, and COSY-spectra, it was concluded that the hydrogen bond between the 2-AlkNH group and N3 was a strong interaction [75].

2.2. Synthesis of 2-(2′-Aminophenyl)benzothiazole and Derivatives

2.2.1. Parent 2-(2′-Aminophenyl)benzothiazole 1

Meyer et al. first reported the synthesis of 1 in 1934. The rearrangement of N-phenyl-2-benzothiazolamine (4) upon heating in a sealed ampoule in the presence of two equivalents of concentrated HCl gave 1 in a 40% yield (Scheme 1) [88].
For the efficient synthesis of compound 1, the most commonly used starting compound is 2-aminothiophenol (5a), which typically undergoes a condensation reaction with anthranilic acid (6). Hein et al. first developed a simple method for synthesizing heterocyclic compounds, including 1 and pbt [89]. They used polyphosphoric acid (PPA) as both a dehydrating agent and a solvent (Scheme 2), facilitating the intermolecular condensation of various compounds. In the reaction between 5a and 6, the yield of 1 was 52%. They carried out the condensation reaction at a high reaction temperature to avoid side reactions. However, it was later shown that decreasing the reaction temperature from 250 °C to 150 °C does not affect the product yield [84,90]. This synthetic route later became one of the most popular and is actively used in contemporary research [77,86,87,91,92,93,94].
Instead of using acid 6 as a starting material, the corresponding aldehyde, anhydride, or even alcohol can be used as well, as described below. Under the reaction conditions, 6 is formed in situ, and further, the condensation reaction takes place, giving the desired amine.
Another widely used approach to synthesize 1 is to use isatoic anhydride (7) as a starting material under various reaction conditions. Gajare et al. proposed a method which includes using natural kaolinitic clay as a catalyst. Compound 7 reacted with 5a in dried chlorobenzene at 120 °C, yielding 55% of 1 (Table 2, entry 1) [95]. Fadda et al. also used 7 in glacial acetic acid with sodium acetate as a catalyst, forming the corresponding anthranilamide (8), which then underwent ring closure to yield 1 with an 80% yield (Table 2, entry 2) [96]. Finally, Tseng et al. provided a methodology that produced a 68% yield of 1 using ZnCl2 as a catalyst (Table 2, entry 3) [97], although it was later shown that the reaction could proceed without a catalyst with good yields [55,98,99,100,101].
The use of 2-aminobenzaldehyde (9a) was applied in several synthetic methods. Ray et al. developed an eco-friendly catalytic way for synthesizing pbt derivatives in water. The authors modified mesoporous silica material, introducing additional 2-(piperazin-1-yl)pyrimidine functionalized organosilane tails. These tails were responsible for thiolate formation and ring closure steps in the reaction mechanism. The reaction was performed using 2-aminothiophenol disulfide (10) and substituted 9a in a 1 to 2 ratio under O2 at 100 °C for 4 h (Table 3, entry (i)). For compound 1, the reaction yield achieved 89%, while for other pbt derivatives, it was in the range of 85–98% [102].
In 2016, Basha et al. utilized a microwave-assisted method for the synthesis of 1 (Table 3, entry (ii)). The reaction involved irradiation of 5a, aldehyde 9a, and catalytic amounts of glacial acetic acid for 10 min to afford the product in quantitative yield [103].
An alternative synthetic method using a different sulfur source was developed by Park et al. They described a one-pot reaction using substituted iodoanilines 11, aromatic benzaldehydes 9, and NaSH·nH2O as a sulfur source, leading to the synthesis of pbt derivatives with good yields (Scheme 3), even with fluoro-, chloro-, and trifluoromethyl substituents in the 2-benzothiazole (bt) fragment. In this procedure, which was the first example of using NaSH as a sulfur source for forming C-S bonds, CuCl (2 mol%) was used as a catalyst in DMSO at 110 °C for 6 h, providing 1 in a high 82% yield [104]. From a mechanistic point of view, the reaction would start with the condensation of 11a and 9a, affording imine 12 (Scheme 3, step (i)). Then, the copper catalyst would promote the reaction between NaSH·nH2O with 12, giving thiophenol 13 (Scheme 3, step (ii)). Subsequently, intramolecular condensation to 14 (Scheme 3, step (iii)) and dehydrogenation would provide the final product 1 (Scheme 3, step (iv)).
Bakthadoss et al. successfully synthesized about 45 heterocyclic compounds using a molten state reaction, which did not require the use of a catalyst or solvent. Compound 1 and 2-(2′-(N-tosyl)aminophenyl)benzothiazole (15) were obtained in excellent yield from the reaction of 10 with the corresponding aldehyde (9a or 9b) at 180 °C (Scheme 4) [105].
A straightforward method of synthesizing amino derivatives in organic chemistry is via a sequential strategy (Table 4) involving the chemical reduction of the corresponding nitro compound. Several groups have used this approach [28,106,107,108], introducing 2-nitrobenzaldehyde 16a (Table 4, entries 1–3) or nitrobenzoyl chloride 16b (Table 4, entry 4) into the reaction with 5a. The isolated 2-(2′-nitrophenyl)benzothiazole (17) was then reduced using common reducing agents such as Sn(II), Fe(0) in an acidic medium, or hydrazine. Compound 1 was formed with good yields ranging from 70 to 85%.
Another catalytic approach for the synthesis of heterocyclic compounds involved the use of 2-aminobenzyl alcohol (18). Shi et al. utilized substituted primary alcohols 18, 5a, and a catalytic quantity of NaOtBu (20 mol%) for synthesizing a series of bt derivatives. The reaction was carried out in toluene at 100 °C for 24 h using an air balloon (Table 5, entry 1). The authors demonstrated the versatility of the methodology on more than 50 benzazoles in mostly good-to-excellent yields. However, this method gave only 43% of the reaction yield in the case of 1 [109].
Anandaraj et al. introduced a palladium-catalyzed approach for the synthesis of 25 pbt derivatives [110]. The pincer-ligand bearing Pd(II) complex A was employed in the acceptorless dehydrogenative coupling strategy (Table 5, entry 2). In the plausible reaction mechanism (Scheme 5), base treatment of benzyl alcohol 18 would result in the formation of an alkoxide, which would displace triphenylphosphine from palladium complex A, affording new complex B. Dehydrogenation in B, would lead to the hydride complex C and aldehyde 9a. The latter would react with 5a, yielding imine 13, followed by rearrangement into 14. Subsequent reaction of heterocyclic compound 14 with hydride complex C would provide Pd-complex D, which would rapidly dehydrogenate to produce 1 and regenerate complex B upon coordination of another alkoxide.
In 2016, a photocatalytic method for obtaining bt derivatives was described. In this case, a reactive 2-azolyl radical, generated from 2-bromoazole 20 under light irradiation in the presence of an iridium sensitizer and N,N-diisopropyl-N-ethylamine (19), reacted with various arenes. Whereas the reaction selectivity was excellent for most compounds, in the case of aniline (11b), the authors obtained a mixture of two amines, with the amino group placed in ortho- and para-position to the benzothiazole. The total yield of the two isomers 1 and 21 was 60%, which could be separated using column chromatography (Scheme 6) [111].
Nacsa and Lambert described a biaryl rearrangement under mild conditions (Scheme 7), which included the synthesis of pbt derivatives. Sulfonic acid derivatives were used as starting materials in these cross-coupling reactions in the presence of tris(trimethylsilyl)silane (TTMSS), pyridine, and a catalytic amount of iodine. The proposed reaction mechanism was based on the intramolecular transfer of an aryl radical generated at room temperature via a combination of TTMSS and air. Ten pbt derivatives were isolated, with yields ranging from moderate to good. In the case of compounds 1 and 2, the reaction yield reached 90% [112].
2,1-benzisoxazole (25) was used as a starting reagent for the synthesis of 1 (Scheme 8). The authors found that 25 converted into a highly reactive ketene 26 upon refluxing in dimethyl-2-imidazolidinone (DMI), which then reacted with 5a to anthranilamide 8 and further to 1 with a good yield of 75% [113].

2.2.2. Derivatives of 2-(2′-Aminophenyl)benzothiazole

In particular cases, the syntheses of amino derivatives of heterocyclic compounds from parent amine can be challenging due to the relatively high stability and low reactivity of the heterocycle, unfavorable chemical reactions, chemical rearrangements, etc. As a consequence, the direct synthetic methods for these derivatives can also be helpful. Among several catalytic works reported in the literature, the Rh-catalyzed C−H amidation approach developed by Liu et al. enabled the synthesis of a series of amides 30 with various substituents either in the benzothiazole ring or in the phenyl group (Scheme 9). From various dioxazolone derivatives (28) and substituted 2-phenylbenzothiazoles (29) as starting materials, 22 amido derivatives 30 were isolated in moderate-to-high yields. The reaction was proposed to proceed through the in situ generation of a rhodium active species [Cp*Rh(III)], which would react with 29, giving a cyclometalated rhodium intermediate E. Next, dioxazolone derivative 28 would coordinate with the metal center, yielding F. The release of carbon dioxide would lead to the rearrangement of the coordination sphere of the rhodium center, facilitating N-C bond formation. Finally, the protonation of intermediate G would afford the desired product 30 [62].
The selective strategy for the synthesis of a series of pbt derivatives was proposed by Wang et al. They loaded isatins (31) into a sealed tube and heated it at 130 °C with 2-aminothiophenols (5) in DMSO in the presence of copper-catalyst and morpholine as a reaction promoter (1 equiv.) or reagent (2.5 equiv.) (Scheme 10). The utilization of CuCl2 led to the formation of 2-amino-pbt derivatives 32, particularly the reaction yield achieved 74% for 1 and 68% for 3. The CuBr catalyst facilitated further proceeding to a three-component reaction with morpholine, affording urea derivatives 33 in variable yields [79].
Isatin (31) was also used as a precursor for the synthesis of amino-pbt derivatives 32 in another catalytic approach. Singh et al. introduced this compound into a metal-free reaction with substituted anilines (11) and S8 as a sulfur source [114]. The reaction was performed in DMSO for 3–6 h in the presence of potassium iodide (40 mol%) as a catalyst, and a set of compounds 36 was isolated in good to high yields (Scheme 11). The only limitation of this method was the para-halide-substituted anilines, which did not react under these conditions. Mechanistically, the reaction of 31 with 11 would lead to the formation of imine 34, followed by nucleophilic attack by S8, resulting in C-S bond formation. Next, an intramolecular electrophilic attack of the aromatic ring on S8 would generate the spirocyclic intermediate 35 with loss of S7 oligomer. Subsequent nucleophilic addition of H2O, followed by C-N bond cleavage of the isatin part, would provide intermediate 36. Finally, decarboxylation and dehydrogenation would yield products 32.
Click chemistry was shown to be another straightforward method for obtaining 2-amino-pbt derivatives. Draganov et al. applied the boronic acid group both as a catalyst and as a site for further potential functionalization [115]. Aldehydes 37 with a boronic acid group in ortho-position reacted with substituted 2-aminothiophenols 5 (Scheme 12); the full reaction conversion to compound 38 was reached within only 5 min at r.t. The authors showed the versatility of the approach and the possibility to perform the coupling methodologies, such as the formation of C-C or C-N bonds, including the synthesis of the functionalized 2-amino-pbt derivative 39a.
In conclusion, even though a large range of synthetic methods has been developed, the most convenient and simplest route to 1 is the condensation/cyclization reaction between 2-aminothiophenol 5a and either a carboxylic acid 6 or a benzaldehyde 9a, giving the desired product on a gram-scale in moderate to good yields. It is also worth noting that the copper-catalyzed synthetic method (Scheme 10) and the pathways involving isatins as a starting material (Scheme 11) allow for the preparation of derivatives of 1 on a 6 to 60 mmol scale. In some cases, catalytic reactions can be advantageous for synthesizing more complex derivatives. Further functionalization of 1 can be achieved using organic chemistry approaches, as will be described in the following section.

2.2.3. General Pathways Toward Functionalization of 2-(2′-Aminophenyl)benzothiazole

Compound 1 is an aromatic amine and exhibits chemical properties characteristic of this class of compounds. For example, 1 formed cocrystals with planar aromatic systems, such as methyl 2-((3-chloro-4-methyl-2-oxo-2H-chromen-7-yl)oxy)acetate (40) (Figure 3), as it was presented by Al-Amiery et al. [116]. In these cocrystals 1·40, two aromatic molecules formed zigzag chains via hydrogen bonds between the amino group of 1 and the keto-group of coumarin derivative, along with π-π-stacking interactions.
Deprotonation of 1 occurs readily from the reactions with strong bases in organic solvents, providing the corresponding amido alkali salts 41 (Scheme 13) [84,85]. The latter can be further introduced in nucleophilic substitution reactions to synthesize phosphorus and silylated derivatives (see below). Furthermore, the amino group in 1 can be modified by nucleophilic substitution reactions yielding amides 42 (Scheme 13). From a sensor development point of view, these compounds are especially interesting because their photophysical properties, along with chemical properties, allow for their use as selective probes [77,108,117,118,119]. Moreover, the electrophilic substitution of hydrogen in the phenyl ring of 1 to the corresponding iodo and further to the cyano derivatives can be performed in the reaction of 1 with KI in acid medium and then with CuCN in DMAc (Scheme 13, 43, and 44) [97]. Other transformations of the amino group to the azide 45 [84,91], imine 46 [27,120,121,122], and alkyl amine derivatives 47 can be readily achieved from 1 (Scheme 13) [75,92].

2.2.4. Functionalization of 2-(2′-Aminophenyl)benzothiazole with Phosphorous Groups

The phosphorous chemistry of compound 1 has started to develop over the last two decades. The first examples were reported by Mittal et al. in 2008. They synthesized phosphorylated and oxo-/thiophosphorylated derivatives of 1 (Scheme 14, 48af and 49) and studied their antifungal activity. However, the SCXRD data were not given for any of these examples; all derivatives were characterized by IR-, NMR-spectroscopy, and mass spectrometry [41].
A significant contribution to the development of phosphorus derivatives of 1 has been made by the Konchenko group. Several new phosphorus-containing compounds have been synthesized using nucleophilic substitution reactions (Scheme 14, 5052) [82], Staudinger reactions (Scheme 15, 53ab) using potassium salt 41 [84,85] or phospha-Mannich reactions (Scheme 15, 54 and 55ab) [50,83]. α-Aminophosphonates (Scheme 15, 55cd) were obtained by Basha et al. in order to study their biological activity [103]. The chemical structures of these compounds are shown in Scheme 14 and Scheme 15.

2.2.5. Functionalization of 2-(2′-Aminophenyl)benzothiazole with Silyl Group

From a coordination chemistry perspective, another promising bidentate proligand is the silanediamine 56 (Scheme 16), which, in its doubly deprotonated form, can coordinate to one or two metal centers. Generally, silanediamide ligands are capable of stabilizing low-valent oxidation states as well as mixed-valence dimeric complexes M(I)/M(II) [123,124,125,126]. The conformational flexibility of the silanediamide ligand allows for the coordination of metals with large atomic radii [125,127,128,129,130]. The incorporation of a photophysically active core into the silanediamine framework could lead to remarkable properties. To explore this potential, Mironova et al. synthesized compound 56 derived from 1 [48,131]. They used a common approach [132,133]; the deprotonated amine reacted with dimethyldichlorosilane in a nucleophilic substitution reaction, providing 56 in an 84% yield.
Overall, the mentioned compounds are also promising as pro-/ligands for coordination chemistry due to the presence of phosphorus, nitrogen, and, in some cases, chalcogen (O or S) or silicon atoms, which could lead to intriguing chemical reactivity [50,131] or formation of interesting coordination architectures with great potential in luminescent applications, as discussed below.

2.3. Luminescence Properties: ESIPT, AIE

Compared to other organic chromophores, the most intriguing properties of derivatives of 1 are characteristic Excited-State Intramolecular Proton Transfer (ESIPT, Section 2.3.1.) and Aggregation-Induced Emission processes (AIE, Section 2.3.2.). These processes, which are mainly observed in substituted amino derivatives (2-RNH-pbt) bearing electron-withdrawing groups, will be discussed in detail below.

