solvatochromic effect

Upload: parijat-banerjee

Post on 14-Apr-2018

223 views

Category:

Documents


0 download

TRANSCRIPT

  • 7/30/2019 Solvatochromic effect

    1/17

    Solvent influence on absorption and fluorescence spectra of

    merocyanine dyes: a theoretical and experimental study

    I. Baraldi, G. Brancolini1, F. Momicchioli*, G. Ponterini, D. Vanossi

    Dipartimento di Chimica, Universitaa di Modena e Reggio Emilia, Via Campi 183, I-41100 Modena, Italy

    Received 8 November 2002

    Abstract

    The solvatonCS INDO model, previously successfully used to describe the solvatochromic properties of merocya-

    nines, has been extended to the study of the solvent influence on the fluorescence spectra (fluorosolvatochromism) of these

    dyes. A ketocyanine (M1) and a stilbazolium betaine (M2) were chosen as representatives of positively and negatively

    solvatochromic behaviours, respectively. The gap of experimental knowledge concerning the emission properties of M2

    was filled by a spectrofluorometric analysis in a set of solvents covering a large range of the ET30 scale. Solvato- andfluorosolvatochromism were described by calculating the S0eq: ! S1FranckCondon and S1eq: ! S0FranckCondon

    transition energies as a function of a polarity factor related to the static dielectric constant of the solvent, and

    ranging from 0 to 1. The absorbing S0eq: and emitting S1eq: units (solute molecule + solvent cage) were approximatedusing the S0 and S1 geometries of the unsolvated molecule and the respective charge distributions fitted to the current value

    ofke. The calculation results fully confirm that S0 and S1 states of merocyanines can be viewed as a mixture of a neutraland a zwitterionic structure whose composition is controlled by the solvent polarity. The plots of the calculated spectral

    data (absorption and emission maxima and corresponding Stokes shifts) vs ke are in fairly good agreement with those ofthe experimental data over almost the entire range of the normalized ENT values, thus showing that specific solvent in-

    teractions are at least partly simulated within the solvatonCS INDO scheme. The methodological prerequisites for a

    correct prediction of solvatochromic shifts are recalled with reference to previous conflicting theoretical interpretations.

    2003 Published by Elsevier Science B.V.

    1. Introduction

    We have recently shown [1,2] that solvent effects

    on both ground-state properties and absorption

    spectra of classic donoracceptor dyes, such as

    merocyanines, can be fairly well accounted forwithin the CS INDO scheme [3]. Briefly, the sol-

    utesolvent interactions were described by the

    simple solvaton model [4] and were incorporated

    in the CS INDO Hamiltonian according to previ-

    ous basically equivalent all valence electron SCF

    approaches [511]. Our procedure, however, is

    characterized by a peculiar modelling of the solv-

    aton set representing the polarized solvent sur-

    rounding the solute. We followed the basic widely

    accepted idea that the electronic structure of

    Chemical Physics 288 (2003) 309325

    www.elsevier.com/locate/chemphys

    * Corresponding author. Tel.: +39-59-2055081; fax: +39-59-

    373543.

    E-mail address: [email protected] (F. Momicch-

    ioli).1 Present address: Dipartimento di Chimica G.Ciamician,

    Universitaa di Bologna, Via F.Selmi 2, I-40126 Bologna, Italy.

    0301-0104/03/$ - see front matter 2003 Published by Elsevier Science B.V.

    doi:10.1016/S0301-0104(03)00046-6

    http://mail%20to:%[email protected]/http://mail%20to:%[email protected]/
  • 7/30/2019 Solvatochromic effect

    2/17

    merocyanines can be described at the p level in

    terms of resonance between neutral and charge-

    separated forms (Fig. 1) and that the solvato-

    chromic behaviour can be traced back to the rel-ative weights of the two structures in the ground

    state and their change upon vertical transition to

    the electronic excited state [12,13]. Such a scheme

    stresses the key role of the p-electron distribution

    and suggests that an effective solvent field must be

    first of all capable of correctly controlling the drift

    of p-electrons from the donor R2N to the ac-ceptor (CO) group. A very elementary way to

    simulate a polarized environment is to position

    one or more point charges at one or either end of

    the chromophore system, as some authors did by

    the middle of the 1990s [14,15]. In a more realistic

    way, following the solvaton model [4] we associ-

    ated with each atom of the conjugated system,

    with p net charge Qp, a fictive particle with charge

    Qp interacting with all electron and core charges

    of the solute according to Borns law. The so-de-

    fined solvaton set reflects the composition of the

    resonance hybrid (Fig. 1), depending on the nature

    of the donor and acceptor groups, and hence may

    effectively account for the p-electron redistribution

    induced by the solvent polarity. Using such a

    solvaton set within the CS INDO CI scheme [1,2],we were able to provide a satisfactory description

    of both the positive solvatochromism of two vi-

    nylogous streptomerocyanines (Fig. 1, n 2; 4)and the large negative solvatochromism of stil-

    bazolium betaine [13] (Fig. 2, M2).

    Reasonable structural variations with solvent

    polarity, related to variations of the resonance

    hybrid composition, were also predicted. The same

    twofold problem had previously been addressed by

    Albert et al. [16] using the self-consistent reaction

    field (SCRF) model within the INDO method, butno choice of the cavity-size parameter had yielded

    a reasonable prediction for the two opposite sol-

    vatochromic behaviours (for more details see

    [1,2]). To our knowledge, the two solvatochromic

    trends were qualitatively well reproduced only by

    Klamt [17] using the AM1/COSMO method.

    Other theoretical studies introducing the solvent

    dielectric field through either a set of point charges

    [14,15] or the virtual charge model [10] dealt only

    with stilbazolium betaine which has attracted great

    attention in relation to its uncommonly large

    negative solvatochromism and the much-discussed

    solvatochromic reversal at low medium polarity

    [18]. Independently of the specific (continuum)solvent model, the majority of the cited theoretical

    studies [1,2,10,15,17] reproduced qualitatively the

    negative solvatochromism of M2. On the other

    hand, Morley [14] predicted the opposite trend

    combining AM1 structure optimization in the

    presence of the solvent and gas phase CNDOVS

    calculation of the transition energy for the solvent-

    distorted molecular geometry. This result was

    interpreted by Morley as evidence that the zwit-

    terionic (benzenoid) form, obtained in the polar

    medium, absorbs at the red of the neutral (qui-nonoid) form prevailing in non-polar medium.

    Such interpretation, however, is questionable

    since, no matter what geometry is used, the elec-

    tronic structure yielded by an MO calculation

    corresponds to a mixture of VB structures. The

    zwitterionic form exists only in the presence of a

    polar medium and its electronic spectrum can be

    calculated only using the solvent-polarized MOs.

    As a matter of fact, all calculations complying with

    this condition [1,2,10,15,17] predicted solvato-Fig. 1. Neutral and charge-separated mesomeric structures of

    simple streptopolymethine merocyanines.

    Fig. 2. Investigated compounds. M1: 1,9-di-(N-phenyl-N-me-

    thyl)-4,6,dimethylene-nona-1,3,6,8-tetraen-5-one; M2: 40-hy-droxy-1-methylstilbazolium betaine.

    310 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    3/17

    chromic shifts in qualitative agreement with ex-

    periment. 2 Qualitatively equivalent results were

    obtained by Benson and Murrell [19] within an

    SCF p-electron treatment where the effect of sol-vent was simulated by suitable choice of one-cen-

    tre core and electron repulsion integrals for the

    nitrogen and oxygen atoms.

    From the above brief survey, our solvatonCS

    INDO scheme appears to have two main advan-

    tages: (i) both ground and excited-state properties

    are calculated for the solute molecule embedded in

    the solvent, (ii) the use of a solvaton set reflecting

    the net p-electron charges enables the solvent in-

    teraction to be modelled in keeping with the VB

    description of the electronic structure of the solute.

