Download as pdf or txt
Download as pdf or txt
You are on page 1of 30

Author’s Accepted Manuscript

Photophysical Behavior of Some Thymol Based


Schiff bases Using Absorption and Fluorescence
Spectroscopy

Ritu Payal, Manju K. Saroj, Neera Sharma,


Ramesh C. Rastogi
www.elsevier.com/locate/jlumin

PII: S0022-2313(17)31592-2
DOI: https://doi.org/10.1016/j.jlumin.2018.02.007
Reference: LUMIN15354
To appear in: Journal of Luminescence
Received date: 13 September 2017
Revised date: 11 January 2018
Accepted date: 2 February 2018
Cite this article as: Ritu Payal, Manju K. Saroj, Neera Sharma and Ramesh C.
Rastogi, Photophysical Behavior of Some Thymol Based Schiff bases Using
Absorption and Fluorescence Spectroscopy, Journal of Luminescence,
https://doi.org/10.1016/j.jlumin.2018.02.007
This is a PDF file of an unedited manuscript that has been accepted for
publication. As a service to our customers we are providing this early version of
the manuscript. The manuscript will undergo copyediting, typesetting, and
review of the resulting galley proof before it is published in its final citable form.
Please note that during the production process errors may be discovered which
could affect the content, and all legal disclaimers that apply to the journal pertain.
Photophysical Behavior of Some Thymol Based Schiff bases Using Absorption and Fluorescence
Spectroscopy
Ritu Payal, Manju K. Saroj, Neera Sharma, Ramesh C. Rastogi*
Department of Chemistry, University of Delhi, Delhi-110007, India
Abstract

The absorption and fluorescence spectra of some biologically active thymol based Schiff bases have been
studied at room temperature in a series of solvents of varying polarity. The solvent effect on the spectral
properties of Schiff bases has been analyzed using Reichardt and Kawski equations and Kamlet-Taft and
Catálan multi-parametric solvent polarity scales. The substituents and the solvents strongly influence the
excited state behavior of these molecules owing to different electronic arrangements. The excited state dipole
moments(e/g and ) of these molecules have been calculated using solvatochromic shift methods based

on bulk solvent polarity functions f (,n) and  (,n) and microscopic solvent polarity parameter E T ,
N

respectively. The increase in excited state dipole moments (e) indicates a more polar excited state and
intramolecular charge transfer (ICT) characteristics of these molecules. The molecular orbitals (HOMO-1,
HOMO, LUMO and LUMO+1) and molecular electrostatic potential (MEP) surfaces have been generated
from their optimized geometries using semi-empirical (PM3) calculations to account for electronic excitations
and changes in the overall charge distribution of Schiff bases. The Taft and Catalán multi-parametric
equations used in the interpretation of the specific solute-solvent interactions reveal that the most important
contribution to the solute-solvent interaction in the excited state comes from the hydrogen-bond acceptor
capacity of the solvent.

Keywords: Solvent effect, Dipole moment, Solvatochromism, Excited state, Stokes-shift, Molecular Orbitals,
Molecular electrostatic potential.

______________________________
*
Corresponding author: E-mail address: rc.rastogi@gmail.com (R.C. Rastogi).
1 Introduction

Schiff bases possess a conjugated system containing a carbon-nitrogen double bond (-C=N-) bound to
two phenyl rings and are represented by the general formula, viz. R1H-C=N-R2R3 [1]. They are well-known to
have a broad range of biological activities such as antifungal [2], antibacterial [3], antimicrobial [4], genotoxic
[5], antiproliferative [6], antiviral [7], antioxidant [8], anticancer [9], etc. Besides various biological activities,
Schiff bases find promising uses in optical switching devices [10], nonlinear optical devices [11], security and
trace inspection devices [12].
Various Schiff bases having diverse backbones with potent biological activities have been explored and
synthesized by various research groups. One of the publications dealing with synthesis of Schiff bases designed
using thymol moiety has reported antioxidant activity [13]. Thymol, a potent molecule, has been identified as an
essential synthetic building block for the synthesis of a variety of biologically active compounds [8,14,15].
Furthermore, its substitution with the imine moiety significantly increases biologically active properties of
Schiff bases which are very much dependent on their molecular structure. It is well known that the spectral
behavior of an organic molecule is strongly related to its structure in both ground and excited states. Thus, a
change of solvent affects both the ground and excited states of a molecule in a different manner.
The solvents strongly influence the chemical and physical properties of the solutes and induce changes in
the electronic transitions of solutes (solvatochromism) [16-20]. The solvent effects are associated with the
nature and the extent of the solute-solvent interactions developed locally within immediate vicinity of solute and
can be studied by absorption and fluorescence spectroscopy due to their strong dependence on solvent polarity.
Thus, solvatochromism can be used as an efficient tool to study various photophysical properties of molecules in
their ground and excited states. The knowledge of solvent effects on absorption and fluorescence spectra is of
particular importance since it can be used to estimate excited-state dipole moments. The knowledge of excited-
state charge distributions and dipole moments is essential in understanding various photodynamic processes
[21,22].

In the present work, photophysical properties of thymol based Schiff bases have been studied in different
solvents using absorption and fluorescence spectroscopy. The solvatochromic shift methods have been applied
to estimate the excited-state dipole moments of different Schiff bases. The ratio (e/g) as well as the change in
dipole moment () has been calculated using bulk and microscopic solvent polarity functions. The excited-
state dipole moments have been determined using theoretically calculated ground-state dipole moments. The
Solvatochromic Comparison method using two well-known multi-parametric scales: Kamlet-Taft and Catalán
has been used to compliment the relative importance of various parameters of medium on the properties of
interest (Schiff bases). The results have been further quantified using their molecular orbital diagrams and
potential energy surfaces.
2 Experimental

2.1. Materials
The thymol based Schiff bases used in the present investigation were obtained from Prof. D.S. Rawat
(Department of Chemistry, University of Delhi) and were used as such. The solvents used were of spectroscopic
and HPLC grade and were checked for the absence of spectral impurities within the requisite scanning range.
Solvents including n-hexane (≥99.5%, Merck), cyclohexane (≥99.7%, Merck), 1,4-dioxane (≥99.8%, Merck),
tetrahydrofuran (≥99.7%, Merck), ethyl acetate (≥99.5%, Merck), chloroform (≥99.4%, Merck),
dichloromethane (≥99.5%, Merck), N,N-dimethylformamide (≥99.8%, Merck), tert-butanol (≥99.5%, Merck),
dimethyl sulphoxide (≥99.8%, Merck), acetonitrile (≥99.8%, Merck), 2-propanol (≥99.7%, Merck), 1-butanol
(≥99.8%, Merck), ethanol (≥99.9%, Merck), methanol (≥99.7%, Merck), ethylene glycol (≥99.0%, Merck) and
triple distilled water were used to prepare the solutions employed in the spectral measurements.
2.2. Instrumentation

Absorption and fluorescence spectra of Schiff bases were measured in solvents of different polarities and
hydrogen bonding abilities using double-beam Spectrophotometer ANALYTIKA JENA UV WINASPECT
SPECORD PC 250 and CARY ECLIPSE VARIAN FLUORESCENCE Spectrofluorometer, respectively at
room temperature. 0.01 M stock solutions of Schiff bases were prepared by dissolving an accurate amount of the
compound in requisite volume of the solvent. The stock solutions were diluted to 4 × 10-5 M and 5 × 10-6 M
solutions to record absorption and fluorescence spectra, respectively with matching quartz cuvettes.
3 Methods: Dipole Moment determination

3.1. Theoretical estimation of ground state dipole moments


The quantum chemical package HyperChem Release 8.0 Pro was used for the theoretical calculations
reported here [23]. The ground state dipole moments (g) and Onsager cavity radii of all the molecules were
estimated by Parametric Method 3 (PM3) using their optimized geometries. The optimized geometry of Schiff
base Sb-OH is shown in Fig. 1.
3.2. Experimental determination of excited-state dipole moments