2.3.1. ESIPT

This photochemical process occurs due to the difference in proton acidity between the ground state and the excited state (Figure 4).
The photochemical cycle involves four states: excitation of the ground state (NS0) leads to the formation of the normal form of the excited state (NS1), followed by ESIPT, resulting in the tautomer excited state (TS1). Fluorescence then occurs, forming the tautomer ground state (TS0), and reverse proton transfer (RPT) takes place. The energy of the transition from the excited state of the tautomer (TS1) to the ground state of the tautomer (TS0) is smaller than that of the normal form (i.e., NS1→NS0); for this reason, the ESIPT band appears as a red-shifted band in the spectrum. Due to the ultrafast rate of ESIPT (<<1 ps), studying this process exceeds the capabilities of most modern detection systems. This remarkable process is often displayed in various OH-type organic compounds [134,135,136], among which 2-(2′-hydroxyphenyl)benzothiazole derivatives [137,138,139,140,141,142,143] play an important role. Intramolecular hydrogen bonds in OH-type frameworks are generally stronger than those in NH-type systems. This increased hydrogen bond stability facilitates ESIPT in the former compounds. In contrast, in NH-type compounds, the energy barrier for ESIPT is higher [144]. However, the advantage of the 2-RNH-pbt framework over 2-(2′-hydroxyphenyl)benzothiazoles lies in its tunability: the introduction of donor/acceptor substituents both in the phenyl ring and amino group can influence the HOMO-LUMO gap. The ESIPT process was clearly demonstrated in aminopbt derivatives by Tseng et al. [97]. They studied a series of compounds (Table 6, 12, 15, 44, and 5760) and showed that the possibility of ESIPT increases in the following order: 2 < 1 < 57 < 15. The strong acceptor substituent (i.e., Tos) at the amino group accelerates ESIPT, while a donor substituent (i.e., Me) prohibits this process. Ultrafast ESIPT was observed in 15, there is only emission of the tautomer form of 15 with a large Stokes shift (Table 6). While the derivative with a less electron-withdrawing group 57 exhibited a slower process, there were two bands corresponding to fluorescence from NS1 and TS1 states (i.e., normal and tautomer, Table 6).
Examples of tuning through the introduction of donor/acceptor groups in the phenyl ring include compounds bearing NH2- and CN-groups in para-position to the tosyl-substituted amino group in 15: 2-(2′-(N-tosyl)amino-5′-aminophenyl)benzothiazole (Table 6, 58) and 2-(2′-(N-tosyl)amino-5′-cyanophenyl)benzothiazole (Table 6, 59). Even though all compounds in this series exhibit only the ESIPT process and do not demonstrate fluorescence from the NS1 state, the introduction of substituents can tune the maximum of emission: emission bands are shifted towards the red region, in the order 591558 (540 nm→555 nm→649 nm). In contrast, the introduction of a CN substituent in the 5′-position 2-(2′-(N-methyl)amino-5′-cyanophenyl)benzothiazole (Table 6, 60) leads to a weak ESIPT process, while 2 shows only the strictly non-proton transfer spectra. Moreover, a combined approach including kinetics, spectroscopy data and theoretical calculations was applied for the determination of the ESIPT mechanism. The kinetic experiments with non-deuterated 60a (Table 6) and deuterated 60a (Figure 5) showed identical resolved kinetic traces, indicating the absence of a kinetic isotope effect. The authors suggested that the rate-limiting step of ESIPT in aminopbt derivatives would be influenced by a change in the molecular core, with N2′ and N3 moving closer to each other to transfer a proton rather than by the proton transfer itself [146]. As a result of the research, the modification of the 1 core can control the rate and energy barrier of the ESIPT, as well as photophysical properties [97].
As it was shown, the introduction of an amide group at the 2′-position in pbt derivatives facilitates the ESIPT process, as observed in 15 and 57. This behavior was also demonstrated in benzoyl-amides (Table 6), which turned out to be promising compounds for OLED technology. Bright white-light emission with high fluorescence quantum yields (Φf = 23–35%) was observed in 30e, 30f, and 30g in the solid state (Table 6). All three compounds exhibited two-band emission in the spectra [62], while the 3 (Table 6) did not undergo the tautomer form at the excitation neither in solid state nor in solution (CHCl3) [76], as confirmed by the theoretical study [147]. In contrast, Tseng et al. observed a slow ESIPT in 44 (Table 6), where the acceptor group (i.e., CN) was placed in 5′-position. Thus, it could be concluded that only the introduction of acceptor groups in the 5′-position of the phenyl ring and at the amino group can lead to the appearance of a second band (ESIPT-band) in the spectrum.
Tseng’s study attracted significant attention for the investigation of ESIPT-active derivatives of 1 using DFT calculations. Most of the calculations were performed with the TD-DFT method in Gaussian 09 program package. The first theoretical research works demonstrated that ESIPT-active compounds can undergo isomerization around the C2-C1′ bond in the excited state (Table 7). This twisting brings the S1 and T1 potential energy curves close together, enabling their intersection. As a result, ISC occurs from Ts1, populating both TT1 and twisted-TT1. Then, these states relax via nonradiative decay to form Ts0 and twisted-Ts0. The authors calculated that twisted-Ts0 must overcome a significant energy barrier to return to Ns0 form. Based on their data, they concluded that ESIPT-active compounds in the ground state may exist as long-lived twisted-Ts0 species [148,149]. Moreover, another study found that a higher free energy of the ESIPT reaction correlates with a higher reaction rate [136]. Calculations of ESIPT energy barriers for 1 and 2 suggest that the methyl donor group hinders the ESIPT process, making it less favorable [148,149,150]. Obviously, the ESIPT in derivatives of 1 can be tuned by introducing donor or acceptor groups at the amino group or 5′-position. However, the effect of the substituent does not always directly correlate with the strength of the intramolecular hydrogen bond, which means the prediction of the ESIPT trends based on only substituent effects may lead to an incorrect effect [136].
Singh et al. studied the luminescent properties of a set of pbt derivatives (Table 8). A drastic difference in absorption spectra was observed for 33m, 33o, and 33p, whereas for other compounds, the absorption maximum remained around 380 nm. In fact, this study provides an intriguing confirmation of the lack of a correlation between substituents in pbt and the occurrence of ESIPT. In contrast to compound 44, the authors did not observe the dual-band emission of compounds with F- and Cl-substituents at the 5′-position and donor substituents at bt heterocycle (33g33i and 33l) in the range of 250–525 nm in chloroform solution. Interestingly, the introduction of any substituent (whether a donor or acceptor group) at the 5′-position did not result in an emission shift for compounds 33g (440 nm), 33h (440 nm), and 33k (440 nm), while the emission of parent 33j is blue-shifted (425 nm) compared to its substituted derivatives (Table 8). Conversely, replacing the methoxy group with F- or Cl-substituents in 33o, 33i, and 33l leads to a hypochromic shift in the emission spectra. The introduction of two donor groups into the bt core led to a red-shifted emission for 33m-33o compared to 33p (Table 8, 470 nm and 465 nm) [114].
Functionalization of 1 with diphenylphosphine further modifies the luminescent properties of benzothiazoles (Table 9). Specifically, all reported compounds in the solid state demonstrate dual-band emission due to the occurrence of both normal (NS1) and ESIPT-induced (TT1) excited states. In addition, 50, 51, and 52 reveal the possibility of twisted-TS1 as well. According to TD-DFT calculations, the excited-state twist in the compounds occurs around the C–N(H) bond, resulting in the appearance of one more excited state capable of radiation decay. Changing the excitation wavelength provides the possibility to change the relative intensity of the two emission bands, which results in a glowing color change from green through white to blue (Figure 6). The observed lifetime of the excited state lies in the range of 30–700 μs, implying a phosphorescent nature of the emission, which is especially important for high-performance OLED devices. The emission of α-aminophosphonates (55a and 55b) and iminophosphonamine 53a is bathochromically shifted compared to the P(III) derivative 54. The combination of the transitions from the excited states enables tunable emission depending on the excitation wavelength [82,84].

2.3.2. AIE

Not only is ESIPT characteristic of fluorophores based on pbt derivatives, but the 2′-amino-pbt compounds can also have Aggregation-Induced Emission (AIE) properties [98]. AIE is a complex and not fully understood phenomenon, which has been reported in conjugation with different additional photophysical processes, and a complete mechanism of AIE has not been described yet. The compounds exhibiting AIE emit light more intensively by aggregation, which typically occurs when water fractions are added to their organic solutions (Figure 7a) [151]. Intra- and intermolecular interactions, such as π–π, dipole–dipole, hydrogen bonding, halogen bonding, and others, are known to affect the AIE behavior [152]. Pbt derivatives characterized by ESIPT (i.e., those with electron-withdrawing groups at 2′-amino group) tend to demonstrate AIE behavior, which is related to the twisted intramolecular charge transfer (TICT) process. In contrast, the derivatives with electron-donating groups at amino group in the 2′-position, where ESIPT does not occur, exhibit aggregation-caused quenching (ACQ) behavior (Figure 7b) [98].
Hydrogen bonding is one of the frequently encountered examples that promote AIE. Several types of hydrogen bonds (Figure 8) were found in the tetrabutylammonium sulfonate–urea amphiphilic salt 61 derived from 1, which plays an important role in the formation of self-associated dimers [78]. Another example combining ESIPT and AIE in pbt derivatives was discussed by Liang et al. They synthesized a series of 2-RNH-pbt-based semisquaraines with different numbers of electron-deficient chloro-substituents in a four-membered ring (Figure 8, 6264). The greater number of chloro-substituents in the ring promotes easier ESIPT, and the structure has a more distorted configuration. This is an example of the AIE phenomenon, which occurs due to extremely weak π-π interactions in the solid state, enhanced by multiple halogen bonds [153].

2.3.3. Miscellaneous Photophysical Properties

Another way of tuning the emission band of 1 is via the synthesis of organic nanoparticles (Lewis-pair nanoparticles). Considering the amine as a Lewis base, the combination of 1 with another Lewis acid (diphenylborinic anhydride) leads to the formation of nano-Lewis-pair adducts, which have already exhibited dual-emission. The variation in chromophore concentration during the step of nanoparticle synthesis results in different fluorescence colors of products, going from light-blue, yellowish-pink, to orange [154].
Interesting results were obtained by Tseng et al., who studied boron complexation of 4′/5′-substituted 2-MeNH-pbt derivatives with BF3·Et2O (Figure 9, 65). In the synthesized complexes, the ligand had a nearly planar structure, which reduced the non-radiative decay channel associated with structural relaxation. These complexes exhibit higher quantum yields than the starting amines, and their emission changes from green, yellow, to orange depending on the substituents on the 4′/5′-position [155].
Another intriguing study was performed by Kuz’mina et al. They carried out the condensation of 1 with phthalic anhydride and isolated two different crystalline forms and one amorphous form of the synthesized product 66 (Figure 10, top and bottom). Phase transitions were found between the three forms: the local melting of the crystal led, simultaneously, to the crystallization of another form with a higher melting point (Figure 10, left) [81].
Thus, the various 2-amino-pbt derivatives demonstrate remarkable luminescent responses toward changes in their geometry and\or intermolecular contacts. This intriguing feature is generally studied in terms of AIE or ACQ, arising upon transition from a solution to a solid state. Specific mechanisms for AIE or ACQ are still not straightforward, and the appearance of these phenomena in a certain compound is hardly predictable. However, when designing pbt compounds and interpreting their luminescent properties one may consider the following possibilities: (1) hydrogen bonding, responsible for ESIPT, which is generally more pronounced in the case of derivatives with acceptor substituents in 5′-position of the phenyl ring and at the amino group; (2) formation of π-π interactions upon aggregation in compounds with extended conjugate π system also may be responsible for AIE or ACQ; (3) different TICT processes may become more prominent for derivatives with non-rigid moieties, giving rise to low-lying excited states capable of radiation decay.

3. Overview of the Coordination Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives

The coordination chemistry of 1 and its derivatives is still in its early stages, even though it offers a promising way to synthesize highly functional compounds. The mutual influence of metal ions and ligands bearing a photoactive core can enhance the luminescent properties and lead to new attractive effects, especially also for biological applications. However, only several examples of characterized complexes are known, including studies on the coordination compounds of Co(II) [156], Ni(II) [157,158], Zn(II) [145], Re(I) [159,160], Re(V) [161] with 1 as a ligand. A few Pt(II) and Pd(II) complexes of various derivatives of 1 have been synthesized, and a series of coordination compounds of rare-earth metals with silanediamine [48,131] and iminophosphonamide [84,85] derivatives of 1 were recently reported.

3.1. Coordination Complexes with Neutral Ligands Based on 2-(2′-Aminophenyl)benzothiazole

The first examples of metal–organic compounds with the parent ligand 1 were the picrate complexes [Co(1)2(PIC)2] (Figure 11, 67) [156] and [Ni(1)2(PIC)2] (PIC = (NO2)3C6H2–O) (Figure 11, 68) [157], for which the authors proposed two possible cis- and trans-isomers. None of the known Co(II) and Ni(II) complexes have been crystallized, so their structures were confirmed by combined physicochemical methods. In all cases, 1 coordinates in a chelate manner via the N3-N2′ atoms. Another heteroligand [Ni(1)(H2O)(L)] (Figure 11, 69a—L1 = Gly, 69a—L2 = Ala) complex has been studied for application against pathogenic fungi, such as Aspergillus niger and Fusarium oxysporum [158].
Later, two tricarbonyl complexes of Re(I) with 1 (Figure 12, 70a and 70b) were synthesized [159,160]. The luminescent properties of fac-70a were investigated. The absorption spectra showed transitions characterized by LLCT and IL transitions involving 1. DFT calculations were used to correlate the transitions in both absorption and emission spectra. The obtained data allowed the authors to suggest potential applications in optoelectronic devices [160].
Two more metal–organic compounds with parent 1 have been reported, both of which are oxorhenium(V) complexes (Figure 12, 71a and 71b). In both complexes, the central Re(V) atom is placed in a six-coordination sphere with distorted octahedral geometries, where 1 and two chloro ligands occupy the equatorial plane. The absorption spectra revealed bands in the range of 300–450 nm, which were assigned to metal-to-ligand charge transfer (MLCT) transitions involving 1, while highly intense absorption bands in the range of 250–300 nm originate from LMCT and LLCT transitions [161].
Recently, two zinc halide complexes with 1 were published [145]. Even though the reactions were carried out in the same way—reaction in ethyl acetate with a zinc halide (ZnCl2 or ZnBr2) to 1 ratio of 1:1 (Figure 12, 72a and 72b)—two different structures were defined via SCXRD. Compound 1 coordinates to zinc in a chelating manner, but the organic ligand in the chloro complex 72a exhibits a planar structure, whereas, in the bromo 72b complex, the planes of the benzothiazole fragment and the phenyl ring are positioned at an angle of 23–31° relative to each other. Despite the conjugation between the heterocyclic fragment and the phenyl ring in the ligand, 1 demonstrates the flexibility of the organic backbone around the C1′-C2 bond (Table 1). The reason for the structural difference remains unclear. Authors compared the mentioned complexes with a similar series of halide complexes with 55b ligand (Figure 12, 73a and 73b). The oxidized phosphine ligand coordinated with the zinc center exclusively via the oxygen of the P=O moiety, while the heterocycles did not participate in the coordination. In all complexes, Zn(II) is in a tetrahedral environment. The pbt core in the ligand is not plane in both 73a and 73b (X = Cl (3.6◦), Br (15.0◦)) with smaller angles between the heterocycle and aniline moiety than in bromo complex 72b. The photoluminescence properties were studied for these complexes. The emission spectra corresponding to the 72a and 72b complexes exhibit a hypsochromic shift in the emission maxima compared to the free ligand; a redshift is also observed as the ligand transitions from chloride to bromide. In contrast to amine complexes, 73a and 73b exhibit two-band emission with maxima at 450 and 600 nm. The authors assumed the first band corresponded to a radiative transition for the regular amine species, while the second band can be explained either with the ESIPT or TICT.
In work carried out by Afonin et al., the reactivity of [M(COD)Cl2] (M=Pd(II), Pt(II), COD = cycloocta-1,5-diene) with ligand 54 ligand derived from 1 was investigated. It was found that reactions between Pd(II) and Pt(II) complexes (74a and 74b) with 54 resulted in the activation/cyclization reactions of the P-C bond, and further P−C bond cleavage led to the formation of PPh2 species. This rearrangement of 54 afforded the formation of a heterocyclic fragment, representing a cation with the cycle {N-C-N-C-C-C} (Figure 13, 75a and 75b). Consequently, the ligand could not be introduced into the inner coordination sphere of the metals. A mechanism for the P-C bond activation/cyclization was suggested using quantum chemical calculations [50].

3.2. Coordination Complexes with Anionic Ligands Based on 2-(2′-Aminophenyl)benzothiazole

Other biologically active complexes were synthesized with a methylthioacetamide ligand attached to 1. The Pd(II) and Pt(II) complexes with this ligand (Figure 14, 76a and 76b) were obtained in moderate yields and structurally [162]. Consequently, the discovery of an alternative method for synthesizing a palladium complex with an S,N,N-pincer ligand was reported (Figure 14, 77). This method employed a mechanochemical approach that yielded the complex quantitatively within less than 7 min by grinding in a vibrational ball mill, compared to up to 2 days for solution synthesis [163]. These complexes 76 and 77 were designed to enhance binding with biological targets through stacking interactions. Moreover, cytotoxicity tests were performed on various human cancer cell lines (human colon, breast, and prostate cancer), yielding promising results. The study showed that the coordination to Pd(II) or Pt(II) was crucial for the complexes to function as effective anticancer agents [162,163].
Due to the known effectiveness of palladium complexes against tumor cells, a series of Pd(II) metal–organic compounds with salicylaldimine-based benzothiazole derivatives with different substituents in phenyl ring were synthesized, and their cytotoxicity was studied (Figure 14, 78). The interaction of each complex with circulating tumor DNA was studied in a combined approach involving absorption spectroscopy, competitive fluorescence quenching assays, and molecular docking simulations. The complex with the 5-methoxysubstituted ligand (78a) exhibited the best cytotoxicity against the liver cancer cells, while the complex with the 3,5-difluorosubstituted ligand (78b) demonstrated the highest anticancer activity against the breast cancer cells. These findings highlight the potential of designing 1 derivatives, allowing flexible cytotoxicity tuning of the corresponding Pd(II) complexes [72].
Similarly to 78a and 78b complexes (Figure 15), bis-ligand and hetero-ligand Ni(II) (79a and 80a) and Co(III) (79b and 80b) complexes were synthesized with ligand based on 2,5-dichlorosalicylaldehyde and 1. The structures of these complexes were confirmed through a combination of physicochemical methods (elemental analysis, electronic absorption, mass and NMR spectral studies, and thermal analysis). The research suggests the use of mixed ligand complexes with 1,10-phenanthroline as antibacterial agents against both Grampositive and Gram-negative bacterial strains. Moreover, at increasing concentrations of mixed ligand complexes, the number of cancerous lung tissue cells decreased in vitro [51].
In contrast to zinc complexes with neutral 1 (Figure 12, 72a and 72b) [145], it turned out that two deprotonated [2-RN-pbt] ligands can bind in the inner coordination sphere of Zn(II). For this, zinc chloride was treated with the sodium salts of the ligands (NaL81−85 and NaL15) in a 1:2 ratio. These compounds were used for the preparation of the OLED devices with the blue color of the resulting diode (Table 10) [164,165,166,167]. The authors used the new complexes as emission layers or electron-transport layers with various hole-transport layers (i.e., PTA, NPD, or CBP). The designed cells showed good brightness at low voltage.
Two studies were performed on the luminescent properties of trivalent rare earth (Ln) complexes with an iminophosphonamide ligand (L53a and L53b) by Sinitsa et al. (Figure 16a,b). The reaction of 53 with potassium hydride in THF led to the formation of potassium salt of iminophosphonamide (L53a− and L53b−) and addition of lanthanoid halides in stochiometric amounts afforded complexes 92 (ratio LnCl3: L53a = 1:3) and 93 (ratio YCl3: L53b− = 1:1). Emission spectra with narrow bands were obtained for the samarium(III) coordination compound with ‘symmetrical’ (i.e., having same substituents at the N atoms) L53a− ligand, indicating a low-symmetry coordination environment and that the iminophosphonamide ligand exhibits an antenna effect. In the series of lanthanoid complexes with the ‘symmetrical’ iminophosphonamide ligand (Figure 16, 92ad), it was demonstrated that both the ligand and the lanthanoid atom influence the emission behavior, with the emission color ranging from turquoise to orange depending on the metal [84]. In contrast to known synthetic methods for silanediamines, it is possible to obtain ‘non-symmetrical’ iminophosphonamines, which can be used for studying the structural features of coordination compounds. Thus, 53b based on 1 and mesitylaniline (NH2-Mes) and its yttrium complex 93 were synthesized (Figure 16) [85]. Due to the presence of chloro ligands in the inner sphere of the yttrium complex, it is a suitable starting reagent for the design of various heteroligand complexes in the future.