    A disadvantage, shared by all continuum models,

    is that it formally leaves out specific solvent in-

    teractions. In principle, such effects can be ac-

    counted for only within semicontinuum type

    theories [20,21] or fully discrete type approaches

    as, for example, those based on statistical me-

    chanics techniques [2224]. However, our very

    simple scheme, where the solvent interaction is

    introduced essentially through a variation of the

    diagonal elements of the Fock matrix corre-

    sponding to the AOs of the p system, may im-

    plicitly account for specific solvent interactions asfirst argued by Benson and Murrell [19].

    In the present work, the solvatonCS INDO

    scheme was subjected to further validation by

    studying the solvent effects on the fluorescence

    spectra (fluorosolvatochromism) 3 which have to

    date received relatively little attention from a theo-

    retical point of view. As test compounds we chose

    two merocyanines, a ketocyanine dye (M1) and

    stilbazolium betaine (M2) (Fig. 2), and the solvent-

    induced spectral shifts of both absorption and flu-

    orescence emission were investigated. Abundantexperimental data concerning the absorption spec-

    tra of these dyes in solvents of different polarities are

    available ([25], and references cited therein). On the

    other hand, to our knowledge solvent effects on the

    fluorescence spectra have been reported only for M1[26,27]. Thus, we first of all carried out an experi-

    mental exploration of the absorption and emission

    solvatochromism of M2. In summary, the experi-

    mental data as a whole show that: (i) M1 exhibits

    strong positive solvatochromism in both absorption

    and emission [2527], (ii) M2, the absorption of

    which is characterized by one of the strongest neg-

    ative solvatochromisms ever observed, features a

    markedly weaker negative fluorosolvatochromism.

    The theoretical interpretation of the entire body

    of experimental evidence was undertaken by ap-

    plying the solvatonCS INDO method within a

    usual scheme where the fluorescence emission takes

    place from the equilibrium geometry of the lowest

    excited singlet state reached very quickly after ver-

    tical S0 ! S1 excitation of the equilibrium groundstate. In principle, this should require geometry

    optimization of the solvatedsolute molecule in both

    the ground state S0 and the emitting S1 (pHpL;

    H HOMO, L LUMO) state. In practice, wesimply used S0 and S1 geometries optimized in the

    gas-phase approximation and calculated the S0

    eq:

    ! S1Franck

    Condon and S1eq: ! S0Franck

    Condon transition energies as functions of thesolvent polarity using solvaton sets reflecting the p

    net charges of S0eq: and S1eq:, respectively. Thecalculation results will first be thoroughly analysed

    by reference to the basic characteristics of the the-

    oretical model and will then be subjected to a de-

    tailed comparison with experimental observations

    in solvents covering the whole scale of solvation

    power. It will be shown that both the solvatochro-

    mic and fluorosolvatochromic behaviours of M1

    and M2 are qualitatively well described within thesolvatonCS INDO scheme.

    2. Experimental investigation on dye M2

    2.1. Materials, instrumentation and details of

    experiments

    M2 (4-[(1-methyl-4(1H)-pyridinylidene)ethylid-

    ene]-2,5-cyclohexadien-1-one) was purchased from

    2 A calculation procedure like the Morley one was applied in

    [16] where INDOSCRF optimized geometries were used in gas

    phase INDO/S type calculations of the spectra. This may

    explain the positive solvatochromism erroneously predicted for

    M2 in [16] when using physically reliable values of the cavity

    radius.3 Hereafter, the solvent dependence of the position of the

    absorption and fluorescence bands will be termed solvatochro-

    mism and fluorosolvatochromism, respectively [25].

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 311

  • 7/30/2019 Solvatochromic effect

    4/17

    Aldrich and was used as received. All solvents

    (Merck and Lab-Scan) were of spectroscopic grade

    and were dehydrated with activated molecular

    sieves before use. The polar and hygroscopic oneswere treated with solid KOH so as to dissolve the

    M2 betaine in the unprotonated form. For the

    same reason, measurements in water were carried

    out in 102 M NaOH.Absorption spectra were recorded on a Perkin

    Elmer k15 spectrophotometer, while a Spex-Jobin

    Yvon Fluoromax 2 spectrofluorometer was em-

    ployed for the fluorescence measurements. All

    experiments were carried out at room temperature

    (1821 C). Maximum optical densities were,

    typically, between 0.1 and 1 (corresponding to

    sample concentrations from 3 106 to 3 105 mol dm3) for absorption measurements,and around 0.10.15 for fluorescence measure-

    ments. In the latter, each sample was excited at, at

    least, two different wavelengths on the high-energy

    side of the absorption band. Emission and exci-

    tation spectra were corrected for the instrumental

    spectral response. Due to the very weak fluores-

    cence emission of M2, especially in low-polarity

    solvents, wide monochromator slits were used

    (68 nm spectral resolution) to improve the signal-

    to-noise ratio. Fluorescence quantum yields UFwere determined in methanol and water with re-

    spect to cresyl violet in methanol (UF 0:65 [28])and eosin in methanol (UF 0:60 [29]) accordingto the usual expression: UF UF;rA=Arn2=n2r ODr=OD, where r refers to the reference, A arethe areas of the corrected emission bands, n are the

    solvent refractive indexes and the optical-density

    (OD) ratios at the excitation wavelengths were

    adjusted to unity. Inner filter effects were deemed

    negligible on both emission spectra and fluores-

    cence quantum yields because of the optical thin-ness of the samples employed and of the low to

    very low spectral overlap between absorption and

    emission.

    2.2. Results and discussion

    The choice of test merocyanine dyes suitable to

    check thoroughly the capability of the solvaton

    CS INDO method of accounting for solvent effects

    on both absorption and fluorescence spectra was

    not a trivial affair. First, in compliance with the

    previous studies limited to effects on the absorp-

    tion spectra [1,2], we needed two compounds

    characterized by opposite solvatochromisms. In[1,2] simple streptomerocyanines (Fig. 1) and stil-

    bazolium betaine (M2 in Fig. 2) were taken as

    typical dyes with strong positive and negative

    solvatochromism, respectively. In both cases the

    intense colour band is due to the lowest p ! p(essentially pH ! pL) transition and the observedsolvent shifts are not affected by the presence of

    any forbidden np state at low energies. On theother hand, in order that the study may be ex-

    tended to fluorosolvatochromism the test com-

    pounds should be characterized by an efficiently

    emitting lowest-lying pHpL state. Previous theo-retical and photophysical studies [3032] have

    pointed out that in general streptomerocyanine

    derivatives have no such characteristics because of

    the presence of lowest-lying np states and theoccurrence of fast transcis isomerization follow-

    ing direct S0 ! S1pHpL excitation. Thus, in thepresent work we replaced the streptomerocyanines

    (Fig. 1) with the ketocyanine M1 (Fig. 2) which is

    known to be an efficient fluorophore exhibiting

    large red shifts in both absorption and emission

    spectra upon increasing the solvent polarity [25].In this case, the theoretical study (Section 3) was

    based on the abundant experimental data reported

    in the literature [2527].

    The suitability of stilbazolium betaine M2 as

    test compound for negative solvatochromism in

    both absorption and fluorescence required on the

    contrary further experimental investigation. As is

    well known, dye M2 exhibits a very large negative

    solvatochromism in absorption ($6500 cm1 ongoing from CHCl3 to H2O solution) [25]. How-

    ever, the absorption spectrum of M2 shows achange in shape on moving from low-polarity

    solvents, where a pronounced structure is ob-

    served, to highly polar and protic solvents, where

    the structure is blurred out. For merocyanines

    closely related to M2, the structured spectra ob-

    served in different solvents of relatively low po-

    larity were attributed to the presence in solution of

    two variously identified species ([33], and refer-

    ences cited therein). Similar observations had led

    other authors to question the reliability of any

    312 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    5/17

    analysis of the solvatochromism of these dyes [34].

    In order to check the importance of any extra

    contribution to the absorption of M2 in solvents of

    low polarity, we carried out a spectrofluorometricanalysis in a 1:1 dichloromethanetoluene mixture.

    The emission and excitation spectra showed only a

    very modest dependence on, respectively, the ex-

    citation and monitoring wavelength. Representa-

    tive spectra are shown in Fig. 3. A close

    correspondence was observed between the ab-

    sorption and excitation spectra, apart from a weak

    solvent Raman band (around 560 nm in Fig. 3) ca.

