Вы находитесь на странице: 1из 12

Article

Cite This: J. Phys. Chem. A 2019, 123, 3937−3948 pubs.acs.org/JPCA

Dynamic Excited-State Intramolecular Proton Transfer Mechanisms


of Two Novel 3‑Hydroxyflavone-Based Chromophores in Two
Different Surroundings
Jiaojiao Hao and Yang Yang*
State Key Laboratory of Molecular Reaction Dynamics, Dalian Institute of Chemical Physics, Chinese Academy of Sciences,
Dalian 116023, China
*
S Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: The dynamic excited-state intramolecular proton transfer (ESIPT)


mechanisms of two novel 3-hydroxyflavone-based chromophores (1 and 2) in different
Downloaded via UNIV OF GOTHENBURG on October 23, 2019 at 16:32:19 (UTC).

surroundings (nonpolar cyclohexane and polar acetonitrile solvents), which are reported
in the recent work (Chou et al. J. Phys. Chem. A. 2010, 114, 10412), are explored in
terms of the density functional theory (DFT) and time-dependent DFT theoretical
methods. The computational absorption and emission spectra for the work rendered
here were in reasonable agreement with the relevant experiment. In order to present the
molecular-level exposition of the ESIPT reactions for these compounds in two different
solvents, we calculated the hydrogen bond (HB) parameters, corresponding infrared
vibrational frequencies, frontier molecular orbitals, and maps of electron density
difference between the S0 and S1 states, and the HB strengthening tendency in S1 states
was verified, giving the probability of ESIPT reactions. In addition, to definitely expose
the ESIPT mechanisms of compounds 1 and 2, we built the potential energy curves and
potential energy surfaces in the S0 and S1 states. Calculated results exhibited that the
ESIPT reaction of compound 1 in nonpolar cyclohexane solvent was more susceptible
than that in polar acetonitrile solvent. For the asymmetric compound 2, only single-
ESIPT processes could occur in both the solvents, and double-ESIPT processes were prohibitive due to high potential energy
barriers. Moreover, the single-ESIPT processes [I (6.26 kcal/mol) and II (6.62 kcal/mol)] in cyclohexane were more
susceptible than that [I′ (6.91 kcal/mol) and II′ (6.90 kcal/mol)] in acetonitrile. Furthermore, the single-ESIPT process I had a
little advantage over the process II in cyclohexane, while the probabilities of processes I′ and II′ were roughly the same in
acetonitrile.

1. INTRODUCTION deeply from the experimental and theoretical points of view in


order to create more valuable compounds.
The hydrogen bond (HB), as one of the remarkable weak
As a rule, the majority of the ESIPT processes contain single
interactions, is generally deemed to play a pivotal role in phys-
PT via the presence of an intramolecular HB between the
ical, chemical, and biological processes, such as water solution,1,2 proton donor (H atom from −OH or −NH2 group) and the
DNA double-strand,3,4 α-helix,5,6 pharmacophore,7,8 and so proton acceptor (O atom from −CO group or N atom
forth.9−13 In recent decades, as the excited-state has been from −CN group). In addition, the intramolecular HB is
understood in depth, it is confirmed that the HB can enable enhanced after light excitation, which facilitates the occurrence
excited-state intramolecular proton transfer (ESIPT), which of the ESIPT reactivity. In fact, such a single PT reaction is not
gives rise to the production of tautomer with the different elec- competent to simulate complicated multiple PT reactions
tronic construction from the initial excited form.14−20 Universally, in biochemical fields, particularly the double PT reaction.
the ESIPT reaction takes place along with the four-level photo- Accordingly, the double-ESIPT reaction exerts a tremendous
cycle, i.e., absorption → ESIPT → emission → ground-state fascination on researchers due to their important features
intramolecular PT (GSIPT). The great difference of charge in materials and biological fields. During the past few decades,
arrangement between the phototautomers in ESIPT process an army of typical and excellent organic compounds with
leads to the result that ESIPT fluorophores possess peculiar two intramolecular HBs (i.e., double proton) has been
photophysical properties, such as the prominent Stokes shift,21 designed and synthesized in experiment;35−39 meanwhile, the
dual fluorescence,22 fastest chemical reaction,23 and so on.24,25 double-ESIPT mechanisms widely have been explored by
Subsequently, these properties trigger crucial applications such
as molecular switches,26 light-emitting diodes,27 fluorescence Received: January 29, 2019
imaging,28 and the like29−34 in organic new materials and Revised: March 28, 2019
devices. Consequently, the ESIPT behaviors have been probed Published: March 29, 2019

© 2019 American Chemical Society 3937 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