3.3. Coordination Complexes with Doubly Deprotonated NSiN Ligands Based on 2-(2′-Aminophenyl)benzothiazole

The design of pincer ligands by introducing two 1 fragments enables the polydentate metal coordination, which is particularly suitable for larger ions, such as lanthanoids. The coordination via four nitrogen atoms provides the stabilization of rare-earth metals (Ln(III)) complexes in both solid state and solution. Additionally, the synthesis of polydentate ligands bearing fluorophoric groups aims to produce bis-ligand complexes [LnL2]q (where q represents a charge of lanthanoid inner sphere). The increased rigidity of the deprotonated proligand in these complexes is beneficial for minimizing non-radiative luminescence decay. Thus, silanediamide ligand (L56 2−) was synthesized and introduced into a series of complexes with Y(III) (96a), La (III) (95a and 96b), Nd(III) (96c), Gd(III) (95b), and Ho(III) (94) (Figure 17). 56 was initially deprotonated with nBuLi or KBz in THF at −90 °C; then, lanthanoid chlorides reacted with the alkali salt. As a result, the smaller Ho(III) formed a polymeric complex 97 with two bridged chloro ligands, while the larger cations (such as Y(III), La(III), Nd(III), Gd(III)) coordinated two L56 2− (Figure 17). In these bis-ligand complexes, two ligands interacted with each other via π-π interactions because of a helical arrangement of flat pbt fragments.
However, in the case of Gd(III), the use of nBuLi for deprotonation led to the unexpected formation of product 97 (Scheme 17), as the benzothiazole underwent nucleophilic addition of the n-butyl group [131]. To confirm that this was due to a local excess of nBuLi in the reaction mixture, the same reaction with YCl3 was performed. NMR data supported the formation of a similar yttrium complex, although crystallization was unsuccessful. In the reaction of 56, YCl3 and nBuLi in a 1:1:2 ratio, 98 yttrium complex with bridged µ-OnBu was isolated in small quantities, which was attributed to the presence of a LiOnBu impurity in the starting material. Transitioning from nBuLi to Li(NTMS2) in the reaction yielded the formation of the desired complex 99a or 99b.
The photoluminescent properties of the obtained compounds were studied, and although the estimated energy of triplet level was suitable for sensitizing the luminescence of Nd3+ and Ho3+ cations in the NIR region, no f-f bands were observed, except a weak band for holmium complex 97 [48]. Additionally, the triplet level of the silanediamide ligand L56 2− was estimated using the Gaussian fitting of the emission spectrum of 95b (16,905 cm−1) and 99b (16,600 cm−1), which were recorded at 77 K [48,131].
In short, several transition-metal complexes with the neutral 2-amino-pbt ligand 1 are known, and their properties have been explored. Additionally, complexes with derivatives of 1 have also been studied, all of them bearing either deprotonated or doubly deprotonated ligands. In all cases, 1 and its derivatives typically coordinate in a chelate manner (i.e., N3 and N2′). An exception is the oxidized 1,3-aminophosphine 55b, which coordinates to the Zn(II) center via oxygen atom (73a and 73b). However, not all reactions lead to the expected complexes. For example, the 1,3-aminophosphine 54 can undergo rearrangement when reacting with Pd(II) or Pt(II), forming the ate complex 75a and 75b. Additionally, the benzothiazole heterocycle of silyldiamine 56 can react with strong bases, leading to nucleophilic addition. This makes the coordination chemistry of 1 and its derivatives an interesting area for further studies, including the use of other metal centers or the preparation of heterometallic compounds.

4. Overview of the Applications of 2-(2′-Aminophenyl)benzothiazole Derivatives

4.1. Applications as Small Molecule Sensors

Published studies on organic derivatives of 1 are mostly focused on the development of sensors for various ions (Figure 18). Extensive research on the detection of cations such as Fe3+ [27], Cu2+ [118,119,168], Zn2+ [77,169,170,171], Hg2+ [28,122], Pd2+ [108], as well as HSO4 [27], pyrophosphate (PPI) [168], F [117], CNO [172,173], and cysteine [99], PhSH [174], carboxylic acids [175], CH2Cl2 [176], phosgene [177] clearly demonstrate these capabilities. Although the sensor mechanism for detecting transition metal cations is attributed to the coordination of the 1 derivative with the cation, no structures of coordination compounds have been described in these works. The mentioned sensors are typically amides of carboxylic acids, Schiff bases, or N-alkyl/aryl-substituted compounds. Two general mechanisms of sensor action are put forward. The first one (lavender box in Figure 18) involves the sensor acting as a polydentate ligand and coordinating through several additional donor atoms, leading to a change in luminescence intensity. In the case of anions, the strong non-covalent interactions between the sensor and sensing species lead to a luminescent response. The second mechanism (blue box in Figure 18) involves the removal of a fluorescence quencher in 2-RNH-pbts (see Figure 18b) or the cleavage of the C-N bond of the amide group (see Section 4.2.), which also leads to a luminescent signal change. For example, a dinitrobenzenesulfonate group quenches the luminescence of compound 101 due to its two strong electron-withdrawing groups (Figure 18b) [178]. In other cases, the sensing mechanism could involve various chemical reactions with sensing molecules [30,168,172,177].

4.2. Applications in APIs and in Biosensing

Several examples of biological applications of derivatives of 1 have been described (Figure 19). A number of articles show their significant potential for use as antifungal [107], antimicrobial, and antitumor agents and antioxidants [103,158], fluorescent probes [55,93,101,179,180,181,182,183] for bioimaging and sensors [99,100,121,176,184,185,186] in living cells.
In a comparative study [107], 21 compounds were evaluated against seven fungi (Colletotrichum orbiculare, Rhizoctonia solani, Phytophthora infestans (Mont.) De Bary, Pythium aphanidermatum, Fusarium moniliforme Sheld, Botryosphaeria berengeriana, and Botrytis cinerea) in vitro. Only one derivative of 1 was tested in this research. Difluorinated compound 102 (Figure 19) showed only moderate antifungal activity. In contrast, the compound bearing an indazole group instead of a bt fragment exhibited the highest antifungal activity, being, on average, 1.5 to 3.5 times more effective.
In a study by Thaslim Basha et al., compounds 55c and 55d (Figure 19) were tested for their antifungal (Candida albicans, Candida non-albicans, Aspergillus niger, and Penicillium Chrysogenumby) and antibacterial activity (Bacillus subtilis, Staphylococcus aureus, Pseudomonas aeruginosa, and Escherichiacolib) in vitro. Compounds 55d exhibited the highest activity in both antifungal and antibacterial assays. Moreover, 55d were the best antioxidant agents (especially those that bear the pyridinyl, thiophenyl, and imidazolyl substituents in the R4 position) [103].
Sensors 103 have been applied in the imaging of HeLa cells (Figure 19). Due to the ethyleneglycol tail, the developed compounds can enter the live cell membrane and form micelles, which significantly increases the intensity of its fluorescence [179].
Compound 104 was designed for separate or simultaneous detection of ClO and ONOO in the Inflammatory RAW 264.7 Cells and Zebrafish (Figure 19). The phenothiazine-based coumarin serves as a ClO sensor, and a boronate ester part of the 104 is responsible for tracking ONOO. When detecting ClO, the fluorescence color changes from red to green, while for ONOO, it shifts from red to purple. If both anions are present, the color changes from red to blue. Due to high selectivity and wide applicability, this fluorescent probe was utilized in vivo and in vitro [180].
Two fluorescent probes 105 (Figure 19, R = H and NO2) were synthesized for β-galactosidase detection. These probes consist of pbt core linked via a self-immolative spacer to a hydrophilic galactopyranoside group, making them water-soluble. The β-galactosidase cleaves the chemical bond between galactopyranoside moiety and aromatic spacer, giving the intermediate 106 (Figure 20). Then, spacer self-elimination proceeded to yield a compound 58. As discussed above, 58 is an ESIPT-active molecule (Table 6) with poor water solubility. Compound 58 exhibits an intensive NIR fluorescence signal upon AIE behavior (Figure 7a). Moreover, probes 105 have been successfully applied for long-term tracking in living cancer cells SCOV-3 (ovarian adenocarcinoma) and A549 (lung cancer) [101].
An extensive theoretical screening was performed by Kanlayakan and Kungwan, evaluating over 70 compounds as potential fluorescent probes for physiology, pharmacology, and molecular biology sensing in the visible region, including twenty 2-RNH-pbt derivatives. Authors establish several photophysical criteria for optimal ESIPT-based probes: (i) λabs should be close to the visible region (~380 nm); (ii) λem of tautomer form should range from 580 nm to 900 nm covering the NIR region; (iii) a large Stokes shift is essential to avoid self-reabsorption, which improves fluorescence efficiency of fluorescent probes; (iv) the kinetic and thermodynamic parameters, which provide the occurrence the ESIPT. Among the studied compounds, acetyl-containing derivatives met the first three criteria but exhibited a high proton transfer (PT) barrier, leading to a slow ESIPT. Additionally, the ESIPT in these molecules was endothermic in nature (Table 11). Moving from the acetyl group to a tosyl group lowered the PT barrier, making the ESIPT process more efficient and thermodynamically favorable—except for 108Ts, which did not follow this trend. The results suggest that incorporating a strong electron-withdrawing group (CN) at the 6-position, alongside an electron-donating group (NMe2) at the 5′-position, enhances the strength of the intramolecular hydrogen bond. However, the replacement of the cyano group with NMe2 at the 5′-position lowers the PT barrier, further promoting ESIPT [187,188].
However, the authors do not specify any limitations in terms of bioimaging, such as photobleaching or chemical stability in the excited state, despite the absorption maxima of the developed biological fluorophores located in the blue region (up to 400 nm). Nonetheless, the possibility of introducing different groups into the 2-amino-pbt core allows for tuning the photophysical properties, making these compounds promising candidates for biological studies. It is possible that in the near future, 2-amino-pbt derivatives could become actively used bio-tools.

4.3. Applications in Catalysis

Among all catalysts reported in the literature, only several examples of Ir(III), Rh(III), and Pd(II) complexes with pbt derivatives have been described. These complexes have been used in cycloadditions [189,190], asymmetric Friedel–Crafts reactions, Michael additions, oxidation of alcohols [191], dehalogenation of aryl bromides [192], water reduction [43], and asymmetric transfer hydrogenation [193,194,195]. Until recently, 1 and its derivatives had not been used in catalysis. However, new visible-light organic photocatalysts based on the 1 framework have been developed for [2+2] photocycloadditions based on triplet-triplet energy-transfer (TTEnT) processes [90]. TTEnT covers a range of important chemical processes such as isomerization (E/Z, cyclopropanes, allenes, sulfoxides) [196], deracemization [197], photocycloaddition reactions ([2+2], [4+4]) [198,199,200,201] or homolytic σ-bond cleavage [202]. Particularly, the reaction of [2+2] photocycloaddition is an important tool in organic chemistry, which allows to synthesize products via the formation of two or several new bonds in one reaction step to obtain a variety of organic compounds with selective structure [203,204]. In a recent study [90], the catalytic properties of 2-RNH-pbt were explored in the reaction of cinnamyl imidazole 110 with a large excess of styrenes 111 (10 equiv.), giving 115 (Scheme 18). The proposed catalytic cycle for this [2+2] photocycloaddition reaction was suggested as follows (Scheme 18). Irradiation of the photocatalyst 1 at 450 nm would lead to the formation of the excited state 1[1]* of the photocatalyst, followed by intersystem crossing (ISC) to the triplet excited state 3[1]*. The latter would undergo triplet-triplet energy transfer with substrate 110, providing an excited 1,4-biradical state 112. Subsequent addition of the olefin 111 would afford 113, and intersystem crossing would lead to the open-shell singlet 114. Finally, radical recombination gives the cyclobutane product 115. By varying the substituents on the amino-group in 2-RNH-pbt, it was demonstrated that the catalytic efficiency of pbt derivatives depends on the compounds’ light-absorbing properties and their tendency to undergo the ESIPT process. The triplet energy of 2-RNH-pbt compounds was calculated using TD-DFT, revealing that compounds have compatible triplet levels for the transfer of triplet energy to cinnamyl imidazole 110 under visible-light irradiation (450 nm). Photocatalysis was shown to proceed through the noncovalent interaction and triplet–triplet energy transfer between photocatalyst and substrate, as confirmed by NMR spectroscopy, time-resolved luminescence studies, and DFT calculations. A series of cyclobutane products with different functional groups were obtained in moderate to good yields, indicating that 2-NHR-pbt is a promising class of organic photosensitizers [90].
In summary, several potential applications have been discussed (i.e., as small molecule sensors, in APIs, in biosensing, and in catalysis). However, the practical implementation of these compounds is still in the early stages. We hope this review will encourage further interest in the active investigation of 1 and its derivatives, with practical applications remaining a goal for future research.

5. Conclusions

This review highlights the many synthetic possibilities to access 2-(2′-aminophenyl)benzothiazole 1 and its derivatives as well as their increasingly studied applications in coordination chemistry, in material sciences, for instance, in OLED technology, in the detection of cations, anions, and small molecules, in bioimaging, in the development of active pharmaceutical ingredients, and even in catalysis.
Compound 1 is a promising fluorescent core belonging to the NH-type of ESIPT systems. Although the intramolecular hydrogen bond in NH-type compounds is weaker than in other ESIPT systems, its main advantage is the ease of tuning the ESIPT rate. In this context, one can manipulate the relative probability of two processes (‘normal’ and ESIPT radiation decay) by external stimuli, thus tuning the emission color. Further studies of benzothiazoles may use the reviewed results to design molecular switches based on altered emission mechanisms.
The coordination chemistry of 1 and its derivatives has been actively studied since 2000 but is still limited to a small amount of different transition and f-elements. Based on SCXRD data, aminopbt derivatives typically coordinate in a chelate manner (i.e., N3 and N2′), however, other coordination modes are also envisageable. This makes coordinating chemistry of 1 and its derivatives an interesting area for further studies, including the use of other metal centers, the preparation of heterometallic compounds, and applications in catalysis. In terms of luminescent applications, incorporation of an ESIPT-active pbt-based molecule in a coordination sphere of a metal ion seems promising since the resulting complex would feature one more radiation pathway, viz. phosphorescence while keeping the ESIPT site. The spectral profile of the luminescence, and consequently the emission color, can be varied by controlling the probability of three possible pathways: ‘normal’ fluorescence, ESIPT fluorescence, or phosphorescence. Compared to purely organic derivatives, the properties of such metal coordination compounds are much less explored, and thus more attention may be given to their design, including the study of the relationship between the emission processes.

Author Contributions

Conceptualization, F.J. and E.K.P.; validation, F.J., T.S.S., A.P. and S.N.K.; writing—original draft preparation, E.K.P.; writing—review and editing, F.J., T.S.S., A.P. and S.N.K.; visualization, E.K.P.; funding acquisition, T.S.S., F.J. and S.N.K.; supervision, F.J. and S.N.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the Russian Science Foundation (project no. 25-23-00389).

Acknowledgments

E.K.P., T.S.S. and S.N.K. thank the Ministry of Science and Higher Education of the Russian Federation, N 125021302132-4. E.K.P. thanks the French Embassy in Russia for a Vernadski PhD scholarship.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the writing of the manuscript; or in the decision to publish the results.

