You are on page 1of 10

pubs.acs.

org/JACS Article

Chapter Open for the Excited-State Intramolecular Thiol Proton


Transfer in the Room-Temperature Solution
Chun-Hsiang Wang,§ Zong-Ying Liu,§ Chun-Hao Huang,§ Chao-Tsen Chen,* Fan-Yi Meng,
Yu-Chan Liao, Yi-Hung Liu, Chao-Che Chang, Elise Y. Li,* and Pi-Tai Chou*
Cite This: J. Am. Chem. Soc. 2021, 143, 12715−12724 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: We report here, for the first time, the experimental observation on
Downloaded via INDIAN INST OF TECH BOMBAY on June 28, 2023 at 05:29:55 (UTC).

the excited-state intramolecular proton transfer (ESIPT) reaction of the thiol


proton in room-temperature solution. This phenomenon is demonstrated by a
derivative of 3-thiolflavone (3TF), namely, 2-(4-(diethylamino)phenyl)-3-
mercapto-4H-chromen-4-one (3NTF), which possesses an SH···O intra-
molecular H-bond (denoted by the dashed line) and has an S1 absorption at 383
nm. Upon photoexcitation, 3NTF exhibits a distinctly red emission maximized at
710 nm in cyclohexane with an anomalously large Stokes shift of 12 230 cm−1.
Upon methylation on the thiol group, 3MeNTF, lacking the thiol proton, exhibits
a normal Stokes-shifted emission at 472 nm. These, in combination with the
computational approaches, lead to the conclusion of thiol-type ESIPT
unambiguously. Further time-resolved study renders an unresolvable (<180 fs)
ESIPT rate for 3NTF, followed by a tautomer emission lifetime of 120 ps. In sharp
contrast to 3NTF, both 3TF and 3-mercapto-2-(4-(trifluoromethyl)phenyl)-4H-
chromen-4-one (3FTF) are non-emissive. Detailed computational approaches indicate that all studied thiols undergo thermally
favorable ESIPT. However, once forming the proton-transferred tautomer, the lone-pair electrons on the sulfur atom brings non-
negligible nπ* contribution to the S1′ state (prime indicates the proton-transferred tautomer), for which the relaxation is dominated
by the non-radiative deactivation. For 3NTF, the extension of π-electron delocalization by the diethylamino electron-donating group
endows the S1′ state primarily in the ππ* configuration, exhibiting the prominent tautomer emission. The results open a new chapter
in the field of ESIPT, covering the non-canonical sulfur intramolecular H-bond and its associated ESIPT at ambient temperature.

■ INTRODUCTION
Along the group 16 (VIA) family of the periodic table, the
the electronically excited state.11−14 Upon photoexcitation,
driven by the changes of the acidity and basicity in the excited
conventional wisdom learned from textbooks tells us that only state, the transfer of a proton from the OH or NH
H2O possesses a hydrogen bond to explain its abnormally high proton-donating site to the O or N proton-accepting site
boiling point that cannot fit into the correlation of increasing becomes feasible. When it takes place within the molecular
boiling point with increasing molecular weight from H2S and unit, the process is dubbed as the excited-state intramolecular
H2Se to H2Te. However, this does not rule out the existence of proton transfer (ESIPT), which has been receiving consid-
H-bonds for other group 16 family members. The formation of erable attention in both fundamental15−23 and applied
an H-bond incorporating a sulfur atom as either the proton researches.24−30 On the one hand, the result of ESIPT forms
donor (SH) or acceptor (S) has long been a research topic a proton-transfer isomer, namely, the tautomer, in the excited
of great interest. In fact, nature does show sulfur containing H- state, which does not exist in the ground state thermally and
bonds. For example, some previous reports have evidenced accordingly gives rise to an anomalously large Stokes shift
that the methionine-containing dipeptides can form amide emission (cf. the non-proton transfer, normal emission). Such
NH···S H-bonds that are even stronger than amideN phototautomerism offers ample fundamental insights and can
be specified with a barrierless process, a barrier induced, or a
H···OC H-bonds.1,2 In artificial DNA or RNA researches,3
sulfur H-bonds have been popularly applied in RNAi resistant
genes4 and protein engineering.5 Moreover, recent high- Received: May 31, 2021
resolution spectroscopy in combination with quantum Published: August 6, 2021
chemical calculations has revealed that sulfur-containing H-
bonds can be as strong as conventional H-bonds.6−10
Among a variety of H-bond associated research directions,
one field should be credited to the proton transfer reaction in