combined experimental time-resolved spectroscopies and density gradient (RDG) isosurfaces for two compounds were
theoretical calculations.40−49 To date, the reaction pathways executed from the Multiwfn package.79 In addition, a series of
of the double-ESIPT processes can be outlined as follows: the DFT functional has been tested in this work and it was
stepwise and simultaneous reaction pathways.50−54 Nevertheless, verified that the B1B95 function procures the most satisfied
the previous research mainly has poured attention into the accordance with the experimental date (see Table S1 in
symmetric ESIPT compounds, but rarely has attempted to the Supporting Information).
asymmetric.
More recently, Chou et al. have synthesized two novel 3. RESULTS AND DISCUSSION
3-hydroxyflavone (3-HF)-based chromophores (shown in 3.1. 2-(4-(Diphenylamino)phenyl)-3-hydroxy-4H-
Figure 1 and 7), 2-(4-(diphenylamino)phenyl)-3-hydroxy- chromen-4-one (Compound 1). 3.1.1. Analysis of Geo-
4H-chromen-4-one (compound 1 involving one intramolecular metric Constructions. In this subsection, the configurations of
HB) and 2,2′-((phenylazanediyl)bis(4,1-phenylene))bis(3- the normal (1-N) and tautomer (1-T) forms for compound 1
hydroxy-4H-chromen-4-one) (asymmetric compound 2 in- in the S0 and S1 states have been optimized by means of the
volving two intramolecular HBs).55 They have determined DFT/TDDFT methods combined with the B1B95/TZVP
ESIPT properties for both the compounds in different solvent level in nonpolar cyclohexane and polar acetonitrile solvents,
environments by means of femtosecond fluorescence spectros- respectively (see Figure 1). The labeled numbers of the
copy. It is well known that the 3-HF and its derivatives have
been a subject of extensive investigation within the past few
decades.56−61 Until now, a multitude of studies has ascertained
the ESIPT mechanisms and dynamics of 3-HF chromophores,
such as that its normal form (N*) changes into the tautomer
form (T*) in the S1 state under light excitation, and it leads
to generate dual fluorescence in this process.62−64 Although
the compounds 1 and 2 have shown resemblance to the
excited-state behaviors of 3-HF and its derivatives with the
aid of spectral measurements, the Chou’s study regarding
these compounds is considered incomplete due to the two Figure 1. Optimized normal (1-N) and tautomer (1-T) forms of
PT sites of compound 2. The first question involves that this compound 1 by the B1B95/TZVP method in nonpolar cyclohexane
experimental work did not provide the adequate informa- and polar acetonitrile solvents.
tion concerning optical properties. The second problem
relates to the absence of the distinct and particular discussions intramolecular HB (O1−H2···O3) were needful in order
on the ESIPT reactions, especially on the more complex to facilitate the discussion about the PT behavior for this
ESIPT processes for the asymmetric compound 2. The third compound. The key structural parameters of 1-N and 1-T
aspect involves the shortage of the in-depth explorations of forms in the S0 and S1 states in two solvents are tabulated in
solvent polarity effect on the ESIPT processes for compounds the Table 1. As seen from Table 1, for the 1-N form, the BL of
1 and 2. Hence, the aim of this work is to shed light on the
detailed ESIPT mechanisms of these compounds and their Table 1. Major Calculated BLs (Å) and BAs (°) of Normal
changes in the different polar solvents in virtue of theoretical and Tautomer Forms for the Compound 1 in the S0 and S1
methods. States Used the DFT/TDDFT/B1B95-TZVP Methods in
Both Nonpolar Cyclohexane and Polar Acetonitrile
2. COMPUTATIONAL DETAILS
Solvents, Respectively
To achieve the useful data derived from the geometric con-
structions, the stable constructions of compounds 1 and 2 in cyclohexane acetonitrile
the ground (S0) and excited (S1) states were optimized using S0 S1 S0 S1
the density functional theory (DFT) and time dependent DFT 1-N O1−H2 0.975 0.986 0.975 0.984
(TDDFT) methods in conjunction with the B1B95 functional65,66 H2−O3 1.977 1.851 1.977 1.878
and TZVP basis set,67−74 as performed in Gaussian09 program.75 δ(O1−H2···O3) 119.5 124.7 119.5 123.6
Moreover, to comply with the previous experiment, the polar- 1-T O1−H2 1.787 1.903 1.815 1.935
izable continuum model (PCM) using the integral equation H2−O3 0.998 0.983 0.994 0.980
formalism variant76−78 with the nonpolar cyclohexane and δ(O1−H2···O3) 125.4 121.5 124.5 120.5
polar acetonitrile solvents was considered when theoretically
computing. The bond lengths (BLs) and angles (BAs) were
not limited in the optimization processes for these compounds. O1−H2 is enlarged from 0.975 (S0) to 0.986 Å (S1), and
Furthermore, the infrared (IR) vibrational frequencies of these the distance of H2···O3 is lessened from 1.977 (S0) to 1.851
compounds were simulated, and it could be confirmed that all Å (S1) in cyclohexane. Moreover, the BA δ (O1−H2···O3) is
optimized constructions were the most stable (no imaginary). increased from 119.5° (S0) to 124.7° (S1), which evidences
Additionally, the vertical excitation energies were calculated that the δ tended to 180° in the S1 state. They indicated
based on the S0 state-optimized geometries. Besides, to expound that the HB (O1−H2···O3) of compound 1 is enhanced in the
the dynamic ESIPT processes of these compounds in two S1 state in nonpolar cyclohexane solvent. A similar trend
different solvents, the potential energy curves (PECs) and appeared in polar acetonitrile solvent: the BL of O1−H2 is
potential energy surfaces (PESs) of the S0 and S1 states were changed from 0.975 (S0) to 0.984 Å (S1); the distance of
constructed, respectively. What is more, the maps of electron H2···O3 is changed from 1.977 (S0) to 1.878 Å (S1); the BA
density difference between the S0 and S1 states and the reduced δ(O1−H2−O3) is changed from 119.5° (S0) to 123.6° (S1).
3938 DOI: 10.1021/acs.jpca.9b00879
J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

Figure 2. RDG isosurfaces for the compound 1: the S0 (1a) and S1 (2a) states in nonpolar cyclohexane; the S0 (1b) and S1 (2b) states in polar
acetonitrile.

That is to say, the HB is strengthened in the S1 state in of around −0.03 au, wherefore the distinct intramolecular HB
both the solvents, which can promote the ESIPT process for of compound 1 is determined in two solvents, respectively.
compound 1. 3.1.2. IR Vibration Spectra, Electronic Spectra, and
As shown in Figure 2, the intramolecular HB (O1−H2···O3) Frontier Molecular Orbitals. First, the IR vibration spectra
between the proton donor (O1 atom) and proton acceptor of compound 1 in nonpolar cyclohexane and polar acetonitrile
(O3 atom) of compound 1 is pictured. It could be plotted solvents have been simulated and are illustrated in Figure 3.
via the RDG versus sign(λ2)ρ and lower panel low-gradient At present, the detection of IR vibrational spectral shift is the
(s = 0.5 au) isosurfaces, which has been executed by the effective way to confirm the changes of the S1 state HB.80−83
Multiwfn package.79 Figure 2(1a,2a,1b,2b) exhibited spikes Figure 3 shows that the stretching vibrational frequency
3939 DOI: 10.1021/acs.jpca.9b00879
J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

Figure 3. Calculated IR spectra of the compound 1 in (a) cyclo-


hexane and (b) acetonitrile solvents based on the B1B95/TZVP Figure 4. Theoretical absorption and fluorescence humps for the 1-N
theoretical level. and 1-T forms in (a) cyclohexane and (b) acetonitrile solvents using
the B1B95/TZVP method.

of O1−H2 bond in compound 1 is charged from 3611.1 cm−1


(S0) to 3437.7 cm−1 (S1) in cyclohexane, which exhibits a
large red-shift value of 173.4 cm−1. Analogously, a large red-
shift value of 166.5 cm−1 is produced in the excitation
process [from 3613.3 cm−1 (S0) to 3446.8 cm−1 (S1)] in
acetonitrile. Aforementioned results again manifested that
intramolecular HB of compound 1 is intensified in the S1
state in both the solvents, which can facilitate the ESIPT
reactivity.
Next, we set out to determine whether the theoretical method
chosen is trustworthy. Therefore, the electronic spectra of com-
pound 1 in two solvents were simulated and are plotted in
Figure 4. For nonpolar cyclohexane solvent, the simulative
absorption hump of compound 1 is situated at 420 nm, which
was rather comparable with the experiment (408 nm).55
Furthermore, the simulative double fluorescence humps of
compound 1 are situated at 474 and 578 nm, respectively,
which were agree very well with experiment (434 and 568 nm). Figure 5. MOs (HOMO and LUMO) of the 1-N form and its
In addition, the simulative absorption (422 nm) and double transition energy in nonpolar cyclohexane (left) and polar acetonitrile
fluorescence (499 and 620 nm) humps were in better agree- (right) solvents.
ment with experiment (400, 480, and 590 nm) in polar aceto-
nitrile solvent. It can be significant to conclude that the Additionally, it is noteworthy that the enhancements of
B1B95/TZVP computing method was equal for the probe of electron density on O3 atoms implied the increase of
compound 1. intramolecular HB in two solvents. For the reason that the
Finally, we explored the charge distribution from the S0 to S1 frontier orbital energy gap (Egap) relates to the high chemical
states based on the analysis of the molecular orbitals (MOs) reactivity and low kinetic stability,84 the Egap values of HOMO
(see Figure 5). As plotted in Figure 5, the MOs of the 1-N → LUMO for compound 1 in both the solvents were
form exhibited the obvious π → π* characteristic transition calculated and are displayed in Figure 5. It can be seen that the
from the highest occupied molecular orbital (HOMO) to the Egap values were 3.620 and 3.585 eV in nonpolar cyclohexane
lowest unoccupied molecular orbital (LUMO) in two solvents. and polar acetonitrile solvents, respectively, which clarified that
3940 DOI: 10.1021/acs.jpca.9b00879
J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