The dipole moment of a molecule in the excited state is determined by the effect of electric field
(internal or external) on the position alteration of the absorption and fluorescence bands. Two methods
depending on the internal electric field (solvatochromism) have been employed in the present analysis.
3.2.1. Dipole moments using bulk polarity parameters: f (,n) and  (,n)

The solvent dependence on absorption and fluorescence band maximum has been used for the estimation
of excited state dipole moments of different probes using methods given by Bilot and Kawski based on the bulk
polarity parameters (f (,n) and (,n)) [24]. Using the simplest quantum mechanical second order perturbation
theory of the interaction between solvent and spherical solute and the Onsager’s model [25], Kawski et al. [24,
26-31] have obtained the following expressions for (  a  f ) and ( a  f ) in different solvents of varying
dielectric constant () and refractive index (n):

ν a  ν f  m1 f (, n)  const (1)

ν a  ν f  m 2 (, n)  const (2)

where  a and f are the wavenumbers (cm-1) of the absorption and fluorescence maxima,  and n are

dielectric constant and the refractive index of the solvent, respectively. The slopes m1 and m2 are expressed as:

2( e   g ) 2  2( e2   g2 )
m1  and m2  (3)
hca3 hca3
where e and g are the dipole moments of excited and ground states of the solute molecule, respectively and 'a'
is the Onsager cavity radius. f (,n) and  (,n), the bulk solvent polarity parameters, are taken as:
2n 2  1    1 n 2  1 
f (, n )    
n 2  2    2 n 2  2 
(4)

3(n 4  1)
 (, n )  f (, n )  2 (5)
(n  1) 2
Plots of the Stokes-shifts (  a  f ) and ( a  f ) against the bulk solvent polarity functions, f (,n)
and  (,n) for different solvents yield the slopes m1 and m2, respectively. Using Eqs. (1-3) and assuming that
the angle between μe and μg is small and the cavity radius 'a' is same in both the ground and the excited states,
the ratio of the dipole moments in the excited and ground states is given by the relation:

μ e  m1  m 2 
  (m2 > m1)
μ g  m 2  m1 
(6)

It may be noted that most theories [32-34] of the solvent effect describing the location of the absorption (

 a ) and fluorescence (  f ) bands lead to similar expressions for (  a   f ) and (  a  f ) (Eqs. 1 and 2),
inspite of different assumptions, with the difference, however, that the applied solvent polarity parameters,
f (,n) and  (,n) differ significantly.

3.2.2. Dipole moments using microscopic solvent polarity parameters ETN


In this method, Ravi et al. [35], explained the theoretical basis for the correlation of spectral shifts with

E TN . The main source of error in the estimation of dipole moment is the limited accuracy of the assessed value
for the radius 'a' of the solute molecule as  is a cubic function of a. The formulation used in this method offers
a partial cancellation of the problems associated with Onsager radius, since the ratio of two Onsager radii (a B/a)
is used (Eq. 7). Hence, the problem associated with the error in estimation of the Onsager radius 'a' is
minimized. The excited-state dipole moment is determined by the following equation:

a  f = 11307.6 [(∆μ / ∆μB)2(aB / a)3] E TN + constant (7)

where (  a  f ) is the Stokes-shift. μB (= 9D) and aB (= 6.2Å) are the dipole moment change on excitation
and the Onsager cavity radius, respectively, of betaine dye [36]. Δμ and 'a' are the corresponding quantities for
the molecule of interest (Schiff bases).

E TN is the solvent polarity parameter proposed by Reichardt [37], which is based on the absorption
wavenumber of standard betaine dye in the solvent. Its values for different solvents are given in Table 1. The

E TN value for any solvent can be expressed as:


E T (solvent )  E T (TMS)
E TN  (8)
E T ( water)  E T (TMS)
where ET(solvent), ET(TMS) and ET(water) are defined according to the solvent polarity parameter, ET(30)

proposed by Dimroth and Reichardt [38,39]. The plot of (  a  f ) (Stokes-shifts) vs. E TN (Eq.7) is used to

obtain the change in the dipole moment (Δμ) from the slope 'm' upon excitation.
   2  a 3 
slope (m)  11307.6    B   (9)
  B   a  

It is difficult to estimate the exact cavity radii for elongated molecules having ellipsoidal form. Therefore,
the Onsager cavity radii of all the molecules have been taken as 40% of the distance between two extreme high
charge centers in the optimized geometries using semi-empirical computations [18,35,40,41]. This is consistent
with the fact that these atoms provide the strongest dipole vector component in all the molecules considered.
Thus, knowing ground state dipole moment ‘g’ from theoretical calculations, excited state dipole moment, e
can be determined.

3.2.3. Specific Interactions: Kamlet-Taft and Catalán Treatments

Solvatochromic parameters have proved to be fruitful in correlating a wide range of chemical and
physical properties involving solute-solvent interactions. A number of polarity scales for such parameters can be
found in the literature however, two well-known multiparametric scales: Kamlet-Taft [42,43] and Catalán
[44,45] solvatochromic methods or Linear Solvation Energy Relationships (LSER) quantify best the nature of
different solute-solvent specific interactions.

The empirical solvent scales of Kamlet-Taft and Catalán that are based on polarity/polarizability
(SPP/*), acidity (SA/) and basicity (SB/) of a solvent have been used to interpret various photophysical
properties in all types of solution equilibria by measuring a specific interaction that is local to the solvation shell
around a dissolved solvatochromic solute. The expressions for Kamlet-Taft (Eq. 10) and Catalán (Eq. 11) scales
are as follows:

ν  ν  s * π *  a α α b ββ
o
(10)

ν  ν  s SPP SPP  a SASA  b SBSB


o
(11)

The coefficients s*/SSPP, a/aSA and b/bSB are interpreted as solute properties. The coefficient, s*/SSPP
is related to the solute’s polarity/polarizability character, a/aSA describes its tendency to accept hydrogen bond
from the solvent and b/bSB measures its property to donate hydrogen bond to the solvent.
The corresponding parameters for different solvents were taken from literature [46,47] and are given in
Table 1.
4 Results and discussion
Schiff bases consist of two phenyl rings (1 and 2) bridged via an azomethine group (-C=NH-). Schiff
bases have been selected with different electron-donating or electron-withdrawing substitution groups on phenyl
ring 2 to study their photophysical behavior. Absorption and fluorescence spectra of Schiff bases have been
studied in a series of solvents of different polarities. The absorption spectra of Sb-OH and Sb-CF3 are shown in
Figs. 2a and 3a, respectively and their absorbance band maxima (A1 and A2) are given in Table 2. The
fluorescence band maxima of Schiff bases have been obtained from Gaussian fitted curves of their
fluorescence spectral bands using Microcal Origin 6.0 professional software [48] and are shown in Fig. 2c for
Sb-OH. The corresponding fluorescence band maxima (F) of Schiff bases have been compiled in Table 3.

4.1. Absorption spectral properties


The absorption spectra of all Schiff bases consist of two well-resolved bands A1 and A2. Their
absorption band maxima (A1 and A2) lie in the range of ~240-285 nm and ~330-410 nm, respectively with the
change in solvents as well as substitution at phenyl ring (2) (Fig. 2a, Table 2). They are found to show small
bathochromic shifts in A1 (~3-11, -* transition) nm and A2 (~6-15 nm, -* transition) with the increase in
polarity of solvents from n-hexane to water. Moreover, the absorption bands become slightly broad and
unstructured on increasing the polarity of the medium (Table 2, Fig. 2a). These results show that the ground
state of these Schiff bases is slightly affected by the polarity of the solvents, which confirms the presence of
reasonable solute-solvent interactions.

