Next Article in Journal
Phosphatidylinositol-4,5-biphosphate (PIP2)-Dependent Thermoring Basis for Cold-Sensing of the Transient Receptor Potential Melastatin-8 (TRPM8) Biothermometer
Previous Article in Journal
Effect of Molecular Architecture of Surface-Active Organosilicon Macromers on Their Colloidal Properties in Relation to Heterophasic Radical Polymerization of Styrene and Methyl Methacrylate
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Ground-State Tautomerism and Excited-State Proton Transfer in 7-Hydroxy-4-methyl-8-((phenylimino)methyl)-2H-chromen-2-one as a Potential Proton Crane

by
Daniela Nedeltcheva-Antonova
1,2 and
Liudmil Antonov
1,*
1
Institute of Electronics, Bulgarian Academy of Sciences, 1784 Sofia, Bulgaria
2
Institute of Organic Chemistry with Centre of Phytochemistry, Bulgarian Academy of Sciences, 1113 Sofia, Bulgaria
*
Author to whom correspondence should be addressed.
Physchem 2024, 4(1), 91-105; https://doi.org/10.3390/physchem4010007
Submission received: 11 December 2023 / Revised: 10 January 2024 / Accepted: 6 March 2024 / Published: 11 March 2024
(This article belongs to the Section Experimental and Computational Spectroscopy)

Abstract

:
The tautomerism in the title compound as a potential long-range proton transfer (PT) switch has been studied by using the DFT and TD-DFT approaches. The data show that in aprotic solvents, the enol tautomer dominates, while the increase in the content of the keto tautomer (short-range PT) rises as a function of polarity of the solvent. In ethanol, due to specific solute–solvent stabilization through intermolecular hydrogen bonding, a substantial amount of the keto forms exists in solution. The irradiation leads to two competitive processes in the excited state, namely ESIPT and trans/cis isomerization around the azomethine bond as in other structurally similar Schiff bases. The studied compound is not suitable for bistable tautomeric switching, where long-range PT occurs, due to the difficult enolization of the coumarin carbonyl group.

1. Introduction

Excited-state intramolecular proton transfer (ESIPT) is a feature of organic molecules, containing strong proton-donating/accepting groups, which, due to the increased acidity/basicity in the excited state, exchange a proton over the short distance of pre-existing hydrogen bonding [1,2,3,4,5,6,7,8,9,10]. ESIPT-exhibiting molecules have become a field of active research in recent decades, due to their applications as light-emitting materials and laser dyes [11,12,13,14,15], optical sensors [16,17,18,19,20,21,22,23,24,25,26,27], organic light-emitting diodes [20,21,28,29,30,31,32,33,34,35,36], optical storage devices [37] and photo stabilizers [38]. The same process underlies a special class of bistable photo switches [39], called proton cranes, where the ESIPT is accompanied by intramolecular rotation [40], leading to the transport of a movable proton from one side of the molecule to another.
The overall process of switching in proton cranes is outlined in Scheme 1. The proton, which is located at the proton donor site of a PT frame (PTF), undergoes ESIPT to the proton crane unit (PCU) upon irradiation. In this way, the proton is separated from the PTF, in which two free proton acceptor sites remain. The competition between them is the driving force of the rotation of the PCU, which can lead to the delivery and release of the proton to the other side of the PTF. The overall process can be classified as long-range intramolecular proton transfer.
Depending on the extent of conjugation between PTF and PCU, and more importantly, on the possibility for tautomeric proton exchange between PTF and PCU, the switching mechanism can be very different [41,42]. In 1 [43,44], the ESIPT (i.e., the transition from I* to II*) leads to tautomeric structural re-arrangement, which changes the nature of the axle from a single to a double bond. This causes twisting around the axle to a structure where the PFT and PCU are perpendicular each other, and relaxation to the ground state occurs through conical intersection. As a result, II and III are populated simultaneously, leading (at suitable relative energies) to a population of I and IV, respectively. In 2 [42,45,46,47], the initial ESIPT leads to II*, which has a zwitterionic structure (i.e., intramolecular deprotonation occurs). The nature of the axle remains unchanged and the long-range PT occurs entirely in the excited state, which leads to IV*, where the PTF is protonated again. Then, the relaxation to the ground state populates IV. In 3 [48,49,50], there is also possibility for tautomeric re-arrangement in the first stage of the process, but due to the highly unstable zwitterionic nature of IV, the end state has never been reached [49,50].
7-hydroxy-quinoline is one of the most frequently studied examples for long-range intramolecular proton transfer [51,52,53,54,55,56,57,58,59,60,61] and is a very suitable structure for PTF, where the proton donor OH group and the proton acceptor N atom are far apart to provide conditions for truly intramolecular proton exchange. The enol tautomer is substantially more stable, while the NH tautomer has been experimentally observed only in protic organic solvents or in the presence of water and results from the intermolecular, solvent-assisted PT mechanism [43,61]. Previous computational studies suggested that the reversible long-range PT process could be observed for a series of 7-hydroxy-quinoline-based switches and oxazine [41,62] and other six-membered ring heterocycles [62], pyridines [62,63,64], the carbaldehyde group [43], and CO-BF2 [65,66] as proton cranes units. Recently, we have reported several classes of proton cranes, where the azomethine group plays the role of a PCU, providing intramolecular PT to III, and in limited cases to IV, upon irradiation [67,68].
7-hydroxy-coumarin and its derivatives (including 4, Scheme 2), also possess a proton donor group (O-H) at a large distance from the proton acceptor group (C=O), making direct intramolecular PT impossible. The substantially more stable enol form is observed in organic solvents, while the keto form appears in the presence of alcohols, water and organic bases as for 7-hydroxy-coumarin [69,70,71,72,73,74] as well as for 4 [74,75,76,77,78,79]. As reported for 4, concentration-assisted PT takes place in the gas phase [80] as well. Quite interestingly, in 7-hydroxy-4-(trifluoromethyl)coumarin in toluene, the addition of 1-methylimidazole leads to excited-state intermolecular long-range PT in the picosecond time scale, where the base acts as a proton-transfer shuttle from the hydroxyl group to the carbonyl [81].
In spite of its structural suitability as a PTF, 7-hydroxycoumarine and its derivative 4 have never been considered in the design of proton cranes. Very recently, the synthesis and spectral properties of 5 have been reported, along with some photochemical information [82,83]. This creates a solid background for detailed theoretical studies that can reveal the ground-state tautomeric properties and excited-state PT in 6 as simplified models of 5. Bearing in mind that the Ph rings in the Schiff bases are partially out of planarity, the effect of weak substituents like OMe on the tautomeric state is negligible [84]. In the current investigation, the tautomerism of 6 in the ground and excited state is studied in a variety of solvents, and the applicability of 7-hydroxy-coumarine Schiff bases as proton cranes is considered.

2. Computational Details

Quantum-chemical calculations in the ground state were performed using the Gaussian 16 C.01 program suite [85]. All structures (in both the ground and excited state) were optimized without restrictions, using tight optimization criteria and an ultrafine grid in the computation of two-electron integrals and their derivatives. The true minima were verified by performing frequency calculations in the corresponding environment. The implicit solvation was described using the Polarizable Continuum Model [86] (the integral equation formalism variant, IEFPCM, as implemented in Gaussian 16). The transition states were estimated using the Berny algorithm [87] and again verified by performing frequency calculations in the corresponding environment.
The M06-2X [88,89] density functional with the TZVP [90] basis set was used for the structure optimizations in the ground state. This fitted hybrid meta-GGA functional with 54% HF exchange was especially developed to describe main-group thermochemistry and non-covalent interactions. The use of M06-2X provides very good predictability of the ground-state tautomerism [68,91,92] in tautomeric compounds and proton cranes in solution, as well as the E/Z isomerization ratio in rotary switches.
The TD-DFT method [93,94] was used for singlet excited-state optimizations. CAM-B3LYP [95] with the TZVP basis set was used for the optimizations (NStates = 6). The selection of CAM-B3LYP was based on its better performance (in comparison with a variety of density functionals, including M06-2X) in describing electronic excitation energies, excited-state geometries, dipole moments and oscillator strengths in a variety of systems [96,97,98,99], including ESIPT ones [100], as well as from our own previous experience [68,101].
Bearing in mind that M06-2X systematically underestimates vertical transition energies [102], the UV-Vis spectral data were predicted by the B3LYP [103] functional (6-311 + G(2d,p) basis set) using the M06-2X optimized ground-state geometries. The corresponding spectral curves (presented as molar absorptivities) were constructed from the theoretically obtained band positions and oscillator strengths by using Gauss band shape [104] according to Equation (1) as follows:
ε λ = i = 1 n ε m a x , i · e ln 2 · 10 14 · 1 λ m a x , i 1 λ Δ ν 1 / 2 2 2
where n is the number of the transitions taken into account, λ m a x , i is the predicted position of the band of the ith transition in nm and ε m a x , i is the calculated maximal intensity of the ith band according to Equation (2) from the oscillator strength f i [105].
ε m a x , i = f i 7.04 × 10 9 · Δ ν 1 / 2
A half-band width ( Δ ν 1 / 2 ) value of 3500 cm−1 was used for all bands.