Abbreviations

2-NH2-pbt2-(2′-aminophenyl)benzothiazole
absAbsorption
ACQAggregation-caused quenching
AIEAggregation-Induced Emission
AlaAlaninate
AlkAlkyl substituent
ArAryl substituent
bt2-benzothiazole
CBP4,4′-bis(N-carbozolyl)-1,1′-biphenyl
CCDCCambridge Crystallographic Data Center
CODCycloocta-1,5-diene
Cp*1,2,3,4,5-pentamethylcyclopentadienyl
DBU1,8-Diazabicyclo(5.4.0)undec-7-ene
DCE1,2-dichloroethane
DFTDensity-functional theory
DMIDimethyl-2-imidazolidinone
DMSODimethylsulfoxide
DNADeoxyribonucleic acid
EElement
emEmission
equiv.Equivalent
ESIPTExcited State Intramolecular Proton Transfer
FurFurfuryl
GlyGlycinate
HEPES buffer(4-(2-hydroxyethyl)-1-piperazineethanesulphonic acid)
hetHeterocyclic fragment
IRInfrared
ISCIntersystem crossing
IUPACInternational Union of Pure and Applied Chemistry
LLigand
LEDLight emitting diode
LLCTLigand-to-ligand charge transfer
LMCT Ligand-to-metal-charge-transfer
LnRare-earth metal
MesMesityl
MLCTMetal-to-ligand charge transfer
MWMicrowave irradiation
NIRnear-infrared region
NMRNuclear magnetic resonance
NPDN,N’-di(1-naphthyl)-N,N’-diphenyl-(1,1′-biphenyl)-4,4′-diamine
NS0Ground state of the normal form
NS1Excited state of the normal form
OFEDOrganic field-effect transistors
OLEDOrganic light-emitting device
p-TolPara-tolyl group
pbt2-phenylbenzothiazole
PETPhotoinduced electron transfer
PICPicrate ligand
PPAPolyphosphoric acid
PPIPyrophosphate
PTProton transfer
PTAOligo(4,4′-(4″-methyl)triphenylamine)
RPTReverse proton transfer
S1 First singlet excited state
SCXRDSingle crystal X-ray diffraction
T1First triplet excited state
TD-DFTTime-dependent density-functional theory
TICTTwisted intramolecular charge transfer
TosTosyl
TS0Ground state of the tautomer
TS1Tautomer excited state
TTEnTTriplet–triplet energy-transfer