While the first absorption band A1 can be assigned to the -* transition of the phenyl ring, the second
absorption band A2 must be due to the interaction of the phenyl ring with the whole molecular moiety [49,50].
No clearly visible n-* transitions have been observed, since they are possibly masked by the more intense -*
type transition ( ≈ 105 M-1cm-1) [51,52]. This assignment of the bands finds support from the literature [53].
Substituent effects in these molecules are evident from the A2 values of Sb-CF3 and Sb-NO2. Both these
molecules show bathochromic shifts in A2 relative to other molecules (Sb-OH, Sb-OCH3, Sb-CH3 & Sb-C2H5)
(Table 2). The outcome can be assigned to the highly electron-withdrawing nature of Sb-CF3 and Sb-NO2.
Exceptionally though, it can be seen that both of these molecules (Sb-CF3 and Sb-NO2) show blue shift from
methanol to water. This may be due to extensive hydrogen bonding among water molecules and non-bonding
electrons present on CF3 and NO2 groups, which increase the energy difference between the Franck-Condon
ground and excited states.
4.2. Fluorescence spectral properties

The wavelength of absorption band, A2 was used as the excitation wavelength to record the fluorescence
spectra of these molecules.
All Schiff bases except Sb-NO2 (non-fluorescent) show a single structured fluorescence band with high
fluorescence intensity in all the solvents (Fig. 2b). Fluorescence band maxima (F) of these molecules as
observed in different solvents lie in the range of 395-455 nm (Fig. 2c, Table 3). All Schiff bases show
bathochromic shifts of ~39-53 nm in F with increase in solvent polarity from n-hexane to water. These results
indicate strong interaction of these molecules with the solvents in the excited state.

A large bathochromic shift with high fluorescence intensity in the fluorescence spectra of these molecules
with increase in the solvent polarity indicates the stabilization of the excited state. These findings suggest a
strong intramolecular charge transfer (ICT) in the excited state relative to the ground state. The absence of
proton donating substituent group in the close proximity of azomethine nitrogen atom rules out the possibility of
excited state intramolecular proton transfer (ESIPT). Therefore, the molecules are found to show ICT due to
their donor-acceptor properties which is also confirmed by the previous solvatochromic studies on some similar
compounds i.e. p-hydroxybenzylideneimines, quinoline N-oxides, benzaldazines and aminotriazole based Schiff
bases [50,54-57].
Schiff bases with different substituents can show two possible resonance structures i.e. RS-I and RS-II in
their excited states as a result of the ICT (Fig. 4). In case of RS-I, delocalization of the π-electrons takes place
from R group of phenyl ring 2 to the nitrogen atom of azomethine moiety [55]. However, in case of RS-II,
π-electron delocalization occurs from hydroxyl group of phenyl ring 1 to phenyl ring 2 [57]. Schiff bases
having electron-donating (-OH, -OCH3) substituents on phenyl ring 2 can follow RS-I. A similar electron
delocalization (RS-I) can take place in case of substituents with positive inductive effect (-CH3, -C2H5).
However, Schiff bases with electron-substituents (-CF3, -NO2) can follow the resonance structure RS-II in the
excited state.
All Schiff bases except Sb-CF3 have been found to show large bathochromic shifts of ~14-31 nm from
ethylene glycol to water because of hydrogen bonding between the-OH group present on the phenyl ring 1 and
water molecules. This type of hydrogen bonding stabilizes the excited state more than the ground state resulting
in a decrease of energy gap between the ground and excited states [37]. Sb-CF3 is showing bathochromic shift of
only 4 nm from ethylene glycol to water as -OH group present on the phenyl ring 1 is involved in a resonance
structure (Fig. 4b, RS-II).
The fluorescence spectra of Sb-CF3 show decrease in fluorescence intensity in all the solvents unlike
other molecules. This is due to the presence of trifluoromethyl group, which quenches the fluorescence intensity
of Sb-CF3 [58]. Further, Sb-NO2 molecule is found to be completely non-fluorescent, which must be a
consequence of strong electron-withdrawing capacity of the nitro group. It is known that the efficient non-
radiative decay processes such as singlet-triplet intersystem crossing and internal conversion impart non-
fluorescent or weakly fluorescent characteristics to the aromatic nitro- compounds [59,60].
4.3. Electronic Properties of Schiff bases: Charge Transfer Prediction

The mechanism of electronic excitations (i.e. donor and acceptor relationship) and changes in the overall
charge distribution of Schiff bases can be understood through molecular orbitals diagrams [61]. For this
purpose, HOMO-1, HOMO, LUMO and LUMO+1 were generated from their structures optimized using PM3
method. The molecular orbitals for Sb-OH and Sb-CF3 are presented in Fig. 5, which provide an idea about the
charge distribution in the ground as well as in the excited state.
HOMO and LUMO of Schiff bases are almost localized on the whole molecule, whereas HOMO-1 and
LUMO+1 show different electron density arrangements. Molecular orbitals, HOMO-1 and LUMO+1 give a
clear indication about the resonance structures followed by these molecules in their excited state (Fig. 4). In case
of Sb-OH, Sb-OCH3, Sb-CH3 and Sb-C2H5, HOMO-1 and LUMO+1 are localized on the phenyl rings 2 and 1,
respectively. This implies electron density transfer from 2 to 1. Hence, RS-I is followed by these molecules.
For Sb-CF3 and Sb-NO2, HOMO-1 and LUMO+1 show the localization of electron densities on phenyl rings 1
and 2, respectively. Therefore, electron density gets transferred from 1 towards 2 [62,63]. Consequently,
resonance structure RS-II is favored for these molecules.

To get a clear insight into interaction of Schiff bases with solvents, their Molecular Electrostatic Potential
(MEP) surfaces were obtained from their optimized geometries by using PM3 method (Fig. 6) [64,65].
The MEP at a point in the vicinity of a molecule gives information about the net electrostatic effect
produced at that point by total charge distribution of the molecule. The difference in electrostatic potential at the
surface is represented by different colors. Red, blue and green colors represent regions of the most negative
electrostatic potential, the most positive electrostatic potential and close to zero potential, respectively.

Hence, the negative potential region (Fig. 6, red color) in MEP lies on electronegative atoms (oxygen
atom of -OH, nitrogen atom of azomethine group, oxygen atoms of NO2 and fluorine atoms of CF3). On the
other hand, positive potential region lies on the hydrogen atoms of the molecules.
4.4. Dipole moments of Schiff bases

Solvatochromic shifts caused by general (non-specific) solvent effects are often described by
solvatochromic methods, which relate the energy difference between absorption and fluorescence maxima to the
dielectric constant and refractive index.

Absorption and fluorescence spectral properties of Schiff bases have shown that the protic and aprotic
solvents affect the molecules in different ways because of the presence of specific solute-solvent interactions in
protic solvents. Therefore, solvents were divided into two sets i.e. protic (tert-butanol & from 2-Propanol to
water) and aprotic (from n-hexane to DMSO and acetonitrile) solvents for the calculation of dipole moments.
For the purpose of estimation of dipole moment, the spectral shifts for the Schiff bases, (  a  f ) and (
a  f ), have been calculated from their absorption (A2) and fluorescence band (F) maxima in different

solvents (Tables 2 & 3). The correlation coefficients obtained from the plots of (  a  f ) and ( a  f ) vs.
bulk solvent polarity functions, f (,n) and (,n) (Eqs. 1 & 2) as well as microscopic solvent polarity parameter

E TN (Eq. 7) show good correlation (Table 4 and Fig. 7). The ratio of dipole moments (eg) and the change in
dipole moments () have been calculated for all the Schiff bases (except Sb-NO2) and are compiled in Tables 5
& 6 [24,35]. The ground state dipole moments (g) were theoretically calculated and used to compute excited-
state dipole moments by considering and eg values.

The eg values of Schiff bases were calculated in protic and aprotic solvents by using slopes, m1 and
m2 obtained from the plots of Eqs. 1 & 2 (Table 5, Fig. 7). The  values of Schiff bases were calculated using
slope (m, Eq. 9, Fig. 7) and Onsager cavity radii (theoretically calculated) in protic and aprotic solvents
(Table 6, Eq. 9) [40].