the compound 1 possesses the low kinetic stability and high


chemical reactivity in both the solvents.
3.1.3. Analysis of PECs. The main objective of this section
was to elucidate the dynamic ESIPT behaviors of compound 1
in two different polarity solvents. Accordingly, we scanned the
length of O1−H2 bond ranging from 0.975 to 1.975 with a step
size of 0.1 Å to construct the PECs on the S0 and S1 states,
respectively (see Figure 6). As can be noted in Figure 6(1,2),
the high potential barriers (PBs) are produced in the GSIPT
process in corresponding cyclohexane (11.52 kcal/mol) and
acetonitrile (12.74 kcal/mol) solvents, respectively, which testi-
fied that the PT process was difficult to occur in the S0 state in
both the solvents. However, the PBs that have been reduced
to 5.32 (in cyclohexane) and 7.27 kcal/mol (in acetonitrile)
verified that the PT process could happen in the S1 state. What
is more, the PB relationship with the (1) (5.32 kcal/mol) <
(2) (7.27 kcal/mol) in the S1 state denoted that the ESIPT
reaction was more likely to take place in nonpolar cyclohexane
solvent for compound 1.
3.2. 2,2′-((Phenylazanediyl)bis(4,1-phenylene))bis(3-
hydroxy-4H-chromen-4-one) (Compound 2). 3.2.1. Anal-
ysis of Geometric Constructions. On account of the structural
asymmetry of compound 2, we optimized out the four kinds of
stable configurations in the S0 and S1 states for the two solvents
using the B1B95/TZVP method, i.e. 2-N (normal form), 2-T1
(tautomer form: only H5 from O4 to O6), 2-T2 (tautomer
form: only H8 from O7 to O9), and 2-T3 (tautomer form: both
the H5 from O4 to O6 and H8 from O7 to O9), respectively (see
Figure 6. PECs of S0 and S1 states for compound 1 in cyclohexane (1) Figure 7). In addition, the important structural parameters of
and acetonitrile (2) solvents combined with the TDDFT/B1B95/TZVP these forms for compound 2 are recorded in Table 2. As shown
method. The inset presents the corresponding optimized constructions. in Table 2, for the nonpolar cyclohexane solvent, the BLs of
O4−H5 and O7−H8 for the 2-N form both are increased from
S0 (0.976 and 0.976 Å) to S1 (0.982 and 0.981 Å) states, and
the both separations of H5···O6 and H8···O9 are shortened
from S0 (1.961 and 1.956 Å) to S1 (1.892 and 1.911 Å)
accordingly. Moreover, both the BAs δ (O4−H5···O6 and O7−
H8···O9) in the S1 (122.8 and 121.7°) state were closer to 180°
than that in the S0 state (120.1 and 120.2°). It has been eluci-
dated that both the HBs (O4−H5···O6 and O7−H8···O9) of
compound 2 are heightened in the S1 state. The same verdict
could be gated for acetonitrile: the BLs of O4−H5 and O7−H8
from 0.975 and 0.975 Å (S0) changed to 0.979 and 0.980 Å
(S1); the separations of H5···O6 and H8···O9 from 1.977 and
1.970 Å (S0) changed to 1.927 and 1.923 Å (S1); the BAs δ
from 119.5 and 119.7° (S0) changed to 121.3 and 121.5° (S1).
To sum up, both the intramolecular HBs (O4−H5···O6 and
O7−H8···O9) are intensified in the S1 state in two solvents,
which can boost the ESIPT reactions for compound 2.
Furthermore, Figure 8 drew the RDG isosurfaces of com-
Figure 7. Optimized constructions of the 2-N, 2-T1, 2-T2, and 2-T3 pound 2 in the S0 and S1 states in two solvents, for ease of
forms by the B1B95/TZVP method in nonpolar cyclohexane and revelation of the interactions of two intramolecular HBs in the
polar acetonitrile solvents. real space. Evidently, the all spikes of around −0.03 au implied
Table 2. Primary Calculated BLs (Å) and BAs (°) of 2-N, 2-T1, 2-T2, and 2-T3 Constructions in the S0 and S1 States Based on
the DFT/TDDFT/B1B95-TZVP Methods in Cyclohexane and Acetonitrile Solvents, Respectively
2-N 2-T1 2-T2 2-T3
S0 S1 S0 S1 S0 S1 S0 S1
cyclohexane O4−H5 0.976 0.982 1.785 1.905 0.976 0.977 1.785 1.832
H5−O6 1.961 1.892 0.999 0.983 1.961 1.949 0.999 0.995
O7−H8 0.976 0.981 0.976 0.978 1.782 1.909 1.782 1.805
H8−O9 1.956 1.911 1.957 1.943 0.999 0.983 0.999 1.003

3941 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

Table 2. continued
2-N 2-T1 2-T2 2-T3
S0 S1 S0 S1 S0 S1 S0 S1
δ(O4−H5···O6) 120.1 122.8 125.4 121.5 120.1 120.3 125.4 125.2
δ(O7−H8···O9) 120.2 121.7 120.2 120.5 125.5 121.3 125.5 122.2
acetonitrile O4−H5 0.975 0.979 1.811 1.939 0.975 0.976 1.810 1.893
H5−O6 1.977 1.927 0.995 0.980 1.977 1.966 0.995 0.985
O7−H8 0.975 0.980 0.975 0.976 1.808 1.941 1.806 1.890
H8−O9 1.970 1.923 1.971 1.961 0.995 0.980 0.996 0.985
δ(O4−H5···O6) 119.5 121.3 124.5 120.3 119.5 119.7 124.6 121.7
δ(O7−H8···O9) 119.7 121.5 119.6 119.8 124.7 120.2 124.7 121.8

Figure 8. RDG isosurfaces for the compound 2: the S0 (1A) and S1 (2A) states in cyclohexane; the S0 (1B) and S1 (2B) states in acetonitrile.

3942 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

Figure 9. Calculated IR spectra of compound 2 in (a) cyclohexane Figure 10. Theoretical absorption and fluorescence humps for the
and (b) acetonitrile solvents based on the B1B95/TZVP theoretical compound 2 in (a) nonpolar cyclohexane and (b) polar acetonitrile
level. solvents using the B1B95/TZVP method.