    3000 cm1 from the monitoring wavenumber, andsome blurring of the vibronic structure possibly

    due to poor instrumental spectral resolution (8

    nm). We conclude that the absorption and emis-

    sion spectra of M2 in low-polarity solvents are due

    to essentially one and the same species. Moreover,

    the rather small Stokes shift and the mirror image

    relationship between the absorption and emission

    peaks suggest that fluorescence emission takes

    place from the local minimum of the initially ex-

    cited 1pHpL state. This is in keeping with therecently reported femtosecond dynamics of dye

    M2 [35]. According to [35], direct S0 ! S1 excita-

    tion of unprotonated stilbazolium betaine M2

    does not result in transcis isomerization but it is

    followed by fast thermally activated relaxation (1.1

    ps) to a non-emitting conformational intermediate.Thus, although emission is predicted to occur from

    the local minimum of the S1pHpL state, the flu-orescence quantum yield should be very small. As

    a matter of fact, we found the fluorescence quan-

    tum yield of M2 to be only 2:0 103 in methanol,1:5 103 in water and even lower values weremeasured in solvents of low polarity. From the

    foregoing considerations, it ensues that M2, even

    though weakly fluorescent, may be used to inves-

    tigate solvatochromism in both absorption and

    emission. For this purpose, the absorption and

    fluorescence emission maxima of M2 were mea-

    sured in nine solvents with ET30 solvent po-larity parameter [25] ranging from 39.1

    (chloroform) to 63.1 (water), and the results are

    reported in Table 1. Both absorption and emission

    spectra exhibit a negative solvatochromism, and

    shift quite regularly to the blue with increasing the

    value of ET30 which represents the overall sol-vation power of the solvents. On the other hand,

    the static dielectric constant e, reflecting only

    Fig. 3. Absorption (full line), fluorescence excitation (dashed line, kem 660 nm) and emission (dotted line, kexc 580 nm) spectra ofM2 in a 1:1 dichloromethanetoluene mixture.

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 313

  • 7/30/2019 Solvatochromic effect

    6/17

    non-specific solvent effects, fails to account for

    the quite different band shifts induced by aprotic

    and protic solvents of comparable polarity. The

    maximum solvatochromic shift found for the

    absorption, Dmabsmax mchloroform mwater 6260 cm1, is in agreement with data of otherauthors ([25], and references cited therein). Much

    smaller hypsochromic shifts were observed foremission, Dmflumax 1340 cm1. We note that aqualitatively similar, yet less pronounced, solvent

    dependence of the absorption and fluorescence

    spectra has recently been reported for another

    negatively solvatochromic merocyanine (MC 540)

    [36].

    3. Theoretical investigation

    3.1. Method and calculation details

    Incorporation of the solvaton model into the

    CS INDO method has already been described in

    detail [1,2]. Here, we only summarize the main

    points concerning the calculation of the solvato-

    chromic shifts. Briefly, the Fock matrix general

    element is expressed as

    Fslm F0lm dlmke

    XN

    B0QB0cAB0 l 2 A; 1

    where the first term F0lm is the CS INDO matrixelement in the absence of solvent [3], while the

    second term accounts for the electronsolvaton

    interaction effects. In particular, ke is a solventpolarity factor related to the static dielectric con-

    stant e. In keeping with Costanciel and Tapias

    virtual charge model [9], where the polarity factor

    is set equal to ffiffiep 1= ffiffiep, we assumed ke torange from 0 (when e is 1) to 1 (for e ! 1). QB0 isthe charge of the solvaton associated with atom B

    and cAB0 is taken equal to the electron repulsion

    integral cAB. As pointed out in Section 1 (see also

    [1,2]), solvatons were only associated with the at-

    oms contributing to the p system (C, N, O) and

    their charges were taken equal to the negative of

    the respective net p-electron charges.

    The determination of the solvent effects on the

    position of the first absorption bands required the

    HartreeFock equations incorporating solutesolvaton interactions to be resolved for the

    equilibrium ground-state geometries and the cor-

    responding vertical S0 ! S1pHpL transitions tobe calculated at different values of ke. As sug-gested by the experimental observations (Section

    2.2) fluorescence emission of both M1 and M2 dyes

    takes place from the excited S1pHpL state aftergeometrical relaxation of the solute and reorgani-

    zation of the solvent molecules in the solvation

    shell. According to this scheme, the fluorescence

    Table 1

    Absorption and emission maxima of M2 in solvents of different characteristics

    Solvent ET30a (kcal/mol) ENT b ec Absorptiond, mmax cm1 Fluorescenced, mmax cm1

    Chloroform 39.1 0.259 4.8 16,310 15,850Dichloromethane 40.7 0.309 8.9 16,470 15,820

    Acetone 42.2 0.355 20.7 17,040 16,130

    DMFe 43.2 0.386 37.0 17,150 16,140

    DMSOf 45.1 0.444 46.7 17,390 16,160

    Acetonitrile 45.6 0.460 37.5 17,590 16,260

    2-Propanol 50.7 0.546 19.9 18,450 16,700

    Methanol 55.4 0.762 32.7 20,700 17,120

    Water 63.1 1.000 78.4 22,570 17,190

    Dmmax cm1 )6260 )1340a Solvent polarity parameter of Dimroth and Reichardt measured at 25 C and 1 bar [25].b Normalized ENT value.c Dielectric constant at room temperature (20 or 25 C) [13].d Typical uncertainties: 30 cm1 on absorption, 60 cm1 on emission.e Dimethylformamide.fDimethyl sulfoxide.

    314 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    7/17

    transitions were calculated at different ke valuesusing geometries and solvaton charges optimized

    for the excited state. For the sake of simplicity, in

    the present work geometrical distortions inducedby the solvaton field on the solute molecule were

    disregarded. In other words, we simply used ge-

    ometries optimized in the absence of solvent

    ke 0. This simplification appears to be ac-ceptable, since solvatochromic shifts are mainly

    determined by polarization and electrostatic effects

    that are not very sensitive to small geometry

    modifications. However, the solvaton charges cor-

    responding to the equilibrium geometries of the

    ground and the excited states were adjusted by an

    iterative procedure, i.e., by performing a sequence

    of SCF (for S0) and SCF CI (for S1) calculations

    until convergence of the QB0 values.

    To summarize, we carried out the following set

    of calculations:

    (i) The S0-state equilibrium geometries of both

    dyes were calculated by minimizing the total CS

    INDO SCF energy of the isolated molecule, as a

    function of the internal coordinates of the chro-

    mophoric group. With these geometries, SCF cal-

    culations were carried out, for different ke values,using solvaton charges corresponding to the S0 net

    p-electron charges. The MOs produced at eachke value by the iteratively adjusted solvaton setwere then used for the CI calculation of the ab-

    sorption spectrum, thus determining the absorp-

    tion spectral shifts.

    (ii) An analogous procedure was carried out to

    calculate fluorosolvatochromic shifts. First, for

    each molecule, the equilibrium geometry of the

    S1pHpL state of the isolated molecule was cal-culated by minimizing the excited-state energy

    ES1 ES0 DES0 S1 derived from the CIcalculation as a function of the internal coordi-nates of the chromophoric group. The so obtained

    geometries were used to generate, at different kevalues, optimized solvaton sets corresponding to

    the excited-state net p-electron charges and the

    MOs to be used in the CI calculation of the

    emission spectra and solvent-induced fluorescence

    shifts.

    Since the excited state responsible for both the

    absorption colour band and the fluorescence

    emission, S1pHpL, is fully represented by the

    singly excited 1UpHpL configuration, we limitedourselves to standard S-CI calculations. The MO

    active spaces included all p and p molecular or-

    bitals of the chromophoric groups and, for M1, ofthe benzene rings. In view of the model character

    of this study, the solvaton charges suitable for the

    equilibrium excited state were simply derived from

    the net p-electron charges of the 1UpHpL con-figuration.