Large eg andevalues for the Schiff bases show that the dipole moments of these molecules
increase in their excited states. This increase in dipole moment indicates a considerable amount of
intramolecular charge transfer in these molecules via resonance structures I and II. eg ande values in
protic solvents are found to be somewhat higher when compared with the aprotic solvents, which must be due to
additional stabilization of RS-I and RS-II in protic solvents caused by specific interactions.
Also, the dipole moment of Sb-CF3 is found to be greater than that of other molecules because of greater
delocalization of charge within the molecule as shown by RS-II. The other Schiff bases follow RS-I and show
partial charge transfer and lower dipole moments than Sb-CF3.

4.5. Multiparametric Solvatochromic Kamlet-Taft and Catalán Treatments

Schiff bases exhibit a distinct solvatochromic behavior in different solvents i.e. protic and aprotic
solvents. The solvent effects on the absorption and fluorescence bands in different solvents have been analyzed
using multiparametric Kamlet-Taft (, , ) and Catalán (SPP, SA, SB) treatments (Eqs. 10 & 11) [66-68]. The
results obtained using Kamlet-Taft and Catalán approaches on Schiff bases are presented in Tables 7 & 8.
The possible specific solute-solvent interactions of Schiff bases are shown in Fig. 8. These molecules can
show three types of hydrogen bonding interactions in the solvent environment: (a) Type X: hydrogen bond
accepting (HBA) solvents and hydroxyl group, (b) Type Y: hydrogen bond donating (HBD) solvents and
nitrogen atom, and (c) Type Z: HBD solvents and electronegative atoms: fluorine and oxygen of Schiff bases
[69,70].

Various interactions of Schiff bases with different solvents are also justified by the MEP surfaces (Fig. 6)
which predict the various hydrogen bonding sites within the molecules. The region of MEP showing negative
potential includes electronegative atoms (oxygen atom of hydroxyl group, nitrogen atom of azomethine group as
well as oxygen and fluorine atoms of -NO2 and -CF3 groups). Therefore, these can form hydrogen bonds with
hydrogen-donating solvents. The presence of positive potential around the hydrogen atoms in Schiff bases
enables their donation to the hydrogen-accepting solvents.
Excitation Processes

The large negative values of b and bSB in ground state reveal the higher HBD ability of the molecules, as
the hydroxyl group participates in the formation of hydrogen bond with the solvents (type X). Large negative
values indicate stabilization because hydrogen bonding stabilizes the excited state and decreases the energy gap
and a bathochromic shift is observed (Table 7). Large negative values for coefficients, b and bSB for Sb-CF3 and
Sb-NO2 are a consequence of the strong electron-withdrawing capacity of -CF3 and -NO2 groups [65]. The
negative coefficients for s* and SSPP indicate an increase in the solvent dipolarity/polarizability. The positive
values of a and small negative values of aSA show the poor HBA ability of the Schiff bases in the ground state.

Emission Processes

The more negative values of a and aSA values relative to b and bSB in emission process as compared to
the excitation process reveal that the HBA ability of Schiff bases is enhanced in the excited state (type Y)
[59,71]. This shows that the nitrogen atom of azomethine moiety is available for contribution to the hydrogen
donating solvents (Figs.4 & 8, Table 8). Although the nitrogen atom is not available for participation in case of
Sb-CF3, electronegative fluorine atom is available for hydrogen donating solvents (type Z). The molecule Sb-
CF3 has negative b and bSB coefficients while all other molecules have positive value; this is attributed to
electron-withdrawing capacity of CF3 group.
The decrease in b and bSB values in the emission process as compared to excitation process reveals that
the strength of hydrogen bonding ability of Schiff bases is reduced because of higher contribution of ICT
character. The more negative values of s*/SSPP and a and aSA in emission process as compared to the excitation
process reveal that the excited state has higher dipolarity/polarizability than the ground state [72]. This validates
the excited state to be more polar and environment sensitive than the ground state as is evident from large μe,
μe/μg and Δμ values (Tables 5-8).
Thus, a detailed analysis of solvatochromism and preferential solvation provides insights into how solute-
solvent and solvent-solvent interactions govern the solvation shell of these Schiff bases.
Conclusion
The photophysical properties of Schiff bases in different solvents with variation of polarity and hydrogen
bonding ability of solvents have been discussed. Schiff bases showed high fluorescent intensity and large Stokes
shifts for all solvents. These studies have brought into focus some important features observed in the spectral
properties due to ICT characteristics of these molecules. Also, the solvation shell of the solute is largely affected
by the solute-solvent interactions. The experimentally calculated e values of these molecules reveal an increase
in the excited state dipole moment due to a large charge separation in the excited state. The change in dipole
moment () is found to be positive and implies a more polar excited state in comparison to ground state. Both,
Kamlet-Taft and Catalan approaches reveal highly acidic and basic characters of Schiff bases in ground and
excited states, respectively. The more negative values of s*/SSPP and a and aSA in emission process as
compared to the excitation process validate large ICT characteristics of Schiff bases in the excited state.
Acknowledgements