© 2021 American Chemical Society https://doi.org/10.1021/jacs.1c05602


12715 J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

reversible process.31−36 On the other hand, the diversification


in the ESIPT mechanism renders different emission properties
in terms of single, dual, and even multiple emissions that can
be fine-tuned by varying the functionality, offering the
versatility in applications such as in sensing,37−40 bioimag-
ing, 41−43 organic lasing, 44−46 and light-emitting di-
odes.34,44,47,48
The above OH (NH)···N (or O) type H-bond, in
theory, is formed via a static dipole−dipole interaction where
both the proton-donating/-accepting elements possess large
electronegativities (>3.0 in Pauling scale). Driven by the
changes of electronic configuration, ESIPT along the H-bond
then takes place. In comparison, the H-bonds involving thiol Figure 1. Molecular structures of 3HF and its derivatives of 3TF and
SH (proton donor) or thioneS (proton acceptor) are a 3MeNTF.
result of their inherent dispersity of the electron cloud.
Therefore, whether ESIPT takes place via the sulfur H-bond is molecular H-bond. 3-HF and its derivatives undergo
of prime fundamental importance. In this regard, in the prominent OH-type ESIPT, which have been widely
cryogenic temperature, the matrix-isolated thiotropolone studied.13,55−58
seems to be capable of undergoing proton transfer via We then report herein, for the first time, the observation of
vibrational excitation in the electronic ground state.49 Also, ESIPT in 3NTF incorporating thiol proton, forming a proton-
in the low-temperature matrixes, Waluk and co-workers have transferred thione tautomer, which gives rise to a prominent
reported the occurrence of ESIPT in β-thioxoketones via red emission maximized at 710 nm in solution and solid state
vibration analyses.9,50−52 One report proposed the occurrence at room temperature. We also infer that ESIPT may not be
of ESIPT in a four-membered ring H-bonding system with a uncommon in the thiol H-bonded polyaromatic molecules,
thiol group as a proton donor; however, the lack of direct opening up a new chapter of excited-state intramolecular thiol
evidence and the calculated rather high barrier make the proton transfer reaction. Details of the results and discussions
conclusion elusive.53 We recently made attempts by synthesiz- are elaborated below.
ing the thione (S)···HO H-bonded molecule 7-hydroxy-
2,2-dimethyl-2,3-dihydro-1H-indene-1-thione (DM-7HIT)54
and examining its possibility of spawning ESIPT. The results,
■ RESULTS AND DISCUSSION
Synthesis and Characterization. One may be skeptical
on the one hand, were disappointing when we found no about why reports on the photophysics of thiol H-bonded
ESIPT. We tentatively rationalized the results by the fact that molecules are so scarce. Chemically, one major reason lies in
the lowest lying excited singlet state is dominated by the nπ* its stability. Thiol contained molecules, similar to the natural
character, i.e., an S1(nπ*) state, where n is ascribed to the amino acid cysteine, are prone to undergo oxidation to form a
higher lying lone-pair electrons of thione (S). This results in disulfide SS bond and thus are difficult to obtain the pure
the drastic decrease of the thione electron-accepting strength thiol compounds and perform photophysical studies. It
and hence the prohibition of ESIPT after photoexcitation. On requires special caution to exclude air during the synthesis
the other hand, the decrease of thione basicity ruptures the ( and purification to prevent the occurrence of oxidation.
S)···HO intramolecular H-bond such that the H-bond on/ Syntheses and characterization of 3TF and its derivatives are
off reaction takes place with respect to light on/off, elaborated below with pertinent characterization data listed in
demonstrating a prototype of photoinduced ultrafast molecular Figures S1−S40.
switch.54 Synthesis and Characterization of 3TF and its
Therefore, up to this stage, on the basis of the absorption Derivatives. Two different synthetic routes (shown in
and emission spectroscopies, unambiguous experimental Scheme 1) were attempted to synthesize 3TF. The first
evidence clearly addressing the ESIPT of sulfur-containing route, in part, was according to the literature,59 where flavone
H-bonded molecules at room temperature, which is practical in (2) molecular framework was constructed first by reacting 2′-
applications, is still pending. Amid the quest of ESIPT hydroxyacetophenone (1) with either benzoyl chloride60 or
associated with sulfur H-bond, in light of the aforementioned benzaldehyde61 and then followed by LDA-mediated lithiation
higher lying energy of the thione lone-pair electrons, we then and nucleophilic addition with sulfur powder to yield 3TF
strategically switched the sulfur H-bond from the OH along with the 3TF dimer with an SS linkage, namely, SSF
(donor)/thione (acceptor) pair to the configuration of SH (see Figure S1) as a minor product. It was suspected that the
(thiol, donor)/CO (acceptor) pair. In comparison to thione, oxidation of 3TF to SSF occurred during the workup
the lower lying lone-pair electrons of thiol, in theory, should procedure as well as during the course of purification. Thus,
make the S2(nπ*) state appreciably higher in energy than that several freeze−pump−thaw cycles were taken to degas all the
of the S1(ππ*) state. Accordingly, in this study, a series of solvents used in step b (see Scheme 1) to prevent the
sulfur intramolecular H-bonded molecules, 3-thiolflavone oxidation. However, these efforts turned out to be in vain.
(3TF) and its derivatives 3FTF and 3NTF, were strategically Therefore, the alternative synthetic route was sought, where
designed and synthesized (see Figure 1). Also, the methylated reacting 1 with benzaldehyde gave conjugated ketone (3) via
(SCH3) derivative for 3NTF, namely, 3MeNTF (see Figure Claisen−Schmidt condensation, then followed by intra-
1), was synthesized for comparison. 3MeNTF lacks a labile molecular Michael addition to yield chromanone (4). The
proton to ensure no occurrence of ESIPT. Note that the bromination of 4 led to diastereomers (5), which were used
replacement of thiol by the hydroxyl group of 3TF forms 3- directly for the subsequent nucleophilic substitution with
hydroxyflavone (3HF) possessing the OH···OC intra- potassium thioacetate in the presence of 18-crown-6 ether as a
12716 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

Scheme 1. Two Synthetic Routes Used to Synthesize 3TF and the Synthesis of 3FTF, 3NTF, and 3MeTFa

a
Reagents and conditions: (a) for 2a, (i) benzoyl chloride, K2CO3, pyridine, acetone, reflux, 18 h, (ii) AcOH, acetone, reflux, 24 h; for 2b and 2c,
(i) 4-(trifluoromethyl)benzaldehyde (2b) or 4-(diethylamino)benzaldehyde (2c), KOH, EtOH, rt, 3 h, (ii) I2, DMSO, 130 °C, 1 h; (b) (i) freshly
prepped LDA, THF, −78 °C, 30 min, (ii) S powder, −78 °C, 1 h, (iii) 10% HCl(aq), 58% (3TF), 5% (SSF); (c) Benzaldehyde, KOH(aq), EtOH, rt,
3 h; (d) NaOAc, EtOH, reflux, 24 h; (e) Br2, CCl4, rt, 2 h; (f) KSAc, 18-crown-6, MeCN, 30 min; (g) DDQ, 1,4-dioxane, reflux, 16 h; (h) (i)
NaOMe, MeOH, rt, 1 h, (ii) 10% HCl, 0 °C, 30 min, 67% (3TF), 12% (SSF), yield was based on 1H-NMR; (i) I2, TEA, MeOH, 30 min; (j) (i)
NaBH4 (3 equiv), EtOH/CHCl3, 0 °C, 1 h, (ii) citric acid (5 equiv), 0 °C, 30 min; (k) MeI (3 equiv), TEA, THF, rt, 1 h. (l) NaBH4 (3 equiv),
EtOH/CHCl3, 0 °C, 1 h, MeI (5 equiv).

phase transfer agent to yield thioacetate chromanone (6). The


oxidation of 6 with dichloro-5,6-dicyano-p-benzoquinone
(DDQ) afforded thioacetate chromenone (7). It was expected
to obtain the desired 3TF after the hydrolysis of 7 under basic
conditions (MeONa/MeOH) and exhausted degassed solvents
as well as an inert atmosphere. However, the mixture of flavone
3TF and SSF was again isolated (see Figure S2).
Recently, Angeli et al. reported a convenient synthetic
procedure to obtain a series of aryl selenols (SeHs) from their
diselenides derivatives in high yield via the reduction with
sodium borohydride (NaBH4) and followed by acidification
especially with citric acid.62 This method was thus attempted
for the synthesis of 3TF. Starting with the mixture of 3TF and
SSF acquired from route 1, SSF was obtained via oxidation
with iodine and triethylamine (TEA), and then followed by
reduction with NaBH4 and acidification with citric acid Figure 2. Mid-IR spectrum of 3NTF.
successfully to afford pure 3TF in good yield. If non-reductive
acids such as HCl were used for acidification, SSF would
appear along with 3TF (see Figure S3). With this successful cm−1.10,66,67 Additionally, a previous report also claimed that
case, 3NTF and 3FTF were synthesized as the same fashion the vibration frequency of an H-bonded thiol is ca. 2500
with acceptable yields. Notably, 3TF, 3FTF, and 3NTF cm−1.67 Therefore, the formation of intramolecular H-bond is
displayed sharp singlet peaks corresponding to the thiol unambiguous for all three title compounds. To further confirm
protons at δ = 5.24, 5.34, and 5.37 ppm in nuclear magnetic the chemical structures, the single crystals of 3TF, 3NTF, and
resonance (NMR) spectra, respectively (see Figure S4), which 3FTF were grown in degassed ethyl acetate (EtOAc) at −4 °C
are more downfield shifted compared with the non-H-bonded under argon and their structures were solved at room
sulfhydryl proton (δ is around 3.4 ppm).63−65 These results temperature, respectively. Their X-ray crystallographic struc-
indicate the existence of intramolecular hydrogen bonding in tures were shown in Figure 3 and Figure S39 and S40. 3NTF
titled molecules. This viewpoint can also be further supported exhibits shorter a SO distance (2.925 Å) than 3TF (2.996
by the mid-IR spectra of 3TFs, where a broad and weak SH Å) and 3FTF (2.956 Å), which, empirically, indicates that
stretching mode absorption can be observed at 2493, 2502, 3NTF possesses the strongest H-bond among the three
and 2500 cm−1 for 3NTF (see Figure 2), 3TF, and 3FTF (see molecules.33 Indirectly, this is verified by the computational
Figure S38), respectively. These frequencies are smaller than simulation in which 3NTF indeed has the shortest HO
the free SH stretching frequency around 2600−2550 bond length (2.049 Å) than that of 3TF (2.064 Å) and 3FTF
12717 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 3. X-ray crystallographic structure of 3NTF. The displacement


ellipsoids are drawn at the 50% probability level, and the H atoms are
drawn as spheres of arbitrary radii.