that the intramolecular HB interactions existed in the S0 and S1


states in both the solvents for compound 2.
3.2.2. IR, Electronic Spectra and MOs. The IR vibration
spectra of compound 2 in two solvents are calculated and
shown in Figure 9. As Figure 9 displays, it is seen that the
two red-shift values of 89.1 (O4−H5) and 84.4 cm−1 (O7−H8)
occurred in process of S0 → S1 states in nonpolar cyclohexane
solvent, based on the variations of the stretching vibrational
frequencies of O4−H5 [3609.4 cm−1 (S0) → 3520.3 cm−1 (S1)]
and O7−H8 [3602.9 cm−1 (S0) → 3518.5 cm−1 (S1)]. In addi-
tion, it is also concluded that the two red-shift values of 83.0
(O4−H5) and 81.2 cm−1 (O7−H8) happened from S0 to S1
states in polar acetonitrile solvent combining with the changes
of the stretching vibrational frequencies of O4−H5 [3612.5 cm−1
(S0) → 3529.5 cm−1 (S1)] and O7−H8 [3605.7 cm−1 (S0) → Figure 11. MOs (HOMO and LUMO) of the 2-N form and its transi-
tion energy in cyclohexane (left) and acetonitrile (right) solvents.
3524.5 cm−1 (S1)]. Hence, we could determine that the two
intramolecular HBs (O4−H5···O6 and O7−H8···O9) are
enhanced in the S1 state in both the solvents, which can experiment (429 nm).55 The fluorescence hump of 2-N is
promote the ESIPT reactions for compound 2. situated at 512 nm, which generated a Stokes shift of 83 nm.
Figure 10 depicts the calculative electronic spectra and Besides, the simulative fluorescence humps of 2-T1, 2-T2, and
corresponding experimental values (shown in parentheses) for 2-T3 are situated at 628, 627, and 668 nm, respectively. It is
compound 2, in which Figure 10a,b indicates these in nonpolar worthy to note that emission spectra manifest the visible
cyclohexane and polar acetonitrile solvents, respectively. double fluorescent phenomena for compound 2 in both the
For the cyclohexane solvent, the absorption hump of 2-N is solvents. However, there were no complete data of emission
situated at 447 nm, which was rather comparable with the spectra, which stemmed from the experimental report.55 There
experimental value (415 nm).55 The fluorescence hump of 2-N were two reasons for pursuing this: one was that the difference
is situated at 484 nm, which demonstrated a Stokes shift of between fluorescence humps of 2-T1 and 2-T2 in acetonitrile
37 nm. Additionally, the calculated fluorescence humps of 590, solvent was intensely small; second was that the 2-T3 cannot
587, and 717 nm are attributed to 2-T1, 2-T2, and 2-T3 forms, be formed due to high PEBs in two solvents. As a matter of
respectively. For the acetonitrile solvent, the value of absorp- fact, the reality may be more complicated, and we will put
tion hump of 2-N is situated at 450 nm and was close to forward the in-depth exploration in Section 3.2.3.
3943 DOI: 10.1021/acs.jpca.9b00879
J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

Table 3. Electronic Excitation Energy (nm), the


Corresponding Oscillator Strengths, and the Corresponding
Compositions of the Low-Lying Singlet Excited States of the
Compound 2
transition λ (nm) f composition CI (%)
cyclohexane S0 → S1 447.8 1.1722 H →L 96.5
S0 → S2 398.7 0.3417 H →L+1 97.0
S0 → S3 330.7 0.1123 H −1→L 90.4
S0 → S4 323.4 0.0178 H →L+4 59.4
acetonitrile S0 → S1 450.6 1.1632 H →L 96.7
S0 → S2 399.7 0.3237 H →L+1 97.4
S0 → S3 329.8 0.1169 H −1→L 90.2
S0 → S4 321.9 0.0177 H →L+4 58.8

Figure 13. PESs of (a) S1 and (b) S0 states for compound 2 in


nonpolar cyclohexane solvent.

diminution. Figure 12 presents the perfectly obvious fact that


the electron density on O6 and O9 was increased after photo-
excitation, while the electron density on O4 and O7 atoms was
decreased. Meanwhile, the Egap values of HOMO → LUMO
for compound 2 in cyclohexane (4.345 eV) and acetonitrile
(4.354 eV) solvents are shown in Figure 11. It is obvious that
compound 2 also has the low kinetic stability and high
chemical reactivity in both the solvents.
Figure 12. Electron density difference map between the S0 and S1 3.2.3. Analysis of PESs. As the above discussion shows, we
states of compound 2 at the TDDFT/B1B95/TZVP level in (a) nonpolar could forecast that the ESIPT processes can take place for
cyclohexane and (b) polar acetonitrile solvents. The orange zones indicate compound 2 in two solvents. Here, to better shed light on the
electron density loss, whereas the green zones correspond to a gain in ESIPT mechanisms of compound 2 in two solvents, the PESs
density following photon absorption.
as functions of O4−H5 and O7−H8 bonds were structured
by feat of the restricted optimization method in Figures 13
To illustrate the charge distributions for compound 2 after and 14. In addition, we tabulated the relevant PEBs for ESIPT
optical-excitation, the MOs and Egap in two solvents are processes in Tables S2 and S3. It is worth noting that scanning
described in Figure 11, and the calculated electronic transition the PESs along the O4−H5 and O7−H8 BLs can offer tentative
energies, corresponding oscillator strengths (f), and composi- insights into the ESIPT mechanism, but the reaction barrier
tions are listed in Table 3. The HOMO and LUMO of 2 does not necessarily coincide with the “true” barrier between
compound were observed in two solvents (see Figure 11), the reactant and the saddle point for the large molecules with
implying that their charge-transfer transitions were π → π* by much geometrical freedom. As can be noted in Figure 13b and
nature and accompanying the diminution of electron density Table S2, the PEBs illustrated that both single and double PT
on O4 and O7 atoms and increase of electron density on O6 reactions for compound 2 were impracticable in the S0 state on
and O9 atoms. Here, to reveal this phenomenon more cyclohexane. For acetonitrile, the PT reactions likewise were
intuitively, the maps of the electron density difference between infeasible in the S0 state (see Figure 14b and Table S3).
the S0 and S1 states for compound 2 are plotted in Figure 12, As Figure 13a displays, the four stable constructions with
in which the green range shows the increase of the electron O4−H5/O7−H8 lengths are denoted as A (0.982/0.981 Å)
density after excitation, and the orange range shows the (2-N), B (1.905/0.978 Å) (2-T1), C (0.977/1.909 Å) (2-T2),
3944 DOI: 10.1021/acs.jpca.9b00879
J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