Ritu Payal is thankful to the University Grants Commission (UGC), New Delhi for the financial support.
The financial support received from the University of Delhi under the Scheme “To strengthen R & D Doctoral
Research Program” is gratefully acknowledged. We gratefully acknowledge Prof. D.S. Rawat, University of
Delhi for providing the required samples.
References
1. Schiff H (1864) Mittheilungen aus dem universitätslaboratorium in pisa: Eine neue Reihe organischer basen.
Justus Liebigs Ann Chem 131(1):118-119
2. Ramesh R, Maheswaran S (2003) Synthesis, spectra, dioxygen affinity and antifungal activity of Ru(III)
Schiff base complexes. J Inorg Biochem 96(4):457-462
3. Esmadi FT, Khabour OF, Abbas K, Mohammad AE, Obeidat RT, Mfady D2 (2016) Synthesis,
characterization and biological activity of some unsymmetrical Schiff base transition metal complexes. Drug
Chem Toxicol 39(1):41-47
4. Kavitha P, Reddy KL (2016) Synthesis, spectral characterisation, morphology, biological activity and DNA
cleavage studies of metal complexes with chromone Schiff base. Arabian J Chem 9(4):596–605
5. Ispir E, Toroğlu S, Kayraldız A (2008) Syntheses, characterization, antimicrobial and genotoxic activities
of new Schiff bases and their complexes. Transition Met Chem 33(8):953-960
6. Neelakantan MA, Esakkiammal M, Mariappan SS, Dharmaraja J, Jeyakumar T (2010) Synthesis,
characterization and biocidal activities of some Schiff base metal complexes. Indian J Pharm Sci 72(2):
216-222
7. Kumar KS, Ganguly S, Veerasamy R, De EC (2010) Synthesis, antiviral activity and cytotoxicity
evaluation of Schiff bases of some 2-phenyl quinazoline-4(3)H-ones. Eur J Med Chem 45(11):5474-5479
8. Sashidhara KV, Rosaiah JN, Bhatia G, Saxena JK (2008) Novel keto-enamine Schiffs bases from
7-hydroxy-4-methyl-2-oxo-2H-benzo[h]chromene-8,10-dicarbaldehyde as potential antidyslipidemic and
antioxidant agents. Eur J Med Chem 43(11):2592-2596
9. Chen W, Ou W, Wang L, Hao Y, Cheng J, Li J, Liu YN (2013) Synthesis and biological evaluation of
hydroxyl-substituted Schiff-bases containing ferrocenyl moieties. Dalton Trans 42(44):15678-15686
10. Zakerhamidi MS, Nejati K, Golghasemi Sorkhabi Sh, Saati M (2013) Substituent and solvent effects on the
spectroscopic properties and dipole moments of hydroxyl benzaldehyde azo dye and related Schiff bases.
J Mol Liq 180:225-234
11. Ziółek M, Burdziński G, Karolczak J (2009) Influence of intermolecular hydrogen bonding on the
photochromic cycle of the aromatic Schiff base N,Nʼ-Bis(salicylidene)-p-phenylenediamine in solution.
J Phys Chem A 113(12):2854-2864
12. Han T, Hong Y, Xie N, Chen S, Zhao N, Zhao E, Lam JWY, Sung HHY, Dong Y, Tong B, Tang BZ
(2013) Defect-sensitive crystals based on diaminomaleonitrile-functionalized Schiff base with aggregation-
enhanced emission. J Mater Chem C 1(44):7314-7320
13. Beena, Kumar D, Rawat DS (2013) Synthesis and antioxidant activity of thymol and carvacrol based Schiff
bases. BioOrg Med Chem Lett 23(3):641-645
14. Kavoosi G, Dadfar SM, Purfard AM (2013) Mechanical, physical, antioxidant, and antimicrobial properties
of gelatin films incorporated with thymol for potential use as nano wound dressing. J Food Sci 78 (2):
244-250
15. Pan K, Chen H, Davidson PM, Zhong Q (2014) Thymol nanoencapsulated by sodium caseinate: Physical
and antilisterial properties. J Agric Food Chem 62(7):1649-1657
16. Kumar S, Jain SK, Rastogi RC (2001) An experimental and theoretical study of excited-state dipole
moments of some flavones using an efficient solvatochromic method based on the solvent polarity
parameter, E TN . Spectrochim. Acta A 57(2):291-298
17. Sharma N, Jain SK, Rastogi RC (2003) Excited-state dipole moments of indoles using solvatochromic shift
methods: An experimental and theoretical study. Bull Chem Soc Jpn 76(9):1741-1746
18. Sharma N, Jain SK, Rastogi RC (2007) Solvatochromic study of excited state dipole moments of some
biologically active indoles and tryptamines. Spectrochim Acta A 66(1):171-176
19. Saroj MK, Sharma N, Rastogi RC (2012) Photophysical study of some 3-benzoylmethyleneindol-2-ones
and estimation of ground and excited states dipole moments from solvatochromic methods using solvent
polarity parameters. J Mol Struct 1012:73-86
20. Raghavendra UP, Basanagouda M, Melavanki RM, Fattepur RH, Thipperudrappa J (2015) Solvatochromic
studies of biologically active iodinated 4-aryloxymethyl coumarins and estimation of dipole moments. J
Mol Liq 202:9-16
21. Desai VR, Hunagund SM, Pujar MS et al (2017) Photophysical properties of a novel and biologically
active 3(2H)-pyridazinone derivative using solvatochromic approach. J Fluoresc doi: 10.1007/s10895-017-
2117-z
22. Zang L, Zhao H, Ji X, Cao W, Zhang Z, Meng P (2017) Photophysical properties, singlet oxygen
generation efficiency and cytotoxic effects of aloe emodin as a blue light photosensitizer for photodynamic
therapy in dermatological treatment. Photochem Photobiol Sci 16(7):1088-1094
23. HyperChem Release 8.0 professional (2011). Hypercube, Inc., USA
24. Kawski A (2002) On the estimation of excited-state dipole moments from solvatochromic shifts of
absorption and fluorescence spectra. Z Naturforsch 57a(5):255-262
25. Onsager L (1936) Electric moments of molecules in liquids. J Am Chem Soc 58:1486-1493
26. Kawski A, Stefanowska U (1965) The anomalous red shift of the absorption and fluorescence spectra of
4-aminophthalimide in dependence on the ratio of homo- and heteropolar solvents. Acta Phys Polon
28(6):809-822
27. Kawski A (1965) Abnormal stokes shift of the absorption and of the fluorescence maximum of
4-aminophthalimide in dioxane-water mixtures. Acta Phys Polon 28(5):647-652
28. Kawski A (1964) Effect of polar molecules on electronic spectrum on 4-amino-phthalimide Acta Phys
Polon 25(2):285-290
29. Bilot L, Kawski A (1963) Dipole moments of some phthalimide derivatives in the first excited singlet state.
Z Naturforsch 18a:256
30. Bilot L, Kawski A (1963) Effect of the solvent on the electronic spectrum of luminescent molecules.
Z Naturforsch 18a:10-15
31. Bilot L, Kawski A (1962) Theory of the effect of solvents on the electron spectra of molecules.
Z Naturforsch 17a:621-627
32. Ooshika Y (1954) Absorption spectra of dyes in solution. J Phys Soc. Jpn. 9(4):594-602
33. McRae EG (1957) Theory of solvent effects on molecular electronic spectra: Frequency shifts. J Phys
Chem 61:562-572
34. Bakhshiev NG (1961) Universal intermolecular interactions and their effect on the position of the
electronic spectra of the molecules in two-component solutions I: Theory (liquid solutions). Opt Spectrosk
10:717-726
35. Ravi M, Samanta A, Radhakrishnan TP (1994) Excited state dipole moments from an efficient analysis of
solvatochromic stokes shift data. J Phys Chem 98(37):9133-9136
36. Reichardt C (1994) Solvatochromic dyes as solvent polarity indicators. Chem Rev 94(8):2319-2358
37. Reichardt C (1998) Solvents and solvent effects in organic chemistry. VCH, Weinheim
38. Dimroth K, Reichardt C, Siepmann T, Bohlmann F (1963) Pyridinium N-phenolbetaines and their use for
the characterization of the polarity of solvents. Justus Liebigs Ann Chem 661:1-37
39. Reichard C (1971) Pyridinium-N-phenol betaines and their application for the characterization of solvent
polarities. VI. Extension of the solvent polarity scale by determination of new molar transition energies (E T
values). Justus Liebigs Ann Chem 752:64-67
40. Lippert EZ (1955) Dipole moment and electronic structure of excited molecules. Z Naturforsch 10a:
541-545
41. Hermant RM, Bakker NAC, Scherer T, Krijnen B, Verhoeven JW (1990) Systematic study of a series of
highly fluorescent rod-shaped donor-acceptor systems, J Am. Chem Soc 112(3):1214-1221
42. Kamlet MJ, Abboud JM, Taft RW (1981) Progress in Physical Organic Chemistry. Wiley, New York
43. Mataga N, Kubata T (1970) Molecular Interactions and Electronic Spectra. Marcel Dekker, New York
44. Catalán J (1995) On the * Solvent Scale. J Org Chem 60(25):8315-8317
45. Catalán J (1997) On the ET(30), *, Py, Sʼ and SPP empirical scales as descriptors of nonspecific solvent
effects. J Org Chem 62(23):8231-8234
46. Taft RW, Abboud JLM, Kamlet MJ (1981) Solvatochromic comparison method. 20: Linear solvation
energy relationships-12: The dδ term in the solvatochromic equations. J Am Chem Soc 103(5):1080-1086
47. Catalán J, López V, Pérez P, Villamil RM, Rodriguez JG (1995) Progress towards a generalized solvent
polarity scale: The solvatochromism of 2-(dimethylamino)-7-nitrofluorene and its homomorph 2-fluoro-7-
nitrofluorene. Liebigs Ann 2:241-252
48. Microcal Origin 6.0 professional (1991-1999) Microcal software, Inc., USA
49. Ebara N (1960) Benzylideneaniline. II. Iodine complexes of benzylideneaniline and its derivatives. Bull
Chem Soc Jpn 33(4):540-543
50. Alexander PW, Sleet RJ (1970) Solvent effects on the ultraviolet absorption spectra of o-, m-, and
p-hydroxybenzylideneimines. Aust J Chem 23(6):1183-1190
51. Etaiw SEH, Awad MK, Fayed TA, El-Hendawy MM (2009) Effect of N-methylation on both ground and
excited states properties of 1-(9-anthryl)-2-(2-benzothiazolyl) ethane. J Mol Struct 919(1-3):12-20
52. Jaffe HH, Yeh S, Gardner RW (1958) The electronic spectra of azobenzene derivatives and their conjugate
acids. J Mol Spectrosc 2:120-136
53. Ebara N (1960) Benzylideneaniline. I. Structure and ultraviolet absorption spectrum of benzylideneaniline.
Bull Chem Soc Jpn 33(4):534-539
54. Okafor EC (1980) Aspects of acyclic azine chemistry-I. The effect of structure and solvents on the
electronic absorption spectra of benzaldazine. Spectrochim Acta A 36(2):207-212
55. Astudillo M, Chokotho NCJ, Jarvis TC, Johnson CD, Lewis CC, McDonnell PD (1985) Hydroxy schiff
base-oxazolidine tautomerism: Apparent breakdown of baldwinʼs rules. Tetrahedron 41(24):5919-5928
56. Andreev VP, Batotsyrenova EG, Ryzhakov AV, Rodina LL (1998) Intramolecular charge transfer
processes in a series of styryl derivatives of pyridine and quinoline n-oxides. Chem Heterocyc Compd
34(8):941-949
57. Issa YM, Hassib HB, Abdelaal HE, Kenawi IM (2011) Spectral investigation of the intramolecular charge-
transfer in some aminotriazole Schiff bases. Spectrochim Acta A 79(5):1364-1374
58. Nagib DA, MacMillan DWC (2011) Trifluoromethylation of arenes and heteroarenes by means of
photoredox catalysis. Nature 480:224-228
59. Saroj MK, Sharma N, Rastogi RC (2011) Solvent effect profiles of absorbance and fluorescence spectra of
some indole based chalcones. J Fluoresc 21(6):2213-2227
60. Sonoda Y, Tsuzuki S, Goto M, Tohnai N, Yoshida M (2010) Fluorescence spectroscopic properties of
nitro-substituted diphenylpolyenes: Effects of intramolecular planarization and intermolecular interactions
in crystals. J Phys Chem A 114(1):172-182
61. Desai VR, Hunagund SM, Basangouda M, Kadadevarmath JS, Sidarai AH (2016) Solvent effects on the
electronic absorption and fluorescence spectra of HNP: Estimation of ground and excited state dipole
moments. J Fluoresc 24(4):1391-1400.
62. Sánchez AJ, Rodríguez M, Métivier R, Ortíz GR, Maldonado JL, Réboles N, Farfán N, Nakatanic K,
Santillan R (2014) Synthesis and crystal structures of a series of Schiff bases: A photo-, solvato- and
acidochromic compound. New J Chem 38:730-738
63. Rančić M, Trišović N, Milčić M, Ušćumlić G, Marinković A (2012) Substituent and solvent effects on
intramolecular charge transfer of 5-arylidene-2,4-thiazolidinediones. Spectrochim Acta A 86:500-507
64. Muthu S, Prasath M (2013) Quantum chemical studies, vibrational analysis, molecular structure, first order
hyper polarizability, NBO and HOMO-LUMO analysis of 3-hydroxybenzaldehyde and its cation.
Spectrochim Acta A 115:789-799
65. Jayabharathi J, Thanikachalam V, Vennila M, Jayamoorthy K (2012) DFT based ESIPT process of
luminescent chemosensor: Taft and Catalan Solvatochromism. Spectrochim Acta A 95:589-595
66. Saha SK, Purkayastha P, Das AB (2008) Photophysical characterization and effect of pH on the twisted
intramolecular charge transfer fluorescence of trans-2-[4-(dimethylamino)styryl]benzothiazole.
J PhotoChem Photobiol A 195(2-3):368-377
67. Jayabharathi J, Thanikachalam V, Perumal MV, Jayamoorthy K (2012) Solvatochromic studies of
fluorescent azo dyes:Kamlet-Taft (π*, α and β) and Catalan (SPP, SA and SB) solvent scales approach.
J Fluoresc 22(2):213-221
68. Stock RI, Nandi LG, Nicoleti CR, Schramm ADS, Meller SL, Heying RS, Coimbra DF, Andriani KF,
Caramori GF, Bortoluzzi AJ, Machado VG (2015) Synthesis and Solvatochromism of Substituted
4-(Nitrostyryl)phenolate Dyes. J Org Chem 80(16):7971-7983
69. Arbeloa TL, Arbeloa FL, Tapia MJ, Arbeloa IL (1993) Hydrogen-bonding effect on the photophysical
properties of 7-aminocoumarin derivatives. J Phys Chem 97:4704-4707
70. Liu X, Cole JM, Low KS (2013) Solvent effects on the uv-vis absorption and emission of optoelectronic
coumarins: A comparison of three empirical solvatochromic models. J Phys Chem C 117:14731-14741
71. Li M, Huang J, Zhou X, Luo H (2010) Synthesis, characterization and spectroscopic investigation of a
novel phenylhydrazone Schiff base with solvatochromism. Spectrochim Acta A 75(2):753-759
72. Naderi F, Farajtabar A (2016) Solvatochromism of fluorescein in aqueous aprotic solvents. J Mol Liq
221:102-107
Figure Captions
Fig. 1 (a) Structures of different Schiff bases.
(b) Optimized geometry of Sb-OH in ground state.
Fig. 2 (a) Absorption spectra, (b) Fluorescence spectra and (c) Gaussian fitted fluorescence spectra of Sb-OH
in different solvents.
Fig. 3 (a) Absorption spectra and (b) Fluorescence spectra of Sb-CF3 in different solvents.