(2.057 Å) in the ground-state geometry (see Figure S41). In


addition, a smaller dihedral angle (28°) between the
chromenone and the phenyl ring was noted in 3NTF
compared to 46° (3TF) and 41° (3FTF), suggesting that
the delocalization of electrons from the diethylamino group to Figure 4. (a) Absorption (circle and line) and emission (solid line)
the carbonyl group is feasible. Detailed data can be found in spectra of 3NTF (black) and 3MeNTF (orange) in cyclohexane at
the Cambridge Data Bank (CCDC No. 2084999, 2085000, 298 K. Also shown are the excitation spectrum monitored at emission
and 2085002). The control compound 3MeNTF was wavelength of 710 (green line) and 770 nm (dashed blue line). The
synthesized either from the reduction of the associated SS solid-state emission of 3NTF is shown in a solid red line. Due to
linked 3NTF dimer (NSSF, see Scheme 1), followed by rather weak emission for 3MeNTF, which is strongly interfered by
quenching with methyl iodide, to afford 86% yield, or reacting intense scattering light, the emission profile was fitted by a Gaussian
3NTF with methyl iodide to afford 91% yield. function (dashed orange line). Note: All spectra are measured in
cyclohexane with a concentration of ∼10−5 M. λex = 380 and 340 nm
Spectroscopic Studies. Figure 4a shows the absorption for 3NTF and 3MeNTF, respectively. Inset: the corresponding
and emission spectra of 3NTF in cyclohexane, where the absorption and fluorescence images of 3NTF in cyclohexane. (b)
absorption shows the lowest lying transition (S0 →S1) at 383 Proposed ESIPT of 3NTF, forming a zwitterionic-type tautomer with
nm. Upon UV excitation, 3NTF exhibits a prominent red two possible canonical structures. Note that the red arrow indicates
emission maximized at 710 nm in cyclohexane (see the inset of the moving direction of the proton.
Figure 4a). The excitation spectra monitored at an entire
emission band (e.g., 710 (green line) and 770 nm (dashed blue with an anomalously large Stokes shift of 12 230 cm−1
line), see Figure 4a) are identical, which are also the same as (difference between absorption peak and emission peak) for
the absorption spectrum. The results clearly indicate that the 3NTF can be unambiguously ascribed to the tautomer
710 nm emission band originates from the absorption of pure emission resulting from SH···O → S···HO
3NTF. In comparison, replacing SH of 3NTF by OH, ESIPT process, forming a zwitterionic-type tautomer shown in
forming 3NHF (see Figure 1 for the structure58), 3NHF Figure 4b.
exhibits an absorption peak at 400 nm and undergoes OH- The occurrence of ESIPT can be further supported by the
type ESIPT, resulting in a tautomer emission at 555 nm in emission spectra in various solvents, which are shown in Figure
nonpolar solvent,58 in which both absorption and emission are 5. With increasing solvent polarity, a new emission band can be
drastically different from that of 3NTF in this study. The result observed at the 400−500 nm region, which shifts bath-
discards the possibility that the origin of the 710 nm emission
for 3NTF is from any trace contaminated OH derivative
3NHF that undergoes ESIPT. In fact, there was no origin for
yielding 3NHF during the synthesis of 3NTF (see Scheme 1).
Also, a possible origin of the 710 nm emission from any
aggregation effect is eliminated by the observation of
concentration-independent absorption and emission spectra
within the studied concentration range 10−5−10−4 M (see
Figure S42). Another possibility is that the observed emission
band at 710 nm may originate from the thiyl radical due to the
photodissociation after UV radiation. This mechanism can be
safely excluded by the TDDFT calculation and transient
absorption measurement (Figures S43−S45 and Table S1;
also, please see detailed discussion in the Supporting
Information (SI)). Upon the methylation of 3NTF, forming
the SCH3 derivative 3MeNTF (see Figure 1), the
absorption and rather weak normal emission maxima appear
at 370 and 472 nm, respectively. 3MeNTF lacks a proton for Figure 5. Absorption (circle and line) and emission (solid line)
transfer and hence renders normal Stokes-shifted emission (see spectra for 3NTF in toluene (black), dichloromethane (red), and
Figure 4a). In brief, in cyclohexane, the 710 nm emission band acetonitrile (blue) at 298 K. λex = 390 nm.