    The geometry optimizations were carried out

    introducing in our CS INDO CI programs an

    optimization code where the gradient is calculated

    numerically within a FletcherPowell-type opti-

    mization algorithm [37]. The way the gradient is

    computed makes the program easily adaptable for

    calculations of excited-state geometries.

    The parameters peculiar to the CS INDO

    method [3] were chosen as follows: (i) the screening

    constants, klm, where the indexes refer to two hy-

    brid atomic orbitals of the CS INDO basis set,

    were given the values: krr 1, kpp 0:50,krp 0:65, knp 0:60, knr 0:72, knn 0:68; (ii)the atomic pair parameters, aAB (a.u.) and R

    0AB (

    AA),

    entering the calculation of the core repulsion en-

    ergy, were assigned the values aCC 1:35,aCN

    1:40, aCO

    1:20, R0CC

    1:7, R0CO

    2:10.

    The remaining pair parameters were given theusual values, aAB 1:50 and R0AB 2:00.

    Finally, two-electron repulsion integrals were

    calculated according to OhnoKlopman [38].

    3.2. Results and discussion

    3.2.1. Ground- and excited-state equilibrium geom-

    etries

    After exhaustive test calculations on ground

    and excited state geometries of prototypic systems,

    not reported here for conciseness [39], the CSINDO based approach described in the Section 3.1

    was applied to M1 and M2 in both S0 and S1states. The optimized bond lengths are reported in

    Table 2. The structural differences between the two

    dyes in the ground state and the geometry changes

    induced by the S0 ! S1 excitation can be illus-trated by analysing the extent of bond length al-

    ternation (BLA) in the conjugated bridge

    connecting the donor (amino) and the acceptor

    (carbonyl) groups, which reflects the relative

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 315

  • 7/30/2019 Solvatochromic effect

    8/17

    weights of the neutral polyenic and zwitterionic

    structures in the VB resonance hybrid (Fig. 1). All

    such characteristics are well expressed by the value

    of the so called BLA parameter (last row of Table

    2), defined as the difference between the average

    length of the single bonds and that of thedouble bonds [16]. The single and double

    bonds were identified by reference to the neutral

    VB structures (see Fig. 2), so high positive (nega-

    tive) value of the BLA parameter indicates pre-

    dominance of the neutral (zwitterionic) mesomeric

    structure. A vanishing BLA parameter, on the

    other hand, indicates a cyanine-like structure

    characterized by similar weights of the two forms.

    On this basis, the calculation results of Table 2

    indicate that: (i) in the ground state, the charge-

    separated form is more important for M2 than forM1 BLA M1;S0 > BLA M2; S0, (ii) in bothmerocyanines bond-length alternation decreases

    upon electronic excitation, i.e., the contribution of

    the neutral VB structure becomes lower on moving

    from S0 to S1 BLA M1; S0 > BLA M1;S1;BLA M2;S0 > BLA M2;S1. The differences

    between the two dyes become more evident if theBLA parameters of M2 are calculated in terms of

    the bonds (d, e, f) of the central polymethinic

    fragment. As a matter of fact, in this case BLA is

    already rather small in the ground state (0.017)

    and changes sign in S1, thus reflecting an opposite

    bond alternation with respect to S0. The goodness

    of our theoretical predictions should now be veri-

    fied by comparing the calculated ground-state ge-

    ometries of Table 2 with the experimentally

    determined structures as well as the results of

    previous theoretical studies. However, as far as we

    know, structure determinations and calculations

    have been reported only for dye M2. From the

    crystal structure determination of M2 trihydrate

    [40] the molecule was found to be almost planar

    and to have a zwitterionic (benzenoid) structure

    with an essentially double CC central bond (1.346AA) connected to the six-membered rings by bonds

    of $1.44 AA, and a CO bond (1.304 AA) ratherlonger than a double C@O bond (1.215 AA). As is

    evident, such structure is at variance with the CS

    INDO optimized geometry of the isolated mole-

    cule (Table 2) which corresponds to a mixture ofthe two VB forms, with a slight prevalence of the

    neutral (quinonoid) one (Fig. 4). This is not sur-

    prising since in the crystal the zwitterionic form is

    imposed by the formation of hydrogen bonds be-

    tween the carbonyl end of the chromophore and

    the crystal water. Thus, the crystal structure is

    assimilable to that of M2 in aqueous solution.

    The CS INDO optimized geometry of M2 is in

    qualitative agreement with previous theoretical

    studies [10,14,16,19] in that all calculations predict

    the quinonoid form to be (more or less) prevailingin the isolated molecule. However, with reference

    to the central polymethine fragment (e, d, f CC

    bonds), our geometry exhibits the smallest BLA

    Fig. 4. Neutral (quinonoid) and zwitterionic (benzenoid) forms of dye M2.

    Table 2

    Bond lengths and bond length alternation (BLA) parameters

    for the ground state S0 and the first pp excited state S1 ofdyes M1 and M2

    Bonds M1 M2

    S0 S1 S0 S1

    a 1.374 1.368 1.365 1.362

    b 1.398 1.411 1.396 1.402

    c 1.446 1.430 1.447 1.444

    d 1.398 1.415 1.419 1.427

    e 1.458 1.450 1.436 1.421

    f 1.245 1.253 1.418 1.430

    g 1.455 1.447

    h 1.401 1.406

    i 1.450 1.447

    l 1.250 1.254

    BLA 0.055a

    0.027a

    0.038b

    0.024b

    0.017c )0.008c

    For labelling of the bonds, see Fig. 2.a BLA M1 c e b d=2.b BLA M2 c e g i b d f h=4.c BLA M2 e d f=2.

    316 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    9/17

    value (present work: 0.017; [19]: 0.030; [16]: 0.038;

    [14]: 0.058; [10]: 0.077). In other words, our de-

    scription reflects a resonance hybrid with similar

    weights of the two limiting structures, while theother theoretical predictions favour more decid-

    edly the quinonoid structure. We do not deny that

    our calculation may have underestimated the

    contribution of the quinonoid form, but we can

    say that this result is consistent with the observed

    solvatochromic behaviour of M2, i.e., an initial

    small red shift followed by a large blue shift of the

    first absorption band on going from low polar to

    highly polar solvents [1,2,13,18].

    3.2.2. Absorption and fluorescence solvatochromism

    Using the S0 and S1 optimized geometries, the

    vertical S0eq: ! S1 and S1eq: ! S0 transitionenergies of M1 and M2 were calculated in a me-

    dium of changing polarity (ke 00.9) by the CSINDO CI procedure described in Section 3.1. The

    results DE are collected in Tables 36, togetherwith the oscillator strengths of the transitions (f),

    the electric dipole moments (l), and the net char-

    ges on the nitrogen and oxygen atoms in the

    S0 QN;O and S1 QN;O states.First of all, let us point out some essential

    characteristics of the theoretical description.Tables 3 and 4 show that on going from ke 0(gas phase) to ke 0:9 (highly polar medium),both the absorption and fluorescence maxima of

    M1 are predicted to undergo marked bathochro-

    mic shifts (DmS0 ! S1 3550 cm1 andDmS1 ! S0 5000 cm1) 4, while the oscillatorstrengths are found to be both quite high (>2) andnot very sensitive to solvent polarity. The batho-

    chromic shifts are related to the increase of the

    solute dipole moment on the S0 ! S1 transition,resulting in a solvent-induced stabilization of theexcited state relative to the ground state, which

    increases with increasing solvent polarity. The

    polarization effects induced by the solvaton field

    on the electronic structure of the solute result in a

    quite regular growing of both lS0 and lS1 on going

    from ke 0 to 0.9. Such growing is little higherwhen using the equilibrium geometry of the excited

    state (Table 4), but Dl lS1 lS0 is predicted to

    be almost the same in absorption (Table 3) and

    emission (Table 4) at all values ofke. However, itis worth noting that the permanent dipole moment

    of M1 is directed along the twofold axis containingthe C@O bond (see Fig. 2), while the S0S1 tran-

    sition is polarized perpendicularly to the twofold

    axis 1B, i.e., is associated to charge transfer alongthe longitudinal molecular axis. Thus, the dipole-

    moment change occurring on the S0S1 transition

    is not as effective a parameter of the solvatochro-

    mic effect as in the case of the parent streptopen-

    tamethinemerocyanine [1,2] where both

    permanent and transition dipole moments are

    aligned with the chromophoric chain. As a matter

    of fact, the S0 and S1 dipole moments of the ke-

    tocyanine M1 (and their increase with increasing

    ke) turn out to be decidedly smaller than those ofits pentamethinemerocyanine moiety, while the

    contrary occurs for oscillator strength and solvent

    shift of the S0 ! S1 transition (see [2]). For a moreprecise comparison, we carried out CS INDO SCI

    test calculations for pentamethinemerocyanine

    (Fig. 1; n 2, R Me) and its N-phenyl derivativein the absence of solvent. The calculation results