Fig. 4 Possible resonance structures of Schiff bases.

Fig. 5 Various molecular orbitals of (a) Sb-OH and (b) Sb-CF3. The isosurfaces correspond to the wave
function value of ±0.015.

Fig. 6 Molecular electrostatic potential surfaces for different Schiff bases: (a) Sb-OH, (b) Sb-OCH3,
(c) Sb-OCH3, (d) Sb-C2H5, (e) Sb-CF3, (f) Sb-NO2 . The red and blue colors correspond to negative
and positive potential regions, respectively. Green color corresponds to the neutral region.
Fig. 7 Plots of Stokes-shifts ( a  f ) and ( a  f ) of Sb-OH vs. f (,n) and (,n), bulk solvent polarity
parameters for (a) protic and (b) aprotic solvents.
Fig. 8 Possible solute-solvent interactions of Schiff bases (S and SH = solvent). Hydrogen bonds may be
formed via: hydrogen atom of -OH group (type X) by hydrogen-bond accepting solvents, nitrogen atom
of azomethine moiety (type Y) by hydrogen-bond donating solvents, and hydrogen bonding between
fluorine atom and hydrogen donating solvents (type Z in excited state in case of Sb-CF3 only).
Table 1 Different polarity parameters used in the solvatochromic studies for
different solvents
Kamlet-Taft
Catalán parametersb
parametersb
f 
Solven 
n E TN (, (,
ta   S
n) n) π S S
  * P
A B
P
n-Hexane 1.88 1.37 0.00 - 0.254 0. 0. - 0.0 0.0 0.5
49 9 0.003 0 0 0. 0 5 1
Cyclohex 2.02 1.42 0.00 - 0.288 0. 0 0. 0 0.
0 0.0 0 0.0 6 0.5 9
ane 62 6 0.003 0 0 1 0 0 7 5
1,4- 2.21 1.42 0.16 0.040 0.310 0. 0 0. 0 0. 0 0.0 0 0.4 3 0.7 7
Dioxane 24 4 0 3 4 0 4 0
THF 7.58 1.40 0.20 0.549 0.551 0. 0 0. 7 0. 9 0.0 0 0.5 4 0.8 1
72 7 0 5 5 0 9 3
Ethyl 6.02 1.37 0.22 0.489 0.498 0. 0 0. 5 0. 5 0.0 0 0.5 1 0.7 8
acetate 24 8 0 4 4 0 4 9
Chlorofor 4.81 1.44 0.25 0.370 0.490 0. 0 0. 5 0. 5 0.0 0 0.0 2 0.7 5
m 59 9 4 0 5 4 7 8
DCM 6.93 1.42 0.30 0.590 0.583 0. 4 0. 5 0. 8 0.0 7 0.1 1 0.8 3
42 9 1 1 7 4 7 7
DMF 6.71 1.43 0.38 0.840 0.710 0. 3 0. 0 0. 3 0.0 0 0.6 8 0.9 6
05 6 0 6 8 3 1 5
tert- 12.4 1.38 0.38 0.688 0.608 0. 0 0. 9 0. 8 0.1 1 0.9 3 0.8 4
Butanol 0 6 9 4 9 4 4 2 2
DMSO 47.2 1.47 0.44 0.841 0.744 0. 2 0. 3 1. 1 0.0 5 0.6 8 1.0 9
4 90 4 0 7 0 7 4 0
Acetonitril 35.9 1.34 0.46 0.860 0.664 0. 0 0. 6 0. 0 0.0 2 0.2 7 0.8 0
e 4 41 0 1 4 6 4 8 9
2- 19.9 1.37 0.54 0.780 0.646 0. 9 0. 0 0. 6 0.2 4 0.8 6 0.8 5
Propanol 2 72 6 7 8 4 8 3 4
1-Butanol 19.9 1.37 0.58 0.750 0.647 0. 6 0. 4 0. 8 0.3 3 0.8 0 0.8 8
2 72 6 8 8 4 4 0 3
Ethanol 24.3 1.36 0.65 0.813 0.653 0. 4 0. 4 0. 7 0.4 1 0.6 9 0.8 7
0 10 4 8 7 5 0 5 5
Methanol 32.6 1.32 0.76 0.855 0.650 0. 3 0. 7 0. 4 0.6 0 0.5 8 0.8 3
6 84 2 9 6 6 0 4 5
Ethylene 37.7 1.48 0.79 0.836 0.710 0. 8 0. 6 0. 0 0.7 5 0.5 5 0.9 7
glycol 0 35 0 9 5 9 1 3 3
Water 78.3 1.33 1.00 0.913 0.683 1. 0 0. 2 1. 2 1.0 7 0.0 4 0.9 1
6 30 0 1 4 0 6 2 6
a 7 7 9 2 5 2
Solvents are listed in order of increasing E N values.
T
b
From literature [46,47].
Table 2 Absorption maxima of Schiff bases in different solvents
Sb-
Sb- O
Sb- Sb- Sb- Sb-
O C
CH3 C2H5 CF3 NO2
H H
Solve
3
nta A A A A A A A A A A A A
b
1 2 1 2 1 2 1 2 1 2 1 2