12718 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

ochromically. On the basis of the previous report on the OH 0.14%, the radiative rate constant kr is deduced to be 1.2 × 107
derivative 3NHF,58 the emission at the 400−500 nm region s−1. In addition, the radiative rate constant (kr) for normal
can be ascribed to the normal form emission, which appears to form species was calculated to be 2.24 × 108 s−1 by Strickler
result from the solvent polarity induced barrier. The and Berg formula,69,70 which is an order of magnitude larger
pronounced solvatochromic property observed for normal than that (1.2 × 107 s−1) for the tautomer state. This can be
form emission indicates the charge transfer nature of the attributed to the mixing of ππ* with a small degree of nπ*
normal form species. This implies that the proton-coupled transition character in the tautomer S1 state, which makes the
electron transfer (PCET) mechanism is operative under the emission less allowed, hence reducing the value of kr (vide
perturbation of solvent polarity. A similar emission profile can infra). The early relaxation dynamics of the 710 nm emission
be observed in ethanol solution (Figure S46). However, the was then further measured by the femtosecond fluorescence
excitation spectrum monitored at 500 nm is slightly blue- up-conversion technique (see the SI for the experimental
shifted compared with the absorption spectrum. Therefore, the detail). The results, shown in the inset of Figure 6b, reveal a
emission bands at 500 and 700 nm can be attributed to the fast decay time of 1.58 ps, followed by a population decay of
non-intramolecular H-bonded normal species and the proton- 120 ps; however, the rise time corresponding to ESIPT cannot
transferred tautomer, respectively.68 Overall, the results all be resolved in our measurement. Also, the kinetic trace
indicate the occurrence of ESIPT, where the normal form monitored at 470 nm, which is expected to be in the normal
emission can be observed under polar or protic solvents’ form emission region, demonstrates a system unresolvable fast
perturbation. To the best of our knowledge, this provides the decay kinetics (<180 fs, see the inset of Figure 6b). The result
first spectroscopic evidence of ESIPT for the intramolecular is consistent with the unresolvable rise kinetics observed upon
thiol H-bonding system in a room-temperature environment. monitoring at the tautomer emission. These results indicate
Similarly, a prominent red emission of 3NTF resulting from that the ESIPT rate for 3NTF is ultrafast and beyond the
ESIPT is also observed in solid with a peak located at 660 nm detection limit of our instrument (temporal resolution ∼180
(red line in Figure 4a). The slightly blue-shifted emission in fs), for which the associated barrier must be negligibly small
solid (cf. solution) is plausibly due to the influence of lattice (vide infra). The fast decay time gradually decreases as the
energy. monitored wavelength of the tautomer emission decreases; on
To gain more insights into the thiol ESIPT kinetics, the contrary, the amplitude and the value of the fast decay time
picosecond time-correlated single photon counting (TCSPC) exhibit an opposite trend. This indicates that certain structural
was carried out. The result of TCSPC measurement for 3NTF relaxation takes place in the excited state (see Figure S47 and
in cyclohexane showed a single exponential decay constant of Table S2). This viewpoint also can be supported by our
120 ps (see Figure 6a). Taking the emission quantum yield of computational results, which show that the dihedral angle
between the phenyl group and the chromenone moiety
decreases from 33.91° to 25.96° during the ESIPT process.
Hence, we rationally assigned the fast decay component to the
structural relaxation of the phenyl ring (see Figure S48).
As for the experimental progress, lastly, we also intend to
extend our observation to two other analogues, i.e., 3TF and
3FTF, standing for the parent (H) and the electron-
withdrawing (CF3) substituents, respectively. Unfortunately,
to our best efforts in improving the detection sensitivity, we
were not able to resolve any meaningful signal that can be
ascribed to the emission of 3TF and 3FTF in cyclohexane at
room temperature. Note that a certain weak emission band
maximized at 540 and 550 nm was observed for 3TF and
3FTF, respectively. However, the excitation spectra monitored
at 540 (3TF) and 550 nm (3FTF) are entirely different from
that of the corresponding absorption spectra (see Figure S49).
As elaborated in detail in the Supporting Information, the
observed emission bands are more likely to originate from the
ion-pair emission of 3TF and 3FTF,71,72 where the trace of
3TF and 3FTF anions exists due to the high acidity of the thiol
group. Such an ion-pair, though ≪1% population and is not
shown in the absorption spectra, has a high emission yield and
thus is non-negligible in the contribution of the emission when
we maximize the detectivity. This conjecture is supported by
purging a trace of the HCl gas into cyclohexane solution,
where the ion-pair emission disappears (see Figure S50). Such
ion-pair emission is negligible in 3NTF due to the stronger
Figure 6. (a) Decay dynamics of the 710 nm emission for 3NTF in
intramolecular H-bonding strength. Note that ion-pair
cyclohexane (black line) and its fitting curve (blue line). Also shown
is the instrument response profile (red solid line) of the TCSPC (see emission spectrum also can be observed for 3NTF in
text for details). (b) Kinetic trace at 700 nm emission acquired by the cyclohexane in the presence of a trace amount of triethylamine
fluorescence up-conversion method. Inset: the kinetic trace at 470 nm (Figure S51), which shows the emission centered at 550 nm,
emission (black circle) and instrument response function (blue line). distinct from the tautomer emission (see Figure 4a). TD-DFT
λex = 390 nm. computations also confirm a strong ππ* emission at 597 nm
12719 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

from the ion pair formed by the deprotonated 3NTF and a On the contrary, according to nearly barrierless ESIPT, 3TF
Na+ as the counterion (see Figures S52 and S53 and Tables S3 and 3FTF are expected to populate at the tautomer state as
and S4). In a brief summary, no emission originating from 3TF well. Experimentally, however, the tautomer emission could
and 3FTF can be resolved in cyclohexane at room temper- not be resolved at this stage. The results indicate the dominant
ature, which is in sharp contrast to the remarkable 710 nm radiationless deactivation pathways for the tautomer S1′ state
proton-transfer tautomer emission observed in 3NTF. Such for both 3TF and 3FTF. As for the rationalization, first, the
discrepancy infers perhaps different relaxation pathways tautomer emissions for 3TF and 3FTF were estimated to be
between 3NTF and 3TF (or 3FTF) in the excited state. As 901 and 951 nm, respectively, according to the calculation (see
elaborated below, this viewpoint is firmly supported by the Figure 8). Compared to the observed 710 nm tautomer
computational approach. emission for 3NTF, the much red-shifted 901 and 951 nm
Computational Approach. In this computational ap- emissions predicted for 3TF and 3FTF, respectively, are to be
proach, the ground-state and excited-state geometries are subject to tremendous radiationless deactivation due to the
optimized by density functional theory (DFT) and time- electronic-state (S1′)−vibration-state (in S0′) coupling, i.e., the
dependent DFT (TD-DFT),73−75 respectively, under B3LYP76 consequence of the energy gap law.78 Moreover, as shown in
/6-311++g(3df,3pd)77 for all title compounds. Figure 7 and Tables S7 and S8, in sharp contrast to the strongly ππ*
dominating nature of the S1′ state for 3NTF, the S1′ for 3TF
and 3FTF is composed of a mixing of ππ* (∼80%) and
relatively high nπ* (∼20%) configuration. Note that all three
compounds exhibit a pair of closely lying S1′ and S2′ states
(within 0.3 eV) at the tautomer ground state, showing
complementary ππ* or nπ* characters (see Figure 7 and
Figures S54−56). At the relaxed tautomer S1′ geometry, these
two states are brought to even closer (within 0.2 eV) such that
certain extent of mixing occurs between the two. For clarity,
Figure 8 succinctly depicts the states and their electronic
configuration involved in ESIPT. When anchoring an electron-
withdrawing group such as 3FTF, the mixing of nπ* clearly
increases in the lowest singlet excited state (see Figure 8 and
Table S8). Once nπ* becomes non-negligible in the tautomer
lowest singlet excited state (S1′ state), the large spin−orbit
coupling parameter of the sulfur atom of the thiocarbonyl
group may also trigger the intersystem crossing to populate at
T1′ (ππ*) state according to Figure S55 (for 3TF) and Figure
S56 (for 3FTF).79 Such a lower energy T1′ state (<1 eV) in
the near IR II (>1000 nm) region should also undergo
dominant radiationless deactivation governed by the emission
energy gap law.78 This viewpoint is supported by the non-
emissive 3TF and 3FTF even in the 77 K cyclohexane solid
Figure 7. Calculated various energy levels for 3NTF in the optimized
S0 and S1 geometries at the normal form and the proton-transferred
matrix. Therefore, despite the theoretical prediction of ESIPT
tautomer form (the tautomer states are denoted with a prime sign). for both 3TF and 3FTF, they are virtually non-emissive in
Please see the text and SI for the details of computational room temperature and cryogenic matrixes. As for 3NTF, the
methodology. electron-donating diethylamino group at the 4′-position
elongates the π-conjugation by coupling with the carbonyl
Figure S54 show the energy diagram of 3NTF (for 3TF and electron acceptor (see Figure 1), retaining most of the
3FTF, see Figures S55 and S56) at different states under dominating ππ* character in the S1′ geometry (Figure 7)
and hence the remarkable 710 nm tautomer emission.