    (Table 7) indicate that the phenyl substitution

    produces a small red shift and intensity increase of

    the transition, while lS0 and lS1 remain almostunchanged and substantially higher than those of

    the ketocyanine (Table 3, first row). The absorp-

    tion data in dichloromethane, also reported in

    Table 7, corroborate the theoretical prediction

    even if transition energies and spectral shift are a

    little overestimated because of the adopted sim-

    plifications (essentially, singly excited CI and co-

    planarity of the phenyl ring). A more correct

    correlation between the solvatochromic behaviour

    of ketocyanine M1 and that of its parent chro-

    mophore emerges by changing from the dipolemoments (i.e., global electrical properties) to the

    atomic net charges reflecting the local charge dis-

    tributions. As a matter of fact, Table 3 shows that

    the net charges on the nitrogen and oxygen atoms

    in the ground state, as well as their changes due to

    an increase of ke and/or S0 ! S1 excitation,closely resemble those previously found for strep-

    topentamethinemerocyanine [2]. In particular, in

    S0 an increase ofke results in a drift of electronsfrom the two nitrogen atoms (donors) to the4 Dm mnon-polar solvent mpolar solvent.

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 317

  • 7/30/2019 Solvatochromic effect

    10/17

    Table 4

    Calculated S1 ! S0 emission properties of M1 as functions of the solvent polarity factor ke (see caption of Table 3 for the meaning ofsymbols)

    ke DE (eV) f lS0 (D) lS1 (D) QN QO QN QO0.0 3.118 2.411 5.67 8.10 )0.12 )0.73 )0.03 )0.78

    0.2 2.968 2.384 6.44 8.99 )0.08 )0.83 +0.01 )0.87

    0.4 2.823 2.362 7.17 9.81 )0.04 )0.91 +0.05 )0.95

    0.6 2.686 2.345 7.86 10.58 0.00 )0.98 +0.10 )1.01

    0.8 2.558 2.331 8.53 11.30 +0.05 )1.05 +0.14 )1.07

    0.9 2.498 2.326 8.86 11.64 +0.07 )1.08 +0.16 )1.10

    Table 5

    Calculated S0 ! S1 absorption properties of M2 as functions of the solvent polarity factor ke (see caption of Table 3 for the meaningof symbols)

    ke DE (eV) f lS0 (D) lS1 (D) QN QO QN QO0.0 2.782 2.075 17.92 18.36 )0.04 )0.73 +0.01 )0.71

    0.2 2.766 1.904 22.09 21.30 +0.02 )0.84 +0.05 )0.81

    0.4 2.790 1.702 26.88 24.73 +0.09 )0.94 +0.11 )0.91

    0.6 2.944 1.425 33.80 30.58 +0.21 )1.04 +0.22 )1.02

    0.8 3.184 1.234 39.91 37.54 +0.35 )1.13 +0.37 )1.11

    0.9 3.291 1.163 43.12 41.81 +0.45 )1.16 +0.47 )1.15

    Table 3

    Calculated S0 ! S1 absorption properties of M1 as functions of the solvent polarity factor ke: vertical transition energy DE, os-cillator strength (f), ground and excited-state dipole moments lS0, lS1 and net charges on nitrogen QN;QN and oxygen QO;QOatoms

    ke DE (eV) f lS0 D lS1 D QN QO QN QO0.0 3.258 2.378 5.59 8.03 )0.13 )0.72 )0.03 )0.77

    0.2 3.163 2.367 6.11 8.65 )0.10 )0.81 0.00 )0.86

    0.4 3.060 2.353 6.64 9.27 )0.06 )0.91 +0.04 )0.94

    0.6 2.952 2.321 7.10 9.87 )0.01 )1.01 +0.09 )1.03

    0.8 2.855 2.263 7.55 10.47 +0.07 )1.09 +0.16 )1.10

    0.9 2.818 2.223 7.79 10.78 +0.12 )1.13 +0.21 )1.14

    Table 6

    Calculated S1 ! S0 emission properties of M2 as functions of the solvent polarity factor ke (see caption of Table 3 for the meaning ofsymbols)

    ke DE (eV) f lS0 (D) lS1 (D) QN QO QN QO0.0 2.732 2.227 19.37 17.56 )0.02 )0.74 +0.01 )0.70

    0.2 2.725 2.126 21.67 19.51 +0.02 )0.83 +0.05 )0.79

    0.4 2.724 2.007 24.49 21.79 +0.08 )0.92 +0.10 )0.89

    0.6 2.732 1.884 27.56 24.27 +0.15 )1.02 +0.16 )1.00

    0.8 2.750 1.757 30.91 26.98 +0.23 )1.11 +0.23 )1.09

    0.9 2.763 1.699 32.53 28.33 +0.28 )1.15 +0.28 )1.13

    318 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    11/17

    oxygen atom (acceptor), traceable to an increased

    weight of the two equivalent zwitterionic forms

    giving rise to a nonamethinecyanine-like elec-

    tronic structure relative to the neutral form (M1

    in Fig. 1). Moreover, the comparison between QN,QO and QN, QO provides an estimate of the realcharge transfer occurring during the S0 ! S1transition at each ke. Table 4 shows that, in-verting the direction of the charge transfer, similar

    considerations apply to the transition from the

    relaxed excited state to the FranckCondonground state, S1eq: ! S0. In conclusion, the ke-tocyanine M1 provides an example of a system

    where the solvatochromic effects may be described

    within the solvaton model better than by using

    SCRF models adopting the Onsager dipolar ap-

    proximation [16].

    Merocyanine M2 is predicted to have a very

    different solvatochromic behaviour both in ab-

    sorption and in emission (Tables 5 and 6). In going

    from ke 0 to 0.9, the absorption maximum isfound to undergo a small bathochromic shift (untilke 0:2) followed by a large hypsochromic shiftleading to a global strongly negative solvatochro-

    mism Dm ffi 4100 cm1 (Table 5). This behav-iour follows that found in our previous theoretical

    study [1,2], even if the transition energies and the

    solvatochromic shift turn out to be somewhat re-

    duced because of some differences in the calcula-

    tion procedure concerning the parametrization

    (kpp 0:5 instead of 0.55) and the use of fixedgeometries optimized at ke 0. The blue-shift of

    the absorption maximum of M2 with increasing

    ke, is accompanied by a substantial decrease of

    the oscillator strength attributable to the increas-

    ing importance of the zwitterionic form (Fig. 4). In

    this case, where permanent and transition dipole

    moments are both parallel to the longitudinal

    molecular axis, the S0 and S1 dipole moments

    should be fully representative of the solvatochro-

    mic behaviour. Briefly, Table 5 shows that lS0 is

    quite high at k

    e

    0, due to a substantial con-

    tribution of the zwitterionic VB structure, andgrows rapidly with increasing ke owing togrowing solvent-induced polarization. At ke 0,lS1 is slightly higher than lS0 but, from ke 0:2on, the FranckCondon S1 state is 23 D less di-

    polar than S0. Thus, both the small red shift be-

    tween ke 0 and 0.2, and the successive strongblue shift can be rationalized in terms of net sta-

    bilization of S1 or S0 induced simply by dipolar

    solutesolvent interactions, as expected for a polar

    solute in a polar solvent [20,25]. Similar consider-

    ations can be made by analysing the net atomiccharges. Table 5 shows, in particular, that in both

    S0 and S1 an increase in solvent polarity gives rise

    to an increase in electron population on the oxy-

    gen atom, while the vertical S0 ! S1 transitionresults in a reduction of such electron-charge ac-

    cumulation. The latter phenomenon, which is the

    opposite of that found with M1 (see Table 3),

    suggests that S0 ! S1 excitation produces an en-richment of the resonance hybrid in the quinonoid

    form. The same set of calculations performed for

    Table 7

    Calculated and experimental properties of the S0 ! S1 transition of the streptopentamethinemerocyanine and its N-phenyl derivativeMolecule Calculateda f lS0 (D) lS1 (D) Experimental

    b 103emax

    DE (eV) DE (eV)

    Me2NCH@CH2CHO 4.087 1.211 8.64 13.43 3.430 51PhNMeCH@CH2CHOc 3.797 1.360 8.49 13.17 3.337 65a At ke 0.b In dichloromethane, [41].c The phenyl ring was assumed to be coplanar with the chomophore chain and to have the benzene structure.