were higher than 11 kcal/mol. In addition, the single-ESIPT


processes (I and II) were competing relationships, and process
I had a little advantage over the II. In addition, reverse PT
(RPT) reactions for nonpolar cyclohexane solvent in the S1
state could happen due to the low PBEs among these four stable
constructions of less than 6.10 kcal/mol (see in Table S2).
As seen in Figure 14a, the four minimum points also
appeared on the S1 state PES in polar acetonitrile solvent: A′
(0.979/0.980 Å), B′ (1.939/0.976 Å), C′ (0.976/1.941 Å), and
D′ (1.893/1.890 Å), and the energy-variant sequence was EA′
≈ EB′ ≈ EC′ < ED′. Therefore, the possible kind of PT paths in
S1 state was the same as that in cyclohexane: four single-ESIPT
processes [I′ (A′ → B′), II′ (A′ → C′), III′ (B′ → D′), and IV′
(C′ → D′)] and three double-ESIPT processes [V′ (A′ →
B′ → D′), VI′ (A′ → C′ → D′), and VII′ (A′ → D′)]. Based
on the calculated results from Table S3, it is noteworthy that
only the single-ESIPT processes I′ (6.91 kcal/mol) and II′
(6.90 kcal/mol) could occur in acetonitrile, and their proba-
bility were roughly the same. Moreover, the double-ESIPT
processes (V′, VI′, and VII′) are embargoed in the S1 state
due to these high PBEs. The RPT processes for acetonitrile
solvent in the S1 state could happen due to PBEs of less than
9.08 kcal/mol (seen in Table S3).
Ultimately, it can be significant to conclude that only single-
ESIPT processes could occur for compound 2, and double-
ESIPT processes are prohibited due to higher PBEs, whether
in nonpolar cyclohexane or polar acetonitrile solvents. Further-
more, the single-ESIPT processes (I and II) in nonpolar
cyclohexane were more sensitive than that (I′ and II′) in polar
acetonitrile. Besides, the process I had a little advantage over
the process II in cyclohexane, while the probabilities of processes
I′ and II′ were roughly the same in acetonitrile.
Figure 14. PESs of (a) S1 and (b) S0 states for compound 2 in polar
acetonitrile solvent.
4. CONCLUSIONS
and D (1.832/1.805 Å) (2-T3), respectively. What is more, the In summary, the ESIPT mechanisms of compounds 1 and 2 in
relative energies of these stable constructions in cyclohexane nonpolar cyclohexane and polar acetonitrile solvents were given a
solvent are demonstrated in Table 4. Therefore, the conceiv- molecular-level exposition based on theoretical investigation.
able four single-ESIPT processes [shown as I (A → B), II Upon the photoexcitation, the intramolecular HBs were inten-
(A → C), III (B → D), and IV (C → D)] and three double- sified in the S1 state for both compounds 1 and 2, which
ESIPT processes [shown as V (A → B → D), VI (A → C → D), presents the feasibility of ESIPT reactions in two solvents.
and VII (A → D)] were presented due to asymmetric config- Additionally, the charge redistribution further exposed the
uration of compound 2. As seen in Table 4, the relationship of ESIPT trend for these compounds. It is worthy to note that
PE between four stable constructions was EA ≈ EB ≈ EC < ED, the ESIPT reaction of compound 1 in the nonpolar cyclohexane
which exhibits that the ESIPT reactions may occur for the solvent was more susceptible than that in polar acetonitrile
nonpolar cyclohexane solvent. It is clear that the single-ESIPT solvent. For asymmetric compound 2, it is concluded that only
processes I (A → B) and II (A → C) are allowed in the S1 state single-ESIPT processes [I (6.26 kcal/mol), I′ (6.62 kcal/mol)],
due to low PEBs (6.26 and 6.62 kcal/mol) (seen in Table S2) [II (6.91 kcal/mol), and II′ (6.90 kcal/mol)] could occur in
for cyclohexane solvent. In contrast, the high PEBs produced both the solvents due to low PEBs, and double-ESIPT pro-
the following paths to be banned: single-ESIPT processes III cesses were prohibitive due to high PBEs. Besides, the single-
(B → D) (11.68 kcal/mol) and IV (C → D) (12.02 kcal/mol) ESIPT processes (I and II) in nonpolar cyclohexane were more
and double-ESIPT processes V (A → B → D) (6.26 and susceptible than those (I′ and II′) in polar acetonitrile. In addi-
11.68 kcal/mol), VI (A → C → D) (6.62 and 12.02 kcal/mol), tion, the process I had a little advantage over the process II in
and VII (A → D) (16.45 kcal/mol). Thus, it is concluded that cyclohexane, while the possibilities of processes I′ and II′ were
only single-ESIPT processes (I and II) could take place in roughly the same in acetonitrile. Here, our sincere wish is that
cyclohexane for compound 2, and double-ESIPT processes (V, this study not only illustrates the detailed ESIPT mechanisms
VI, and VII) were forbidden in the S1 state because these PBEs of compounds 1 and 2 in two different solvents but also maybe

Table 4. Potential Energy (Hartree) of the Stable Constructions in the S0 and S1 States
2-N 2-T1 2-T2 2-T3
S0 S1 S0 S1 S0 S1 S0 S1
cyclohexane 0.00 61.41 11.52 61.93 11.50 62.18 23.02 73.62
Acetonitrile −6.40 52.09 4.44 52.39 4.40 52.42 15.25 60.04

3945 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

boosts the relevant utilization and development about these Poisons with Promising Antiproliferative Activity. J. Med. Chem. 2018,
chromophores in the future. 61, 1375−1379.


(9) Cramer, C. J.; Truhlar, D. G. A Universal Approach to Solvation
ASSOCIATED CONTENT Modeling. Acc. Chem. Res. 2008, 41, 760−768.
(10) Gao, D.; Zegkinoglou, I.; Divins, N. J.; Scholten, F.; Sinev, I.;
*
S Supporting Information
Grosse, P.; Roldan Cuenya, B. Plasma-Activated Copper Nanocube
The Supporting Information is available free of charge on the Catalysts for Efficient Carbon Dioxide Electroreduction to Hydro-
ACS Publications website at DOI: 10.1021/acs.jpca.9b00879. carbons and Alcohols. ACS Nano 2017, 11, 4825−4831.
Comparing theoretical absorption data for compound 1 (11) Mesele, O. O.; Thompson, W. H. A “Universal” Spectroscopic
Map for the OH Stretching Mode in Alcohols. J. Phys. Chem. A 2017,
in cyclohexane solvent based on different functionals
121, 5823−5833.
(B3LYP, CAM-B3LYP, M06, B1B95, PBEPBE, (12) Calixto, A. R.; Ramos, M. J.; Fernandes, P. A. Influence of
WB97XD); calculated potential energy barriers (kcal/mol) Frozen Residues on the Exploration of the PES of Enzyme Reaction
among the four stable points on both S0 and S1 state Mechanisms. J. Chem. Theory Comput. 2017, 13, 5486−5495.
PESs for compound 2 in nonpolar cyclohexane solvent; (13) Zhang, Q.; Chen, S.; Wang, H.; Yu, H. Exquisite Enzyme-
calculated potential energy barriers (kcal/mol) among Fenton Biomimetic Catalysts for Hydroxyl Radical Production by
the four stable points on both S0 and S1 state PESs for Mimicking an Enzyme Cascade. ACS Appl. Mater. Interfaces 2018, 10,
compound 2 in polar acetonitrile solvent; Cartesian 8666−8675.
coordinates of constructions (1-N and 1-T) in S0 and S1 (14) Zhao, G.-J.; Han, K.-L. Role of Intramolecular and
states in nonpolar cyclohexane and polar acetonitrile Intermolecular Hydrogen Bonding in Both Singlet and Triplet
solvents; Cartesian coordinates of constructions (2-N, Excited States of Aminofluorenones on Internal Conversion,
2-T1, 2-T2, and 2-T3) in S0 and S1 states in nonpolar Intersystem Crossing, and Twisted Intramolecular Charge Transfer.
cyclohexane and polar acetonitrile solvents (PDF) J. Phys. Chem. A 2009, 113, 14329−14335.


(15) Houari, Y.; Chibani, S.; Jacquemin, D.; Laurent, A. D. TD-DFT
Assessment of the Excited State Intramolecular Proton Transfer in
AUTHOR INFORMATION Hydroxyphenylbenzimidazole (HBI) Dyes. J. Phys. Chem. B 2015,
Corresponding Author 119, 2180−2192.
*E-mail: light_yang@dicp.ac.cn. Phone: +86-411-84379037. (16) Padalkar, V. S.; Seki, S. Excited-state intramolecular proton-
transfer (ESIPT)-inspired solid state emitters. Chem. Soc. Rev. 2016,
Fax: +86-411-84675584.
45, 169−202.
ORCID (17) Ma, Y.; Yang, Y.; Lan, R.; Li, Y. Effect of Different Substituted
Yang Yang: 0000-0002-3387-2417 Groups on Excited-State Intramolecular Proton Transfer of 1-
Notes (Acylamino)-anthraquinons. J. Phys. Chem. C 2017, 121, 14779−
14786.
The authors declare no competing financial interest.