different optimized geometries. Two remarks can be pointed
out from the computational results. First, for all title
compounds, we failed to obtain the S1 geometry of the normal CONCLUSION
state. In other words, the optimization beginning at the As elaborated above, we provide unambiguous spectroscopic
Franck−Condon excited state of the normal form always ends evidence for the excited-state intramolecular thiol proton
up in the energy minimum having the proton transfer tautomer transfer in the room-temperature solution and solid, which is
structure (S1′ state in Figure 7, where prime indicates also firmly supportive by theoretical approaches. This finding
tautomer). This may not be surprising as it indicates that the not only opens a new chapter in the field of ESIPT but also
potential energy surface along the ESIPT coordinate is nearly may stimulate the research activity in probing the relationship
barrierless, which is consistent with the consequence of time- between the non-classical thiol H-bonding strength and
resolved measurement of ultrafast (<180 fs) ESIPT in 3NTF. molecular structure in ESIPT, which is of great fundamental
Second, the assignment of the optimized S1′ state for 3NTF is interest in terms of chemical bonding and reaction. From the
primarily in the ππ*(>91%) configuration (see Figure 8 and viewpoint of applications, the facile preparation of the −S−S−
Figure S57 and Table S6). Such an allowed transition (f ∼ linkage makes feasible the strategic design of the thiol H-
0.13) is expected to undergo a radiative emission. The vertical bonded ESIPT molecules in an −S−S− oxidation form.
S1′ (ππ*/nπ*) →S0′ transition is estimated to be 770 nm Subsequently, the analytes induced reduction to the thiol
(Table S5), which, within the calculation uncertainty, matches molecules having ESIPT property may offer an excellent
the experimentally observed 710 nm emission. sensing platform where the anomalously large Stokes shifted
12720 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

Figure 8. Proposed ESIPT, the electronic configuration of the lowest tautomer singlet excited state and its relaxation processes. The percentage
denotes the ratio of two different configurations. IC: internal conversion

tautomer emission can be exploited as a unique signal Authors


transducer free from interference. Relevant researches have Chun-Hsiang Wang − Department of Chemistry, National
been initiated and are currently in progress. Taiwan University, Taipei 10617, Taiwan (R.O.C.)

■ ASSOCIATED CONTENT
* Supporting Information

Zong-Ying Liu − Department of Chemistry, National Taiwan
University, Taipei 10617, Taiwan (R.O.C.); orcid.org/
0000-0003-1254-7153
The Supporting Information is available free of charge at Chun-Hao Huang − Department of Chemistry, National
https://pubs.acs.org/doi/10.1021/jacs.1c05602. Taiwan Normal University, Taipei 11677, Taiwan (R.O.C.)
Fan-Yi Meng − Department of Chemistry, National Taiwan
Discussions of experimental procedures, synthetic University, Taipei 10617, Taiwan (R.O.C.)
methods, photophysical properties, possibility of the Yu-Chan Liao − Department of Chemistry, National Taiwan
radical emission, and computational results, figures of University, Taipei 10617, Taiwan (R.O.C.)
NMR spectra, IR spectra, X-ray crystallographic Yi-Hung Liu − Department of Chemistry, National Taiwan
structures, computed ground-state geometries, absorp- University, Taipei 10617, Taiwan (R.O.C.)
tion and emission spectra, transient absorption spec- Chao-Che Chang − Department of Chemistry, National
trum, transient spectral profile, kinetic traces, calculated Taiwan University, Taipei 10617, Taiwan (R.O.C.)
molecular orbitals, and calculated various energy levels,
and tables of calculated excitation energies, time- Complete contact information is available at:
resolved data, and theoretical absorption and lumines- https://pubs.acs.org/10.1021/jacs.1c05602
cence data (PDF)
Author Contributions
Accession Codes §
C.-H.W., Z.-Y.L., and C.-H.H. contributed equally.
CCDC 2084999−2085000 and 2085002 contain the supple- Notes
mentary crystallographic data for this paper. These data can be The authors declare no competing financial interest.


obtained free of charge via www.ccdc.cam.ac.uk/data_request/
cif, or by emailing data_request@ccdc.cam.ac.uk, or by ACKNOWLEDGMENTS
contacting The Cambridge Crystallographic Data Centre, 12
Union Road, Cambridge CB2 1EZ, UK; fax: +44 1223 336033. The authors appreciate the Ministry of Science and


Technology (MOST 109-2639-M-002-001-ASP, MOST 106-
AUTHOR INFORMATION 2113-M-002-012-MY3, and MOST 109-2113-M-003-009),
Taiwan and National Taiwan University, for their generous
Corresponding Authors support. The authors are grateful for the computational
Chao-Tsen Chen − Department of Chemistry, National resources provided by the Center for Cloud Computing in
Taiwan University, Taipei 10617, Taiwan (R.O.C.); National Taiwan Normal University.


Email: chenct@ntu.edu.tw
Elise Y. Li − Department of Chemistry, National Taiwan REFERENCES
Normal University, Taipei 11677, Taiwan (R.O.C.); (1) Biswal, H. S.; Wategaonkar, S. Nature of the N-H···S Hydrogen
orcid.org/0000-0003-1206-1110; Email: eliseytli@ Bond. J. Phys. Chem. A 2009, 113, 12763−12773.
ntnu.edu.tw (2) Chand, A.; Sahoo, D. K.; Rana, A.; Jena, S.; Biswal, H. S. The
Pi-Tai Chou − Department of Chemistry, National Taiwan Prodigious Hydrogen Bonds with Sulfur and Selenium in Molecular
University, Taipei 10617, Taiwan (R.O.C.); orcid.org/ Assemblies, Structural Biology, and Functional Materials. Acc. Chem.
0000-0002-8925-7747; Email: chop@ntu.edu.tw Res. 2020, 53, 1580−1592.