References

  1. Abdel-Wahab, B.F.; Mohamed, H.A. Pyrazolothiazoles: Synthesis and Applications. Phosphorus Sulfur Silicon Relat. Elem. 2013, 188, 1680–1693. [Google Scholar] [CrossRef]
  2. Neto, B.A.D.; Carvalho, P.H.P.R.; Correa, J.R. Benzothiadiazole Derivatives as Fluorescence Imaging Probes: Beyond Classical Scaffolds. Acc. Chem. Res. 2015, 48, 1560–1569. [Google Scholar] [CrossRef] [PubMed]
  3. Drakontaeidi, A.; Papanotas, I.; Pontiki, E. Multitarget Pharmacology of Sulfur–Nitrogen Heterocycles: Anticancer and Antioxidant Perspectives. Antioxidants 2024, 13, 898. [Google Scholar] [CrossRef] [PubMed]
  4. Zhylko, V.; Saliyeva, L.; Slyvka, N.; Grozav, A.; Yakovychuk, N.; Melnyk, D.; Melnyk, O.; Litvinchuk, M.; Vovk, M. Synthesis, Antimicrobial and Antioxidant Activity Evaluation, DFT-Calculation, and Docking Studies of 3-Aryl-5,6-Dihydroimidazo [2,1-b][1,3]Thiazoles. Curr. Chem. Lett. 2025, 14, 107–118. [Google Scholar]
  5. Leung-Toung, R.; Tam, T.F.; Wodzinska, J.M.; Zhao, Y.; Lowrie, J.; Simpson, C.D.; Karimian, K.; Spino, M. 3-Substituted Imidazo[1,2-d][1,2,4]-Thiadiazoles: A Novel Class of Factor XIIIa Inhibitors. J. Med. Chem. 2005, 48, 2266–2269. [Google Scholar]
  6. Dekker, F.J.; Ghizzoni, M.; van der Meer, N.; Wisastra, R.; Haisma, H.J. Inhibition of the PCAF Histone Acetyl Transferase and Cell Proliferation by Isothiazolones. Bioorg. Med. Chem. 2009, 17, 460–466. [Google Scholar]
  7. Bhatti, R.S.; Shah, S.; Suresh; Krishan, P.; Sandhu, J.S. Recent Pharmacological Developments on Rhodanines and 2,4-Thiazolidinediones. Int. J. Med. Chem. 2013, 2013, 793260. [Google Scholar]
  8. Kaur, N. Palladium-Catalyzed Approach to the Synthesis of S-Heterocycles. Catal. Rev.—Sci. Eng. 2015, 57, 478–564. [Google Scholar]
  9. Shao, Q.; Pan, W.; Liang, Q.; Li, C.; Zhang, F.; Yang, Y.; Liu, S.; Chen, G.; Fan, B. Design of a Dual-Function Prodrug Fluorescence Probes for Melanoma Detection and Anticancer Drug Release. Chin. J. Anal. Chem. 2025, 53, 100500. [Google Scholar]
  10. Lamansky, S.; Djurovich, P.; Murphy, D.; Abdel-Razzaq, F.; Lee, H.-E.; Adachi, C.; Burrows, P.E.; Forrest, S.R.; Thompson, M.E. Highly Phosphorescent Bis-Cyclometalated Iridium Complexes: Synthesis, Photophysical Characterization, and Use in Organic Light Emitting Diodes. J. Am. Chem. Soc. 2001, 123, 4304–4312. [Google Scholar]
  11. Adachi, C.; Baldo, M.A.; Thompson, M.E.; Forrest, S.R. Nearly 100% Internal Phosphorescence Efficiency in an Organic Light-Emitting Device. J. Appl. Phys. 2001, 90, 5048–5051. [Google Scholar]
  12. Fan, C.; Zhu, L.; Jiang, B.; Li, Y.; Zhao, F.; Ma, D.; Qin, J.; Yang, C. High Power Efficiency Yellow Phosphorescent OLEDs by Using New Iridium Complexes with Halogen-Substituted 2-Phenylbenzo[d]Thiazole Ligands. J. Phys. Chem. C 2013, 117, 19134–19141. [Google Scholar] [CrossRef]
  13. Neto, B.A.D.; Lapis, A.A.M.; Da Silva Júnior, E.N.; Dupont, J. 2,1,3-Benzothiadiazole and Derivatives: Synthesis, Properties, Reactions, and Applications in Light Technology of Small Molecules. Eur. J. Org. Chem. 2012, 2013, 228–255. [Google Scholar]
  14. Sukhikh, T.S.; Komarov, V.Y.; Konchenko, S.N.; Benassi, E. The Hows and Whys of Peculiar Coordination of 4-Amino-2,1,3-Benzothiadiazole. Polyhedron 2018, 139, 33–43. [Google Scholar]
  15. Li, T.-Y.; Wu, J.; Wu, Z.-G.; Zheng, Y.-X.; Zuo, J.-L.; Pan, Y. Rational Design of Phosphorescent Iridium(III) Complexes for Emission Color Tunability and Their Applications in OLEDs. Coord. Chem. Rev. 2018, 374, 55–92. [Google Scholar]
  16. Jang, J.-H.; Park, H.J.; Park, J.Y.; Kim, H.U.; Hwang, D.-H. Orange Phosphorescent Ir(III) Complexes Consisting of Substituted 2-Phenylbenzothiazole for Solution-Processed Organic Light-Emitting Diodes. Org. Electron. 2018, 60, 31–37. [Google Scholar]
  17. Hu, Y.-X.; Xia, X.; He, W.-Z.; Tang, Z.-J.; Lv, Y.-L.; Li, X.; Zhang, D.-Y. Recent Developments in Benzothiazole-Based Iridium(Ⅲ) Complexes for Application in OLEDs as Electrophosphorescent Emitters. Org. Electron. 2019, 66, 126–135. [Google Scholar]
  18. Sukhikh, T.S.; Ogienko, D.S.; Bashirov, D.A.; Konchenko, S.N. Luminescent Complexes of 2,1,3-Benzothiadiazole Derivatives. Russ. Chem. Bull. 2019, 68, 651–661. [Google Scholar]
  19. Prajapati, M.J.; Solanki, J.D.; Machhi, H.K.; Soni, S.S.; Sen, P.; Surati, K.R. Yellowish-Orange Phosphorescent Iridium(III) Complexes of Bis-Cyclometalated Ligand with Pyrazolone Derivatives: Synthesis, Characterization, Photophysical and Thermal Properties. J. Mater. Sci.: Mater. Electron. 2020, 31, 13778–13786. [Google Scholar]
  20. Li, Y.; Yao, J.; Wang, C.; Zhou, X.; Xu, Y.; Hanif, M.; Qiu, X.; Hu, D.; Ma, D.; Ma, Y. Highly Efficient Deep-Red/near-Infrared D-A Chromophores Based on Naphthothiadiazole for OLEDs Applications. Dye. Pigment. 2020, 173, 107960. [Google Scholar]
  21. Zhang, Y.; Song, J.; Qu, J.; Qian, P.-C.; Wong, W.-Y. Recent Progress of Electronic Materials Based on 2,1,3-Benzothiadiazole and Its Derivatives: Synthesis and Their Application in Organic Light-Emitting Diodes. Sci. China Chem. 2021, 64, 341–357. [Google Scholar] [CrossRef]
  22. Sukhikh, T.S.; Khisamov, R.M.; Konchenko, S.N. Unexpectedly Long Lifetime of the Excited State of Benzothiadiazole Derivative and Its Adducts with Lewis Acids. Molecules 2021, 26, 2030. [Google Scholar] [CrossRef] [PubMed]
  23. Dias, G.G.; Souto, F.T.; Machado, V.G. 2,1,3-Benzothiadiazoles Are Versatile Fluorophore Building Blocks for the Design of Analyte-Sensing Optical Devices. Chemosensors 2024, 12, 156. [Google Scholar] [CrossRef]
  24. Shi, K.; Zhang, W.; Zhou, Y.; Wei, C.; Huang, J.; Wang, Q.; Wang, L.; Yu, G. Chalcogenophene-Sensitive Charge Carrier Transport Properties in A–D–A′′–D Type NBDO-Based Copolymers for Flexible Field-Effect Transistors. Macromolecules 2018, 51, 8662–8671. [Google Scholar]
  25. Bulumulla, C.; Gunawardhana, R.; Kularatne, R.N.; Hill, M.E.; McCandless, G.T.; Biewer, M.C.; Stefan, M.C. Thieno[3,2-b]Pyrrole-Benzothiadiazole Banana-Shaped Small Molecules for Organic Field-Effect Transistors. ACS Appl. Mater. Interfaces 2018, 10, 11818–11825. [Google Scholar]
  26. Kinik, F.P.; Ortega-Guerrero, A.; Ongari, D.; Ireland, C.P.; Smit, B. Pyrene-Based Metal Organic Frameworks: From Synthesis to Applications. Chem. Soc. Rev. 2021, 50, 3143–3177. [Google Scholar] [PubMed]
  27. Bhardwaj, V.K.; Saluja, P.; Hundal, G.; Hundal, M.S.; Singh, N.; Jang, D.O. Benzthiazole-Based Multifunctional Chemosensor: Fluorescent Recognition of Fe3+ and Chromogenic Recognition of HSO4. Tetrahedron 2013, 69, 1606–1610. [Google Scholar] [CrossRef]
  28. Dhaka, G.; Singh, J.; Kaur, N. Benzothiazole Based Chemosensors Having Appended Amino Group(s): Selective Binding of Hg2+ Ions by Three Related Receptors. Inorganica Chim. Acta 2017, 462, 152–157. [Google Scholar] [CrossRef]
  29. Medeiros, G.A.; Correa, J.R.; de Andrade, L.P.; Lopes, T.O.; de Oliveira, H.C.B.; Diniz, A.B.; Menezes, G.B.; Rodrigues, M.O.; Neto, B.A.D. A Benzothiadiazole-Quinoline Hybrid Sensor for Specific Bioimaging and Surgery Procedures in Mice. Sens. Actuators B Chem. 2021, 328, 128998. [Google Scholar]
  30. Mao, P.; Song, Y.; Zhao, X.; Wu, W.; Wang, Y. A Ratiometric Benzimidazole-Based Fluorescent Probe for The Recognition of Phosgene in Solution and Gaseous Phases. J. Fluoresc. 2024, 1–9. [Google Scholar] [CrossRef]
  31. Velusamy, M.; Justin Thomas, K.R.; Lin, J.T.; Hsu, Y.-C.; Ho, K.-C. Organic Dyes Incorporating Low-Band-Gap Chromophores for Dye-Sensitized Solar Cells. Org. Lett. 2005, 7, 1899–1902. [Google Scholar] [CrossRef] [PubMed]
  32. Mishra, A.; Bäuerle, P. Small Molecule Organic Semiconductors on the Move: Promises for Future Solar Energy Technology. Angew. Chem. Int. Ed. Engl. 2012, 51, 2020–2067. [Google Scholar] [CrossRef] [PubMed]
  33. Mishra, A.; Bäuerle, P. Niedermolekulare Organische Halbleiter Auf Dem Vormarsch—Ausblick Auf Künftige Solartechniken. Angew. Chem. 2012, 124, 2060–2109. [Google Scholar] [CrossRef]
  34. Yuan, Y.-J.; Zhang, J.-Y.; Yu, Z.-T.; Feng, J.-Y.; Luo, W.-J.; Ye, J.-H.; Zou, Z.-G. Impact of Ligand Modification on Hydrogen Photogeneration and Light-Harvesting Applications Using Cyclometalated Iridium Complexes. Inorg. Chem. 2012, 51, 4123–4133. [Google Scholar] [CrossRef]
  35. Knyazeva, E.A.; Rakitin, O.A. Influence of structural factors on the photovoltaic properties of dye-sensitized solar cells. Russ. Chem. Rev. 2016, 85, 1146–1183. [Google Scholar] [CrossRef]
  36. Yang, C.; Mehmood, F.; Lam, T.L.; Chan, S.L.-F.; Wu, Y.; Yeung, C.-S.; Guan, X.; Li, K.; Chung, C.Y.-S.; Zhou, C.-Y.; et al. Stable Luminescent Iridium(Iii) Complexes with Bis(N-Heterocyclic Carbene) Ligands: Photo-Stability, Excited State Properties, Visible-Light-Driven Radical Cyclization and CO2 Reduction, and Cellular Imaging. Chem. Sci. 2016, 7, 3123–3136. [Google Scholar] [CrossRef]
  37. Neto, B.A.D.; Correa, J.R.; Spencer, J. Fluorescent Benzothiadiazole Derivatives as Fluorescence Imaging Dyes: A Decade of New Generation Probes. Chem. Eur. J. 2022, 28, e202103262. [Google Scholar] [CrossRef]
  38. Pozharskii, A.F.; Soldatenkov, A.T.; Katritzky, A.R. Heterocycles in Life and Society, 2nd ed.; John Wiley & Sons, Ltd.: Hoboken, NJ, USA, 2011; pp. 209–246. ISBN 9781119998372. [Google Scholar]
  39. Shermolovich, Y.; Pazenok, S. Topics in Heterocyclic Chemistry, 1st ed.; Springer: Berlin, Germany, 2011; pp. 101–138. [Google Scholar]
  40. Leung-Toung, R.; Tam, T.F.; Zhao, Y.; Simpson, C.D.; Li, W.; Desilets, D.; Karimian, K. Synthesis of 3-Substituted Bicyclic Imidazo[1,2-d][1,2,4]Thiadiazoles and Tricyclic Benzo[4,5]Imidazo[1,2-d][1,2,4]Thiadiazoles. J. Org. Chem. 2005, 70, 6230–6241. [Google Scholar] [CrossRef]
  41. Mittal, S.; Samota, M.K.; Kaur, J.; Seth, G. Synthesis, Spectral, and Antifungal Evaluation of Phosphorylated and Thiophosphorylated Benzothiazole Derivatives. Phosphorus Sulfur Silicon Relat. Elem. 2007, 182, 2105–2113. [Google Scholar] [CrossRef]
  42. Carvalho, T.O.; Carvalho, P.H.P.R.; Correa, J.R.; Guido, B.C.; Medeiros, G.A.; Eberlin, M.N.; Coelho, S.E.; Domingos, J.B.; Neto, B.A.D. Palladium Catalyst with Task-Specific Ionic Liquid Ligands: Intracellular Reactions and Mitochondrial Imaging with Benzothiadiazole Derivatives. J. Org. Chem. 2019, 84, 5118–5128. [Google Scholar] [CrossRef]
  43. Kitamura, T.; Yamanishi, K.; Inoue, S.; Yan, Y.-N.; Yano, N.; Kataoka, Y.; Handa, M.; Kawamoto, T. Clamshell Palladium(II) Complexes: Suitable Precursors for Photocatalytic Hydrogen Production from Water. Eur. J. Inorg. Chem. 2022, 2022, e202200259. [Google Scholar]
  44. Rigotti, T.; Schwinger, D.P.; Graßl, R.; Jandl, C.; Bach, T. Enantioselective Crossed Intramolecular [2+2] Photocycloaddition Reactions Mediated by a Chiral Chelating Lewis Acid. Chem. Sci. 2022, 13, 2378–2384. [Google Scholar]
  45. Ma, J.; Zhang, X.; Huang, X.; Luo, S.; Meggers, E. Preparation of Chiral-at-Metal Catalysts and Their Use in Asymmetric Photoredox Chemistry. Nat. Protoc. 2018, 13, 605–632. [Google Scholar]
  46. Grell, Y.; Demirel, N.; Harms, K.; Meggers, E. Chiral Bis(Oxazoline) Ligands as C2-Symmetric Chiral Auxiliaries for the Synthesis of Enantiomerically Pure Bis-Cyclometalated Rhodium(III) Complexes. Organometallics 2019, 38, 3852–3859. [Google Scholar]
  47. Ilichev, V.A.; Balashova, T.V.; Polyakova, S.K.; Rogozhin, A.F.; Kolybalov, D.S.; Bashirov, D.A.; Konchenko, S.N.; Yablonskiy, A.N.; Rumyantcev, R.V.; Fukin, G.K.; et al. Synthesis, Structure, and Luminescence Properties of Sodium and Ytterbium Complexes with 2-(Benzothiazol-2-Yl)Selenophenolate Ligands. Russ. Chem. Bull. 2022, 71, 298–305. [Google Scholar] [CrossRef]
  48. Mironova, O.A.; Ryadun, A.A.; Sukhikh, T.S.; Pushkarevsky, N.A.; Konchenko, S.N. Synthesis and Photophysical Properties of Rare Earth (La, Nd, Gd, Y, Ho) Complexes with Silanediamido Ligands Bearing a Chelating Phenylbenzothiazole Chromophore. New J. Chem. 2023, 47, 3406–3416. [Google Scholar] [CrossRef]
  49. Wang, M.-H.; Tsai, M.-Y.; Su, Y.-C.; Chiu, S.-T.; Lin, P.-H.; Long, J. Zero-Field Single-Molecule Magnet Behavior in a Series of Dinuclear Dysprosium(III) Complexes Based on Benzothiazolyl-Based Ligands and β-Diketonates. Cryst. Growth Des. 2024, 24, 422–431. [Google Scholar] [CrossRef]
  50. Afonin, M.Y.; Martynenko, P.A.; Kolybalov, D.S.; Khisamov, R.M.; Konchenko, S.N.; Sukhikh, T.S. Pd(II)- and Pt(II)-Assisted P–C Activation/Cyclization Reactions with a Luminescent α-Aminophosphine. Inorg. Chem. 2024, 63, 369–380. [Google Scholar] [CrossRef]
  51. Ibrahim, S.; Naik, N.; Shivamallu, C.; Raghavendra, H.L.; Shati, A.A.; Alfaifi, M.Y.; Elbehairi, S.E.I.; Amachawadi, R.G.; Kollur, S.P. Synthesis, Structure, and in Vitro Biological Studies of Benzothiazole Based Schiff Base Ligand and Its Binary and Ternary Co(III) and Ni(II) Complexes. Inorganica Chim. Acta 2024, 559, 121792. [Google Scholar]
  52. Ilichev, V.A.; Silantyeva, L.I.; Rogozhin, A.F.; Bochkarev, M.N. Lanthanide Coordination Compounds with Benzothiazolate Ligands: Synthesis, Structure and Luminescent Properties: Review. J. Struct. Chem. 2024, 65, 1825–1864. [Google Scholar]
  53. Mathis, C.A.; Wang, Y.; Holt, D.P.; Huang, G.-F.; Debnath, M.L.; Klunk, W.E. Synthesis and Evaluation of 11C-Labeled 6-Substituted 2-Arylbenzothiazoles as Amyloid Imaging Agents. J. Med. Chem. 2003, 46, 2740–2754. [Google Scholar] [PubMed]
  54. Djuidje, E.N.; Sciabica, S.; Buzzi, R.; Dissette, V.; Balzarini, J.; Liekens, S.; Serra, E.; Andreotti, E.; Manfredini, S.; Vertuani, S.; et al. Design, Synthesis and Evaluation of Benzothiazole Derivatives as Multifunctional Agents. Bioorg. Chem. 2020, 101, 103960. [Google Scholar]
  55. Liu, Y.; Feng, B.; Cao, X.; Tang, G.; Liu, H.; Chen, F.; Liu, M.; Chen, Q.; Yuan, K.; Gu, Y.; et al. A Novel “AIE + ESIPT” near-Infrared Nanoprobe for the Imaging of γ-Glutamyl Transpeptidase in Living Cells and the Application in Precision Medicine. Analyst 2019, 144, 5136–5142. [Google Scholar] [PubMed]
  56. Zhan, Z.; Su, Z.; Chai, L.; Li, C.; Liu, R.; Lv, Y. Multimodal Imaging Iridium(III) Complex for Hypochlorous Acid in Living Systems. Anal. Chem. 2020, 92, 8285–8829. [Google Scholar] [CrossRef]
  57. Xu, Z.-H.; Gao, H.; Zhang, N.; Zhao, W.; Cheng, Y.-X.; Xu, J.-J.; Chen, H.-Y. Ultrasensitive Nucleic Acid Assay Based on Cyclometalated Iridium(III) Complex with High Electrochemiluminescence Efficiency. Anal. Chem. 2021, 93, 1686–1692. [Google Scholar]
  58. Terpstra, K.; Huang, Y.; Na, H.; Sun, L.; Gutierrez, C.; Yu, Z.; Mirica, L.M. 2-Phenylbenzothiazolyl Iridium Complexes as Inhibitors and Probes of Amyloid β Aggregation. Dalton Trans. 2024, 53, 14258–14264. [Google Scholar]
  59. Sadek, O.; Galán, L.A.; Gendron, F.; Baguenard, B.; Guy, S.; Bensalah-Ledoux, A.; Le Guennic, B.; Maury, O.; Perrin, D.M.; Gras, E. Chiral Benzothiazole Monofluoroborate Featuring Chiroptical and Oxygen-Sensitizing Properties: Synthesis and Photophysical Studies. J. Org. Chem. 2021, 86, 11482–11491. [Google Scholar]
  60. He, P.; Chen, Y.; Li, X.-N.; Yan, Y.-Y.; Liu, C. AIPE-Active Cationic Ir(Iii) Complexes for Efficient Detection of 2,4,6-Trinitrophenol and Oxygen. Dalton Trans. 2023, 52, 128–135. [Google Scholar]
  61. Hu, Y.-X.; Xia, X.; He, W.-Z.; Chi, H.-J.; Dong, Y.; Xiao, G.-Y.; Lv, Y.-L.; Li, X.; Zhang, D.-Y. Novel Ir(III) Complexes Ligated with 2-(2,6-Difluoropyridin-3-Yl)Benzo[d]Thiazole for Highly Efficient OLEDs with Mild Efficiency Roll-Off. Dye. Pigment. 2019, 166, 254–259. [Google Scholar]
  62. Liu, D.; Ding, Q.; Fu, Y.; Song, Z.; Peng, Y. Rh-Catalyzed C–H Amidation of 2-Arylbenzo[d]Thiazoles: An Approach to Single Organic Molecule White Light Emitters in the Solid State. Org. Lett. 2019, 21, 2523–2527. [Google Scholar]
  63. Paul, R.; Chandra Shit, S.; Mandal, H.; Rabeah, J.; Kashyap, S.S.; Nailwal, Y.; Shinde, D.B.; Lai, Z.; Mondal, J. Benzothiazole-Linked Metal-Free Covalent Organic Framework Nanostructures for Visible-Light-Driven Photocatalytic Conversion of Phenylboronic Acids to Phenols. ACS Appl. Nano Mater. 2021, 4, 11732–11742. [Google Scholar] [CrossRef]
  64. Song, Y.; Hu, L.; Cheng, Q.; Chen, Z.; Su, H.; Liu, H.; Liu, R.; Zhu, S.; Zhu, H. Benzothiazole Derivatives with Varied π-Conjugation: Synthesis, Tunable Solid-State Emission, and Application in Single-Component LEDs. J. Mater. Chem. C Mater. 2022, 10, 6392–6401. [Google Scholar] [CrossRef]
  65. Song, Y.; He, Y.; Hu, L.; Cheng, Q.; Chen, Z.; Liu, R.; Zhu, S.; Zhu, H. Panchromatic Luminescent D–π–A Benzothiazoles with Different π-Bridging Modulation: Design, Synthesis and Application in WLED Devices. Mater. Chem. Front. 2023, 7, 2860–2870. [Google Scholar] [CrossRef]
  66. Chou, P.-T.; Studer, S.L.; Martinez, M.L. Studies of the Triplet State of 2-(2′-Hydroxyphenyl) Benzothiazole. Chem. Phys. Lett. 1991, 178, 393–398. [Google Scholar] [CrossRef]
  67. Yi, P.; Peng, H.; Wang, Z.; Yu, X.; Li, X.; Liang, Y. Theoretical Study of Structure, Energetic, Hydrogen Bonds Strength and Intramolecular Proton Transfer of 2-(2-R (ROH, NH2, SH) Phenyl (or Pyridyl)) Benzoxazoles (or Benzothiazoles). Chin. J. Chem. 2011, 29, 650–654. [Google Scholar] [CrossRef]