(n (n (n (n (n (n (n (n (n (n (n (n
m) m) m) m) m) m) m) m) m) m) m) m)
n-
275 339 275 338 260 340 259 342 272 353 266 383
Hexane
Cyclohe
276 339 275 340 260 341 261 344 272 355 266 385
xane
1,4-
277 341 279 341 256 344 262 347 276 361 268 393
Dioxane
THF 279 342 280 344 256 348 264 348 276 363 270 402
Ethyl
277 340 278 341 - 344 - 346 - 347 267 397
acetate
Chlorofo
279 337 280 339 259 346 263 343 - 344 270 398
rm
DCM 273 335 279 335 256 341 262 341 - 343 271 392
DMF 282 344 281 346 - 347 - 351 278 364 270 407
tert-
283 345 280 346 256 346 259 350 279 368 269 407
Butanol
DMSO 284 346 282 347 - 348 - 352 281 366 271 410
Acetonitr
275 334 277 338 256 337 256 340 273 357 267 392
ile
2-
280 340 278 343 258 345 257 347 280 367 269 404
Propanol
1-
281 340 278 344 260 345 259 349 280 367 270 405
Butanol
Ethanol 280 340 277 343 259 344 258 346 278 364 271 402
Methano
278 341 277 342 261 344 260 346 279 361 268 399
l
Ethylene
278 342 277 343 262 344 260 346 282 363 272 401
glycol
Water 278 342 277 343 262 346 260 354 278 358 272 395
a
Solvents are listed in order of increasing E TN values.
b
A1 and A2 indicate the two excitation wavelengths.
Table 3 Fluorescence maxima ( Fmax ) of Schiff bases in different solvents
obtained from Gaussian fitted Curves
(A2 is excitation wavelength.)
Sb-
Sb-
O Sb- Sb- Sb- Sb-
O
C CH3 C2H5 CF3 NO2d
Solventa H
H3
Fb
F (nm) F (nm) F (nm) F (nm) F (nm)
(nm)
n-Hexane 399 401 402 402 401 NF
( (6 ( (6 (
Cyclohexane 4006 4046) 4001 4061) 3982 NF
( (2 ( (6 (
1,4-Dioxane 4059 40751 4080 4107) 4246 NF
)1( 1 3)
c
(5 8( (9 (
THF 4062 8 40804) 4102 2) 4129) 4260
2) NF
6( (5 6( (8 (
4058)8 40738) 4093)7 4123) 4219
Ethyl acetate 4( 2( 3) NF
4 (5) (1 (
Chloroform 4018) 41320 4150 2) 41551 4305
1) NF
7( (4 7( (2 (
DCM 4126 6) 41310
) 415 9
2) )
42001 4241
8) NF
9( (4 4( (6 (
DMF 4090 )7 41326) 4142 2) )
4238) 4341) NF
9( (4 9( (8 (
4108)0 41118) 4121)5 4144) 4316
tert-Butanol 7( 8( 1) NF
1 (4) (7 (
DMSO 4157) 41751 4179 2) 4314) 4397
2) NF
4( (4 8( (4 (
Acetonitrile 4118 7) 41130
) 415 3
1) 4210) 4255
3) NF
2( (5 3( (7 (
2-Propanol 4123 )8 41323) 4169 3) 4181) 4343
3) NF
0( (4 3( (7 (
4147)5 41452) 4172)3 4206) 4364
1-Butanol 6( 0( 2) NF
9 (4) (4 (
Ethanol 4148) 41434 4170 1) 4228) 4297
2) NF
7( (4 2( (1 (
4168)3 41749) 4221)8 42816 4345
Methanol 9( 6( 6) NF
(4 (2 (
Ethylene 4210 )7 42640) 4259 1) )
42924 4384
2) NF
glycol 7( (9 5( (3 (
4522)4 4407)) 4418)7 )
4445) 4497
Water 6( 6( 3) NF
9) (1 9) (1 8(
a
Solvents 32are listed in order7)of increasing E TN3) values 4) 4)
6) 9 8
b
F indicates
) the fluorescence wavelength. ) )
c
Fluorescence intensities are given in parenthesis.
d
Not used in dipole moment calculations due to non-fluorescent nature.
Table 4 Correlation coefficients for the fits of  a   f and  a  f of Schiff bases
against the solvent polarity functions f (,n),  (,n) and E TN
Correlation coefficients
Molecule Protic solvents Aprotic solvents
f (,n)  (,n) E TN f(,n)  (,n) E TN
Sb-OH 0.901 0.901 0.940 0.900 0.905 0.907
Sb-OCH3 0.908 0.907 0.908 0.902 0.901 0.903
Sb-CH3 0.904 0.931 0.983 0.897 0.903 0.905
Sb-C2H5 0.916 0.906 0.960 0.894 0.901 0.906
Sb-CF3 0.910 0.905 0.944 0.903 0.896 0.905
Table 5 Ratio of dipole moments (eg) obtained from the solvatochromic method based
on bulk polarity parameters for protic and aprotic solvents
Protic solvents Aprotic solvents
Molecule g 
a
Slope (cm )-1
Dipole moment (D) Slope (cm-1) Dipole moment (D)
m1 b
-m2 b
eg c
e d
m1 b
-m2b
egc ed
Sb-OH 1.76 7406 4238 3.31 5.83 878 1953 2.63 4.64
Sb-OCH3 1.80 7335 3802 3.15 5.69 535 1223 2.56 4.61
Sb-CH3 2.24 6565 2823 2.51 5.53 522 1469 2.10 4.64
Sb-C2H5 1.04 6533 4398 5.12 5.35 1015 1656 4.16 4.35
Sb-CF3 3.86 4370 1433 1.98 7.62 622 2512 1.66 6.40

a
Ground state dipole moment calculated theoretically using PM3 method [23].
b
Calculated from the plots of Stokes-shifts vs. f ( , n) and  ( , n) , E TN (Eqs. 1 and 2).
c
Calculated using ratio of m1 and m2, μ e   m1  m 2  (Eq. 6).
μ g  m 2  m1 
d
Calculated using g.

