theoretical study of the effects of solvent environment on...

10
Theoretical study of the effects of solvent environment on photophysical properties and electronic structure of paracyclophane chromophores Artëm Masunov a! and Sergei Tretiak Los Alamos National Laboratory, Theoretical Division and Center for Nonlinear Studies, Los Alamos, New Mexico 87545 Janice W. Hong, Bin Liu, and Guillermo C. Bazan Department of Chemistry, University of California, Santa Barbara, California 93106 sReceived 6 December 2004; accepted 31 January 2005; published online 14 June 2005d We use first-principles quantum-chemical approaches to study absorption and emission properties of recently synthesized distyrylbenzene sDSBd derivative chromophores and their dimers stwo DSB molecules linked through a f2.2gparacyclophane moietyd. Several solvent models are applied to model experimentally observed shifts and radiative lifetimes in Stokes nonpolar organic solvents stoluened and water. The molecular environment is simulated using the implicit solvation models, as well as explicit water molecules and counterions. Calculations show that neither implicit nor explicit solvent models are sufficient to reproduce experimental observations. The contact pair between the chromophore and counterion, on the other hand, is able to reproduce the experimental data when a partial screening effect of the solvent is taken into account. Based on our simulations we suggest two mechanisms for the excited-state lifetime increase in aqueous solutions. These findings may have a number of implications for organic light-emitting devices, electronic functionalities of soluble polymers and molecular fluorescent labels, and their possible applications as biosensors and charge/energy conduits in nanoassemblies. © 2005 American Institute of Physics. fDOI: 10.1063/1.1878732g I. INTRODUCTION The interest in the molecules studied in this paper is twofold. First, f2.2gparacyclophane sPCPd-based chro- mophores can serve as a model of intermolecular contacts and solid-state interactions in photoluminescent conducting polymers, widely used in organic light-emitting devices. 1 The PCP core holds two p-conjugated systems cofacially sscheme 1d at a distance that is shorter than the van der Waals distance, and thus provides strong through-space in- teractions between those systems which can strongly affect the photophysical properties of chromophores and serves as a conduit for delocalization of excitations between chains. Second, water-soluble conjugated oligomers are promising fluorescent labels for biomedical applications. 2 Molecules used for these applications should not be sensitive to envi- ronmental perturbations and retain their emission properties and high quantum yield. Detailed experimental studies ac- companied by careful theoretical analysis can provide a valu- able insight into the nature of underlying fundamental pho- tophysical processes and, ultimately, may help in designing robust novel fluorescent labels. The electronic structure of unsubstituted PCP was stud- ied theoretically using semiempirical approaches by Canuto and Zerner more than a decade ago. 3 They analyzed the ex- cited states in terms of two weakly interacting p systems. The time-dependent Hartree–Fock technique was then used to study large PCP derivatives with extended p systems. 4 It was found that fluorescent properties of these derivatives, unlike unsubstituted PCP, are strongly dependent on the na- ture of the lowest excited state. PCP derivatives and related compounds were subject of a large number of subsequent combined experimental and theoretical studies, 5 including their nonlinear optical properties. 6,7 In particular, the recent experimental investigation 8 of PCP and distyrylbenzene sDSBd derivatives reports that water-soluble ionically derivatized PCP chromophores sPCPnw and PCPdw, see scheme 1d exhibit significant sol- vatochromic shifts in their emission spectrum s,50 nmd and about 20 ns si.e., an order of magnituded longer lifetimes in aqueous solution compared to neutrally derivatized analogs ad Author to whom correspondence should be addressed. Electronic mail: [email protected] SCHEME 1. Chromophores studied in this work: PCPn sR =CH 2 CH 3 d, PCPd sR =OCH 3 d, PCPnt fR = sCH 2 d 6 Brg, PCPnw fR = sCH 2 d 6 NMe 3 + ·Br - g, PCPdt fR =OSisCH 3 d 2 tBug, PCPdw fR =OsCH 2 d 4 SO 3 - ·NBu 4 + g, DSBn sR =CH 2 CH 3 d, DSBd sR =OCH 3 d, DSBnt fR = sCH 2 d 6 Brg, DSBnw fR = sCH 2 d 6 NMe 3 + ·Br - g, DSBdt fR =OSisCH 3 d 2 tBug, and DSBdw fR =OsCH 2 d 4 SO 3 - ·NBu 4 + g. THE JOURNAL OF CHEMICAL PHYSICS 122, 224505 s2005d 0021-9606/2005/122~22!/224505/10/$22.50 © 2005 American Institute of Physics 122, 224505-1 Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Upload: others