  68. Yang, P.; Zhao, J.; Wu, W.; Yu, X.; Liu, Y. Accessing the Long-Lived Triplet Excited States in Bodipy-Conjugated 2-(2-Hydroxyphenyl) Benzothiazole/Benzoxazoles and Applications as Organic Triplet Photosensitizers for Photooxidations. J. Org. Chem. 2012, 77, 6166–6178. [Google Scholar] [CrossRef]
  69. Chen, C.-L.; Chen, Y.-T.; Demchenko, A.P.; Chou, P.-T. Amino Proton Donors in Excited-State Intramolecular Proton-Transfer Reactions. Nat. Rev. Chem. 2018, 2, 131–143. [Google Scholar] [CrossRef]
  70. Katkova, M.A.; Pushkarev, A.P.; Balashova, T.V.; Konev, A.N.; Fukin, G.K.; Ketkov, S.Y.; Bochkarev, M.N. Near-Infrared Electroluminescent Lanthanide [Pr(Iii), Nd(Iii), Ho(Iii), Er(Iii), Tm(Iii), and Yb(Iii)] N,O-Chelated Complexes for Organic Light-Emitting Devices. J. Mater. Chem. 2011, 21, 16611–16620. [Google Scholar] [CrossRef]
  71. Balashova, T.V.; Burin, M.E.; Ilichev, V.A.; Starikova, A.A.; Marugin, A.V.; Rumyantcev, R.V.; Fukin, G.K.; Yablonskiy, A.N.; Andreev, B.A.; Bochkarev, M.N. Features of the Molecular Structure and Luminescence of Rare-Earth Metal Complexes with Perfluorinated (Benzothiazolyl)Phenolate Ligands. Molecules 2019, 24, 2376. [Google Scholar] [CrossRef]
  72. Wongsuwan, S.; Chatwichien, J.; Sirisaksoontorn, W.; Chainok, K.; Songsasen, A.; Chotima, R. Novel Pd(Ii) Pincer Complexes Bearing Salicylaldimine-Based Benzothiazole Derivatives: Synthesis, Structural Characterization, DNA/BSA Binding, and Biological Evaluation. New J. Chem. 2023, 47, 10624–10637. [Google Scholar] [CrossRef]
  73. Plyuta, N.; Barra, A.-L.; Novitchi, G.; Avarvari, N. Six-Coordinated Nickel(Ii) Complexes with Benzothiadiazole Schiff-Base Ligands: Synthesis, Crystal Structure, Magnetic and HFEPR Study. Dalton Trans. 2024, 53, 8835–8842. [Google Scholar] [PubMed]
  74. Palanisamy, S.; Lee, W.-Z.; Wang, Y.-M. CCDC 1483661: Experimental Crystal Structure Determination; Cambridge Crystallographic Data Centre: Cambridge, UK, 2017. [Google Scholar] [CrossRef]
  75. Venkatachalam, T.K.; Pierens, G.K.; Bernhardt, P.V.; Reutens, D.C. Heteronuclear NMR Spectroscopic Investigations of Hydrogen Bonding in 2-(Benzo[d]Thiazole-2′-Yl)-N-Alkylanilines. Magn. Reson. Chem. 2015, 53, 448–453. [Google Scholar] [PubMed]
  76. Chen, Y.; Fang, Y.; Gu, H.; Qiang, J.; Li, H.; Fan, J.; Cao, J.; Wang, F.; Lu, S.; Chen, X. Color-Tunable and ESIPT-Inspired Solid Fluorophores Based on Benzothiazole Derivatives: Aggregation-Induced Emission, Strong Solvatochromic Effect, and White Light Emission. ACS Appl. Mater. Interfaces 2020, 12, 55094–55106. [Google Scholar] [PubMed]
  77. Fang-Lu, F.; Jin-Qiu, J.; Xue-Mei, C. Synthesis, Crystal Structure and Fluorescent Properties of a Novel Benzothiazole-Derived Fluorescent Probe for Zn2+. J. Chem. Res. 2015, 39, 661–664. [Google Scholar]
  78. White, L.J.; Wells, N.J.; Blackholly, L.R.; Shepherd, H.J.; Wilson, B.; Bustone, G.P.; Runacres, T.J.; Hiscock, J.R. Towards Quantifying the Role of Hydrogen Bonding within Amphiphile Self-Association and Resultant Aggregate Formation. Chem. Sci. 2017, 8, 7620–7630. [Google Scholar]
  79. Wang, H.; Xu, Z.; Deng, G.-J.; Huang, H. Selective Formation of 2-(2-Aminophenyl)Benzothiazoles via Copper-Catalyzed Aerobic C−C Bond Cleavage of Isatins. Adv. Synth. Catal. 2020, 362, 1663–1668. [Google Scholar]
  80. Wang, G.; Ding, N.; Hao, H.; Jiang, Q.; Feng, Q.; Liu, K.; Hua, C.; Bian, H.; Fang, Y.; Liu, F. Controlling the Excited-State Relaxation for Tunable Single-Molecule Dual Fluorescence in Both the Solution and Film States. J. Mater. Chem. C Mater. 2022, 10, 1118–1126. [Google Scholar] [CrossRef]
  81. Kuz’mina, L.G.; Bezzubov, S.I.; Kulagin, S.V.; Bolotin, B.M. Structure and Physicochemical Properties of Three Structural Forms of Organic Luminophore 2-((2-Benzo[d]Thiazol-2-Yl)Phenyl)Carbamoyl)Benzoic Acid. Crystallogr. Rep. 2022, 67, 566–574. [Google Scholar] [CrossRef]
  82. Khisamov, R.M.; Ryadun, A.A.; Sukhikh, T.S.; Konchenko, S.N. Excitation Wavelength-Dependent Room-Temperature Phosphorescence: Unusual Properties of Novel Phosphinoamines. Mol. Syst. Des. Eng. 2021, 6, 1056–1065. [Google Scholar] [CrossRef]
  83. Sukhikh, T.S.; Kolybalov, D.S.; Khisamov, R.M.; Konchenko, S.N. Phenyl-2-benzothiazole-based α-aminophosphines: Synthesis, crystal structure, and photophysical properties. J. Struct. Chem. 2022, 63, 1446–1452. [Google Scholar]
  84. Sinitsa, D.K.; Pylova, E.K.; Mironova, O.A.; Bashirov, D.A.; Ryadun, A.A.; Sukhikh, T.S.; Konchenko, S.N. Lanthanide Complexes with a New Luminescent Iminophosphonamide Ligand Bearing Phenylbenzothiazole Substituents. Dalton Trans. 2024, 53, 2181–2192. [Google Scholar] [CrossRef] [PubMed]
  85. Sinitsa, D.K.; Sukhikh, T.S.; Pushkarevsky, N.A.; Konchenko, S.N. New Asymmetric Iminophosphonamide Ligand and Its Yttrium Complex: Synthesis and Crystal Structure. J. Struct. Chem. 2024, 65, 666–675. [Google Scholar] [CrossRef]
  86. Billeau, S.; Chatel, F.; Robin, M.; Faure, R.; Galy, J.-P. 1H and 13C Chemical Shifts for 2-Aryl and 2-N-Arylamino Benzothiazole Derivatives. Magn. Reson. Chem. 2006, 44, 102–105. [Google Scholar] [CrossRef] [PubMed]
  87. Pierens, G.K.; Venkatachalam, T.K.; Reutens, D. A Comparative Study between Para-Aminophenyl and Ortho-Aminophenyl Benzothiazoles Using NMR and DFT Calculations. Magn. Reson. Chem. 2014, 52, 453–459. [Google Scholar] [CrossRef]
  88. Meyer, M.; Molomut, N.; Nowak, M.; Ogur, M. Rearrangement of Aryl Benzothiazoles. Recl. Trav. Chim. Pays-Bas 1934, 53, 37–40. [Google Scholar] [CrossRef]
  89. Hein, D.W.; Alheim, R.J.; Leavitt, J.J. The Use of Polyphosphoric Acid in the Synthesis of 2-Aryl- and 2-Alkyl-Substituted Benzimidazoles, Benzoxazoles and Benzothiazoles1. J. Am. Chem. Soc. 1957, 79, 427–429. [Google Scholar] [CrossRef]
  90. Pylova, E.; Lasorne, B.; McClenaghan, N.; Jonusauskas, G.; Taillefer, M.; Konchenko, S.; Prieto, A.; Jaroschik, F. Visible-Light Organic Photosensitizers Based on 2-(2-Aminophenyl)Benzothiazoles for Photocycloaddition Reactions. Chem. Eur. J. 2024, 30, e202401851. [Google Scholar] [CrossRef]
  91. Zhang, J.; Guo, W. A New Fluorescent Probe for Gasotransmitter H2S: High Sensitivity, Excellent Selectivity, and a Significant Fluorescence off–on Response. Chem. Commun. 2014, 50, 4214–4217. [Google Scholar] [CrossRef]
  92. Venkatachalam, T.K.; Stimson, D.H.R.; Bhalla, R.; Pierens, G.K.; Reutens, D.C. Synthesis, Characterization and 11C-Radiolabeling of Aminophenyl Benzothiazoles: Structural Effects on the Alkylation of Amino Group. J. Label. Compd. Radiopharm. 2014, 57, 566–573. [Google Scholar] [CrossRef]
  93. Yang, T.; Zhang, L.; Shang, Y.; Zhu, Z.; Jin, S.; Guo, Z.; Wang, X. Concurrent Suppression of Aβ Aggregation and NLRP3 Inflammasome Activation for Treating Alzheimer’s Disease. Chem. Sci. 2022, 13, 2971–2980. [Google Scholar] [CrossRef]
  94. Kim, J.K.; Bong, S.Y.; Park, R.; Park, J.; Jang, D.O. An ESIPT-Based Fluorescent Turn-on Probe with Isothiocyanate for Detecting Hydrogen Sulfide in Environmental and Biological Systems. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2022, 278, 121333. [Google Scholar]
  95. Gajare, A.S.; Shaikh, N.S.; Jnaneshwara, G.K.; Deshpande, V.H.; Ravindranathan, T.; Bedekar, A. V Clay Catalyzed Conversion of Isatoic Anhydride to 2-(o-Aminophenyl)Oxazolines. J. Chem. Soc. Perkin Trans. 1 2000, 6, 999–1001. [Google Scholar]
  96. Fadda, A.A.; Refat, H.M.; Zaki, M.E.A.; Monir, E. Reactions of Isatoic Anhydride with Bifunctional Reagents: Synthesis of Some New Quinazolone Fused Heterocycles, 2-Substituted Anilinoheterocyclic Derivatives and Other Related Compounds. Synth. Commun. 2001, 31, 3537–3545. [Google Scholar] [CrossRef]
  97. Tseng, H.-W.; Liu, J.-Q.; Chen, Y.-A.; Chao, C.-M.; Liu, K.-M.; Chen, C.-L.; Lin, T.-C.; Hung, C.-H.; Chou, Y.-L.; Lin, T.-C.; et al. Harnessing Excited-State Intramolecular Proton-Transfer Reaction via a Series of Amino-Type Hydrogen-Bonding Molecules. J. Phys. Chem. Lett. 2015, 6, 1477–1486. [Google Scholar] [CrossRef]
  98. Huang, S.; Feng, B.; Cheng, X.; Huang, X.; Ding, J.; Yu, K.; Dong, J.; Zeng, W. Controlling ESIPT-Based AIE Effects for Designing Optical Materials with Single-Component White-Light Emission. Chem. Eng. J. 2023, 476, 146436. [Google Scholar]
  99. Feng, B.; Liu, Y.; Huang, S.; Huang, X.; Huang, L.; Liu, M.; Wu, J.; Du, T.; Wang, S.; Feng, X.; et al. Highly Selective Discrimination of Cysteine from Glutathione and Homo-Cysteine with a Novel AIE-ESIPT Fluorescent Probe. Sens. Actuators B Chem. 2020, 325, 128786. [Google Scholar] [CrossRef]
  100. Bi, A.; Liu, M.; Huang, S.; Zheng, F.; Ding, J.; Wu, J.; Tang, G.; Zeng, W. Construction and Theoretical Insights into the ESIPT Fluorescent Probe for Imaging Formaldehyde in Vitro and in Vivo. Chem. Commun. 2021, 57, 3496–3499. [Google Scholar]
  101. Feng, B.; Chu, F.; Huang, X.; Fang, Y.; Liu, M.; Liu, M.; Chen, F.; Zeng, W. Debut of a NIR ESIPT-Based Fluorescent Probe with Synergistic Effects for Boosting High-Contrast Imaging of β-Galactosidase in Ovarian Cancer. Sens. Actuators B Chem. 2023, 396, 134541. [Google Scholar]
  102. Ray, S.; Das, P.; Banerjee, B.; Bhaumik, A.; Mukhopadhyay, C. Piperazinylpyrimidine Modified MCM-41 for the Ecofriendly Synthesis of Benzothiazoles by the Simple Cleavage of Disulfide in the Presence of Molecular O2. RSC Adv. 2015, 5, 72745–72754. [Google Scholar]
  103. Thaslim Basha, S.; Sudhamani, H.; Rasheed, S.; Venkateswarlu, N.; Vijaya, T.; Naga Raju, C. Microwave-Assisted Neat Synthesis of α-Aminophosphonate/Phosphinate Derivatives of 2-(2-Aminophenyl)Benzothiazole as Potent Antimicrobial and Antioxidant Agents. Phosphorus Sulfur Silicon Relat. Elem. 2016, 191, 1339–1343. [Google Scholar]
  104. Park, N.; Heo, Y.; Kumar, M.R.; Kim, Y.; Song, K.H.; Lee, S. Synthesis of Benzothiazoles through Copper-Catalyzed One-Pot Three-Component Reactions with Use of Sodium Hydrosulfide as a Sulfur Surrogate. Eur. J. Org. Chem. 2012, 2012, 1984–1993. [Google Scholar] [CrossRef]
  105. Bakthadoss, M.; Selvakumar, R.; Srinivasan, J. An Efficient Protocol for the Synthesis of Benzoheterocyclic Compounds via Solid-State Melt Reaction (SSMR). Tetrahedron Lett. 2014, 55, 5808–5812. [Google Scholar] [CrossRef]
  106. Stevens, M.F.G.; Shi, D.-F.; Castro, A. Antitumour Benzothiazoles. Part 2. Formation of 2,2′-Diaminobiphenyls from the Decomposition of 2-(4-Azidophenyl)Benzazoles in Trifluoromethanesulfonic Acid. J. Chem. Soc. Perkin Trans. 1996, 1, 83–93. [Google Scholar] [CrossRef]
  107. Du, S.; Tian, Z.; Yang, D.; Li, X.; Li, H.; Jia, C.; Che, C.; Wang, M.; Qin, Z. Synthesis, Antifungal Activity and Structure-Activity Relationships of Novel 3-(Difluoromethyl)-1-Methyl-1H-Pyrazole-4-Carboxylic Acid Amides. Molecules 2015, 20, 8395–8408. [Google Scholar] [CrossRef]
  108. Du, M.; Zhang, Y.; Yu, Y.; Zhao, H.; Guo, Y.; Yang, Y. A Highly Sensitive and Selective Fluorescent Probe Based on a Pd-Catalyzed Reaction for Detection of Pd2+. Anal. Methods 2019, 11, 6053–6061. [Google Scholar]
  109. Shi, X.; Guo, J.; Liu, J.; Ye, M.; Xu, Q. Unexpectedly Simple Synthesis of Benzazoles by TBuONa-Catalyzed Direct Aerobic Oxidative Cyclocondensation of o-Thio/Hydroxy/Aminoanilines with Alcohols under Air. Chem. Eur. J. 2015, 21, 9988–9993. [Google Scholar]
  110. Anandaraj, P.; Saranya, S.; Ramesh, R. Synthesis and Structural Characterization of Palladium Pincer Complexes: Sustainable Synthesis of Benzothiazoles. Appl. Organomet. Chem. 2023, 37, e7062. [Google Scholar]
  111. Arora, A.; Weaver, J.D. Photocatalytic Generation of 2-Azolyl Radicals: Intermediates for the Azolylation of Arenes and Heteroarenes via C–H Functionalization. Org. Lett. 2016, 18, 3996–3999. [Google Scholar]
  112. Nacsa, E.D.; Lambert, T.H. Cross-Coupling of Sulfonic Acid Derivatives via Aryl-Radical Transfer (ART) Using TTMSS or Photoredox. Org. Chem. Front. 2018, 5, 64–69. [Google Scholar]
  113. Jiang, J.; Cai, X.; Hu, Y.; Liu, X.; Chen, X.; Wang, S.-Y.; Zhang, Y.; Zhang, S. Thermo-Promoted Reactions of Anthranils with Carboxylic Acids, Amines, Phenols, and Malononitrile under Catalyst-Free Conditions. J. Org. Chem. 2019, 84, 2022–2031. [Google Scholar]
  114. Singh, M.; Vaishali; Paul, A.K.; Singh, V. Isatin as a 2-Aminobenzaldehyde Surrogate: Transition Metal-Free Efficient Synthesis of 2-(2′-Aminophenyl)Benzothiazole Derivatives. Org. Biomol. Chem. 2020, 18, 4459–4469. [Google Scholar] [CrossRef] [PubMed]
  115. Draganov, A.B.; Wang, K.; Holmes, J.; Damera, K.; Wang, D.; Dai, C.; Wang, B. Click with a Boronic Acid Handle: A Neighboring Group-Assisted Click Reaction That Allows Ready Secondary Functionalization. Chem. Commun. 2015, 51, 15180–15183. [Google Scholar] [CrossRef] [PubMed]
  116. Al-Amiery, A.A.; Al-Temimi, A.A.; Kadhum, A.A.H.; Al-Majedy, Y.K.; Al-Bayati, R.I.; Aday, H.A.; Mohamad, A.B. Co-Crystal Structure of Mixed Molecules of Methyl 2-(3-Chloro-4-Methyl-2-Oxo-2H-Chromen-7-Yloxy)Acetate and 2-(2-Aminophenyl)Benzothiazole. J. Struct. Chem. 2013, 54, 648–649. [Google Scholar] [CrossRef]
  117. Dhaka, G.; Singh, J.; Kaur, N. Benzothiazole Possessing Reversible and Reusable Selective Chemosensor for Fluoride Detection Based on Inhibition of Excited State Intramolecular Proton Transfer. Inorganica Chim. Acta 2016, 450, 380–385. [Google Scholar] [CrossRef]
  118. Ryu, H.; Choi, M.G.; Cho, E.J.; Chang, S.-K. Cu2+-Selective Fluorescent Probe Based on the Hydrolysis of Semicarbazide Derivative of 2-(2-Aminophenyl)Benzothiazole. Dye. Pigment. 2018, 149, 620–625. [Google Scholar] [CrossRef]
  119. Gülbaş, H.E.; Bozkurt, S.; Halay, E. Synthesis and Characterization of Dicarboxamide Derivative: Cu+2 Selective Turn-on Fluorometric and Colorimetric Chemosensor. J. Heterocycl. Chem. 2023, 60, 1223–1229. [Google Scholar] [CrossRef]
  120. Arafath, M.A.; Adam, F.; Al-Suede, F.S.R.; Razali, M.R.; Ahamed, M.B.K.; Abdul Majid, A.M.S.; Hassan, M.Z.; Osman, H.; Abubakar, S. Synthesis, Characterization, X-Ray Crystal Structures of Heterocyclic Schiff Base Compounds and in Vitro Cholinesterase Inhibition and Anticancer Activity. J. Mol. Struct. 2017, 1149, 216–228. [Google Scholar] [CrossRef]
  121. Palanisamy, S.; Wang, Y.-L.; Chen, Y.-J.; Chen, C.-Y.; Tsai, F.-T.; Liaw, W.-F.; Wang, Y.-M. In Vitro and in Vivo Imaging of Nitroxyl with Copper Fluorescent Probe in Living Cells and Zebrafish. Molecules 2018, 23, 2551. [Google Scholar] [CrossRef]
  122. Kaur, H.; Kaur, N.; Singh, N. Nitrogen and Sulfur Co-Doped Fluorescent Carbon Dots for the Trapping of Hg(Ii) Ions from Water. Mater. Adv. 2020, 1, 3009–3021. [Google Scholar] [CrossRef]
  123. Tsai, Y.-C.; Lu, D.-Y.; Lin, Y.-M.; Hwang, J.-K.; Yu, J.-S.K. Structural Transformations in Dinuclear Zinc Complexes Involving Zn–Zn Bonds. Chem. Commun. 2007, 40, 4125–4127. [Google Scholar] [CrossRef]
  124. Lu, D.; Yu, J.K.; Kuo, T.; Lee, G.; Wang, Y.; Tsai, Y. Theory-Guided Experiments on the Mechanistic Elucidation of the Reduction of Dinuclear Zinc, Manganese, and Cadmium Complexes. Angew. Chem. Int. Ed. Engl. 2011, 50, 7611–7615. [Google Scholar] [CrossRef] [PubMed]
  125. Pan, C.-L.; Chen, W.; Song, J. Lanthanide(II)−Alkali Sandwich Complexes with Cation−Arene π Interactions: Synthesis, Structure, and Solvent-Mediated Redox Transformations. Organometallics 2011, 30, 2252–2260. [Google Scholar]
  126. Pan, C.; Sheng, S.; Hou, C.; Pan, Y.; Wang, J.; Fan, Y. A New Type of Lanthanide Complex—Two Divalent Ytterbium Species Assembled from Cation–π Interactions. Eur. J. Inorg. Chem. 2012, 2012, 779–782. [Google Scholar]
  127. Zhou, L.; Yao, Y.; Li, C.; Zhang, Y.; Shen, Q. Synthesis and Characterization of a Series of New Lanthanide Derivatives Supported by Silylene-Bridged Diamide Ligands and Their Catalytic Activities for the Polymerization of Methyl Methacrylate. Organometallics 2006, 25, 2880–2885. [Google Scholar]
  128. Zhu, X.; Fan, J.; Wu, Y.; Wang, S.; Zhang, L.; Yang, G.; Wei, Y.; Yin, C.; Zhu, H.; Wu, S.; et al. Synthesis, Characterization, Selective Catalytic Activity, and Reactivity of Rare Earth Metal Amides with Different Metal−Nitrogen Bonds. Organometallics 2009, 28, 3882–3888. [Google Scholar] [CrossRef]
  129. Pan, C.-L.; Pan, Y.-S.; Wang, J.; Song, J.-F. A Heterometallic Sandwich Complex of Europium(Ii) for Luminescent Studies. Dalton Trans. 2011, 40, 6361. [Google Scholar]
  130. Mironova, O.A.; Lashchenko, D.I.; Ryadun, A.A.; Sukhikh, T.S.; Bashirov, D.A.; Pushkarevsky, N.A.; Konchenko, S.N. Synthesis and Photophysical Properties of Rare Earth Complexes Bearing Silanediamido Ligands Me2Si(NAryl)2 2− (Aryl = Dipp, Mes). New J. Chem. 2022, 46, 2351–2359. [Google Scholar]
  131. Mironova, O.A.; Ryadun, A.A.; Pushkarevskii, N.A.; Sukhikh, T.S.; Konchenko, S.N. Silandiamide Complexes of Lanthanides with N-Phenylbenzothiazole Substituents: Synthesis, Unexpected Products, Structure, and Luminescence. J. Struct. Chem. 2024, 65, 399–411. [Google Scholar]