18
Table 6 Change in dipole moment () obtained from the solvatochromic method based on
microscopic polarity parameters for protic and aprotic solvents

Cavity Protic solvents Aprotic solvents


Molecule ga radiusb  c
ed
c ed
a(Å) mc (cm-1) mc (cm-1)
D D D D
Sb-OH 1.76 5.148 3803 3.95 5.71 1990 2.86 4.62

Sb-OCH3 1.80 5.704 3551 3.84 5.64 1273 2.66 4.47

Sb-CH3 2.24 5.199 2585 3.30 5.51 1552 2.56 4.77

Sb-C2H5 1.04 5.797 2711 3.98 5.03 1712 3.17 4.21

Sb-CF3 3.86 5.322 3198 3.56 7.42 1995 2.81 6.67

a
Ground state dipole moment calculated theoretically using PM3 method [23].
b
Onsager cavity radius (40% of the longest axis in the optimized geometry of the molecule) obtained by using
semiempirical calculations [40,41].
c
Calculated from the plots of spectral shifts,  a   f vs. E TN (Eq. 7).
.dCalculated using g.

19
Table 7 Multiparametric correlation of Kamlet-Taft ( ν  ν o  s * π *  a α α b ββ ) and
Catalán ( ν  ν o  sSPP SPP  a SASA  b SBSB ) treatment of spectral properties in the
excitation process of Schiff bases
Kamlet-Taft parameters Catalán parameters
Molecule a
vo s* a b R vo SSPP aSA bSB Ra
Sb-OH 29676 -113 159 -596 0.911 29640 -22 -192 -513 0.917
Sb-OCH3 29632 -89 112 -763 0.944 29784 -254 -172 -627 0.937

Sb-CH3 29368 -244 80 -396 0.930 29649 -316 -55 -561 0.901

Sb-C2H5 29318 -323 42 -589 0.938 29654 -689 -152 -387 0.953

Sb-CF3 28583 -332 0.70 -1817 0.893 28236 -721 -706 -1728 0.902

Sb-NO2 25961 -489 195 -1300 0.911 27167 -1954 96 -1007 0.934

a
Correlation coefficients.

20
Table 8 Multiparametric correlation of Kamlet-Taft ( ν  ν o  s * π *  a α α b ββ ) and
Catalán ( ν  ν o  sSPP SPP  a SASA  b SBSB ) treatment of spectral properties in
the emission process of Schiff bases
Kamlet-Taft Parameters Catalán Parameters
Molecule a
vo s* a b R vo SSPP aSA bSB Ra
Sb-OH 25266 -1248 -1248 181 0.980 26166 -2082 -1509 405 0.941
Sb-OCH3 24973 -1134 -1134 301 0.959 26128 -1261 -773 84 0.948

Sb-CH3 24961 -1133 -1133 190 0.981 26038 -2237 -1063 331 0.984

Sb-C2H5 24936 -1570 -1570 94 0.958 27209 -2829 -879 107 0.931

Sb-CF3 25843 -1583 -1583 -457 0.976 27007 -4060 -575 -146 0.986

Sb-NO2 - - - - - - - - - -

a
Correlation coefficients.

21
(a)


N



HO

Molecules R Abbreviation
2-isopropyl-5-methyl-4-(4-hydroxybenzylideneamino)phenol OH Sb-OH
2-isopropyl-5-methyl-4-(4-methoxybenzylideneamino)phenol OCH3 Sb-OCH3
2-isopropyl-5-methyl-4-(4-methylbenzylideneamino)phenol CH3 Sb-CH3
2-isopropyl-5-methyl-4-(4-ethylbenzylideneamino)phenol C2H5 Sb-C2H5
2-isopropyl-5-methyl-4-(4-(trifluoromethyl)benzylideneamino)phenol CF3 Sb-CF3
2-isopropyl-5-methyl-4-(4-nitrobenzylideneamino)phenol NO2 Sb-NO2

(b)

Fig. 1

22
0.8 n-Hexane
(a) Cyclohexane
0.7 1,4-Dioxane
THF
0.6 Ethylacetate
Chloroform

Absorbance
DCM
0.5 DMF
tert-Butanol
0.4 DMSO
Acetonitrile
0.3 2-Propanol
1-Butanol
0.2 Ethanol
Methanol
0.1 Ethyleneglycol
Water
0
230 280 330 380 430 480 530 580
Wavelength (nm)
1000 n-Hexane
(b) Cyclohexane
900 1,4-Dioxane
Fluorescence Intensity

800 THF
Ethylacetate
700 Chloroform
DCM
600 DMF
tert-Butanol
500
DMSO
400 Acetonitrile
2-Propanol
300 1-Butanol
Ethanol
200 Methanol
100 Ethyleneglycol
Water
0
350 400 450 500 550 600
Wavelength (nm)
1000 n-Hexane
Cyclohexane
(c) 1,4-Dioxane
THF
800
Fluorescence Intensity

Ethylacetate
Chloroform
DCM
600 DMF
tert-Butanol
DMSO
Acetonitrile
400 2-Propanol
1-Butanol
Ethanol
200 Methanol
Ethyleneglycol
Water
0
350 400 450 500 550 600
Wavelength (nm)
Fig. 2

23
0.8 n-Hexane
(a) Cyclohexane
0.7 1,4-Dioxane
THF
0.6 Ethylacetate
Chloroform
DCM
Absorbance
0.5
DMF
tert-Butanol
0.4 DMSO
Acetonitrile
0.3 2-Propanol
1-Butanol
0.2 Ethanol
Methanol
0.1 Ethyleneglycol
Water
0
230 280 330 380 430 480 530 580
Wavelength (nm)

70 n-Hexane
(b) Cyclohexane
1,4-Dioxane
60
THF
Fluorescence Intensity

Ethylacetate
50 Chloroform
DCM
DMF
40
tert-Butanol
DMSO
30 Acetonitrile
2-Propanol
20 1-Butanol
Ethanol
Methanol
10 Ethyleneglycol
Water
0
350 400 450 500 550 600
Wavelength (nm)

Fig. 3

24
R R

N N

HO HO
R = OH, OCH3, CH3, C2H5
RS-I

R R

N N

HO HO
R = CF3, NO2

RS-II

Fig. 4

25
(a) (b)

LUMO+1

LUMO

HOMO

HOMO-1

Fig. 5

26
(a) (b)

(c) (d)

(e) (f)

Fig. 6

27
a) Protic Solvents (b) Aprotic Solvents

7000 7000

a  f / cm 1
a  f / cm 1

6000 6000

5000 5000

4000 y = 7903.6x - 941.35 4000 y = 878.41x + 4577.9


R = 0.901 R = 0.900
3000 3000
0.6 0.7 0.8 0.9 1.0 0.0 0.2 0.4 0.6 0.8 1.0

f (,n) f (,n)

56000 55000
a + f)/ cm-1
a + f)/ cm-1

55000
54000
54000
53000
53000

y = -1953.1x + 55754 52000 y = -4237.9x + 58803


52000
R = 0.905 R = 0.901
51000 51000
0.5 0.7 0.9 1.1 1.3 1.5 1.2 1.25 1.3 1.35 1.4 1.45

 (,n)  (,n)

8000
a  f / cm 1

6000
1
a  f / cm

7000

6000 5000

5000
4000
4000 y = 3802.5x + 2962.5 y = 1989.6x + 4498.2
R = 0.940 R = 0.907
3000 3000
0.3 0.5 0.7 0.9 1.1 0 0.1 0.2 0.3 0.4 0.5

E TN E TN

Fig. 7

28
Z
S R H S
H
Y
N

X
S HO

Fig. 8

29

You might also like