12721 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

(3) Villalobos, A.; Ness, J. E.; Gustafsson, C.; Minshull, J.; (23) Wei, Y. C.; Zhang, Z. Y.; Chen, Y. A.; Wu, C. H.; Liu, Z. Y.; Ho,
Govindarajan, S. Gene Designer: A Synthetic Biology Tool for S. Y.; Liu, J. C.; Lin, J. A.; Chou, P. T. Mechanochromism Induced
Constructing Artificial DNA Segments. BMC Bioinf. 2006, 7, 285. through the Interplay between Excimer Reaction and Excited State
(4) Kumar, D.; Gustafsson, C.; Klessig, D. F. Validation of RNAi Intramolecular Proton Transfer. Commun. Chem. 2019, 2, 10.
Silencing Specificity Using Synthetic Genes: Salicylic Acid-Binding (24) Sedgwick, A. C.; Wu, L. L.; Han, H. H.; Bull, S. D.; He, X. P.;
Protein 2 Is Required for Innate Immunity in Plants. Plant J. 2006, 45, James, T. D.; Sessler, J. L.; Tang, B. Z.; Tian, H.; Yoon, J. Excited-
863−868. State Intramolecular Proton-Transfer (ESIPT) Based Fluorescence
(5) Gustafsson, C.; Govindarajan, S.; Minshull, J. Putting Engineer- Sensors and Imaging Agents. Chem. Soc. Rev. 2018, 47, 8842−8880.
ing Back into Protein Engineering: Bioinformatic Approaches to (25) Okamoto, H.; Itani, K.; Yamaji, M.; Konishi, H.; Ota, H.
Catalyst Design. Curr. Opin. Biotechnol. 2003, 14, 366−370. Excited-State Intramolecular Proton Transfer (ESIPT) Fluorescence
(6) Gonzalez, L.; Mo, O.; Yanez, M. High-Level Ab Initio from 3-Amidophthalimides Displaying RGBY Emission in the Solid
Calculations on the Intramolecular Hydrogen Bond in Thiomalo- State. Tetrahedron Lett. 2018, 59, 388−391.
naldehyde. J. Phys. Chem. A 1997, 101, 9710−9719. (26) Zhang, Z. Y.; Chen, Y. A.; Hung, W. Y.; Tang, W. F.; Hsu, Y.
(7) Jablonski, M.; Kaczmarek, A.; Sadlej, A. J. Estimates of the H.; Chen, C. L.; Meng, F. Y.; Chou, P. T. Control of the Reversibility
Energy of Intramolecular Hydrogen Bonds. J. Phys. Chem. A 2006, of Excited-State Intramolecular Proton Transfer (ESIPT) Reaction:
110, 10890−10898. Host-Polarity Tuning White Organic Light Emitting Diode on a New
(8) Nowroozi, A.; Roohi, H.; Hajiabadi, H.; Raissi, H.; Khalilinia, E.; Thiazolo[5,4-d]thiazole ESIPT System. Chem. Mater. 2016, 28,
Birgan, M. N. O-H···S Intramolecular Hydrogen Bond in 8815−8824.
Thiomalonaldehyde Derivatives; a Quantum Chemical Study. (27) Donner, J. S.; Thompson, S. A.; Kreuzer, M. P.; Baffou, G.;
Comput. Theor. Chem. 2011, 963, 517−524. Quidant, R. Mapping Intracellular Temperature Using Green
(9) Posokhov, Y.; Gorski, A.; Spanget-Larsen, J.; Duus, F.; Hansen, Fluorescent Protein. Nano Lett. 2012, 12, 2107−2111.
P. E.; Waluk, J. The structure of the phototransformation product of (28) Paige, J. S.; Wu, K. Y.; Jaffrey, S. R. RNA Mimics of Green
monothiodibenzoylmethane. Chem. Phys. Lett. 2001, 350, 502−508. Fluorescent Protein. Science 2011, 333, 642−646.
(10) Mishra, K. K.; Borish, K.; Singh, G.; Panwaria, P.; Metya, S.; (29) Kwon, J. E.; Park, S. Y. Advanced Organic Optoelectronic
Madhusudhan, M. S.; Das, A. Observation of an Unusually Large IR Materials: Harnessing Excited-State Intramolecular Proton Transfer
Red-Shift in an Unconventional S−H···S Hydrogen-Bond. J. Phys. (ESIPT) Process. Adv. Mater. 2011, 23, 3615−3642.
Chem. Lett. 2021, 12, 1228−1235. (30) Lim, S.-J.; Seo, J.; Park, S. Y. Photochromic Switching of
(11) Barbara, P. F.; Brus, L. E.; Rentzepis, P. M. Intramolecular Excited-State Intramolecular Proton-Transfer (ESIPT) Fluorescence:
Proton-Transfer and Excited-State Relaxation in 2-(2- A Unique Route to High-Contrast Memory Switching and Non-
Hydroxyphenyl)Benzothiazole. J. Am. Chem. Soc. 1980, 102, 5631− destructive Readout. J. Am. Chem. Soc. 2006, 128, 14542−14547.
(31) Tomin, V. I.; Demchenko, A. P.; Chou, P. T. Thermodynamic
5635.
vs. Kinetic Control of Excited-State Proton Transfer Reactions. J.
(12) Beens, H.; Grellman, Kh; Gurr, M.; Weller, A. H. Effect of
Photochem. Photobiol., C 2015, 22, 1−18.
Solvent and Temperature on Proton Transfer Reactions of Excited
(32) Liu, Z. Y.; Hu, J. W.; Huang, T. H.; Chen, K. Y.; Chou, P. T.
Molecules. Discuss. Faraday Soc. 1965, 39, 183−193.
Excited-State iIntramolecular Proton Transfer in the Kinetic-Control
(13) Mcmorrow, D.; Kasha, M. Intramolecular Excited-State Proton-
Regime. Phys. Chem. Chem. Phys. 2020, 22, 22271−22278.
Transfer in 3-Hydroxyflavone - Hydrogen-Bonding Solvent Perturba-
(33) Liu, Z. Y.; Hu, J. W.; Chen, C. L.; Chen, Y. A.; Chen, K. Y.;
tions. J. Phys. Chem. 1984, 88, 2235−2243. Chou, P. T. Correlation among Hydrogen Bond, Excited-State
(14) Williams, D. L.; Heller, A. Intramolecular Proton Transfer
Intramolecular Proton-Transfer Kinetics and Thermodynamics for
Reactions in Excited Fluorescent Compounds. J. Phys. Chem. 1970, -OH Type Proton-Donor Molecules. J. Phys. Chem. C 2018, 122,
74, 4473−4480. 21833−21840.
(15) Kim, J.; Kim, C. H.; Burger, C.; Park, M.; Kling, M. F.; Kim, D. (34) Tang, K. C.; Chang, M. J.; Lin, T. Y.; Pan, H. A.; Fang, T. C.;
E.; Joo, T. Non-Born-Oppenheimer Molecular Dynamics Observed Chen, K. Y.; Hung, W. Y.; Hsu, Y. H.; Chou, P. T. Fine Tuning the
by Coherent Nuclear Wave Packets. J. Phys. Chem. Lett. 2020, 11, Energetics of Excited-State Intramolecular Proton Transfer (ESIPT):
755−761. White Light Generation in A Single ESIPT System. J. Am. Chem. Soc.
(16) Kim, J.; Heo, W.; Joo, T. Excited State Intramolecular Proton 2011, 133, 17738−17745.
Transfer Dynamics of 1-Hydroxy-2-acetonaphthone. J. Phys. Chem. B (35) Yu, W. S.; Cheng, C. C.; Cheng, Y. M.; Wu, P. C.; Song, Y. H.;
2015, 119, 2620−2627. Chi, Y.; Chou, P. T. Excited-State Intramolecular Proton Transfer in
(17) Lee, J.; Kim, C. H.; Joo, T. Active Role of Proton in Excited Five-Membered Hydrogen-Bonding Systems: 2-Pyridyl Pyrazoles. J.
State Intramolecular Proton Transfer Reaction. J. Phys. Chem. A 2013, Am. Chem. Soc. 2003, 125, 10800−10801.
117, 1400−1405. (36) Tseng, H. W.; Liu, J. Q.; Chen, Y. A.; Chao, C. M.; Liu, K. M.;
(18) Hsieh, C. C.; Chou, P. T.; Shih, C. W.; Chuang, W. T.; Chung, Chen, C. L.; Lin, T. C.; Hung, C. H.; Chou, Y. L.; Lin, T. C.; Wang,
M. W.; Lee, J.; Joo, T. Comprehensive Studies on an Overall Proton T. L.; Chou, P. T. Harnessing Excited-State Intramolecular Proton-
Transfer Cycle of the ortho-Green Fluorescent Protein Chromophore. Transfer Reaction via a Series of Amino-Type Hydrogen-Bonding
J. Am. Chem. Soc. 2011, 133, 2932−2943. Molecules. J. Phys. Chem. Lett. 2015, 6, 1477−1486.
(19) Takeuchi, S.; Tahara, T. Coherent Nuclear Wavepacket (37) Lin, W. C.; Fang, S. K.; Hu, J. W.; Tsai, H. Y.; Chen, K. Y.
Motions in Ultrafast Excited-State Intramolecular Proton Transfer: Ratiometric Fluorescent/Colorimetric Cyanide-Selective Sensor
Sub-30-fs Resolved Pump-Probe Absorption Spectroscopy of 10- Based on Excited-State Intramolecular Charge Transfer-Excited-
Hydroxybenzo[h]quinoline in Solution. J. Phys. Chem. A 2005, 109, State Intramolecular Proton Transfer Switching. Anal. Chem. 2014,
10199−10207. 86, 4648−4652.
(20) Ishii, K.; Takeuchi, S.; Tahara, T. Infrared-Induced Coherent (38) Xu, Y. Q.; Pang, Y. Zn2+-Triggered Excited-State Intra-
Vibration of a Hydrogen-Bonded System: Effects of Mechanical and molecular Proton Transfer: a Sensitive Probe with Near-Infrared
Electrical Anharmonic Couplings. J. Chem. Phys. 2009, 131, 044512. Emission from Bis(benzoxazole) Derivative. Dalton Trans. 2011, 40,
(21) Tahara, T.; Takeuchi, S.; Ishii, K. Observation of Nuclear 1503−1509.
Wavepacket Motion of Reacting Excited States in Solution. J. Chin. (39) Yu, L.; Li, Y. Y.; Yu, H.; Zhang, K.; Wang, X. W.; Chen, X. F.;
Chem. Soc. 2006, 53, 181−189. Yue, J.; Huo, T. X.; Ge, H. W.; Alamry, K. A.; Marwani, H. M.; Wang,
(22) Yoneda, Y.; Sotome, H.; Mathew, R.; Lakshmanna, Y. A.; S. H. A Fluorescence Probe for Highly Selective and Sensitive
Miyasaka, H. Non-condon Effect on Ultrafast Excited-State Intra- Detection of Gaseous Ozone Based on Excited-State Intramolecular
molecular Proton Transfer. J. Phys. Chem. A 2020, 124, 265−271. Proton Transfer Mechanism. Sens. Actuators, B 2018, 266, 717−723.