(18) Li, Y.; Ma, Y.; Yang, Y.; Shi, W.; Lan, R.; Guo, Q. Effects of
different substituents of methyl 5-R-salicylates on the excited state
ACKNOWLEDGMENTS intramolecular proton transfer process. Phys. Chem. Chem. Phys. 2018,
We sincerely thank the financial support from the National 20, 4208−4215.
Natural Science Foundation of China (grants no. 21403226 (19) Hao, J.; Yang, Y. Y. Y. The effects of different heterocycles and
and 21503226). solvents on the ESIPT mechanisms for three novel photoactive mono-


formylated benzoxazole derivatives. Org. Chem. Front. 2018, 5, 2234−
REFERENCES 2243.
(20) Yang, Y.; Chen, Y. P.; Zhao, Y.; Shi, W.; Ma, F. C.; Li, Y. Q.
(1) Ye, Q.; Li, Y.; Zhuo, L.; Zhang, W.; Xiong, W.; Wang, C.; Wang, Under different solvents excited-state intramolecular proton transfer
P. Optimal allocation of physical water resources integrated with mechanism and solvatochromic effect of 2-(2-hydroxyphenyl)
virtual water trade in water scarce regions: A case study for Beijing,
benzothiazole molecule. J. Lumin. 2019, 206, 326−334.
China. Water Res. 2018, 129, 264−276.
(21) Arribat, M.; Rémond, E.; Clément, S.; Lee, A. V. D.; Cavelier,
(2) Debenedetti, P. G.; Klein, M. L. Chemical physics of water. Proc.
F. Phospholyl(borane) Amino Acids and Peptides: Stereoselective
Natl. Acad. Sci. U.S.A. 2017, 114, 13325−13326.
Synthesis and Fluorescent Properties with Large Stokes Shift. J. Am.
(3) Kurokawa, C.; Fujiwara, K.; Morita, M.; Kawamata, I.;
Kawagishi, Y.; Sakai, A.; Murayama, Y.; Nomura, S.-i. M.; Murata, Chem. Soc. 2018, 140, 1028−1034.
S.; Takinoue, M.; et al. DNA cytoskeleton for stabilizing artificial cells. (22) Bhuiya, S.; Haque, L.; Goswami, R.; Das, S. Multispectroscopic
Proc. Natl. Acad. Sci. U.S.A. 2017, 114, 7228−7233. and Theoretical Exploration of the Comparative Binding Aspects of
(4) Pawelczak, K. S.; Gavande, N. S.; VanderVere-Carozza, P. S.; Bioflavonoid Fisetin with Triple- and DoubleHelical Forms of RNA. J.
Turchi, J. J. Modulating DNA Repair Pathways to Improve Precision Phys. Chem. B 2017, 121, 11037−11052.
Genome Engineering. ACS Chem. Biol. 2018, 13, 389−396. (23) Driscoll, E.; Sorenson, S.; Dawlaty, J. M. Ultrafast Intra-
(5) Lin, M. M.; Shorokhov, D.; Zewail, A. H. Dominance of molecular Electron and Proton Transfer in Bis(imino)isoindole
misfolded intermediates in the dynamics of α-helix folding. Proc. Natl. Derivatives. J. Phys. Chem. A 2015, 119, 5618−5625.
Acad. Sci. U.S.A. 2014, 111, 14424−14429. (24) Zhao, J.; Liu, X.; Zheng, Y. Controlling Excited State Single
(6) Kumar, S.; Hamilton, A. D. α-Helix Mimetics as Modulators of versus Double Proton Transfer for 2,2′-Bipyridyl-3,3′-diol: Solvent
Aβ Self-Assembly. J. Am. Chem. Soc. 2017, 139, 5744−5755. Effect. J. Phys. Chem. A 2017, 121, 4002−4008.
(7) Prota, A. E.; Bargsten, K.; Diaz, J. F.; Marsh, M.; Cuevas, C.; (25) Peng, Y.; Ye, Y.; Xiu, X.; Sun, S. Mechanism of Excited-State
Liniger, M.; Neuhaus, C.; Andreu, J. M.; Altmann, K.-H.; Steinmetz, Intramolecular Proton Transfer for 1,2-Dihydroxyanthraquinone:
M. O. A new tubulin-binding site and pharmacophore for micro- Effect of Water on the ESIPT. J. Phys. Chem. A 2017, 121, 5625−
tubule-destabilizing anticancer drugs. Proc. Natl. Acad. Sci. U.S.A. 5634.
2014, 111, 13817−13821. (26) Stricker, L.; Fritz, E.-C.; Peterlechner, M.; Doltsinis, N. L.;
(8) Ortega, J. A.; Riccardi, L.; Minniti, E.; Borgogno, M.; Arencibia, Ravoo, B. J. Arylazopyrazoles as Light-Responsive Molecular Switches
J. M.; Greco, M. L.; Minarini, A.; Sissi, C.; De Vivo, M. in Cyclodextrin-Based Supramolecular Systems. J. Am. Chem. Soc.
Pharmacophore Hybridization To Discover Novel Topoisomerase II 2016, 138, 4547−4554.