3. Results and Discussion

The possible tautomeric forms of 5 and 6 are given in Scheme 2. According to the data, obtained by Liubimov and co-authors [82], the E tautomer of 5, absorbing in the near UV range (340 nm), is substantially more stable in aprotic solvents like toluene. The KE form has an absorption maximum around 450 nm, the intensity of which arises in protic solvents. The same behavior is typical for the tautomeric Schiff bases derived from 2-naphthol (7, Scheme 3) [84] and 7-hydroxy-quinoline (8) [68], which feature intramolecular hydrogen bonding and increased aromaticity in the OH-containing part.
The absorption spectra of the possible tautomers of 6, as a model of 5, are shown in Figure 1 and Figure S1 in toluene and acetonitrile, respectively. In line with the experiment, the E tautomer absorbs around 330–340 nm, which corresponds very well to the experimental data in these two solvents [82,83]. The existence of the keto tautomers (KE, KK and K) should be manifested by the appearance of red-shifted absorbance, each of them with their own signature. According to the relative stabilities of the tautomers, shown in Figure 2 and, in detail, in Table S1, the energy difference between E and KE suggests a very small amount of the latter is presented in toluene and acetonitrile, with more in the latter. The energies of KK, and especially of K, are too high to be observed in solution. According to the spectrum of 5 in toluene [82], a very low intensive tail is observed in the range 400–500 nm, while in acetonitrile [83], the red-shifted band is well defined at 450 nm, which corresponds to the suggested spectra of KE and KK. As discussed above, due to the high relative energy (>6 kcal/mol in respect of E, Table S1), the KK tautomer should be excluded. In the case of 6, the appearance of KE has an individual feature, which differs from the E to KE transitions in 7 and 8. While in 7 and 8, the increase in the content of the KE form leads to a decrease in the intensity of the band assigned to the enol [68,84], in the case of 6, as very clearly seen from Figure 1 and Figure S1, the increase in the keto form should also lead to a rise in the absorption at 340 nm, due to the overlapping between the S0-S1 transition of E and the S0–Sn transitions in KE, with the latter having larger oscillator strength. Actually, this is observed for 5, comparing the molar absorptivities at 340 nm in toluene and ethanol [82]. The intensity of the band in toluene is ~23 000 M−1·cm−1, which nicely corresponds to that in Figure 1. The same value in ethanol, where the content of KE is substantially higher, is much larger. Based on the predicted theoretical data for the KE maximum in ethanol with molar absorptivity around 9 000 M−1·cm−1 (Figure S2) and the measured value in ethanol (4 500 M−1·cm−1), it can be estimated roughly that the amount of 5KE is between 40 and 50%. This is surprisingly high compared to the predicted relative energies in ethanol as a medium (Table 1), suggesting that this tautomeric state can be only attributed to a specific solute–solvent stabilization.
As seen from Table 1, the relative stability of the tautomers in ethanol (as implicit solvation) suggests around 5% of KE, which is much lower than the observed value. Even a very simple explicit solvation model with one molecule solvent attached changes the situation substantially. The relative stability of KE increases to almost 15%, which shows the reasons for substantial stabilization in alcohols. The increase in the number of attached solvent molecules would probably shift the equilibrium further to the keto tautomer, as shown in the case of two ethanol molecules being attached, where the solvent stabilization of KE is substantial. Although the structure of the KK + EtOH and K + EtOH complexes, shown in Table 1, suggest that a channel for the exchange of the proton between KK and K potentially exists, the energy of the latter is too high to be really populated in solution. It is especially interesting to mention that this solvent channel actually works in the direction of the more stable KK, as seen from the 1:2 complex of K. Of course, the data, collected in Table 1, represent a very simplified model of the reality in solution, but they show the trends in the change in the tautomeric state as a result of the specific solvation.
The discussion up to now has been based on the theoretical simulations, which logically raises the concern of how reliable these conclusions are, especially because there are no quantitative data about the position of the tautomeric EKE equilibrium in 5 and 6 to confirm the reliability of using the M06-2X/TZVP approach for the relative stability prediction. On one hand, it was previously proved that the M06-2X/TZVP level of theory predicts, in a reliable way, the tautomeric state and important structural characteristics of azodyes and Schiff bases in aprotic solvents [91]. On the other hand, experimental data are available for 7 and 8, where the same theoretical approach was used [68]. The corresponding information to compare 6–8 is collected in Table 2.
As seen from the data in Table 2, there is reasonable agreement between the theoretically obtained ∆E values and the experimental Gibbs free energies for 7 and 8. Of course, as it was underlined above, this agreement is limited to aprotic solvents, since the implicit solvation model does not account the specific solute–solvent interactions with the protic solvents. This means, for instance, that the theoretical results for alcohols, obtained by using the PCM solvation only, have to be interpreted with care. The information given in Table 1 is a very convincing example.
The data collected in Table 2 allow to make conclusions about the effect of the PTF in compounds 6-8. In all three compounds in toluene, the enol is the most stable form, but the relative stability of the KE form increases from 6 to 8, from an almost negligible amount in 6 to almost 1/3 of the total population in 8. The KK form is strongly destabilized in 7, due to steric hindrance between the tautomeric NH and the C8H proton, and in 6, due to the lack of stabilizing hydrogen bonding as in 8. The K tautomer does not exist in 7 and is very strongly destabilized in 6. The latter is a result of the difficult enolization of the carbonyl group in the coumarin backbone and/or by the strong stabilization of the enol form within the 7-hydroxy-coumarin core. In structurally similar compounds, where 7-hydroxy-coumarin plays a role of potential PTF, and where benzothiazole [106,107] and other conjugated nitrogen-containing heterocycles [22,108] are attached as PCU, there is no evidence of the existence of KK and K switched forms.
According to the data reported by Liubimov and co-authors [82], in toluene 5 exhibits a single emission band around 545 nm. The isolated PTU, 4, also show a single band, but it is substantially blue shifted (at around 380 nm), which is attributed to the local emission of the only existing in this solvent enol tautomer. Therefore, the substantial Stokes shift (11 000 cm−1) in 5 suggests an ESIPT process from E* to KK* as the authors stated. Additional irradiation does not lead to spectral changes in 5 in in this solvent.
In Figure 3, the ground- and first excited-state potential energy surfaces are sketched. According to these, and also to the general understanding of long-range PT in similar compounds [41], the excitation of the enol form through an energetically high Franck–Condon state could lead either spontaneously to conical intersection (CI) relaxation in the C=N bond isomerization region, yielding a mixture of ground-state E and Ecis, or ballistically, through ESIPT, to KE*. Then, the excited KE* can emit or, by twisting around the C-C bond (the axle in Scheme 2), can reach the twisting conical intersection region, relaxing as a mixture of KE and KK. As shown above, KE almost immediately restores to the predominant E, while KK can be accumulated due to the large relaxation barrier back (~25 kcal/mol in respect of KK). Only in such a consequence of events could an emission from KK be observed. Bearing in mind that no spectral changes are observed upon irradiation, the accumulation of KK is less probable to be assumed. Therefore, the observed red-shifted emission could be only attributed to KE*. The large Stokes shift of this emission probably comes from the substantial stabilization of KE* in respect of the enolic Franck–Condon state, as seen from the figure. Schiff base 8 also exhibits an ESIPT emission in toluene around 490 nm with a Stokes shift of 7400 cm−1 [68], which comes from it being lower in energy, compared to here, in the Franck–Condon state. An emission in cyclohexane around 480 nm, attributed to the ESIPT process [109], is measured for 7, but there are no theoretical data to make a more detailed interpretation.
The situation dramatically changes in both 4 and 5 going to an ethanol solution [82]. As seen from Figure 4, the emission spectrum of 6 depends on the excitation wavelength; a single emission band with a maximum at 530 nm is observed when excited at 461 nm (spectrum 3), whereas two emission bands with maxima at 442 nm and 530 nm are observed when excited at 375 nm (spectra 2 and 5). According to the discussion above, in the ground state, a substantial amount of KE exists in ethanol, expressed by the red-shifted band at around 450 nm. Therefore, the emission observed upon excitation at 461 nm can be attributed to the KE* as it is observed in toluene.
Since, in the region around 380 nm, both E and KE are absorbing, the excitation in this region leads to an emission at 530 nm, attributed to the existing KE, and to an emission at 442 nm (missing in toluene), which probably comes from the enol form, where the tautomeric OH group is engaged in intermolecular hydrogen bonding with the solvent. From the existing data, it is impossible to make a conclusion as to whether the intramolecular O-H…N bonding is completely or partially broken by the solvent. As seen from Figure 4, upon irradiation with an external light source, substantial spectral changes are observed, associated with a rise in the emission at 440 nm. Basically, the irradiation can lead to two processes, as shown in Figure 3—trans/cis isomerization to Ecis or switching—leading to the accumulation of KK. The latter process should be more probable in toluene, where switching relaxation back from KK to KE through TS(KE-KK) is expected to be slower. The ground-state twisting transition state is very polar, and its energy decreases from 32 kcal/mol in toluene (Table S1) to 26 kcal/mol in ethanol, considered as an implicit solvent. The real stabilization is probably higher because of the increased possibilities for specific solute–solvent interactions in the transition state, in which the NH proton does not participate in intramolecular hydrogen bonding. It is much more probable to assume trans to cis isomerization with the accumulation of some amounts of Ecis. This form (as seen from Figure S2) is blue shifted compared to E, and the intramolecular hydrogen bonding is broken, which allows additional stabilization by the solvent. The relaxation back to E proceeds through the inversion transition state (as predicted by the theoretical calculations, Figure 3), in which energy does not depend on the solvent polarity. The process of cis/trans isomerization has been already described in structural analogues of 6 by Traven and co-authors [110,111]. A very recent study by Tang and co-workers describes the breakage of the CO-O bond in alcohols under strong external irradiation [112] in 4. The same process is theoretically predicted by Krauter at al [71], which creates an additional possibility in the interpretation of the data from Figure 4.