12722 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

(40) Maity, D.; Kumar, V.; Govindaraju, T. Reactive Probes for (56) Sengupta, P. K.; Kasha, M. Excited-State Proton-Transfer
Ratiometric Detection of Co2+ and Cu+ Based on Excited-State Spectroscopy of 3-Hydroxyflavone and Quercetin. Chem. Phys. Lett.
Intramolecular Proton Transfer Mechanism. Org. Lett. 2012, 14, 1979, 68, 382−385.
6008−6011. (57) Chou, P. T.; Huang, C. H.; Pu, S. C.; Cheng, Y. M.; Liu, Y. H.;
(41) Sytnik, A.; Kasha, M. Excited-State Intramolecular Proton- Wang, Y.; Chen, C. T. Tuning Excited-State Charge/Proton Transfer
Transfer as a Fluorescence Probe for Protein Binding-Site Static Coupled Reaction via the Dipolar Functionality. J. Phys. Chem. A
Polarity. Proc. Natl. Acad. Sci. U. S. A. 1994, 91, 8627. 2004, 108, 6452−6454.
(42) Sytnik, A.; Gormin, D.; Kasha, M. Interplay between Excited- (58) Chou, P. T.; Martinez, M. L.; Clements, J. H. Reversal of
State Intramolecular Proton-Transfer and Charge-Transfer in Excitation Behavior of Proton-Transfer Vs Charge-Transfer by
Flavonols and Their Use as Protein-Binding-Site Fluorescence Probes. Dielectric Perturbation of Electronic Manifolds. J. Phys. Chem.
Proc. Natl. Acad. Sci. U. S. A. 1994, 91, 11968−11972. 1993, 97, 2618−2622.
(43) Fu, Y. H.; Chen, C. Y.; Chen, C. T. Tuning of Hydrogen (59) Costa, A. M. B. S. R. C. S.; Dean, F. M.; Jones, M. A.; Varma, R.
Peroxide-Responsive Polymeric Micelles of Biodegradable Triblock S. Lithiation in Flavones, Chromones, Coumarins, and Benzofuran
Polycarbonates as a Potential Drug Delivery Platform with Derivatives. J. Chem. Soc., Perkin Trans. 1 1985, 1, 799−808.
Ratiometric Fluorescence Signaling. Polym. Chem. 2015, 6, 8132− (60) Wang, X.; Liu, J.; Zhang, Y. An Efficient One-Pot Synthesis and
Anticancer Activity of 4’-Substituted Flavonoids. Russ. J. Gen. Chem.
8143.
2018, 88, 1036−1041.
(44) Hsu, Y. H.; Chen, Y. A.; Tseng, H. W.; Zhang, Z. Y.; Shen, J. Y.;
(61) Chimenti, F.; Fioravanti, R.; Bolasco, A.; Chimenti, P.; Secci,
Chuang, W. T.; Lin, T. C.; Lee, C. S.; Hung, W. Y.; Hong, B. C.; Liu,
D.; Rossi, F.; Yanez, M.; Orallo, F.; Ortuso, F.; Alcaro, S.; Cirilli, R.;
S. H.; Chou, P. T. Locked ortho- and para-Core Chromophores of
Ferretti, R.; Sanna, M. L. A New Series of Flavones, Thioflavones, and
Green Fluorescent Protein; Dramatic Emission Enhancement via Flavanones as Selective Monoamine Oxidase-B Inhibitors. Bioorg.
Structural Constraint. J. Am. Chem. Soc. 2014, 136, 11805−11812. Med. Chem. 2010, 18, 1273−1279.
(45) Chen, K. Y.; Hsieh, C. C.; Cheng, Y. M.; Lai, C. H.; Chou, P. T. (62) Angeli, A.; Tanini, D.; Nocentini, A.; Capperucci, A.; Ferraroni,
Extensive Spectral Tuning of the Proton Transfer Emission from 550 M.; Gratteri, P.; Supuran, C. T. Selenols: a New Class of Carbonic
to 675 nm via a Rational Derivatization of 10-Hydroxybenzo[h]- Anhydrase Inhibitors. Chem. Commun. 2019, 55, 648−651.
quinoline. Chem. Commun. 2006, 4395−4397. (63) Xue, H.; Jing, B.; Liu, S.; Chae, J.; Liu, Y. Copper-catalyzed
(46) Park, S.; Kwon, O. H.; Kim, S.; Park, S.; Choi, M. G.; Cha, M.; direct synthesis of aryl thiols from aryl iodides using sodium sulfide
Park, S. Y.; Jang, D. J. Imidazole-Based Excited-State Intramolecular aided by catalytic 1, 2-ethanedithiol. Synlett 2017, 28, 2272−2276.
Proton-Transfer Materials: Synthesis and Amplified Spontaneous (64) Kumar, R.; Shard, A.; Andhare, N. H.; Sinha, A. K.; et al.
Emission from a Large Single Crystal. J. Am. Chem. Soc. 2005, 127, Thiol−Ene “Click” Reaction Triggered by Neutral Ionic Liquid: The
10070−10074. “Ambiphilic” Character of [hmim] Br in the Regioselective
(47) Mamada, M.; Inada, K.; Komino, T.; Potscavage, W. J.; Nucleophilic Hydrothiolation. Angew. Chem. 2015, 127, 842−846.
Nakanotani, H.; Adachi, C. Highly Efficient Thermally Activated (65) Bandyopadhyay, S.; Dey, A. Convenient detection of the thiol
Delayed Fluorescence from an Excited-State Intramolecular Proton functional group using H/D isotope sensitive Raman spectroscopy.
Transfer System. ACS Cent. Sci. 2017, 3, 769−777. Analyst 2014, 139, 2118−2121.
(48) Park, S.; Seo, J.; Kim, S. H.; Park, S. Y. Tetraphenylimidazole- (66) Richter, W.; Schiel, D.; Wöger, W. Line Merging in the OH and
Based Excited-State Intramolecular Proton-Transfer Molecules for SH Stretching Vibrational Raman Profiles of Ethanol and Ethanethiol