Post on 10-Jun-2020

3 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

Theoretical study of the effects of solvent environment on photophysicalproperties and electronic structure of paracyclophane chromophores

Artëm Masunova! and Sergei TretiakLos Alamos National Laboratory, Theoretical Division and Center for Nonlinear Studies, Los Alamos,New Mexico 87545

Janice W. Hong, Bin Liu, and Guillermo C. BazanDepartment of Chemistry, University of California, Santa Barbara, California 93106

sReceived 6 December 2004; accepted 31 January 2005; published online 14 June 2005d

We use first-principles quantum-chemical approaches to study absorption and emission properties ofrecently synthesized distyrylbenzenesDSBd derivative chromophores and their dimersstwo DSBmolecules linked through af2.2gparacyclophane moietyd. Several solvent models are applied tomodel experimentally observed shifts and radiative lifetimes in Stokes nonpolar organic solventsstoluened and water. The molecular environment is simulated using the implicit solvation models, aswell as explicit water molecules and counterions. Calculations show that neither implicit nor explicitsolvent models are sufficient to reproduce experimental observations. The contact pair between thechromophore and counterion, on the other hand, is able to reproduce the experimental data when apartial screening effect of the solvent is taken into account. Based on our simulations we suggesttwo mechanisms for the excited-state lifetime increase in aqueous solutions. These findings mayhave a number of implications for organic light-emitting devices, electronic functionalities ofsoluble polymers and molecular fluorescent labels, and their possible applications as biosensors andcharge/energy conduits in nanoassemblies. ©2005 American Institute of Physics.fDOI: 10.1063/1.1878732g

I. INTRODUCTION

The interest in the molecules studied in this paper istwofold. First, f2.2gparacyclophane sPCPd-based chro-mophores can serve as a model of intermolecular contactsand solid-state interactions in photoluminescent conductingpolymers, widely used in organic light-emitting devices.1

The PCP core holds twop-conjugated systems cofaciallysscheme 1d at a distance that is shorter than the van derWaals distance, and thus provides strong through-space in-teractions between those systems which can strongly affectthe photophysical properties of chromophores and serves asa conduit for delocalization of excitations between chains.Second, water-soluble conjugated oligomers are promisingfluorescent labels for biomedical applications.2 Moleculesused for these applications should not be sensitive to envi-ronmental perturbations and retain their emission propertiesand high quantum yield. Detailed experimental studies ac-companied by careful theoretical analysis can provide a valu-able insight into the nature of underlying fundamental pho-tophysical processes and, ultimately, may help in designingrobust novel fluorescent labels.

The electronic structure of unsubstituted PCP was stud-ied theoretically using semiempirical approaches by Canutoand Zerner more than a decade ago.3 They analyzed the ex-cited states in terms of two weakly interactingp systems.The time-dependent Hartree–Fock technique was then used

to study large PCP derivatives with extendedp systems.4 Itwas found that fluorescent properties of these derivatives,unlike unsubstituted PCP, are strongly dependent on the na-ture of the lowest excited state. PCP derivatives and relatedcompounds were subject of a large number of subsequentcombined experimental and theoretical studies,5 includingtheir nonlinear optical properties.6,7

In particular, the recent experimental investigation8 ofPCP and distyrylbenzenesDSBd derivatives reports thatwater-soluble ionically derivatized PCP chromophoressPCPnw and PCPdw, see scheme 1d exhibit significant sol-vatochromic shifts in their emission spectrums,50 nmd andabout 20 nssi.e., an order of magnituded longer lifetimes inaqueous solution compared to neutrally derivatized analogs

adAuthor to whom correspondence should be addressed. Electronic mail:[email protected]

SCHEME 1. Chromophores studied in this work: PCPnsR=CH2CH3d,PCPd sR=OCH3d, PCPnt fR=sCH2d6Brg, PCPnw fR=sCH2d6NMe3

+·Br−g,PCPdt fR=OSisCH3d2tBug, PCPdw fR=OsCH2d4SO3

−·NBu4+g, DSBn sR

=CH2CH3d, DSBd sR=OCH3d, DSBnt fR=sCH2d6Brg, DSBnw fR=sCH2d6NMe3

+·Br−g, DSBdt fR=OSisCH3d2tBug, and DSBdw fR=OsCH2d4SO3

−·NBu4+g.