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 319

  • 7/30/2019 Solvatochromic effect

    12/17

    the vertical S1eq: ! S0 transition, i.e., using ge-ometry and solvaton charges of the relaxed excited

    state, led to similar trends of the dipole moments

    and net charges (Table 6). Two points, however,have to be noted: (i) lS1 lS0 is negative over thewhole range ofke, (ii) the increase oflS1 and lS0with increasing ke is less marked than that foundfor the vertical S0eq: ! S1 transition (Table 5).Such trend, which is contrary to that exhibited by

    M1 (Tables 3 and 4) (where lS1 > lS0), is related tothe fact that the solvaton field acting on the

    equilibrium excited state is weak compared with

    that acting on the equilibrium ground state. As a

    consequence, the S1eq: ! S0 transition energyturns out to be little affected by solvent polarity,

    with a global negative solvatochromism of just

    250 cm1 ()0.031 eV) (Table 6).The above discussion has shown how the

    adopted theoretical scheme is reflected in the cal-

    culation results. Now, these results have to be

    compared with the observed solvato- and fluoro-

    solvatochromic behaviours of M1 and M2. In or-

    der to make such comparison as concise and

    comprehensive as possible, we have plotted in a

    same diagram the calculated and experimental

    band frequencies (in cm1) using, respectively, the

    ke and ENT [25] scales, both ranging from 0.0 to1.0. This way, we obtained four diagrams corre-

    sponding to the absorption and fluorescence

    properties of M1 (Figs. 5 and 6) and M2 (Figs. 7

    and 8). Let us comment first on Figs. 5 and 6.

    As is evident, the calculated mmax values are

    overestimated with respect to the experimental

    values. However, the deviation is nearly the same

    for absorption (Fig. 5) and emission (Fig. 6), and

    keeps almost constant over the entire range of sol-

    vent polarity ($0.5 eV at ke, ENT 0:1 and $0.45eV at ke, E

    N

    T 0:9). Thus, apart from the sys-tematic blue-shift of the calculated S0eq: ! S1and S1eq: ! S0 transitions (essentially due to theadopted CI truncation), our simple solvatonCS

    INDO scheme describes fairly well the absorption

    and fluorescence properties of ketocyanine M1

    as well as their dependence on the solvent charac-

    teristics. As a matter of fact, the calculated Dmmaxvalues mk 0 mk 0:9 for absorption3550 cm1 and emission 5000 cm1 are ingood qualitative agreement with the experimental

    ones mENT 0:1 mENT 1:0 ([26]: 4010,3980 cm1; [27]: 3600, 3880 cm1). Interest-ingly, the agreement becomes even better if the

    comparison is made over the same interval (0.10.9)

    of solvent-polarity parameter. In this case, the sol-

    vatochromic shifts, derived by simple linear inter-

    polations, were found to be: Dmabs=cm1 ffi 3170calcd:, 3260 [26], 2800 [27], Dmfluo ffi 4350

    Fig. 5. Wavenumber of the calculated and experimental ab-

    sorption maximum of M1 vs, respectively, the ke and nor-malized ENT values of solvent polarity. The experimental mmaxvalues in eight solvents (toluene, THF, acetone, DMF, 2-pro-

    panol, ethanol, methanol and water) with ENT ranging from

    0.090 to 1.000, are reported from [26,27].

    Fig. 6. Wavenumber of the calculated and experimental emis-

    sion maximum of M1 vs, respectively, the ke and normalizedENT values of solvent polarity. The experimental mmax values are

    reported from [26,27] (see caption of Fig. 5 for the solvent

    characteristics).

    320 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    13/17

    calcd:, 3810 [26], 3710 [27]. Such estimatesindicate that excepting solvents with very high ENT ,

    where specific solutesolvent interactions become

    very important, in M1 the fluorescence is more

    sensitive than the absorption to solvent polarity.

    This behaviour, originating in the marked orien-

    tational stabilization of the emitting state S1 [20]

    is evidenced by the fact that, leaving out solvents

    with ENT > 0:7, the remaining experimental points

    are aligned almost parallely to the theoretical plots

    (Figs. 5 and 6). The same thing emerges from the

    plots of the Stokes shifts (mabsmax mfluomax Dmabsfluo) vs the normalized solvent-polarity parame-ters (Fig. 9). In spite of the large fluctuations of the

    experimental data, it is evident that calculated and

    observed Stokes shifts of M1 exhibit similar plots

    until ENT ffi 0:7 . Beyond this value, the theoreticalplot goes on increasing, while the experimental ones

    first bend and then undergo a dramatic drop down

    in water (ENT 1:0). We can conclude that: (i) due tothe water peculiarity, any discussion considering

    only the overall spectral shifts between water

    (ENT 1:0) and toluene (ENT ffi 0:1) may lead towrong interpretations, (ii) except for EN

    T

    P 0:7, inM1 solvato- and fluorosolvatochromic shifts seem

    to be to a great extent determined by non-specifc

    solutesolvent interactions.

    Now, let us consider the stilbazolium betaine

    M2 (Figs. 7 and 8). In low polarity solvents, the

    absorption and emission maxima of M2 are red-

    shifted with respect to those of M1 of more or less

    0.5 eV. This is correctly predicted by calculations

    (e.g., at ke 0 the calculated red shifts are 0.48and 0.38 eV, respectively; see Tables 3 and 5 and

    Tables 4 and 6) even if the individual transition

    energies are rather overestimated ($0.7 eV at ke,ENT 0:3) due to the adopted CI limitations. Fig. 7shows that both the initial small red shift and the

    Fig. 7. Wavenumber of the calculated and experimental ab-

    sorption maximum of M2 vs, respectively, the ke and ENTvalues of solvent polarity. The experimental data in 25 solvents

    covering the entire range of ENT were taken from [18]. The mmaxvalues of Table 1 are also reported.

    Fig. 8. Wavenumber of the calculated and experimental emis-

    sion maximum of M2 vs, respectively, the ke and normalizedENT values of solvent polarity. The experimental mmax values are

    those of Table 1.

    Fig. 9. Calculated and experimental Stokes shift of M1 vs, re-

    spectively, the ke and normalized ENT values of solvent po-larity. Experimental values were derived from [26,27] (see

    caption of Fig. 5 for the solvent characteristics).

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 321

  • 7/30/2019 Solvatochromic effect

    14/17

    successive strong blue shift of the absorption band

    of M2 observed on increasing the ENT value are

    qualitatively well reproduced by the calculations.