Highly Efficient Blue Electroluminescence. Adv. Funct. Mater. 2008, in the Liquid State due to Conformational Dynamics. Mol. Phys. 1987,
18, 726−731. 60, 691−699.
(49) Nunes, C. M.; Pereira, N. A. M.; Reva, I.; Amado, P. S. M.; (67) Burneau, A.; Perchard, J. P.; Zuppiroli, G.; Limouzi, J.;
Cristiano, M. L. S.; Fausto, R. Bond-Breaking/Bond-Forming Maréchal, E. Effect of Electrical Anharmonicity on the Profiles of H-
Reactions by Vibrational Excitation: Infrared-Induced Bidirectional Bonded OH and SH Stretching Bands. Mol. Phys. 1980, 41, 1373−
Tautomerization of Matrix-Isolated Thiotropolone. J. Phys. Chem. 1386.
Lett. 2020, 11, 8034−8039. (68) Strandjord, A. J. G.; Barbara, P. F. Hydrogen-Deuterium
(50) Pietrzak, M.; Dobkowski, J.; Gorski, A.; Gawinkowski, S.; Kijak, Isotope Effects on the Excited-State Proton-Transfer Kinetics of 3-
M.; Luboradzki, R.; Hansen, P. E.; Waluk, J. Arresting Consecutive Hydroxyflavone. Chem. Phys. Lett. 1983, 98, 21−26.
Steps of a Photochromic Reaction: Studies of beta-Thioxoketones (69) Rohatgi-Mukherjee, K. Fundamentals of photochemistry; New
Combining Laser Photolysis with NMR Detection. Phys. Chem. Chem. Age International, 1978.
(70) Strickler, S. J.; Berg, R. A. Relationship between Absorption
Phys. 2014, 16, 9128−9137.
Intensity and Fluorescence Lifetime of Molecules. J. Chem. Phys.
(51) Gorski, A.; Posokhov, Y.; Hansen, B. K. V.; Spanget-Larsen, J.;
1962, 37, 814.
Jasny, J.; Duus, F.; Hansen, P. E.; Waluk, J. Photochromism in p-
(71) Protti, S.; Mezzetti, A.; Cornard, J. P.; Lapouge, C.; Fagnoni,
Methylbenzoylthioacetone and Related beta-Thioxoketones. Chem.
M. Hydrogen Bonding Properties of DMSO in Ground-State
Phys. 2007, 338, 11−22. Formation and Optical Spectra of 3-Hydroxyflavone Anion. Chem.
(52) Gorski, A.; Posokhov, Y.; Hansen, B. K. V.; Spanget-Larsen, J.; Phys. Lett. 2008, 467, 88−93.
Jasny, J.; Duus, F.; Hansen, P. E.; Waluk, J. Photochromism and (72) Parthenopoulos, D. A.; Kasha, M. Ground-State Anion
Polarization Spectroscopy of p-Methyl(thiobenzoyl) Acetone. Chem. Formation and Picosecond Excitation Dynamics of 3-Hydroxyflavone
Phys. 2006, 328, 205−215. in Formamide. Chem. Phys. Lett. 1990, 173, 303−309.
(53) Makhal, S. C.; Bhattacharyya, A.; Guchhait, N. Thiolactim- (73) Petersilka, M.; Gossmann, U. J.; Gross, E. K. U. Excitation
Thiolactam Photoisomerisation: Sulfur as Proton Donor for Excited energies from time-dependent density-functional theory. Phys. Rev.
State Proton Transfer Process. Chem. Phys. Lett. 2019, 717, 112−118. Lett. 1996, 76, 1212−1215.
(54) Liu, Z. Y.; Hu, J. W.; Huang, C. H.; Huang, T. H.; Chen, D. G.; (74) Stratmann, R. E.; Scuseria, G. E.; Frisch, M. J. An efficient
Ho, S. Y.; Chen, K. Y.; Li, E. Y.; Chou, P. T. Sulfur-Based implementation of time-dependent density-functional theory for the
Intramolecular Hydrogen-Bond: Excited-State Hydrogen-Bond On/ calculation of excitation energies of large molecules. J. Chem. Phys.
Off Switch with Dual Room-Temperature Phosphorescence. J. Am. 1998, 109, 8218−8224.
Chem. Soc. 2019, 141, 9885−9894. (75) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
(55) Mcmorrow, D.; Kasha, M. Proton-Transfer Spectroscopy of 3- Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Petersson,
Hydroxyflavone in an Isolated-Site Crystal Matrix. Proc. Natl. Acad. G. A.; Nakatsuji, H.et al. Gaussian 16, Revision A.03; Gaussian Inc.:
Sci. U. S. A. 1984, 81, 3375−3378. Wallingford, CT, 2016.

12723 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724
Journal of the American Chemical Society pubs.acs.org/JACS Article

(76) Becke, A. D. Density-Functional Thermochemistry 0.3. The


Role of Exact Exchange. J. Chem. Phys. 1993, 98, 5648−5652.
(77) Frisch, M. J.; Pople, J. A.; Binkley, J. S. Self-Consistent
Molecular-Orbital Methods 0.25. Supplementary Functions for
Gaussian-Basis Sets. J. Chem. Phys. 1984, 80, 3265−3269.
(78) Englman, R.; Jortner, J. The Energy Gap Law for Radiationless
Transitions in Large Molecules. Mol. Phys. 1970, 18, 145−164.
(79) Huang, C.-H.; Wu, P.-J.; Chung, K.-Y.; Chen, Y.-A.; Li, E. Y.;
Chou, P.-T. Room-Temperature Phosphorescence from Small
Organic Systems Containing a Thiocarbonyl Moiety. Phys. Chem.
Chem. Phys. 2017, 19, 8896−8901.

12724 https://doi.org/10.1021/jacs.1c05602
J. Am. Chem. Soc. 2021, 143, 12715−12724

You might also like