4. Conclusions

The tautomerism of 7-hydroxy-4-methyl-8-((phenylimino)methyl)-2H-chromen-2-one has been studied by using the density functional theory in ground and excited states, and the theoretical data were considered in comparison with the existing experimental results. Although the studied compound can potentially exist as a mixture of four tautomers, the enol form dominates in the ground state in aprotic solvents. Small amounts of the short-range proton-transferred keto tautomer are observed as a function of the polarity of the solvent. In ethanol, the content of the keto tautomer is substantial as a result of its specific stabilization by the solvent. According to the theoretical data, the long-range PT could not be observed due to the difficulty of the enolization of the carbonyl group in the coumarin frame. Upon excitation, two competitive processes could be expected, namely ESIPT and trans to cis isomerization, around the azomethine bond. However, the existing experimental data and their theoretical interpretation suggest that the latter is much more probable.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/physchem4010007/s1, Figure S1: Predicted absorption spectra of the tautomers of 6 in acetonitrile; Figure S2: Predicted absorption spectra of the tautomers and isomers of 6 in ethanol; Table S1: Relative energies of the tautomers of 6 and their spectral parameters in toluene and in acetonitrile (in brackets) in the ground state.

Author Contributions

Conceptualization, L.A.; methodology, L.A. and D.N.-A.; investigation, L.A. and D.N.-A.; resources, L.A.; writing—original draft preparation, L.A.; writing—review and editing, L.A. and D.N.-A.; visualization, L.A.; supervision, L.A.; project administration, L.A.; funding acquisition, L.A. All authors have read and agreed to the published version of the manuscript.

Funding

The author thanks the Bulgarian National Science Fund, National Research Program VIHREN, by the project T-Motors (contracted as KP-06-DV-9/2019) for the financial support for this investigation. We also acknowledge the provided access to the e-infrastructure of the NCHDC Б—part of the Bulgarian National Roadmap on RIs—with the financial support of Grant No. D01-168/2022.