    Quantitatively speaking, the large negative solva-tochromism of M2 (mabschloroform mabswater 6260cm1 (Table 1), 6490 cm1 [18,25]) is somewhatunderestimated by the solvatonCS INDO calcu-

    lations (mabske0:2 mabske0:9 ffi 4230 cm1, see Table5 and Fig. 7). According to the previous discussion

    on the values and solvent-induced changes of di-

    pole moments, in this case lS1 < lS0 fluorosol-vatochromism should be expected to be weaker

    than solvatochromism, contrary to what occurs in

    the case of M1 lS1 > lS0. This is fully confirmedby our fluorescence study on M2 in various sol-

    vents showing that mfluochloroform mfluowater 1340 cm1(Table 1 and Fig. 8), i.e., about a fifth of the blue-

    shift observed in absorption. Calculations under-

    estimate fluorosolvatochromic effects to the same

    extent as the solvatochromic ones, so mfluoke0:2mfluoke0:9 reduces to $ 300 cm1 (Table 6, Fig. 8).In other words, the calculation results agree with

    experiment as regards the direction of the spectral

    shifts of absorption and fluorescence maxima, but

    in both cases the effects are predicted to be weaker

    than the experimental ones. This allows for a much

    better agreement between calculated and observedStokes shifts (Dmabs fluo), as clearly shown byFig. 10. As is evident, the theoretical and experi-

    mental plots of Dmabs fluo are in very goodagreement over the experimentally investigated

    range of ENT values, and indicate that the Stokes

    shift of M2 increases rapidly on increasing thesolvent polarity. Interestingly, contrary to what is

    found for M1 (Fig. 9), in this case the experimental

    plot displays no inversion 5 and remains parallel to

    the theoretical one until ENT 1:0.Figs. 9 and 10 lend themselves to the following

    concluding remarks. Although dyes M1 and M2

    exhibit opposite solvatochromic and fluorosolva-

    tochromic behaviours, in both systems the Stokes

    shift increases considerably (with the only excep-

    tion of M1 in water) with increasing the solvation

    power of the solvent (as expressed by the EN

    T

    va-

    lue). This occurs for opposite reasons in the two

    dyes: in M1 the Stokes shift raises because the

    bathochromic shift of the fluorescence is greater

    than that of the absorption; on the contrary, in M2

    the same phenomenon is due to the hypsochromic

    shift of the absorption being greater than that of

    the fluorescence. Both phenomena are well de-

    scribed within the solvatonCS INDO scheme and

    can be rationalized in terms of the different free-

    molecule electronic structures (i.e., the different

    relative weights of the neutral and zwitterionic VB

    structures) and of their evolution with increasingthe solvent polarity. Moreover, it should be noted

    (see also Figs. 58) that the solvatonCS INDO

    approach, formally including only electrostatic

    solutesolvent interactions, actually accounts

    fairly well for the effects of media characterized by

    quite high ENT values where specific (H-bond) in-

    teractions are expected to be as important as di-

    polar interactions. The reason for this somewhat

    surprising result lies in the way the solutesolvaton

    interactions were included in the molecular Ham-

    Fig. 10. Calculated and experimental Stokes shift of M2 vs,

    respectively, the ke and normalized ENT values of solvent po-larity. The experimental values derive from the absorption and

    emission data of Table 1.

    5 Fig. 10 shows that the solvation power of hydrogen-bond

    donor solvents towards M2 (in both the ground and the excited

    states) is well described by the ET30 ENT parameter. On thecontrary (Fig. 9), the solvation power of such solvents

    (primarily water) towards M1 is poorly described within the

    ET30 ENT ) solvent scale. This is likely due to the fact that theprobe used to develop this scale (pyridinium N-phenolate

    betaine) [25] is chemically more similar to M2 than to M1. A

    better general description could be achieved with proper

    multiparameter approaches [25,36], but such correlation anal-

    ysis does not fall within the purposes of this work.

    322 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    15/17

    iltonian. As reminded in Sections 1 and 3.1, our

    procedure acts prevalently on the p-electron sys-

    tem by correction terms linearly related to the

    solvation parameter ke (Eq. (1)). Within thisscheme, the modifications of the Fs

    ppelements at

    high ke values may well account for the effects ofH-bond interactions occurring in the molecular

    plane (e.g., those involving the sp2-like oxygen

    lone pairs). This interpretation is plausible since it

    has been shown [19] that the effects of protic sol-

    vents on merocyanine spectra may be described

    within p-electron theory using heteroatom pa-

    rameters corresponding to the molecule proton-

    ated on the oxygen.

    A complete validation of the calculated ground

    and excited-state dipole moments was not possible

    for lack of experimental data. However, using the

    solvatochromic comparison method, Banerjee et

    al. [27] evaluated the ratio lS1=lS0 to be equal to1.4 for dye M1. Table 5 shows that for M1 the

    theoretical prediction is in very good agreement

    with experiment (at ke 0, lS1=lS0 1:44).

    4. Summary and conclusions

    The main purpose of the present paper was toappraise the capability of the solvatonCS INDO

    model to arrange both solvatochromism and flu-

    orosolvatochromism of merocyanine dyes. In or-

    der to carry out an exhaustive test, we searched for

    two sample merocyanines characterized by oppo-

    site solvent effects. As a prototypic dye exhibiting

    large positive solvatochromism on both absorp-

    tion and emission, we chose the ketocyanine M1

    which has been the subject of extensive experi-

    mental work. The stilbazolium betaine M2 was

    chosen for the large negative solvatochromism ofits absorption band, which has been widely studied

    both experimentally and theoretically. Since, to

    our knowledge, the emission properties of M2 had

    never been reported before, we carried out an ex-

    perimental investigation on the matter. We found

    that M2 is rather weakly fluorescent and exhibits

    negative fluorosolvatochromism, even if the emis-

    sion is much less sensitive to a change of solvent

    polarity than the absorption. Moreover, the anal-

    ysis of the fluorescence emission and excitation

    spectra led us to conclude that, as it happens in the

    case of the strongly fluorescent M1, the emission

    of M2 originates from the local minimum of the

    solvated S1pHpL state.According to this body of experimental evidence,

    the theoretical descriptions of solvatochromism and

    fluorosolvatochromism were undertaken within

    the same conventional scheme, i.e., calculating

    the shifts induced by a solvent-polarity change

    on the vertical S0eq: ! S1FranckCondon andS1eq: ! S0FranckCondon transitions. Such ascheme was directly applied within the solvatonCS

    INDO method. The equilibrium absorbing and

    emitting units (molecule + solvaton pattern) and the

    corresponding S0!

    S1 and S1!

    S0 vertical tran-

    sitions were studied for M1 and M2 as a function of

    a polarity factor formally related to the dielectric

    constant, ke ffiffiep 1= ffiffiep. For the sake of sim-plicity, the molecular geometries in the equilibrium

    S0 and S1 states were approximated by those opti-

    mized in the absence of solvent, and the respective

    solvaton patterns, setup using the subset of the net

    p-electron charges, were adjusted iteratively start-

    ing from those corresponding to the unsolvated

    molecules.

    The solvent-dependent composition of the res-

    onance hybrid between the covalent and the zwit-terionic VB structures was taken as the key to

    interpret the calculated ground and excited-state

    properties of the two merocyanines at the various

    ke values. The analysis of the gas-phase opti-mized geometries in terms of BLA parameter

    showed that in the ground state M1 has a markedly

    covalent character while in M2 the covalent char-

    acter is nearly balanced by the zwitterionic one. On

    moving from S0 to S1, the contribution of the

    charge-separated structure increases in both dyes,

    even if the phenomenon is less marked in M2 thanin M1. Introduction of a polar solvent stabilizes the

    zwitterionic forms but, due to the difference in the

    starting electronic structures, it results in opposite

    solvatochromic behaviours in both absorption and

    emission. The use of solvaton sets reflecting the net

    p-electron charges proved to be most effective to

    rule the evolution of the resonance hybrid com-

    position of S0 and S1 as the solvent polarity in-

    creases. The description of the solutesolvent

    interaction in terms of local charge distributions

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 323

  • 7/30/2019 Solvatochromic effect

    16/17

    makes the solvaton model free from symmetry

    imposed restraints, this being at variance with all

    solvaton models (e.g., SCRF) where the solute is

    treated as a point dipole. Thus, the fact that M1 hasmoderate S0 and S1 dipole moments perpendicular

    to the conjugated chain (the longitudinal compo-

    nents being equal to zero by symmetry) while M2

    has large dipole moments aligned with the long

    molecular axis leads to quite different reaction

    fields (and solutesolvent interaction energies)

    within the dipolar approximation in spite of the

    basic similarities of the two systems. On the con-

    trary, the solvaton model provides fully consistent

    descriptions for the M1 and M2 solvatochromisms.