3946 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

(27) Shendre, S.; Sharma, V. K.; Dang, C.; Demir, H. V. Exciton (45) Chen, Y.-T.; Wu, P.-J.; Peng, C.-Y.; Shen, J.-Y.; Tsai, C.-C.; Hu,
Dynamics in Colloidal Quantum-Dot LEDs under Active Device W.-P.; Chou, P.-T. A study of the competitive multiple hydrogen
Operations. ACS Photonics 2018, 5, 480−486. bonding effect and its associated excited-state proton transfer
(28) Yin, C.; Zhen, X.; Zhao, H.; Tang, Y.; Ji, Y.; Lyu, Y.; Fan, Q.; tautomerism. Phys. Chem. Chem. Phys. 2017, 19, 28641−28646.
Huang, W.; Pu, K. Amphiphilic Semiconducting Oligomer for Near- (46) Liu, X.; Zhao, J.; Zheng, Y. Insight into the excited-state double
Infrared Photoacoustic and Fluorescence Imaging. ACS Appl. Mater. proton transfer mechanisms of doxorubicin in acetonitrile solvent.
Interfaces 2017, 9, 12332−12339. RSC Adv. 2017, 7, 51318−51323.
(29) Catalan, J.; del Valle, J. C. Toward the photostability (47) Zhao, J.; Dong, H.; Zheng, Y. Theoretical Insights Into the
mechanism of intramolecular hydrogen bond systems. The photo- Excited State Double Proton Transfer Mechanism of Deep Red
physics of 1′-hydroxy-2′-acetonaphthone. J. Am. Chem. Soc. 1993, Pigment Alkannin. J. Phys. Chem. A 2018, 122, 1200−1208.
115, 4321−4325. (48) Hao, J.; Yang, Y. Insight into the new excited-state
(30) Frey, P.; Whitt, S.; Tobin, J. A low-barrier hydrogen bond in intramolecular proton transfer (ESIPT) mechanism of N,N0-bis-
the catalytic triad of serine proteases. Science 1994, 264, 1927−1930. (salicylidene)-p-phenylenediamine (p-BSP). Chem. Phys. 2018, 501,
(31) Cleland, W.; Kreevoy, M. Low-barrier hydrogen bonds and 53−59.
enzymic catalysis. Science 1994, 264, 1887−1890. (49) Hao, J. J.; Yang, Y. The theoretical study about the ESIPT
(32) Acuña, A. U.; Amat, F.; Catalán, J.; Costela, A.; Figuera, J.; mechanism for 2,4-bis(benzooxazol-2′-yl)hydroquinone: Single or
Munoz, J. M. Pulsed Liquid Lasers from Proton Transfer in the double? J. Phys. Org. Chem. 2018, 32, No. e3903.
Excited State. Chem. Phys. Lett. 1986, 132, 567−569. (50) Zhou, Q.; Du, C.; Yang, L.; Zhao, M.; Dai, Y.; Song, P.
(33) Sakai, K.-i.; Tsuzuki, T.; Itoh, Y.; Ichikawa, M.; Taniguchi, Y. Mechanism for the Excited-State Multiple Proton Transfer Process of
Using Proton-transfer Laser Dyes for Organic Laser Diodes. Appl. Dihydroxyanthraquinone Chromophores. J. Phys. Chem. A 2017, 121,
Phys. Lett. 2005, 86, 081103−081105. 4645−4651.
(34) Kraus, G. A.; Zhang, W.; Fehr, M. J.; Petrich, J. W.; (51) Zhao, J.; Yao, H.; Liu, J.; Hoffmann, M. R. New Excited-State
Wannemuehler, Y.; Carpenter, S. Research at the Interface between Proton Transfer Mechanisms for 1,8-Dihydroxydibenzo[a,h]-
Chemistry and Virology:Development of a Molecular Flashlight. phenazine. J. Phys. Chem. A 2015, 119, 681−688.
Chem. Rev. 1996, 96, 523−536. (52) Yang, D.; Yang, G.; Zhao, J.; Zheng, R.; Wang, Y. A competitive
(35) Randino, C.; Ziółek, M.; Gelabert, R.; Organero, J. A.; Gil, M.; excited state dynamical process for the 2,2′-((1E,1′E)-((3,3′-
Moreno, M.; Lluch, J. M.; Douhal, A. Photo-deactivation pathways of dimethyl-[1,1′-biphenyl]-4,4′-diyl)-bis(azanylylidene))bis-
a double H-bonded photochromic Schiff base investigated by (methanylylidene))-diphenol system. RSC Adv. 2017, 7, 1299−1304.
combined theoretical calculations and experimental time-resolved (53) Yang, D.; Zhao, J.; Yang, G.; Song, N.; Zheng, R.; Wang, Y.
studies. Phys. Chem. Chem. Phys. 2011, 13, 14960−14972. Elaborating the excited-state proton transfer behaviors for novel 3H-
(36) Moroz, V. V.; Chalyi, A. G.; Serdiuk, I. E.; Roshal, A. D.;
MC and P2H-CH. Org. Chem. Front. 2017, 4, 1935−1942.
Zadykowicz, B.; Pivovarenko, V. G.; Wróblewska, A.; Błażejowski, J. (54) Hao, J.; Yang, Y. Exploring the ESIPT dynamical processes of
Tautomerism and Behavior of 3-Hydroxy-2-phenyl-4H-chromen-4-
two novel chromophores: symmetrical structure CHC and asym-
ones (Flavonols) and 3,7-Dihydroxy-2,8-diphenyl-4H,6H-pyrano[3,2-
metric structure CHN. Org. Chem. Front. 2018, 5, 1330−1341.
g]chromene-4,6-diones (Diflavonols) in Basic Media: Spectroscopic
(55) Lin, C.-C.; Chen, C.-L.; Chung, M.-W.; Chen, Y.-J.; Chou, P.-
and Computational Investigations. J. Phys. Chem. A 2013, 117, 9156−
T. Effects of Multibranching on 3-Hydroxyflavone-Based Chromo-
9167.
phores and the Excited-State Intramolecular Proton Transfer
(37) Xiao, H.; Chen, K.; Cui, D.; Jiang, N.; Yin, G.; Wang, J.; Wang,
Dynamics. J. Phys. Chem. A 2010, 114, 10412−10420.
R. Two novel aggregation-induced emission active coumarin-based
(56) Dick, B. AM1 and INDO/S calculations on electronic singlet
Schiff bases and their applications in cell imaging. New J. Chem. 2014,
and triplet states involved in excited-state intramolecular proton
38, 2386−2393.
(38) Jagadesan, P.; Whittemore, T.; Beirl, T.; Turro, C.; McGrier, P. transfer of 3-hydroxyflavone. J. Phys. Chem. 1990, 94, 5752−5756.
L. Excited-State Intramolecular Proton-Transfer Properties of Three (57) Wang, M.; Spataru, T.; Lombardi, J. R.; Birke, R. L. Time
Tris(N-Salicylideneaniline)-Based Chromophores with Extended Resolved Surface Enhanced Raman Scattering Studies of 3-
Conjugation. Chem.Eur. J. 2017, 23, 917−925. Hydroxyflavone on a Ag Electrode. J. Phys. Chem. C 2007, 111,
(39) Che, M.; Gao, Y.-J.; Zhang, Y.; Xia, S.-H.; Cui, G. Electronic 3044−3052.
structure calculations and nonadiabatic dynamics simulations of (58) Postupalenko, V. Y.; Zamotaiev, O. M.; Shvadchak, V. V.;
excited-state relaxation of Pigment Yellow 101. Phys. Chem. Chem. Strizhak, A. V.; Pivovarenko, V. G.; Klymchenko, A. S.; Mely, Y. Dual-
Phys. 2018, 20, 6524−6532. Fluorescence L-Amino Acid Reports Insertion and Orientation of
(40) Mohammed, O. F.; Xiao, D.; Batista, V. S.; Nibbering, E. T. J. Melittin Peptide in Cell Membranes. Bioconjugate Chem. 2013, 24,
Excited-State Intramolecular Hydrogen Transfer (ESIHT) of 1,8- 1998−2007.
Dihydroxy-9,10-anthraquinone (DHAQ) Characterized by Ultrafast (59) Sholokh, M.; Zamotaiev, O. M.; Das, R.; Postupalenko, V. Y.;
Electronic and Vibrational Spectroscopy and Computational Model- Richert, L.; Dujardin, D.; Zaporozhets, O. A.; Pivovarenko, V. G.;
ing. J. Phys. Chem. A 2014, 118, 3090−3099. Klymchenko, A. S.; Mély, Y. Fluorescent Amino Acid Undergoing
(41) Ma, C.; Yang, Y.; Li, C.; Liu, Y. TD-DFT Study of the Double Excited State Intramolecular Proton Transfer for Site-Specific Probing
Excited-State Intramolecular Proton Transfer Mechanism of 1,3- and Imaging of Peptide Interactions. J. Phys. Chem. B 2015, 119,
Bis(2-pyridylimino)-4,7-dihydroxyisoindole. J. Phys. Chem. A 2015, 2585−2595.
119, 12686−12692. (60) Tang, Z.; Wang, Y.; Bao, D.; Lv, M.; Yang, Y.; Tian, J.; Dong, L.
(42) Serdiuk, I. E.; Roshal, A. D. Single and double intramolecular Theoretical Investigation of an Excited-State Intramolecular Proton-
proton transfers in the electronically excited state of flavone Transfer Mechanism for an Asymmetric Structure of 3,7-Dihydroxy-4-
derivatives. RSC Adv. 2015, 5, 102191−102203. oxo-2-phenyl-4H-chromene-8-carbaldehyde: Single or Double? J.
(43) Zhao, J.; Chen, J.; Liu, J.; Hoffmann, M. R. Competitive Phys. Chem. A 2017, 121, 8807−8814.
excited-state single or double proton transfer mechanisms for bis-2,5- (61) Dávila, Y. A.; Sancho, M. I.; Almandoz, M. C.; Gasull, E.
(2-benzoxazolyl)-hydroquinone and its derivatives. Phys. Chem. Chem. Spectroscopic and Electronic Analysis of Chelation Reactions of
Phys. 2015, 17, 11990−11999. Galangin and Related Flavonoids with Nickel(II). J. Chem. Eng. Data
(44) Benelhadj, K.; Massue, J.; Ulrich, G. 2,4 and 2,5-bis- 2018, 63, 1488−1497.
(benzooxazol-2′-yl)hydroquinone (DHBO) and their borate com- (62) Sengupta, P. K.; Kasha, M. Excited state proton-transfer
plexes: synthesis and optical properties. New J. Chem. 2016, 40, spectroscopy of 3-hydroxyflavone and quercetin. Chem. Phys. Lett.
5877−5884. 1979, 68, 382−385.