Data Availability Statement

Data is contained within the article or Supplementary Materials.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Weller, A. Über die Fluoreszenz der Salizylsäure und verwandter Verbindungen. Naturwissenschaften 1955, 42, 175–176. [Google Scholar] [CrossRef]
  2. Weller, A. Innermolekularer Protonenübergang Im Angeregten Zustand. Z. Elektrochem. Berichte Bunsenges. Phys. Chem. 1956, 60, 1144–1147. [Google Scholar] [CrossRef]
  3. Arnaut, L.G.; Formosinho, S.J. Excited-State Proton Transfer Reactions I. Fundamentals and Intermolecular Reactions. J. Photochem. Photobiol. A Chem. 1993, 75, 1–20. [Google Scholar] [CrossRef]
  4. Formosinho, S.J.; Arnaut, L.G. Excited-State Proton Transfer Reactions II. Intramolecular Reactions. J. Photochem. Photobiol. A Chem. 1993, 75, 21–48. [Google Scholar] [CrossRef]
  5. Hynes, J.T.; Klinman, J.P.; Limbach, H.-H.; Schowen, R.L. (Eds.) Hydrogen-Transfer Reactions; Wiley-VCH: Weinheim, Germany, 2006; ISBN 978-3-527-61154-6. [Google Scholar]
  6. 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] [CrossRef]
  7. Antonov, L. (Ed.) Tautomerism: Methods and Theories; Wiley-VCH: Weinheim, Germany, 2014; ISBN 978-3-527-33294-6. [Google Scholar]
  8. Serdiuk, I.E.; Roshal, A.D. Exploring Double Proton Transfer: A Review on Photochemical Features of Compounds with Two Proton-Transfer Sites. Dye. Pigment. 2017, 138, 223–244. [Google Scholar] [CrossRef]
  9. Joshi, H.C.; Antonov, L. Excited-State Intramolecular Proton Transfer: A Short Introductory Review. Molecules 2021, 26, 1475. [Google Scholar] [CrossRef] [PubMed]
  10. Jankowska, J.; Sobolewski, A.L. Modern Theoretical Approaches to Modeling the Excited-State Intramolecular Proton Transfer: An Overview. Molecules 2021, 26, 5140. [Google Scholar] [CrossRef] [PubMed]
  11. Kwon, J.E.; Park, S.Y. Advanced Organic Optoelectronic Materials: Harnessing Excited-State Intramolecular Proton Transfer (ESIPT) Process. Adv. Mater. 2011, 23, 3615–3642. [Google Scholar] [CrossRef] [PubMed]
  12. Tsutsui, M.; Taniguchi, M. Single Molecule Electronics and Devices. Sensors 2012, 12, 7259–7298. [Google Scholar] [CrossRef] [PubMed]
  13. Sakai, K.; Tsuzuki, T.; Itoh, Y.; Ichikawa, M.; Taniguchi, Y. Using Proton-Transfer Laser Dyes for Organic Laser Diodes. Appl. Phys. Lett. 2005, 86, 081103. [Google Scholar] [CrossRef]
  14. Chen, K.-Y.; Hsieh, C.-C.; Cheng, Y.-M.; Lai, C.-H.; Chou, P.-T. Extensive Spectral Tuning of the Proton Transfer Emission from 550 to 675 Nm via a Rational Derivatization of 10-Hydroxybenzo[h]Quinoline. Chem. Commun. 2006, 4395. [Google Scholar] [CrossRef]
  15. Stoerkler, T.; Pariat, T.; Laurent, A.D.; Jacquemin, D.; Ulrich, G.; Massue, J. Excited-State Intramolecular Proton Transfer Dyes with Dual-State Emission Properties: Concept, Examples and Applications. Molecules 2022, 27, 2443. [Google Scholar] [CrossRef]
  16. Sedgwick, A.C.; Wu, L.; Han, H.-H.; Bull, S.D.; He, X.-P.; James, T.D.; Sessler, J.L.; Tang, B.Z.; Tian, H.; Yoon, J. Excited-State Intramolecular Proton-Transfer (ESIPT) Based Fluorescence Sensors and Imaging Agents. Chem. Soc. Rev. 2018, 47, 8842–8880. [Google Scholar] [CrossRef]
  17. Li, Y.; Dahal, D.; Abeywickrama, C.S.; Pang, Y. Progress in Tuning Emission of the Excited-State Intramolecular Proton Transfer (ESIPT)-Based Fluorescent Probes. ACS Omega 2021, 6, 6547–6553. [Google Scholar] [CrossRef]
  18. Chen, L.; Fu, P.; Wang, H.; Pan, M. Excited-State Intramolecular Proton Transfer (ESIPT) for Optical Sensing in Solid State. Adv. Opt. Mater. 2021, 9, 2001952. [Google Scholar] [CrossRef]
  19. Shekhovtsov, N.A.; Bushuev, M.B. Enol or Keto? Interplay between Solvents and Substituents as a Factor Controlling ESIPT. J. Mol. Liq. 2022, 361, 119611. [Google Scholar] [CrossRef]
  20. Shekhovtsov, N.A.; Nikolaenkova, E.B.; Ryadun, A.A.; Vorobyeva, S.N.; Krivopalov, V.P.; Bushuev, M.B. Dual Emission of ESIPT-Capable 2-(2-Hydroxyphenyl)-4-(1 H-Pyrazol-1-Yl)Pyrimidines: Interplay of Fluorescence and Phosphorescence. New J. Chem. 2023, 47, 6361–6377. [Google Scholar] [CrossRef]
  21. Shekhovtsov, N.A.; Ryadun, A.A.; Bushuev, M.B. Luminescence of a Zinc(II) Complex with a Protonated 1-Hydroxy-1 H-imidazole ESIPT Ligand: Direct Excitation of a Tautomeric Form. ChemistrySelect 2021, 6, 12346–12350. [Google Scholar] [CrossRef]
  22. Nikolaeva, O.G.; Popova, O.S.; Dubonosova, I.V.; Karlutova, O.Y.; Dubonosov, A.D.; Bren, V.A.; Minkin, V.I. Spectral-Luminescent and Ionochromic Properties of Azomethine Imine-Coumarin Conjugates. Russ. J. Gen. Chem. 2022, 92, 841–849. [Google Scholar] [CrossRef]
  23. Minkin, V.I.; Tsukanov, A.V.; Dubonosov, A.D.; Bren, V.A. Tautomeric Schiff Bases: Iono-, Solvato-, Thermo- and Photochromism. J. Mol. Struct. 2011, 998, 179–191. [Google Scholar] [CrossRef]
  24. Wang, D.; Li, S.-J.; Cao, W.; Wang, Z.; Ma, Y. ESIPT-Active 8-Hydroxyquinoline-Based Fluorescence Sensor for Zn(II) Detection and Aggregation-Induced Emission of the Zn(II) Complex. ACS Omega 2022, 7, 18017–18026. [Google Scholar] [CrossRef] [PubMed]
  25. Bren, V.A.; Dubonosov, A.D.; Popova, O.S.; Revinskii, Y.V.; Tikhomirova, K.S.; Minkin, V.I. Synthesis and Photo- and Ionochromic and Spectral-Luminescent Properties of 5-Phenylpyrazolidin-3-One Azomethine Imines. Int. J. Photoenergy 2018, 2018, 9746534. [Google Scholar] [CrossRef]
  26. Bren, V.A.; Popova, O.S.; Tolpygin, I.E.; Chernoivanov, V.A.; Revinskii, Y.V.; Dubonosov, A.D. New Ionochromic Azomethinimine Chemosensors. Russ. Chem. Bull. 2015, 64, 668–671. [Google Scholar] [CrossRef]
  27. Popova, O.S.; Bren’, V.A.; Tkachev, V.V.; Utenyshev, A.N.; Revinskii, Y.V.; Tikhomirova, K.S.; Starikov, A.G.; Borodkin, G.S.; Dubonosov, A.D.; Shilov, G.V.; et al. Benzenoid-Quinoid Tautomerism of Azomethines and Their Structural Analogs 56. Azomethine Imines, Derivatives of Salicylic and 2-Hydroxynaphthoic Aldehydes. Russ. Chem. Bull. 2016, 65, 648–653. [Google Scholar] [CrossRef]
  28. Shekhovtsov, N.A.; Vorob’eva, S.; Nikolaenkova, E.B.; Ryadun, A.A.; Krivopalov, V.P.; Gourlaouen, C.; Bushuev, M.B. Complexes on the Base of a Proton Transfer Capable Pyrimidine Derivative: How Protonation and Deprotonation Switch Emission Mechanisms. Inorg. Chem. 2023, 62, 16734–16751. [Google Scholar] [CrossRef]
  29. Fu, P.-Y.; Yi, S.-Z.; Pan, M.; Su, C.-Y. Excited-State Intramolecular Proton Transfer (ESIPT) Based Metal–Organic Supramolecular Optical Materials: Energy Transfer Mechanism and Luminescence Regulation Strategy. Acc. Mater. Res. 2023, 4, 939–952. [Google Scholar] [CrossRef]
  30. Shekhovtsov, N.A.; Ryadun, A.A.; Plyusnin, V.F.; Nikolaenkova, E.B.; Tikhonov, A.Y.; Bushuev, M.B. First 1-Hydroxy-1 H-Imidazole-Based ESIPT Emitter with an O–H⋯O Intramolecular Hydrogen Bond: ESIPT-Triggered TICT and Speciation in Solution. New J. Chem. 2022, 46, 22804–22817. [Google Scholar] [CrossRef]
  31. Shekhovtsov, N.A.; Nikolaenkova, E.B.; Berezin, A.S.; Plyusnin, V.F.; Vinogradova, K.A.; Naumov, D.Y.; Pervukhina, N.V.; Tikhonov, A.Y.; Bushuev, M.B. Tuning ESIPT-Coupled Luminescence by Expanding π-Conjugation of a Proton Acceptor Moiety in ESIPT-Capable Zinc(ii) Complexes with 1-Hydroxy-1 H-Imidazole-Based Ligands. Dalton Trans. 2022, 51, 15166–15188. [Google Scholar] [CrossRef]
  32. Trannoy, V.; Léaustic, A.; Gadan, S.; Guillot, R.; Allain, C.; Clavier, G.; Mazerat, S.; Geffroy, B.; Yu, P. A Highly Efficient Solution and Solid State ESIPT Fluorophore and Its OLED Application. New J. Chem. 2021, 45, 3014–3021. [Google Scholar] [CrossRef]
  33. Shekhovtsov, N.A.; Nikolaenkova, E.B.; Berezin, A.S.; Plyusnin, V.F.; Vinogradova, K.A.; Naumov, D.Y.; Pervukhina, N.V.; Tikhonov, A.Y.; Bushuev, M.B. A 1-Hydroxy-1 H-imidazole ESIPT Emitter Demonstrating anti-Kasha Fluorescence and Direct Excitation of a Tautomeric Form. ChemPlusChem 2021, 86, 1436–1441. [Google Scholar] [CrossRef]