  132. Chen, H.; Bartlett, R.A.; Dias, H.V.R.; Olmstead, M.M.; Power, P.P. Synthesis and Structural Characterization of Manganese(II) Derivatives of the Bulky, Chelating Bis(Amido) Ligands [(NMes)2SiMe2]2- and [DippNCH2CH2NDipp]2-, Their Neutral Amine Precursors, and Their Lithium Salts (Mes = 2,4,6-Me3C6H2; Dipp = 2,6-i-Pr2C6H3). Inorg. Chem. 1991, 30, 2487–2494. [Google Scholar]
  133. Hill, M.S.; Hitchcock, P.B. [Me2Al(THF)2]+[{Me2Si(NDipp)2}2Zr2Cl5]- (Dipp = 2,6-Diisopropylphenyl), an Unusual Zirconium/Aluminum Ion Pair Containing a THF-Stabilized Dimethylaluminum Cation. Organometallics 2002, 21, 3258–3262. [Google Scholar] [CrossRef]
  134. Udhayakumari, D.; Jerome, P.; Vijay, N.; Hwan Oh, T. ESIPT: An Approach and Future Perspective for the Detection of Biologically Important Analytes. J. Lumin. 2024, 267, 120350. [Google Scholar] [CrossRef]
  135. Padalkar, V.S.; Seki, S. Excited-State Intramolecular Proton-Transfer (ESIPT)-Inspired Solid State Emitters. Chem. Soc. Rev. 2016, 45, 169–202. [Google Scholar] [CrossRef] [PubMed]
  136. Liu, Z.-Y.; Wei, Y.-C.; Chou, P.-T. Correlation between Kinetics and Thermodynamics for Excited-State Intramolecular Proton Transfer Reactions. J. Phys. Chem. A 2021, 125, 6611–6620. [Google Scholar] [CrossRef] [PubMed]
  137. Patil, V.S.; Padalkar, V.S.; Tathe, A.B.; Sekar, N. ESIPT-Inspired Benzothiazole Fluorescein: Photophysics of Microenvironment PH and Viscosity. Dye. Pigment. 2013, 98, 507–517. [Google Scholar] [CrossRef]
  138. Padalkar, V.; Ramasami, P.; Sekar, N. TD-DFT Study of Excited-State Intramolecular Proton Transfer (ESIPT) of 2-(1,3-Benzothiazol-2-Yl)-5-(N,N-Diethylamino)Phenol with Benzoxazole and Benzimidazole Analogues. Procedia. Comput. Sci. 2013, 18, 797–805. [Google Scholar] [CrossRef]
  139. Majumdar, P.; Zhao, J. 2-(2-Hydroxyphenyl)-Benzothiazole (HBT)-Rhodamine Dyad: Acid-Switchable Absorption and Fluorescence of Excited-State Intramolecular Proton Transfer (ESIPT). J. Phys. Chem. B 2015, 119, 2384–2394. [Google Scholar] [CrossRef]
  140. Wang, T.; Lv, M.; Zhang, Y.; Gao, Y.; Cai, Z.; Zhang, Y.; Song, J.; Liu, J.; Yin, H.; Shang, F. TDDFT Study on the ESIPT Properties of 2-(2′-Hydroxyphenyl)-Benzothiazole and Sensing Mechanism of a Derived Fluorescent Probe for Fluoride Ion. Molecules 2024, 29, 1541. [Google Scholar] [CrossRef]
  141. Peng, H.; Kong, S.; Deng, X.; Deng, Q.; Qi, F.; Liu, C.; Tang, R. Development of a Ratiometric Fluorescent Probe with Zero Cross-Talk for the Detection of SO2 Derivatives in Foods and Live Cells. J. Agric. Food Chem. 2023, 71, 14322–14329. [Google Scholar] [CrossRef]
  142. Chen, Y.; Lu, S.; Abbas Abedi, S.A.; Jeong, M.; Li, H.; Hwa Kim, M.; Park, S.; Liu, X.; Yoon, J.; Chen, X. Janus-Type ESIPT Chromophores with Distinctive Intramolecular Hydrogen-bonding Selectivity. Angew. Chem. Int. Ed. Engl. 2023, 62, e202311543. [Google Scholar] [CrossRef]
  143. Jain, A.; De, S.; Haloi, P.; Barman, P. The Solvent-Regulated Excited State Reaction Mechanism of 2-(2′-Hydroxyphenyl)Benzothiazole Aggregates. Photochem. Photobiol. Sci. 2024, 23, 65–78. [Google Scholar] [CrossRef]
  144. Kanlayakan, N.; Kerdpol, K.; Prommin, C.; Salaeh, R.; Chansen, W.; Sattayanon, C.; Kungwan, N. Effects of Different Proton Donor and Acceptor Groups on Excited-State Intramolecular Proton Transfers of Amino-Type and Hydroxy-Type Hydrogen-Bonding Molecules: Theoretical Insights. New J. Chem. 2017, 41, 8761–8771. [Google Scholar] [CrossRef]
  145. Sukhikh, T.S.; Kolybalov, D.S.; Pylova, E.K.; Konchenko, S.N. Luminescent Zn Halide Complexes with 2-(2-Aminophenyl)Benzothiazole Derivatives. Inorganics 2022, 10, 138. [Google Scholar] [CrossRef]
  146. Demchenko, A.P.; Tang, K.-C.; Chou, P.-T. Excited-State Proton Coupled Charge Transfer Modulated by Molecular Structure and Media Polarization. Chem. Soc. Rev. 2013, 42, 1379–1408. [Google Scholar] [PubMed]
  147. Ji, S.; Ding, Z.; Zhao, J.; Zheng, D. Substituent Control of Dynamical Process for Excited State Intramolecular Proton Transfer of Benzothiazole Derivatives. Chem. Phys. 2022, 560, 111568. [Google Scholar]
  148. Zhu, Q.; An, B.; Yuan, H.; Li, Y.; Guo, X.; Zhang, J. Computational Studies on Amino-Type Excited-State Intramolecular Proton Transfer and Subsequent Cis–Trans Isomerisation Reactions of Three 2-(2′-Aminophenyl)Benzothiazole Derivatives. Mol. Phys. 2017, 115, 288–296. [Google Scholar]
  149. An, B.; Yuan, H.; Zhu, Q.; Li, Y.; Guo, X.; Zhang, J. Theoretical Insight into the Excited-State Intramolecular Proton Transfer Mechanisms of Three Amino-Type Hydrogen-Bonding Molecules. Spectrochim. Acta A Mol. Biomol. Spectrosc. 2017, 175, 36–42. [Google Scholar]
  150. Yang, D.; Zhang, T.; Jia, M.; Cheng, S. Modulating NH-Based Excited-State Intramolecular Proton Transfer by Different Electron-Donating/Withdrawing Substituents in 2-(2′-Aminophenyl)Benzothiazole Compounds. Chem. Phys. Lett. 2019, 724, 57–66. [Google Scholar] [CrossRef]
  151. Mei, J.; Leung, N.L.C.; Kwok, R.T.K.; Lam, J.W.Y.; Tang, B.Z. Aggregation-Induced Emission: Together We Shine, United We Soar! Chem. Rev. 2015, 115, 11718–11940. [Google Scholar]
  152. Peng, Q.; Shuai, Z. Molecular Mechanism of Aggregation-Induced Emission. Aggregate 2021, 2, e91. [Google Scholar]
  153. Liang, K.; Shao, Q.; Xia, G.; Wang, Y.; Jiang, L.; Hong, L.; Wang, H. Tuning Solid State Emission of Semisquaraines via Trimming Central-Ring Structures. Dye. Pigment. 2020, 173, 107926. [Google Scholar] [CrossRef]
  154. Yao, H.; Minami, H.; Funada, T. Organic Nanoparticles Based on Lewis-Pair Formation: Observation of Prototropically Controlled Dual Fluorescence. Photochem. Photobiol. Sci. 2018, 17, 1376–1385. [Google Scholar] [PubMed]
  155. Tseng, S.-M.; Chao, C.-M.; Chang, K.-H.; Wen, C.-S.; Chou, T.-C.; Tsai, T.-L.; Wu, T.-W.; Haung, X.-C.; Liu, J.-Q.; Hung, C.-H.; et al. Substituent Effects in Six(Anilido)-Five(Thiazole) Membered Ring Boron Difluoride Dyes. ChemPhotoChem 2022, 6, e202100188. [Google Scholar]
  156. Maurya, R.C.; Sharma, P. Synthesis, Magnetic and Spectral Studies of Co(II) Picrate Complexes with Heterocyclic Nitrogen Donors. Indian J. Chem. 1999, 38A, 509–513. [Google Scholar]
  157. Maurya, R.C.; Sharma, P.; Roy, S. Synthesis and Characterization of Some Mixed-Ligand Picrate Complexes of Nickel(II) Involving Heterocyclic Nitrogen Donors. Synth. React. Inorg. Met.-Org. Chem. 2003, 33, 683–698. [Google Scholar] [CrossRef]
  158. Pal, N.; Kumar, M.; Seth, G. Biological Important Ni(II) Ternary Complexes Derived from 2-Substituted Benzothiazoles and Amino Acids. J. Chemistry. 2011, 8, 612847. [Google Scholar]
  159. Booysen, I.N.; Gerber, T.I.A.; Mayer, P. Reactions of the Fac-[Re(CO)3]+ and [ReO]3+ Moieties with Substituted Benzothiazoles. Inorganica Chim. Acta 2010, 363, 1292–1296. [Google Scholar]
  160. Machura, B.; Wolff, M.; Gryca, I.; Palion, A.; Michalik, K. Novel Re(I) Tricarbonyl Complexes of Chelating Ligands with Aromatic N-Heterocycle Ring and Aliphatic Amine Donor—Synthesis, Spectroscopic Characterization, X-Ray Structure and DFT Calculations. Polyhedron 2011, 30, 2275–2285. [Google Scholar]
  161. Machura, B.; Wolff, M.; Gryca, I.; Kruszynski, R. Syntheses, Structures, Spectroscopic Properties and DFT Calculations of Re(V)-Benzothiazole and 2-(2-Aminophenyl)Benzothiazole Complexes. Polyhedron 2012, 40, 93–104. [Google Scholar] [CrossRef]
  162. Aleksanyan, D.V.; Churusova, S.G.; Rybalkina, E.Y.; Nelyubina, Y.V.; Kozlov, V.A. Pd(II) and Pt(II) Pincer Complexes of a Benzothiazole-Appended Methylthioacetamide Ligand: Synthesis and in Vitro Cytotoxicity. INEOS OPEN 2021, 6, 237–242. [Google Scholar]
  163. Aleksanyan, D.V.; Churusova, S.G.; Brunova, V.V.; Peregudov, A.S.; Shakhov, A.M.; Rybalkina, E.Y.; Klemenkova, Z.S.; Kononova, E.G.; Denisov, G.L.; Kozlov, V.A. Mechanochemistry for the Synthesis of Non-Classical N-Metalated Palladium(Ii) Pincer Complexes. Dalton Trans. 2021, 50, 16726–16738. [Google Scholar]
  164. Kaplunov, M.G.; Yakushchenko, I.K.; Krasnikova, S.S.; Pivovarov, A.P.; Balashova, I.O. Electroluminescent Materials Based on New Metal Complexes for Organic Light-Emitting Diodes. High Energ. Chem. 2008, 42, 563–565. [Google Scholar] [CrossRef]
  165. Krasnikova, S.S.; Shestakov, A.F.; Kaplunov, M.G. Experimental and Theoretical Study of Electroluminescence Spectra of the Light-Emitting Devices Based on Zinc Complexes of Amino-Substituted Ligands. Mol. Cryst. Liq. Cryst. 2014, 589, 56–66. [Google Scholar] [CrossRef]
  166. Odod, A.V.; Nikonova, E.N.; Nikonov, S.Y.; Kopylova, T.N.; Kaplunov, M.G.; Krasnikova, S.S.; Nikitenko, S.L.; Yakushchenko, I.K. Electroluminescence of Zinc Complexes in Various OLED Structures. Russ. Phys. J. 2017, 60, 7–13. [Google Scholar] [CrossRef]
  167. Nikitenko, S.L.; Kaplunov, M.G.; Yakushchenko, I.K.; Echmaev, S.B. Electroluminescence and Photovoltaic Properties of Bis{N-[2-(Benzothiazol-2-Yl)Phenyl]-N-(4-Methylphenylsulfonyl)Amido}zinc. Russ. Chem. Bull. 2017, 66, 980–985. [Google Scholar] [CrossRef]
  168. Tang, L.; Xia, J.; Zhong, K.; Tang, Y.; Gao, X.; Li, J. A Simple AIE-Active Fluorogen for Relay Recognition of Cu2+ and Pyrophosphate through Aggregation-Switching Strategy. Dye. Pigment. 2020, 178, 108379. [Google Scholar] [CrossRef]
  169. Tang, L.; Dai, X.; Zhong, K.; Wen, X.; Wu, D. A Phenylbenzothiazole Derived Fluorescent Sensor for Zn(II) Recognition in Aqueous Solution Through “Turn-On” Excited-State Intramolecular Proton Transfer Emission. J. Fluoresc. 2014, 24, 1487–1493. [Google Scholar] [CrossRef]
  170. Janakipriya, S.; Tamilmani, S.; Thennarasu, S. A Novel 2-(2′-Aminophenyl)Benzothiazole Derivative Displays ESIPT and Permits Selective Detection of Zn2+ Ions: Experimental and Theoretical Studies. RSC Adv. 2016, 6, 71496–71500. [Google Scholar] [CrossRef]
  171. Halay, E. Cation Sensing by a Novel Triazine-Cored Intermediate as a Fluorescent Chemosensor Incorporating Benzothiazole Fluorophore. Res. Chem. Intermed. 2021, 47, 4281–4295. [Google Scholar] [CrossRef]
  172. Bozkurt, S.; Halay, E. Synthesis, Application and AIE Properties of Novel Fluorescent Tetraoxocalix[2]Arene[2]Triazine: The Detection of a Hazardous Anion, Cyanate. Tetrahedron 2020, 76, 131647. [Google Scholar] [CrossRef]
  173. Bozkurt, S.; Halay, E.; Durmaz, M.; Topkafa, M.; Ceylan, Ö. A Novel Turn-on Fluorometric “Reporter-Spacer-Receptor” Chemosensor Based on Calix[4]Arene Scaffold for Detection of Cyanate Anion. J. Heterocycl. Chem. 2021, 58, 1079–1088. [Google Scholar] [CrossRef]
  174. Choi, M.G.; Cho, M.J.; Ryu, H.; Hong, J.; Chang, S.-K. Fluorescence Signaling of Thiophenol by Hydrolysis of Dinitrobenzenesulfonamide of 2-(2-Aminophenyl)Benzothiazole. Dye. Pigment. 2017, 143, 123–128. [Google Scholar] [CrossRef]
  175. Halay, E.; Bozkurt, S. Enantioselective Recognition of Carboxylic Acids by Novel Fluorescent Triazine-Based Thiazoles. Chirality 2018, 30, 275–283. [Google Scholar] [CrossRef] [PubMed]
  176. Durmaz, M.; Acikbas, Y.; Bozkurt, S.; Capan, R.; Erdogan, M.; Ozkaya, C. A Novel Calix[4]Arene Thiourea Decorated with 2-(2-Aminophenyl)Benzothiazole Moiety as Highly Selective Chemical Gas Sensor for Dichloromethane Vapor. ChemistrySelect 2021, 6, 4670–4676. [Google Scholar]
  177. Chen, L.; Wu, D.; Kim, J.-M.; Yoon, J. An ESIPT-Based Fluorescence Probe for Colorimetric, Ratiometric, and Selective Detection of Phosgene in Solutions and the Gas Phase. Anal. Chem. 2017, 89, 12596–12601. [Google Scholar]
  178. Tang, Y.; Lee, D.; Wang, J.; Li, G.; Yu, J.; Lin, W.; Yoon, J. Development of Fluorescent Probes Based on Protection–Deprotection of the Key Functional Groups for Biological Imaging. Chem. Soc. Rev. 2015, 44, 5003–5015. [Google Scholar] [CrossRef]
  179. Liu, B.-Q.; Chen, Y.-T.; Chen, Y.-W.; Chung, K.-Y.; Tsai, Y.-H.; Li, Y.-J.; Chao, C.-M.; Liu, K.-M.; Tseng, H.-W.; Chou, P.-T. Ethylene Glycol Modified 2-(2′-Aminophenyl)Benzothiazoles at the Amino Site: The Excited-State N-H Proton Transfer Reactions in Aqueous Solution, Micelles and Potential Application in Live-Cell Imaging. Methods Appl. Fluoresc. 2016, 4, 014004. [Google Scholar] [CrossRef]
  180. Huang, T.; Yan, S.; Yu, Y.; Xue, Y.; Yu, Y.; Han, C. Dual-Responsive Ratiometric Fluorescent Probe for Hypochlorite and Peroxynitrite Detection and Imaging In Vitro and In Vivo. Anal. Chem. 2022, 94, 1415–1424. [Google Scholar]
  181. Leuma Yona, R.; Mazères, S.; Faller, P.; Gras, E. Thioflavin Derivatives as Markers for Amyloid-β Fibrils: Insights into Structural Features Important for High-Affinity Binding. ChemMedChem 2008, 3, 63–66. [Google Scholar]
  182. Li, X.; Tao, R.-R.; Hong, L.-J.; Cheng, J.; Jiang, Q.; Lu, Y.-M.; Liao, M.-H.; Ye, W.-F.; Lu, N.-N.; Han, F.; et al. Visualizing Peroxynitrite Fluxes in Endothelial Cells Reveals the Dynamic Progression of Brain Vascular Injury. J. Am. Chem. Soc. 2015, 137, 12296–12303. [Google Scholar] [CrossRef]
  183. Cellier, M.; Fabrega, O.J.; Fazackerley, E.; James, A.L.; Orenga, S.; Perry, J.D.; Salwatura, V.L.; Stanforth, S.P. 2-Arylbenzothiazole, Benzoxazole and Benzimidazole Derivatives as Fluorogenic Substrates for the Detection of Nitroreductase and Aminopeptidase Activity in Clinically Important Bacteria. Bioorg. Med. Chem. 2011, 19, 2903–2910. [Google Scholar]
  184. Sedgwick, A.C.; Sun, X.; Kim, G.; Yoon, J.; Bull, S.D.; James, T.D. Boronate Based Fluorescence (ESIPT) Probe for Peroxynitrite. Chem. Commun. 2016, 52, 12350–12352. [Google Scholar] [CrossRef]
  185. Zhang, J.; Liu, K.; Li, J.; Xie, Y.; Li, Y.; Wang, X.; Xie, X.; Jiao, X.; Tang, B. Harnessing Se=N to Develop Novel Fluorescent Probes for Visualizing the Variation of Endogenous Hypobromous Acid (HOBr) during the Administration of an Immunotherapeutic Agent. Chem. Commun. 2021, 57, 12679–12682. [Google Scholar] [CrossRef]
  186. Zhang, J.; Liu, S.; Liu, C.; Hu, Y.; Peng, T.; He, Y. A Novel Ratiometric Fluorescent Probe for Hypobromous Acid Detection in Living Cells. ChemistrySelect 2022, 7, e202203278. [Google Scholar] [CrossRef]
  187. Kanlayakan, N.; Kungwan, N. Theoretical Study of Heteroatom and Substituent Effects on Excited-State Intramolecular Proton Transfers and Electronic Properties of Amino-Type Hydrogen Bonding Molecules. J. Lumin. 2021, 238, 118260. [Google Scholar] [CrossRef]
  188. Kanlayakan, N.; Kungwan, N. Molecular Design of Amino-Type Hydrogen-Bonding Molecules for Excited-State Intramolecular Proton Transfer (ESIPT)-Based Fluorescent Probe Using the TD-DFT Approach. New J. Chem. 2021, 45, 12500–12508. [Google Scholar] [CrossRef]
  189. Huang, X.; Quinn, T.R.; Harms, K.; Webster, R.D.; Zhang, L.; Wiest, O.; Meggers, E. Direct Visible-Light-Excited Asymmetric Lewis Acid Catalysis of Intermolecular [2+2] Photocycloadditions. J. Am. Chem. Soc. 2017, 139, 9120–9123. [Google Scholar] [CrossRef]
  190. Daub, M.E.; Jung, H.; Lee, B.J.; Won, J.; Baik, M.-H.; Yoon, T.P. Enantioselective [2+2] Cycloadditions of Cinnamate Esters: Generalizing Lewis Acid Catalysis of Triplet Energy Transfer. J. Am. Chem. Soc. 2019, 141, 9543–9547. [Google Scholar] [CrossRef]
  191. Yao, Z.-J.; Lin, N.; Qiao, X.-C.; Zhu, J.-W.; Deng, W. Cyclometalated Half-Sandwich Iridium Complex for Catalytic Hydrogenation of Imines and Quinolines. Organometallics 2018, 37, 3883–3892. [Google Scholar] [CrossRef]
  192. Mdluli, V.; Diluzio, S.; Lewis, J.; Kowalewski, J.F.; Connell, T.U.; Yaron, D.; Kowalewski, T.; Bernhard, S. High-Throughput Synthesis and Screening of Iridium(III) Photocatalysts for the Fast and Chemoselective Dehalogenation of Aryl Bromides. ACS Catal. 2020, 10, 6977–6987. [Google Scholar] [CrossRef]
  193. Hu, N.; Jung, H.; Zheng, Y.; Lee, J.; Zhang, L.; Ullah, Z.; Xie, X.; Harms, K.; Baik, M.-H.; Meggers, E. Catalytic Asymmetric Dearomatization by Visible-Light-Activated [2+2] Photocycloaddition. Angew. Chem. Int. Ed. Engl. 2018, 57, 6242–6246. [Google Scholar]
  194. Zhang, L.; Meggers, E. Steering Asymmetric Lewis Acid Catalysis Exclusively with Octahedral Metal-Centered Chirality. Acc. Chem. Res. 2017, 50, 320–330. [Google Scholar]
  195. Geng, L.; Sun, R.; Zhang, D.-S.; Yu, M.-H.; Chang, Z.; Bu, X.-H. Harnessing Triplet Excitons: Advances in Luminescence Metal Coordination Compounds. Coord. Chem. Rev. 2024, 518, 216066. [Google Scholar]
  196. Zhang, X.; Rovis, T. Photocatalyzed Triplet Sensitization of Oximes Using Visible Light Provides a Route to Nonclassical Beckmann Rearrangement Products. J. Am. Chem. Soc. 2021, 143, 21211–21217. [Google Scholar]
  197. Huang, M.; Zhang, L.; Pan, T.; Luo, S. Deracemization through Photochemical E/Z Isomerization of Enamines. Science 2022, 375, 869–874. [Google Scholar]
  198. Müller, C.; Bauer, A.; Maturi, M.M.; Cuquerella, M.C.; Miranda, M.A.; Bach, T. Enantioselective Intramolecular [2 + 2]-Photocycloaddition Reactions of 4-Substituted Quinolones Catalyzed by a Chiral Sensitizer with a Hydrogen-Bonding Motif. J. Am. Chem. Soc. 2011, 133, 16689–16697. [Google Scholar] [CrossRef]
  199. Tian, D.; Sun, X.; Cao, S.; Wang, E.-M.; Yin, Y.; Zhao, X.; Jiang, Z. Enantioselective Intermolecular [2 + 2] Photocycloadditions of Vinylazaarenes with Triplet-State Electron-Deficient Olefins. Chin. J. Catal. 2022, 43, 2732–2742. [Google Scholar]