3947 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948
The Journal of Physical Chemistry A Article

(63) Itoh, M.; Tokumura, K.; Tanimoto, Y.; Okada, Y.; Takeuchi, cence Quenching Mechanism. J. Phys. Chem. B 2007, 111, 8940−
H.; Obi, K.; Tanaka, I. Time-resolved and steady-state fluorescence 8945.
studies of the excited-state proton transfer in 3-hydroxyflavone and 3- (83) Yu, F.; Li, P.; Li, G.; Zhao, G.; Chu, T.; Han, K. A Near-IR
hydroxychromone. J. Am. Chem. Soc. 1982, 104, 4146−4150. Reversible Fluorescence Probe Modulated by Selenium for Monitor-
(64) Schwartz, B. J.; Peteanu, L. A.; Harris, C. B. Direct observation ing Peroxynitrite and Imaging in Living Cells. J. Am. Chem. Soc. 2011,
of fast proton transfer: femtosecond photophysics of 3-hydroxy- 133, 11030−11033.
flavone. J. Phys. Chem. 1992, 96, 3591−3598. (84) Luber, S.; Adamczyk, K.; Nibbering, E. T. J.; Batista, V. S.
(65) Becke, A. D. Density-functional thermochemistry. I. V. A new Photoinduced Proton Coupled Electron Transfer in 2-(2′-Hydrox-
dynamical correlation functional and implications for exact-exchange yphenyl)-Benzothiazole. J. Phys. Chem. A 2013, 117, 5269−5279.
mixing. J. Chem. Phys. 1996, 104, 1040−1046.
(66) Becke, A. D. Density-functional exchangeenergy approximation
with correct asymptotic behavior. Phys. Rev. A: At., Mol., Opt. Phys.
1988, 38, 3098−3100.
(67) Lee, C.; Yang, W.; Parr, R. G. Development of the Colle-
Salvetti Correlation-Energy Formula into a Functional of the
Electron-Density. Phys. Rev. B: Condens. Matter Mater. Phys. 1998,
37, 785−789.
(68) Furche, F.; Ahlrichs, R. Adiabatic time-dependent density
functional methods for excited state properties. J. Chem. Phys. 2002,
117, 7433−7447.
(69) Feller, D. The Role of Databases in Support of Computational
Chemistry Calculations. J. Comput. Chem. 1996, 17, 1571−1586.
(70) Li, C.; Yang, Y.; Ma, C.; Liu, Y. Effect of amino group on the
excited-state intramolecular proton transfer (ESIPT) mechanisms of
2-(2′-hydroxyphenyl) benzoxazole and its amino derivatives. RSC
Adv. 2016, 6, 5134−5140.
(71) Kanlayakan, N.; Kerdpol, K.; Prommin, C.; Salaeh, R.;
Chansen, W.; Sattayanon, C.; Kungwan, N. Effects of different
proton donor and acceptor groups on excited-state intramolecular
proton transfers of amino-type and hydroxy-type hydrogen-bonding
molecules: theoretical insights. New J. Chem. 2017, 41, 8761−8771.
(72) Li, C.; Yang, Y.; Li, D.; Liu, Y. A theoretical study of the
potential energy surfaces for the double proton transfer reaction of
model DNA base pairs. Phys. Chem. Chem. Phys. 2017, 19, 4802−
4808.
(73) Huang, J.-D.; Zhang, J.; Chen, D.; Ma, H. Density functional
theoretical investigation of intramolecular proton transfer mechanisms
in the derivatives of 3-hydroxychromone. Org. Chem. Front. 2017, 4,
1812−1818.
(74) Blase, X.; Duchemin, I.; Jacquemin, D. The Bethe−Salpeter
equation in chemistry: relations with TD-DFT, applications and
challenges. Chem. Soc. Rev. 2018, 47, 1022−1043.
(75) Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.;
Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci,
B.; Petersson, G. A.; et al. Gaussian 09, revisionC.01; Gaussian, Inc.:
Wallingford, CT, 2009.
(76) Cammi, R.; Tomasi, J. Remarks on the use of the apparent
surface charges (ASC) methods in solvation problems: Iterative
versus matrix-inversion procedures and the renormalization of the
apparent charges. J. Comput. Chem. 1995, 16, 1449−1458.
(77) Mennucci, B.; Cancès, E.; Tomasi, J. Evaluation of Solvent
Effects in Isotropic and Anisotropic Dielectrics and in Ionic Solutions
with a Unified Integral Equation Method: Theoretical Bases,
Computational Implementation, and Numerical Applications. J.
Phys. Chem. B 1997, 101, 10506−10517.
(78) Cancès, E.; Mennucci, B.; Tomasi, J. A new integral equation
formalism for the polarizable continuum model: Theoretical back-
ground and applications to isotropic and anisotropic dielectrics. J.
Chem. Phys. 1997, 107, 3032−3041.
(79) Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction
analyzer. J. Comput. Chem. 2012, 33, 580−592.
(80) Zhao, G.-J.; Han, K.-L. Hydrogen Bonding in the Electronic
Excited State. Acc. Chem. Res. 2012, 45, 404−413.
(81) Zhao, G.-J.; Han, K.-L. Early Time Hydrogen-Bonding
Dynamics of Photoexcited Coumarin 102 in Hydrogen-Donating
Solvents: Theoretical Study. J. Phys. Chem. A 2007, 111, 2469−2474.
(82) Zhao, G.-J.; Liu, J.-Y.; Zhou, L.-C.; Han, K.-L. Site-Selective
Photoinduced Electron Transfer From Alcoholic Solvents to the
Chromophore Facilitated by Hydrogen Bonding: A New Fluores-

3948 DOI: 10.1021/acs.jpca.9b00879


J. Phys. Chem. A 2019, 123, 3937−3948

Вам также может понравиться