  34. Petdee, S.; Chaiwai, C.; Benchaphanthawee, W.; Nalaoh, P.; Kungwan, N.; Namuangruk, S.; Sudyoadsuk, T.; Promarak, V. Imidazole-Based Solid-State Fluorophores with Combined ESIPT and AIE Features as Self-Absorption-Free Non-Doped Emitters for Electroluminescent Devices. Dye. Pigment. 2021, 193, 109488. [Google Scholar] [CrossRef]
  35. Long, Y.; Mamada, M.; Li, C.; Dos Santos, P.L.; Colella, M.; Danos, A.; Adachi, C.; Monkman, A. Excited State Dynamics of Thermally Activated Delayed Fluorescence from an Excited State Intramolecular Proton Transfer System. J. Phys. Chem. Lett. 2020, 11, 3305–3312. [Google Scholar] [CrossRef] [PubMed]
  36. Shekhovtsov, N.A.; Nikolaenkova, E.B.; Ryadun, A.A.; Samsonenko, D.G.; Tikhonov, A.Y.; Bushuev, M.B. ESIPT-Capable 4-(2-Hydroxyphenyl)-2-(Pyridin-2-Yl)-1H-Imidazoles with Single and Double Proton Transfer: Synthesis, Selective Reduction of the Imidazolic OH Group and Luminescence. Molecules 2023, 28, 1793. [Google Scholar] [CrossRef] [PubMed]
  37. Park, S.; Kim, S.; Seo, J.; Park, S.Y. Application of Excited-State Intramolecular Proton Transfer (ESIPT) Principle to Functional Polymeric Materials. Macromol. Res. 2008, 16, 385–395. [Google Scholar] [CrossRef]
  38. Smith, T.P.; Zaklika, K.A.; Thakur, K.; Walker, G.C.; Tominaga, K.; Barbara, P.F. Ultrafast Studies on Proton Transfer in Photostabilizers. J. Photochem. Photobiol. A Chem. 1992, 65, 165–175. [Google Scholar] [CrossRef]
  39. Feringa, B.L.; Browne, W.R. (Eds.) Molecular Switches, 2nd ed.; Wiley-VCH: Weinheim, Germany, 2011; ISBN 978-3-527-63440-8. [Google Scholar]
  40. Nedeltcheva-Antonova, D.; Antonov, L. Controlled Tautomerism: Is It Possible? In Tautomerism: Concepts and Applications in Science and Technology; Antonov, L., Ed.; Wiley-VCH: Weinheim, Germany, 2016; pp. 273–294. ISBN 978-3-527-69571-3. [Google Scholar]
  41. Sobolewski, A.L. Reversible Molecular Switch Driven by Excited-State Hydrogen Transfer. Phys. Chem. Chem. Phys. 2008, 10, 1243–1247. [Google Scholar] [CrossRef]
  42. van der Loop, T.H.; Ruesink, F.; Amirjalayer, S.; Sanders, H.J.; Buma, W.J.; Woutersen, S. Unraveling the Mechanism of a Reversible Photoactivated Molecular Proton Crane. J. Phys. Chem. B 2014, 118, 12965–12971. [Google Scholar] [CrossRef]
  43. Vetokhina, V.; Nowacki, J.; Pietrzak, M.; Rode, M.F.; Sobolewski, A.L.; Waluk, J.; Herbich, J. 7-Hydroxyquinoline-8-Carbaldehydes. 1. Ground- and Excited-State Long-Range Prototropic Tautomerization. J. Phys. Chem. A 2013, 117, 9127–9146. [Google Scholar] [CrossRef]
  44. Lapinski, L.; Rostkowska, H.; Nowacki, J.; Nowak, M.J. Photoinduced Long-Distance Hydrogen-Atom Transfer in Molecules with a 7-Hydroxyquinoline Frame and a Carbaldehyde or Aldoxime Group as the Intramolecular Hydrogen Transporting Crane. J. Phys. Chem. A 2023, 127, 3104–3113. [Google Scholar] [CrossRef]
  45. de Bekker, E.J.A.; Geerlings, J.D.; Varma, C.A.G.O. Mechanism of a Photoinduced Solvent-Assisted Transfer of a Proton to a Specified Remote Target. J. Phys. Chem. A 2000, 104, 5916–5927. [Google Scholar] [CrossRef]
  46. Jalink, C.J.; Huizer, A.H.; Varma, C.A.G.O. Rotational Brownian Motion of the Proton Crane Operating in a Photoinduced Long-Distance Intramolecular Proton Transfer. J. Chem. Soc. Faraday Trans. 1992, 88, 1643. [Google Scholar] [CrossRef]
  47. Jalink, C.J.; Huizer, A.H.; Varma, C.A.G.O. Rate-Limiting Action of a Proton Crane in Long-Range Intramolecular Proton Transfer. J. Chem. Soc. Faraday Trans. 1992, 88, 2655–2659. [Google Scholar] [CrossRef]
  48. Kaczmarek, L.S. Bipyridines. Part XVIII. On the Synthesis of [2,2′-Bipyridine]-3-Ol and Other Novel 2,2′-Bipyridine Derivatives from [2,2′-Bipyridine]-3,3′-Diamine. Bull. Pol. Acad. Sci. Tech. Sci. 1985, 33, 401–409. [Google Scholar]
  49. Kaczmarek, Ł.; Balicki, R.; Lipkowski, J.; Borowicz, P.; Grabowska, A. Structure and Photophysics of Deazabipyridyls. Excited Internally Hydrogen-Bonded Systems with One Proton Transfer Reaction Site. J. Chem. Soc. Perkin Trans. 2 1994, 1603. [Google Scholar] [CrossRef]
  50. Borowicz, P.; Grabowska, A.; Leś, A.; Kaczmarek, Ł.; Zagrodzki, B. New Phototautomerizing Systems: Non-Symmetric Derivatives of [2,2′-Bipyridyl]-3,3′-Diol. Chem. Phys. Lett. 1998, 291, 351–359. [Google Scholar] [CrossRef]
  51. Ewing, G.W.; Steck, E.A. Absorption Spectra of Heterocyclic Compounds. I. Quinolinols and Isoquinolinols. J. Am. Chem. Soc. 1946, 68, 2181–2187. [Google Scholar] [CrossRef]
  52. Mason, S.F. The Tautomerism of N-Heteroaromatic Hydroxy-Compounds. Part I. Infrared Spectra. J. Chem. Soc. 1957, 4874–4880. [Google Scholar] [CrossRef]
  53. Mason, S.F. The Tautomerism of N-Heteroaromatic Hydroxy-Compounds. Part II. Ultraviolet Spectra. J. Chem. Soc. 1957, 5010–5017. [Google Scholar] [CrossRef]
  54. Mason, S.F.; Philp, J.; Smith, B.E. Prototropic Equilibria of Electronically Excited Molecules. Part II. 3-, 6-, and 7-Hydroxyquinoline. J. Chem. Soc. A 1968, 3051–3056. [Google Scholar] [CrossRef]
  55. Lee, S.-I.; Jang, D.-J. Proton Transfers of Aqueous 7-Hydroxyquinoline in the First Excited Singlet, Lowest Triplet, and Ground States. J. Phys. Chem. 1995, 99, 7537–7541. [Google Scholar] [CrossRef]
  56. García-Ochoa, I.; Bisht, P.B.; Sánchez, F.; Martinez-Atáz, E.; Santos, L.; Tripathi, H.B.; Douhal, A. Experimental and Theoretical Studies of the Proton-Hopping Reaction of 7-Hydroxyquinoline in Viscous Hydroxylic Media. J. Phys. Chem. A 1998, 102, 8871–8880. [Google Scholar] [CrossRef]
  57. Chou, P.-T.; Wei, C.-Y.; Chris Wang, C.-R.; Hung, F.-T.; Chang, C.-P. Proton-Transfer Tautomerism of 7-Hydroxyquinolines Mediated by Hydrogen-Bonded Complexes. J. Phys. Chem. A 1999, 103, 1939–1949. [Google Scholar] [CrossRef]
  58. Kwon, O.-H.; Lee, Y.-S.; Yoo, B.K.; Jang, D.-J. Excited-State Triple Proton Transfer of 7-Hydroxyquinoline along a Hydrogen-Bonded Alcohol Chain: Vibrationally Assisted Proton Tunneling. Angew. Chem. Int. Ed. 2006, 45, 415–419. [Google Scholar] [CrossRef] [PubMed]
  59. Park, S.-Y.; Lee, Y.-S.; Kwon, O.-H.; Jang, D.-J. Proton Transport of Water in Acid–Base Reactions of 7-Hydroxyquinoline. Chem. Commun. 2009, 926. [Google Scholar] [CrossRef]
  60. Park, S.-Y.; Jang, D.-J. Excited-State Hydrogen Relay along a Blended-Alcohol Chain as a Model System of a Proton Wire: Deuterium Effect on the Reaction Dynamics. Phys. Chem. Chem. Phys. 2012, 14, 8885. [Google Scholar] [CrossRef]
  61. Kumpulainen, T.; Lang, B.; Rosspeintner, A.; Vauthey, E. Ultrafast Elementary Photochemical Processes of Organic Molecules in Liquid Solution. Chem. Rev. 2017, 117, 10826–10939. [Google Scholar] [CrossRef] [PubMed]
  62. Rode, M.F.; Sobolewski, A.L. Effect of Chemical Substituents on the Energetical Landscape of a Molecular Photoswitch: An Ab Initio Study. J. Phys. Chem. A 2010, 114, 11879–11889. [Google Scholar] [CrossRef]
  63. Benesch, C.; Rode, M.F.; Čížek, M.; Rubio-Pons, O.; Thoss, M.; Sobolewski, A.L. Switching the Conductance of a Single Molecule by Photoinduced Hydrogen Transfer. J. Phys. Chem. C 2009, 113, 10315–10318. [Google Scholar] [CrossRef]
  64. Georgiev, A.; Antonov, L. 8-(Pyridin-2-Yl)Quinolin-7-Ol as a Platform for Conjugated Proton Cranes: A DFT Structural Design. Micromachines 2020, 11, 901. [Google Scholar] [CrossRef]
  65. Csehi, A.; Illés, L.; Halász, G.J.; Vibók, Á. The Effect of Chemical Substituents on the Functionality of a Molecular Switch System: A Theoretical Study of Several Quinoline Compounds. Phys. Chem. Chem. Phys. 2013, 15, 18048–18054. [Google Scholar] [CrossRef]
  66. Csehi, A.; Halász, G.J.; Vibók, Á. Molecular Switch Properties of 7-Hydroxyquinoline Compounds. Int. J. Quantum Chem. 2014, 114, 1135–1145. [Google Scholar] [CrossRef]
  67. Nakashima, K.; Petek, A.; Hori, Y.; Georgiev, A.; Hirashima, S.; Matsushima, Y.; Yordanov, D.; Miura, T.; Antonov, L. Acylhydrazone Subunits as a Proton Cargo Delivery System in 7-Hydroxyquinoline. Chem. Eur. J. 2021, 27, 11559–11566. [Google Scholar] [CrossRef] [PubMed]