    A thorough comparison with the experimental

    data was made possible by the construction of di-

    agrams where calculated and observed frequencies

    of the absorption and emission maxima were

    plotted against the ke and ENT solvent polarityparameters, both ranging from 0 to 1. Apart from a

    systematic overestimation (0.50.7 eV) of the

    transition energies the solvatonCS INDO calcu-

    lations provided good descriptions of the opposite

    solvatochromic behaviours of M1 and M2 in both

    absorption and emission. Quantitatively speaking,

    the solvent shifts of the absorption and fluores-

    cence spectra were correlated better for the posi-tively solvatochromic ketocyanine than for the

    negatively solvatochromic stilbazolium betaine,

    where the solvatochromic ranges (mnon-polarsolvent mhighly polar solvent) of absorptionand emission were both somewhat underestimated.

    However, in very good agreement with experiment

    the Stokes shifts (Dmabs fluo) of both M1 andM2 were predicted to substantially increase (espe-

    cially in the case of M2) on increasing the solvent

    polarity. We notice that, except for the peculiar

    behaviour of M1 in water, the solvatonCS INDOmodel is capable of accounting for the solvation

    effects on merocyanine spectra over almost the

    entire range of the normalized ENT values. This

    happens since the polarization of the p-electron

    system induced by specific (H-bond) solvent inter-

    actions taking place in the molecular plane is im-

    plicitly accounted for within our formulation of the

    solvaton model.

    To sum up, the solvatonCS INDO method

    was applied to the study of solvent effects on both

    absorption and fluorescence spectra of positively

    and negatively solvatochromic merocyanines.

    Such a severe test has been passed rather well, at

    least as far as the main qualitative aspects areconcerned (but also the sizes of the solvent shifts

    were in general well reproduced). An elementary

    condition underlying these results is that the elec-

    tronic transitions of the solute were generated in

    the presence of the polarized solvent. Quite sur-

    prisingly, the controversy that arose some years

    ago in the interpretation of the solvatochromism

    of M2 (see Section 1) came from theoretical

    treatments devoid of this prerequisite.

    Acknowledgements

    This research was jointly supported by the

    MURST (Roma) and the University of Modena

    and Reggio Emilia within the Programmi di

    Ricerca di Interesse Nazionale. We are indebted

    to Dr. J.P. Flament for providing us with his ge-

    ometry optimization program and to Prof. G.

    Berthier for valuable discussions.

    References

    [1] I. Baraldi, F. Momicchioli, G. Ponterini, D. Vanossi,

    Chem. Phys. 238 (1998) 353.

    [2] I. Baraldi, F. Momicchioli, G. Ponterini, D. Vanossi, Adv.

    Quantum Chem. 36 (1999) 121.

    [3] F. Momicchioli, I. Baraldi, M.C. Bruni, Chem. Phys. 82

    (1983) 229.

    [4] G. Klopman, Chem. Phys. Lett. 1 (1967) 200.

    [5] H.A. Germer, Theor. Chim. Acta 34 (1974) 145.

    [6] H.A. Germer, Theor. Chim. Acta 35 (1974) 273.

    [7] S. Miertus, O. Kysel, Chem. Phys. 21 (1977) 27.

    [8] S. Miertus, O. Kysel, Chem. Phys. Lett. 65 (1979) 395.

    [9] R. Costanciel, O. Tapia, Theor. Chim. Acta 48 (1978)383.

    [10] A. Botrel, A. Le Beuze, P. Jacques, H. Straub, J. Chem.

    Soc. Faraday Trans. II 80 (1984) 1235.

    [11] A. Botrel, B. Aboad, F. Corre, F. Tonnard, Chem. Phys.

    194 (1995) 101.

    [12] J. Griffiths, Colour and Constitution of Organic Molecules,

    Academic Press, London, 1976.

    [13] C. Reichardt, Solvent Effects in Organic Chemistry, Verlag

    Chemie, Weinheim, 1979.

    [14] J.O. Morley, J. Mol. Struct. (Theochem) 304 (1994) 191.

    [15] L. da Silva, C. Machado, M.C. Rezende, J. Chem. Soc.

    Perkin Trans. II (1995) 483.

    324 I. Baraldi et al. / Chemical Physics 288 (2003) 309325

  • 7/30/2019 Solvatochromic effect

    17/17

    [16] I.D.A. Albert, T.J. Marks, M.A. Ratner, J. Phys. Chem.

    100 (1996) 9714.

    [17] A. Klamt, J. Phys. Chem. 100 (1996) 3349.

    [18] P. Jacques, J. Phys. Chem. 90 (1986) 5535.

    [19] H.G. Benson, J.N. Murrell, J. Chem. Soc. Faraday Trans.II 68 (1972) 137.

    [20] L. Salem, Electrons in Chemical Reactions: First Princi-

    ples, Wiley, New York, 1982 (Chapter 8).

    [21] J. Tomasi, M. Persico, Chem. Rev. 94 (1994) 3027.

    [22] J.T. Blair, K. Krogh-Jespersen, R.M. Levy, J. Am. Chem.

    Soc. 111 (1989) 6948.

    [23] K. Coutinho, S. Canuto, M.C. Zerner, J. Chem. Phys. 112

    (2000) 9874.

    [24] K.J. de Almeida, K. Coutinho, W.B. de Almeida, W.R.

    Rocha, S. Canuto, Phys. Chem. Chem. Phys. 3 (2001)

    1583.

    [25] C. Reichardt, Chem. Rev. 94 (1994) 2319.

    [26] M.A. Kessler, O.S. Wolfbeis, Spectrochim. Acta 47A

    (1991) 187.

    [27] D. Banerjee, A. Kumar Laha, S. Bagchi, J. Photochem.

    Photobiol. A: Chem. 85 (1995) 153.

    [28] S.J. Isak, E.M. Eyring, J. Phys. Chem. 96 (1992) 1738.

    [29] G.R. Fleming, A.W.E. Knight, J.M. Morris, R.J.S. Mor-

    rison, G.W. Robinson, J. Am. Chem. Soc. 99 (1977) 4306.

    [30] I. Baraldi, S. Ghelli, Z.A. Krasnaya, F. Momicchioli, A.S.

    Tatikolov, D. Vanossi, G. Ponterini, J. Photochem. Pho-

    tobiol. A: Chem. 105 (1997) 297.

    [31] P. Milliee, F. Momicchioli, D. Vanossi, J. Phys. Chem. B

    104 (2000) 9621.[32] I. Baraldi, F. Momicchioli, G. Ponterini, A.S. Tatikolov,

    D. Vanossi, Phys. Chem. Chem. Phys. 5 (2003).

    [33] J.O. Morley, R.M. Morley, A.L. Fitton, J. Am. Chem. Soc.

    120 (1998) 11479.

    [34] J. Catalaan, E. Mena, W. Meutermans, J. Elguero, J. Phys.

    Chem. 96 (1992) 3615.

    [35] C. Burda, M.H. Abdel-Kader, S. Link, M.A. El-Sayed, J.

    Am. Chem. Soc. 122 (2000) 6720.

    [36] B. CCunderlkovaa, L. SSikurovaa, Chem. Phys. 263 (2001)

    415.

    [37] R. Fletcher, M.J. Powell, Comput. J. 6 (1963) 163.

    [38] K. Ohno, Theor. Chim. Acta 2 (1964) 219;

    G. Klopman, J. Am. Chem. Soc. 86 (1964) 4550.

    [39] G. Brancolini, Doctorate thesis, University of Modena and

    Reggio Emilia, 2002.

    [40] D.J.A. De Ridder, D. Heijdenrijk, H. Schenk, R.A.

    Dommisse, G.L. Lemieere, J.A. Lepoivre, F.A. Alder-

    weireldt, Acta Crystallogr. C 46 (1990) 2197.

    [41] S.S. Malhotra, M.C. Whiting, J. Chem. Soc. (1960) 3812.

    I. Baraldi et al. / Chemical Physics 288 (2003) 309325 325