  200. Miller, Z.D.; Lee, B.J.; Yoon, T.P. Enantioselective Crossed Photocycloadditions of Styrenic Olefins by Lewis Acid Catalyzed Triplet Sensitization. Angew Chem. Int. Ed. Engl. 2017, 56, 11891–11895. [Google Scholar]
  201. Großkopf, J.; Kratz, T.; Rigotti, T.; Bach, T. Enantioselective Photochemical Reactions Enabled by Triplet Energy Transfer. Chem. Rev. 2022, 122, 1626–1653. [Google Scholar]
  202. Patra, T.; Bellotti, P.; Strieth-Kalthoff, F.; Glorius, F. Photosensitized Intermolecular Carboimination of Alkenes through the Persistent Radical Effect. Angew Chem. Int. Ed. Engl. 2020, 59, 3172–3177. [Google Scholar]
  203. Poplata, S.; Tröster, A.; Zou, Y.-Q.; Bach, T. Recent Advances in the Synthesis of Cyclobutanes by Olefin [2+2] Photocycloaddition Reactions. Chem. Rev. 2016, 116, 9748–9815. [Google Scholar] [CrossRef]
  204. Dutta, S.; Erchinger, J.E.; Strieth-Kalthoff, F.; Kleinmans, R.; Glorius, F. Energy Transfer Photocatalysis: Exciting Modes of Reactivity. Chem. Soc. Rev. 2024, 53, 1068–1089. [Google Scholar] [PubMed]
Figure 1. Main applications of pbt derivatives [45,52,54,56,59,60].
Figure 1. Main applications of pbt derivatives [45,52,54,56,59,60].
Molecules 30 01659 g001
Figure 2. Molecular structure of 2-NH2-pbt, 1 obtained by SCXRD (DMSO solvent not shown) [74].
Figure 2. Molecular structure of 2-NH2-pbt, 1 obtained by SCXRD (DMSO solvent not shown) [74].
Molecules 30 01659 g002
Scheme 1. First synthesis of 1 via thermal rearrangement of N-phenyl-2-benzothiazolamine 4 [88].
Scheme 1. First synthesis of 1 via thermal rearrangement of N-phenyl-2-benzothiazolamine 4 [88].
Molecules 30 01659 sch001
Scheme 2. Synthesis of 1 using anthranilic acid 6 as a starting material [89].
Scheme 2. Synthesis of 1 using anthranilic acid 6 as a starting material [89].
Molecules 30 01659 sch002
Scheme 3. Synthesis of 1 using NaSH as a sulfur source with major steps (i)–(iv) of proposed mechanism [104].
Scheme 3. Synthesis of 1 using NaSH as a sulfur source with major steps (i)–(iv) of proposed mechanism [104].
Molecules 30 01659 sch003
Scheme 4. Synthesis of 1 and 15 under molten state reaction conditions [105].
Scheme 4. Synthesis of 1 and 15 under molten state reaction conditions [105].
Molecules 30 01659 sch004
Scheme 5. Plausible mechanism for the synthesis of 1 from 2-aminobenzyl alcohol 18 involving a Pd(II) catalytic system [110].
Scheme 5. Plausible mechanism for the synthesis of 1 from 2-aminobenzyl alcohol 18 involving a Pd(II) catalytic system [110].
Molecules 30 01659 sch005
Scheme 6. Photocatalytic method for synthesizing a mixture of 1 and 21 using 2-bromoazole 20 [111].
Scheme 6. Photocatalytic method for synthesizing a mixture of 1 and 21 using 2-bromoazole 20 [111].
Molecules 30 01659 sch006
Scheme 7. Synthesis of 1 and 2 through an aryl radical transfer mechanism [112].
Scheme 7. Synthesis of 1 and 2 through an aryl radical transfer mechanism [112].
Molecules 30 01659 sch007
Scheme 8. Synthesis of 1 involving 2,1-benzisoxazole 25 [113].
Scheme 8. Synthesis of 1 involving 2,1-benzisoxazole 25 [113].
Molecules 30 01659 sch008
Scheme 9. Synthesis of 2-RNH-pbt derivatives 30 from dioxazolone 28 via a Rh-catalyzed approach [62].
Scheme 9. Synthesis of 2-RNH-pbt derivatives 30 from dioxazolone 28 via a Rh-catalyzed approach [62].
Molecules 30 01659 sch009
Scheme 10. Selective synthesis of 2-aminopbt derivatives 32 and 33 using isatin 31 as a starting material [79].
Scheme 10. Selective synthesis of 2-aminopbt derivatives 32 and 33 using isatin 31 as a starting material [79].
Molecules 30 01659 sch010
Scheme 11. Synthesis of 2-aminopbt derivatives 32 via a metal-free approach from isatin 31 and S8 [114].
Scheme 11. Synthesis of 2-aminopbt derivatives 32 via a metal-free approach from isatin 31 and S8 [114].
Molecules 30 01659 sch011
Scheme 12. Alternative pathway for synthesis of pbt derivatives 38 and 39 using a click chemistry approach [115].
Scheme 12. Alternative pathway for synthesis of pbt derivatives 38 and 39 using a click chemistry approach [115].
Molecules 30 01659 sch012
Figure 3. (a) Chemical structure of cocrystal of 1 with methyl 2-((3-chloro-4-methyl-2-oxo-2H-chromen-7-yl)oxy)acetate 40, involving hydrogen bonds; (b) X-ray structure of known cocrystal 1·40 [116].
Figure 3. (a) Chemical structure of cocrystal of 1 with methyl 2-((3-chloro-4-methyl-2-oxo-2H-chromen-7-yl)oxy)acetate 40, involving hydrogen bonds; (b) X-ray structure of known cocrystal 1·40 [116].
Molecules 30 01659 g003
Scheme 13. Overview of chemical transformations of 1.
Scheme 13. Overview of chemical transformations of 1.
Molecules 30 01659 sch013
Scheme 14. Synthesis of phosphorus derivatives of 1.
Scheme 14. Synthesis of phosphorus derivatives of 1.
Molecules 30 01659 sch014
Scheme 15. Synthesis of 1,3-aminophosphonate and 1,3-aminophosphine derivatives of 1.
Scheme 15. Synthesis of 1,3-aminophosphonate and 1,3-aminophosphine derivatives of 1.
Molecules 30 01659 sch015
Scheme 16. Synthesis of silanediamine framework 56 derived from 1 [48,131].
Scheme 16. Synthesis of silanediamine framework 56 derived from 1 [48,131].
Molecules 30 01659 sch016
Figure 4. ESIPT process in 2-RNH-pbt derivatives [134].
Figure 4. ESIPT process in 2-RNH-pbt derivatives [134].
Molecules 30 01659 g004
Figure 5. Deuterated derivative of 60a [97].
Figure 5. Deuterated derivative of 60a [97].
Molecules 30 01659 g005
Figure 6. (a) Chromacity CIE x,y coordinates of 50 (1), 51 (2), and 52 (3) showing emission colors upon different excitation wavelengths, (b) photographs of crystals on the example of 50 under visible transmitted light, UV light with the maximum at 250 nm and UV light with the maximum at 365 nm, adapted from ref. [82].
Figure 6. (a) Chromacity CIE x,y coordinates of 50 (1), 51 (2), and 52 (3) showing emission colors upon different excitation wavelengths, (b) photographs of crystals on the example of 50 under visible transmitted light, UV light with the maximum at 250 nm and UV light with the maximum at 365 nm, adapted from ref. [82].
Molecules 30 01659 g006
Figure 7. (a) AIE and (b) ACQ phenomena of 2-aminopbt derivatives [151], adapted with permission from reference [98], Copyright@Elsevier 2023.
Figure 7. (a) AIE and (b) ACQ phenomena of 2-aminopbt derivatives [151], adapted with permission from reference [98], Copyright@Elsevier 2023.
Molecules 30 01659 g007
Figure 8. 2-RNH-pbt frameworks showing AIE properties [78,153].
Figure 8. 2-RNH-pbt frameworks showing AIE properties [78,153].
Molecules 30 01659 g008
Figure 9. 2-MeNH-pbt derivatives 78 bearing a boron difluoride moiety [155].
Figure 9. 2-MeNH-pbt derivatives 78 bearing a boron difluoride moiety [155].
Molecules 30 01659 g009
Figure 10. Molecular structures of two crystalline modifications of 2-aminopbt derivative 79 (top and bottom) and change in the glow of a single crystal with an increase in temperature under UV irradiation (left) [81].
Figure 10. Molecular structures of two crystalline modifications of 2-aminopbt derivative 79 (top and bottom) and change in the glow of a single crystal with an increase in temperature under UV irradiation (left) [81].
Molecules 30 01659 g010
Figure 11. Proposed chemical structures of the Co and Ni coordination compounds with neutral 1 [156,157,158].
Figure 11. Proposed chemical structures of the Co and Ni coordination compounds with neutral 1 [156,157,158].
Molecules 30 01659 g011
Figure 12. Structurally characterized Re and Zn coordination compounds of 1 and comparison with phosphine oxide complexes 86 [145,159,160,161].
Figure 12. Structurally characterized Re and Zn coordination compounds of 1 and comparison with phosphine oxide complexes 86 [145,159,160,161].
Molecules 30 01659 g012
Figure 13. Activation/cyclization reactions of the P-C bond in reaction of 54 with 74a or 74b [50].
Figure 13. Activation/cyclization reactions of the P-C bond in reaction of 54 with 74a or 74b [50].
Molecules 30 01659 g013
Figure 14. Structurally characterized metal–organic compounds of Pd- and Pt-based on S,N,N and O,N,N pincer derivatives of 1 [72,162,163].
Figure 14. Structurally characterized metal–organic compounds of Pd- and Pt-based on S,N,N and O,N,N pincer derivatives of 1 [72,162,163].
Molecules 30 01659 g014
Figure 15. Proposed structure of metal–organic compounds of Ni(II) and Co(III) based on salicylaldimine-based derivatives of 1 [51].
Figure 15. Proposed structure of metal–organic compounds of Ni(II) and Co(III) based on salicylaldimine-based derivatives of 1 [51].
Molecules 30 01659 g015
Figure 16. Structurally characterized rare-earth metal complexes with (a) anionic ‘symmetrical’ NPN derivative of 1 and (b) anionic ‘non-symmetrical’ NPN derivative of 1 [84,85].
Figure 16. Structurally characterized rare-earth metal complexes with (a) anionic ‘symmetrical’ NPN derivative of 1 and (b) anionic ‘non-symmetrical’ NPN derivative of 1 [84,85].
Molecules 30 01659 g016
Figure 17. Structurally characterized rare-earth complexes with silanediamide ligand based on 1 [48].
Figure 17. Structurally characterized rare-earth complexes with silanediamide ligand based on 1 [48].
Molecules 30 01659 g017
Scheme 17. Unexpected formation of rare-earth complexes with silanediamide ligand L56 2− [131].
Scheme 17. Unexpected formation of rare-earth complexes with silanediamide ligand L56 2− [131].
Molecules 30 01659 sch017
Figure 18. Overview of derivatives of 1 used for sensor applications and sensing mechanisms: (a) sensing mechanism, involving metal coordination or non-covalent interactions; (b) sensing mechanism, involving the removal of a quencher group of fluorescence [169,174].
Figure 18. Overview of derivatives of 1 used for sensor applications and sensing mechanisms: (a) sensing mechanism, involving metal coordination or non-covalent interactions; (b) sensing mechanism, involving the removal of a quencher group of fluorescence [169,174].
Molecules 30 01659 g018
Figure 19. Some examples of compounds derived from 1 used in biological applications.
Figure 19. Some examples of compounds derived from 1 used in biological applications.
Molecules 30 01659 g019
Figure 20. The mechanism of interaction of β-galactosidase with bio-probe 105 leading to a fluorescence signal change [101].
Figure 20. The mechanism of interaction of β-galactosidase with bio-probe 105 leading to a fluorescence signal change [101].
Molecules 30 01659 g020
Scheme 18. [2+2] photocycloaddition reaction using 1 as organic photocatalyst and proposed mechanism [90], [1]* is the photocatalyst in the excited state.
Scheme 18. [2+2] photocycloaddition reaction using 1 as organic photocatalyst and proposed mechanism [90], [1]* is the photocatalyst in the excited state.
Molecules 30 01659 sch018
Table 1. 1H and 13C NMR data of 1 in different deuterated solvents [87].
Table 1. 1H and 13C NMR data of 1 in different deuterated solvents [87].
Molecules 30 01659 i001
1H NMR Chemical Shifts, ppm13C NMR Chemical Shifts, ppm
1HBenzeneChloroformAcetoneDMSO13CBenzeneChloroformAcetoneDMSO
3′6.326.786.926.901′115.7115.2115.0113.1
4′7.007.227.227.222′147.8146.7148.9147.6
5′6.556.756.686.653′117.3116.7117.6116.5
6′7.707.717.697.634′132.1131.5132.8131.7
47.887.977.998.005′117.1116.8116.9 115.6
57.127.457.507.506′131.0130.3130.9129.9
66.977.357.417.412170.2 169.2170.2168.8
77.407.868.028.073a154.7154.7154.8153.2
NH26.156.407.117.344123.0122.4123.1122.0
5126.6125.9127.2126.3
6125.3124.8125.4125.0
7121.7121.1122.3121.7
7a134.0133.2133.9132.4
Table 2. Synthesis of 1 using isatoic anhydride 7 as a starting material.
Table 2. Synthesis of 1 using isatoic anhydride 7 as a starting material.
Molecules 30 01659 i002
EntryReaction Conditionst, °CTime, hYield, %Ref.
1acidic kaolinitic clay (20% w/w), dried PhCl, argon atmosphere1202055[95]
2CH3COONa, glacial acetic acid120380[96]
3ZnCl2 (30 mol%), PhCl140568[97]
Table 3. Synthesis of 1 using 2-aminobenzaldehyde 9 as a starting material.
Table 3. Synthesis of 1 using 2-aminobenzaldehyde 9 as a starting material.
Molecules 30 01659 i003
EntryReaction Conditionst, °CTime, hYield, %Ref.
(i)modified silica dioxide catalyst (cat.), H2O, O2, I2100489[102]
(ii)CH3COOH, MW (490 W)-10 min99[103]
Table 4. Two-step sequential synthesis of 1, involving (i) formation and (ii) reduction of 2-(2′-nitrophenyl)benzothiazole 17.
Table 4. Two-step sequential synthesis of 1, involving (i) formation and (ii) reduction of 2-(2′-nitrophenyl)benzothiazole 17.
Molecules 30 01659 i004
EntryStarting ReagentStepReaction Conditions Time, hYield, %Ref.
116a(i)K3[Fe(CN)6], toluene, reflux1072[107]
(ii)NH2NH2·H2O, Pd-C, MeOH, reflux885
216a(i)640 W10 min78[28]
(ii)SnCl2·2H2O, abs. EtOH, 70 °C170
316a(i)EtOH, 80 °C371[108]
intermediate stepchloranil, CH2Cl2overnight86
(ii)Fe, HCl, NH4Cl, THF/H2O, 60 °C0.585
416b(i)pyridine, r.t.182[106]
(ii)SnCl2·2H2O, EtOH, reflux485
Table 5. Synthesis of 1 using 2-aminobenzyl alcohol 18 as a starting material.
Table 5. Synthesis of 1 using 2-aminobenzyl alcohol 18 as a starting material.
Molecules 30 01659 i005
EntryReaction Conditionst, °CTime, hYield, %Ref.
1NaOtBu (50 mol %), air, toluene1002443[109]
2Pd(II) catalyst (1 mol%), m-xylene, KOH1201271[110]
Table 6. ESIPT and non-ESIPT compounds derived from 1 and their photophysical properties in solid state and/or in CH2Cl2.
Table 6. ESIPT and non-ESIPT compounds derived from 1 and their photophysical properties in solid state and/or in CH2Cl2.
Molecules 30 01659 i006
Compoundλ abs, nmλ em, nmTS1→TS0, nmRef.
1360455-[145]
2 1392458-[97]
3392 *469-[76]
57 1335395565[97]
15 1330-555[97]
44 1363419588[97]
58 1362-649[97]
59 1336-540[97]
60 1393436623[97]
30en.d.420550[62]
30fn.d.435550[62]
30gn.d.440565[62]
1 data are presented for solution in CH2Cl2, *—compound absorbs in whole visible spectrum (400–800 nm).
Table 7. The calculated barriers for ESIPT (NS1→TS1) and twisting of tautomer form around C1′-C2 [148,149].
Table 7. The calculated barriers for ESIPT (NS1→TS1) and twisting of tautomer form around C1′-C2 [148,149].
Molecules 30 01659 i007
CompoundNS1→TS1, eVTwisted-TS0→TS0, eV
10.38not calculated
20.39not calculated
570.300.33
150.120.27
440.340.34
590.120.83
Table 8. Luminescent properties of pbt derivatives with donor substituents in bt cycle in chloroform solution.
Table 8. Luminescent properties of pbt derivatives with donor substituents in bt cycle in chloroform solution.
Molecules 30 01659 i008
Compoundλ abs, nmλ em, nmRef.
441363419, 588[97]
33g265, 378440[114]
33h267, 379440[114]
33i270, 306, 378440[114]
33j265, 374425[114]
33k265, 314, 377440[114]
33l244, 270, 380440[114]
33m303, 385470[114]
33n300, 377470[114]
33o245, 336, 390470[114]
33p270, 322465[114]
1 data are presented for solution in CH2Cl2.
Table 9. Photophysical properties in solid state of known organic phosphorus derivatives of 1.
Table 9. Photophysical properties in solid state of known organic phosphorus derivatives of 1.
Molecules 30 01659 i009
Compoundλabs, nmλem, nmτ, μsRef.
50255, 263, 291, 330–374 (br)455, 55083[82]
51253, 262, 292, 307, 340–384 (br)450, 575 212[82]
52254, 261, 284, 296, 316–379 (br)455, 555 33, 700[82]
54260−320 (br.), 390445, 600-[50]
55a300, 380450, 610 -[83]
55b300, 380450, 610 -[83]
53a-475, 634 -[84]
Table 10. 2-RNH-pbt proligands for Zn complexes with applications in OLED devices [164,165,166,167].
Table 10. 2-RNH-pbt proligands for Zn complexes with applications in OLED devices [164,165,166,167].
Molecules 30 01659 i010
Compoundλabs, nmλem, nmBrightness *, cd·m−2 (at 8 V)
Zn(L81)2 (86)385463100
Zn(L82)2 (87)380445230
Zn(L15)2 (88)390440140
Zn(L83)2 (89)390430-
Zn(L84)2 (90)380445270
Zn(L85)2 (91)380440-
*—data are presented for devices based on corresponding complexes.
Table 11. Calculated ESIPT barriers for 2-amino-pbt derivatives [188].
Table 11. Calculated ESIPT barriers for 2-amino-pbt derivatives [188].
Molecules 30 01659 i011
CompoundR = Ac, ESIPT Barrier, eVR = Ts, ESIPT Barrier, eV
1070.230.08
1080.330.13
1090.340.07
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Pylova, E.K.; Sukhikh, T.S.; Prieto, A.; Jaroschik, F.; Konchenko, S.N. Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives: Syntheses, Photophysical Properties and Applications. Molecules 2025, 30, 1659. https://doi.org/10.3390/molecules30081659

AMA Style

Pylova EK, Sukhikh TS, Prieto A, Jaroschik F, Konchenko SN. Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives: Syntheses, Photophysical Properties and Applications. Molecules. 2025; 30(8):1659. https://doi.org/10.3390/molecules30081659

Chicago/Turabian Style

Pylova, Ekaterina K., Taisiya S. Sukhikh, Alexis Prieto, Florian Jaroschik, and Sergey N. Konchenko. 2025. "Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives: Syntheses, Photophysical Properties and Applications" Molecules 30, no. 8: 1659. https://doi.org/10.3390/molecules30081659

APA Style

Pylova, E. K., Sukhikh, T. S., Prieto, A., Jaroschik, F., & Konchenko, S. N. (2025). Chemistry of 2-(2′-Aminophenyl)benzothiazole Derivatives: Syntheses, Photophysical Properties and Applications. Molecules, 30(8), 1659. https://doi.org/10.3390/molecules30081659

Article Metrics

Back to TopTop