  68. Georgiev, A.; Yordanov, D.; Ivanova, N.; Deneva, V.; Vassilev, N.; Kamounah, F.S.; Pittelkow, M.; Crochet, A.; Fromm, K.M.; Antonov, L. 7-OH Quinoline Schiff Bases: Are They the Long Awaited Tautomeric Bistable Switches? Dye. Pigment. 2021, 195, 109739. [Google Scholar] [CrossRef]
  69. Grzywacz, J.; Taszner, S.; Kruszewski, J. Further Study on the Forms of Umbelliferone in Excited State. Z. Naturforschung A 1978, 33, 1307–1311. [Google Scholar] [CrossRef]
  70. Taszner, S.; Grzywacz, J. Further Study on the Forms of Umbelliferone in Excited State. Part II. Z. Naturforschung A 1979, 34, 548–550. [Google Scholar] [CrossRef]
  71. Krauter, C.M.; Möhring, J.; Buckup, T.; Pernpointner, M.; Motzkus, M. Ultrafast Branching in the Excited State of Coumarin and Umbelliferone. Phys. Chem. Chem. Phys. 2013, 15, 17846. [Google Scholar] [CrossRef]
  72. Jacquemin, D.; Perpète, E.A.; Scalmani, G.; Frisch, M.J.; Assfeld, X.; Ciofini, I.; Adamo, C. Time-Dependent Density Functional Theory Investigation of the Absorption, Fluorescence, and Phosphorescence Spectra of Solvated Coumarins. J. Chem. Phys. 2006, 125, 164324. [Google Scholar] [CrossRef]
  73. Sakurai, T.; Mizoguchi, H.; Yamada, S.; Inoue, H. Proton-Transfer Reactions between 7-Hydroxycoumarin and Tertiary Amines. Formation of the Ground-State Tautomer Anion. Ber. Bunsenges. Phys. Chem. 1996, 100, 505–507. [Google Scholar] [CrossRef]
  74. Moriya, T. Excited-State Reactions of Coumarins in Aqueous Solutions. I. The Phototautomerization of 7-Hydroxycoumarin and Its Derivative. Bull. Chem. Soc. Jpn. 1983, 56, 6–14. [Google Scholar] [CrossRef]
  75. Georgieva, I.; Trendafilova, N.; Aquino, A.; Lischka, H. Excited State Properties of 7-Hydroxy-4-Methylcoumarin in the Gas Phase and in Solution. A Theoretical Study. J. Phys. Chem. A 2005, 109, 11860–11869. [Google Scholar] [CrossRef] [PubMed]
  76. Georgieva, I.; Trendafilova, N.; Aquino, A.J.A.; Lischka, H. Excited-State Proton Transfer in 7-Hydroxy-4-Methylcoumarin along a Hydrogen-Bonded Water Wire. J. Phys. Chem. A 2007, 111, 127–135. [Google Scholar] [CrossRef] [PubMed]
  77. De Silva, N.; Minezawa, N.; Gordon, M.S. Excited-State Hydrogen Atom Transfer Reaction in Solvated 7-Hydroxy-4-Methylcoumarin. J. Phys. Chem. B 2013, 117, 15386–15394. [Google Scholar] [CrossRef]
  78. Abdel-Mottaleb, M.S.A.; El-Sayed, B.A.; Abo-Aly, M.M.; El-Kady, M.Y. Fluorescence-Properties and Excited State Interactions of 7-Hydroxy-4-Methylcoumarin Laser Dye. J. Photochem. Photobiol. A Chem. 1989, 46, 379–390. [Google Scholar] [CrossRef]
  79. Schulman, S.G.; Rosenberg, L.S. Tautomerization Kinetics of 7-Hydroxy-4-Methylcoumarin in the Lowest Excited Singlet State. J. Phys. Chem. 1979, 83, 447–451. [Google Scholar] [CrossRef]
  80. Stamm, A.; Schwing, K.; Gerhards, M. Investigation of the Hydrated 7-Hydroxy-4-Methylcoumarin Dimer by Combined IR/UV Spectroscopy. J. Chem. Phys. 2014, 141, 194304. [Google Scholar] [CrossRef] [PubMed]
  81. Westlake, B.C.; Paul, J.J.; Bettis, S.E.; Hampton, S.D.; Mehl, B.P.; Meyer, T.J.; Papanikolas, J.M. Base-Induced Phototautomerization in 7-Hydroxy-4-(Trifluoromethyl)Coumarin. J. Phys. Chem. B 2012, 116, 14886–14891. [Google Scholar] [CrossRef]
  82. Liubimov, A.V.; Venidiktova, O.V.; Valova, T.M.; Shienok, A.I.; Koltsova, L.S.; Liubimova, G.V.; Popov, L.D.; Zaichenko, N.L.; Barachevsky, V.A. Photochromic and Luminescence Properties of a Hybrid Compound Based on Indoline Spiropyran of the Coumarin Type and Azomethinocoumarin. Photochem. Photobiol. Sci. 2018, 17, 1365–1375. [Google Scholar] [CrossRef]
  83. Zaichenko, N.L.; Valova, T.M.; Venidiktova, O.V.; Lyubimov, A.V.; Shienok, A.I.; Koltsova, L.S.; Ayt, A.O.; Lyubimova, G.V.; Popov, L.D.; Barachevsky, V.A. Excited State Proton Transfers in Hybrid Compound Based on Indoline Spiropyran of the Coumarin Type and Azomethinocoumarin in the Presence of Metal Ions. Molecules 2021, 26, 6894. [Google Scholar] [CrossRef]
  84. Antonov, L.; Fabian, W.M.F.; Nedeltcheva, D.; Kamounah, F.S. Tautomerism of 2-Hydroxynaphthaldehyde Schiff Bases. J. Chem. Soc. Perkin Trans. 2 2000, 1173–1179. [Google Scholar] [CrossRef]
  85. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 16 Rev. C.01, Gaussian, Inc.: Wallingford, CT, USA, 2019.
  86. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef] [PubMed]
  87. Schlegel, H.B. Optimization of Equilibrium Geometries and Transition Structures. J. Comput. Chem. 1982, 3, 214–218. [Google Scholar] [CrossRef]
  88. Zhao, Y.; Truhlar, D.G. Density Functionals with Broad Applicability in Chemistry. Acc. Chem. Res. 2008, 41, 157–167. [Google Scholar] [CrossRef]
  89. Zhao, Y.; Truhlar, D.G. The M06 Suite of Density Functionals for Main Group Thermochemistry, Thermochemical Kinetics, Noncovalent Interactions, Excited States, and Transition Elements: Two New Functionals and Systematic Testing of Four M06-Class Functionals and 12 Other Functionals. Theor. Chem. Acc. 2008, 120, 215–241. [Google Scholar] [CrossRef]
  90. Schäfer, A.; Huber, C.; Ahlrichs, R. Fully Optimized Contracted Gaussian Basis Sets of Triple Zeta Valence Quality for Atoms Li to Kr. J. Chem. Phys. 1994, 100, 5829–5835. [Google Scholar] [CrossRef]
  91. Antonov, L. Tautomerism in Azo and Azomethyne Dyes: When and If Theory Meets Experiment. Molecules 2019, 24, 2252. [Google Scholar] [CrossRef] [PubMed]
  92. Rayne, S.; Forest, K. A Comparative Examination of Density Functional Performance against the ISOL24/11 Isomerization Energy Benchmark. Comput. Theor. Chem. 2016, 1090, 147–152. [Google Scholar] [CrossRef]
  93. Improta, R. UV-Visible Absorption and Emission Energies in Condensed Phase by PCM/TD-DFT Methods. In Computational Strategies for Spectroscopy; Barone, V., Ed.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2011; pp. 37–75. ISBN 978-1-118-00872-0. [Google Scholar]
  94. Adamo, C.; Jacquemin, D. The Calculations of Excited-State Properties with Time-Dependent Density Functional Theory. Chem. Soc. Rev. 2013, 42, 845–856. [Google Scholar] [CrossRef]
  95. Yanai, T.; Tew, D.P.; Handy, N.C. A New Hybrid Exchange–Correlation Functional Using the Coulomb-Attenuating Method (CAM-B3LYP). Chem. Phys. Lett. 2004, 393, 51–57. [Google Scholar] [CrossRef]
  96. Li, R.; Zheng, J.; Truhlar, D.G. Density Functional Approximations for Charge Transfer Excitations with Intermediate Spatial Overlap. Phys. Chem. Chem. Phys. 2010, 12, 12697. [Google Scholar] [CrossRef]
  97. Wang, J.; Durbeej, B. How Accurate Are TD-DFT Excited-state Geometries Compared to DFT Ground-state Geometries? J Comput Chem 2020, 41, 1718–1729. [Google Scholar] [CrossRef]
  98. Mahato, B.; Panda, A.N. Assessing the Performance of DFT Functionals for Excited-State Properties of Pyridine-Thiophene Oligomers. J. Phys. Chem. A 2021, 125, 115–125. [Google Scholar] [CrossRef] [PubMed]
  99. Sarkar, R.; Boggio-Pasqua, M.; Loos, P.-F.; Jacquemin, D. Benchmarking TD-DFT and Wave Function Methods for Oscillator Strengths and Excited-State Dipole Moments. J. Chem. Theory Comput. 2021, 17, 1117–1132. [Google Scholar] [CrossRef]
  100. Louant, O.; Champagne, B.; Liégeois, V. Investigation of the Electronic Excited-State Equilibrium Geometries of Three Molecules Undergoing ESIPT: A RI-CC2 and TDDFT Study. J. Phys. Chem. A 2018, 122, 972–984. [Google Scholar] [CrossRef]
  101. Rode, M.F.; Nedeltcheva-Antonova, D.; Antonov, L. Long-Range Proton Transfer in 7-Hydroxy-Quinoline-Based Azomethine Dyes: A Hidden Reason for the Low Efficiency. Molecules 2022, 27, 8225. [Google Scholar] [CrossRef] [PubMed]
  102. Antonov, L.; Kawauchi, S.; Okuno, Y. Prediction of the Color of Dyes by Using Time-Dependent Density Functional Theory. Bulg. Chem. Commun. 2014, 46, 228–237. [Google Scholar]
  103. Becke, A.D. Density-functional Thermochemistry. III. The Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef]
  104. Antonov, L. Drawbacks of the Present Standards for Processing Absorption Spectra Recorded Linearly as a Function of Wavelength. TrAC Trends Anal. Chem. 1997, 16, 536–543. [Google Scholar] [CrossRef]
  105. Antonov, L. (Ed.) Absorption UV-Vis Spectroscopy and Chemometrics: From Qualitative Conclusions to Quantitative Analysis. In Tautomerism: Methods and Theories; Wiley-VCH: Weinheim, Germany, 2013; pp. 25–47. ISBN 978-3-527-65882-4. [Google Scholar]
  106. Xie, L.; Chen, Y.; Wu, W.; Guo, H.; Zhao, J.; Yu, X. Fluorescent Coumarin Derivatives with Large Stokes Shift, Dual Emission and Solid State Luminescent Properties: An Experimental and Theoretical Study. Dye. Pigment. 2012, 92, 1361–1369. [Google Scholar] [CrossRef]
  107. Miksa, B. Fluorescent Dyes Used in Polymer Carriers as Imaging Agents in Anticancer Therapy. Med. Chem. 2016, 6, 611–639. [Google Scholar] [CrossRef]
  108. Karcz, D.; Starzak, K.; Ciszkowicz, E.; Lecka-Szlachta, K.; Kamiński, D.; Creaven, B.; Jenkins, H.; Radomski, P.; Miłoś, A.; Ślusarczyk, L.; et al. Novel Coumarin-Thiadiazole Hybrids and Their Cu(II) and Zn(II) Complexes as Potential Antimicrobial Agents and Acetylcholinesterase Inhibitors. Int. J. Mol. Sci. 2021, 22, 9709. [Google Scholar] [CrossRef] [PubMed]
  109. Joshi, H.; Kamounah, F.S.; Gooijer, C.; van der Zwan, G.; Antonov, L. Excited State Intramolecular Proton Transfer in Some Tautomeric Azo Dyes and Schiff Bases Containing an Intramolecular Hydrogen Bond. J. Photochem. Photobiol. A Chem. 2002, 152, 183–191. [Google Scholar] [CrossRef]
  110. Traven, V.F.; Miroshnikov, V.S.; Pavlov, A.S.; Ivanov, I.V.; Panov, A.V.; Chibisova, T.A. Unusual E/Z-Isomerization of 7-Hydroxy-4-Methyl-8-[(9H-Fluoren-2-Ylimino)Methyl]-2H-1-Benzopyran-2-One in Acetonitrile. Mendeleev Commun. 2007, 17, 88–89. [Google Scholar] [CrossRef]
  111. Traven’, V.F.; Ivanov, I.V.; Panov, A.V.; Safronova, O.B.; Chibisova, T.A. Solvent-Induced E/Z(C=N)-Isomerization of Imines of Some Hydroxy-Substituted Formylcoumarins. Russ. Chem. Bull. 2008, 57, 1989–1995. [Google Scholar] [CrossRef]
  112. Tang, Q.; Xian, Z.; Li, L.; Cao, H.; Wang, L.; Zheng, X. Study on the Photochemical Reaction Process of 4-Methyl-7-Hydroxycoumarin and Its Mechanism by Multi-Spectroscopic Technologies. J. Phys. Conf. Ser. 2021, 2112, 012023. [Google Scholar] [CrossRef]
Scheme 1. Sketch of a molecular switch, based on a long-range intramolecular PT. The red color indicates proton donor sites, while blue is used for proton acceptor sites. The twisting axle is shown in green. Below, three compounds, as experimentally studied examples, are shown.
Scheme 1. Sketch of a molecular switch, based on a long-range intramolecular PT. The red color indicates proton donor sites, while blue is used for proton acceptor sites. The twisting axle is shown in green. Below, three compounds, as experimentally studied examples, are shown.
Physchem 04 00007 sch001
Scheme 2. Enol (4E) and keto (4K) tautomers of 7-hydroxy-4-methylcoumarin (up). Possible tautomeric forms and mechanism for intramolecular long-range PT in the Schiff bases, discussed in the current study (down). The skeleton important for the switching is given in bold lines. The twisting angle a is indicated in magenta.
Scheme 2. Enol (4E) and keto (4K) tautomers of 7-hydroxy-4-methylcoumarin (up). Possible tautomeric forms and mechanism for intramolecular long-range PT in the Schiff bases, discussed in the current study (down). The skeleton important for the switching is given in bold lines. The twisting angle a is indicated in magenta.
Physchem 04 00007 sch002
Scheme 3. Schiff bases, possessing intramolecular hydrogen bonding.
Scheme 3. Schiff bases, possessing intramolecular hydrogen bonding.
Physchem 04 00007 sch003
Figure 1. Predicted absorption spectra of the tautomers of 6 in toluene (in acetonitrile, see Figure S1).
Figure 1. Predicted absorption spectra of the tautomers of 6 in toluene (in acetonitrile, see Figure S1).
Physchem 04 00007 g001
Figure 2. Potential energy surface of 6 in the ground state. DE is the relative total energy, while DG is the relative Gibbs free energy. Detailed information is provided in Table S1.
Figure 2. Potential energy surface of 6 in the ground state. DE is the relative total energy, while DG is the relative Gibbs free energy. Detailed information is provided in Table S1.
Physchem 04 00007 g002
Figure 3. Ground-(M06-2X) and first excited-state (CAM-B3LYP) potential energy surface of 6 in toluene. The stable structures are indicated in blue; the transition states are indicated in red. The Franck–Condon state of the enol form is shown in green. The relative energies are given in kcal/mol units. In the excited states, the oscillator strength is listed after the energy, separated by a slash. The values of the angles α (describing the twisting motion around the axle, see Scheme 2) and β (describing the trans to cis isomerization of the azomethine bond) are given in brackets; the second one is underlined.
Figure 3. Ground-(M06-2X) and first excited-state (CAM-B3LYP) potential energy surface of 6 in toluene. The stable structures are indicated in blue; the transition states are indicated in red. The Franck–Condon state of the enol form is shown in green. The relative energies are given in kcal/mol units. In the excited states, the oscillator strength is listed after the energy, separated by a slash. The values of the angles α (describing the twisting motion around the axle, see Scheme 2) and β (describing the trans to cis isomerization of the azomethine bond) are given in brackets; the second one is underlined.
Physchem 04 00007 g003
Figure 4. Absorption spectrum (1), fluorescence spectra at λex = 332 (5), 375 nm (2, 7), and 461 nm (3), and fluorescence excitation spectrum at λreg = 461 nm (4) before (1–5) and after (6, 7) UV irradiation with the external source of high intensity of compound (4) in ethanol. Reproduced from [82] with permission from the Springer Nature BV.
Figure 4. Absorption spectrum (1), fluorescence spectra at λex = 332 (5), 375 nm (2, 7), and 461 nm (3), and fluorescence excitation spectrum at λreg = 461 nm (4) before (1–5) and after (6, 7) UV irradiation with the external source of high intensity of compound (4) in ethanol. Reproduced from [82] with permission from the Springer Nature BV.
Physchem 04 00007 g004
Table 1. Relative energies of the 1:1 and 1:2 complexes of 6 with ethanol molecules in the ethanol environment. The corresponding values for the single tautomers in the same environment are given in brackets.
Table 1. Relative energies of the 1:1 and 1:2 complexes of 6 with ethanol molecules in the ethanol environment. The corresponding values for the single tautomers in the same environment are given in brackets.
StructureRelative Energetics
[kcal/mol]
(M06-2X/TZVP)
StructureRelative Energetics
[kcal/mol]
(M06-2X/TZVP)
∆E∆∆G ∆E∆∆G
Physchem 04 00007 i001
E + EtOH
0.0
(0.0)
0.0
(0.0)
Physchem 04 00007 i002
E + 2EtOH
1.340.24
Physchem 04 00007 i003
KE + EtOH
0.99
(1.80)
1.36
(2.50)
Physchem 04 00007 i004
KE + 2EtOH
0.00.0
Physchem 04 00007 i005
KK + EtOH
3.4
(6.0)
3.9
(6.8)
Physchem 04 00007 i006
KK + 2EtOH
0.932.63
Physchem 04 00007 i007
K + EtOH
17
(30)
15
(29)
K + 2EtOHspontaneously goes to KK
Table 2. Theoretically predicted ∆E and ∆G values compared with the experimentally determined ∆G values at room temperature in toluene and acetonitrile (in brackets).
Table 2. Theoretically predicted ∆E and ∆G values compared with the experimentally determined ∆G values at room temperature in toluene and acetonitrile (in brackets).
StructureRelative Energetics
[kcal/mol]
678
M06-2X/TZVPM06-2X/TZVP[84]M06-2X/TZVP[68]
∆E∆G∆E∆G∆Gexp∆E∆G∆Gexp
E0.0
(0.0)
0.0
(0.0)
0.0
(0.0)
0.0
(0.0)
0.0
(0.0)
0.0
(0.37)
0.0
(0.10)

(0.52)
KE2.8
(1.76)
3.5
(2.4)
1.70
(0.55)
1.80
(0.60)
1.42 *
(0.35)
0.52
(0.0)
0.62
(0.0)

(0.0)
KK7.3
(6.0)
8.0
(6.7)
9.0
(7.1)
9.8
(8.3)
-1.9
(0.70)
2.05
(0.63)
-
(1.0)
K32
(30)
32
(29)
---7.6
(4.8)
7.5
(4.7)
-
* in cyclohexane.
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

Nedeltcheva-Antonova, D.; Antonov, L. Ground-State Tautomerism and Excited-State Proton Transfer in 7-Hydroxy-4-methyl-8-((phenylimino)methyl)-2H-chromen-2-one as a Potential Proton Crane. Physchem 2024, 4, 91-105. https://doi.org/10.3390/physchem4010007

AMA Style

Nedeltcheva-Antonova D, Antonov L. Ground-State Tautomerism and Excited-State Proton Transfer in 7-Hydroxy-4-methyl-8-((phenylimino)methyl)-2H-chromen-2-one as a Potential Proton Crane. Physchem. 2024; 4(1):91-105. https://doi.org/10.3390/physchem4010007

Chicago/Turabian Style

Nedeltcheva-Antonova, Daniela, and Liudmil Antonov. 2024. "Ground-State Tautomerism and Excited-State Proton Transfer in 7-Hydroxy-4-methyl-8-((phenylimino)methyl)-2H-chromen-2-one as a Potential Proton Crane" Physchem 4, no. 1: 91-105. https://doi.org/10.3390/physchem4010007

Article Metrics

Back to TopTop