THE JOURNAL OF CHEMICAL PHYSICS122, 224505s2005d

0021-9606/2005/122~22!/224505/10/$22.50 © 2005 American Institute of Physics122, 224505-1

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 2: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

sPCPnt and PCPdtd in toluene solution. Side-chain function-ality was found to have no effect on the excited states of thecore-conjugated system, since both ionic and neutral chro-mophoressPCPdw and PCPdtd have practically indistin-guishable photophysical properties, for example, when dis-solved in the same solventfdimethyl sulfoxidesDMSOdg. Inthe contrast, DSB monomer analogsssee scheme 1d displaymuch less pronounced difference in aqueous versus nonpolarsolutionss,5 nm and 1 ns, respectivelyd.

This paper is aimed to elucidate electronic mechanismsleading to these differences. We employ first-principlesquantum-chemical calculations and different theoreticalmodels accounting for solvent environment, and report bothsuccesses and failures.

II. COMPUTATIONAL DETAILS

All quantum-chemical calculations were performed us-ing theGAUSSIAN03 suite.9 This package implements a non-equilibrium integral equation formalism polarizable con-tinuum model sIEFPCMd,10 which we utilized in oursimulations. Saturated molecular side-chains were truncatedsPCPn, PCPd, DSBn, and DSBd, see scheme 1d, as they donot participate in the conjugated electronic system. Explicitwater molecules and counterions were added in some calcu-lations, as described in Sec. III. Ground-state optimized ge-ometries have been obtained using the Hartree–FocksHFdtheory level combined with the Slater-type orbitals with threeGaussianssSTO-3Gd basis set. For ground-state geometries,we previously found that HF method is superior to thedensity-functional theorysDFTd approach based on the hy-brid Becke 3 Lee–Yang–ParrsB3LYPd functional by repro-ducing accurately the bond length alternation parameters insimilar conjugated systems when compared to available ex-perimental structures.1 The molecular geometries of the low-est excited states were optimized at the configuration inter-action singlessCISd / STO-3G theory level. Time-dependentDFT sTD-B3LYPd optimizations for the large conjugatedmolecules were recently found to be unstable and often con-verging to unphysical charge-transfer states with vanishingoscillator strengths.12 The excited-state electronic structureswere calculated using the time-dependent DFTsTD-B3LYPdapproach,13 which was found to be accurate for excited statesin conjugated molecules,14 and, in particular, for both linearand nonlinear spectra in centrosymmetrically substitutedchromophores.7,11 At the excited-state optimal geometry,transition frequenciesVge

sfd and dipolesmgesfd corresponding to

the vertical fluorescence process were used to calculate theapproximate radiative lifetimet0 as15 scgs unitsd

1

t0=

4

"c3sn2 + 2dVge

sfd3mgesfd2, s1d

where"" is the Plank’s constant,c is the speed of light, andn is the refractive index.

To analyze the nature of the excited states involved inthe photophysical processes we further used the natural tran-sition orbitalssNTOsd for excited states,16 derived from thecalculated transition densities. This analysis offers the mostcompact representation of a given transition density in terms

of an expansion into single-particle transitions. This ap-proach is particularly attractive for given molecules, sinceeach excited state in question can be well represented as atransition between a single dominant pair of transition orbit-als swith 84% or more contributiond. Figures of transitionorbitals were drawn using theXCRYSDEN graphicalpackage.17

III. RESULTS AND DISCUSSION

A. Molecular structures in vacuum

As we pointed out previously,11 even the most accurateelectronic structure method will fail if used with an inaccu-rate molecular geometry. Therefore, we need to select atheory level, both feasible for routine geometry optimiza-tions, and reliable enough to accurately reproduce the detailsof the molecular geometry. Two geometric parameters areespecially important for the electronic properties of conju-gated molecules: planarity which can be defined as dihedralangle along the backbone, and bond length alternationfBLA,or the difference in length between the single and doublebonds,rsC–Cd–rsCvCdg. The molecular planarity for stil-bene derivatives was a matter of controversy in the past threedecades.11,18 Apparently, it is determined by the balance oftwo opposite trends. Conjugation of thep system leads toplanarization of the molecule, while steric repulsion betweenthe ethylenic and theortho-hydrogen atoms of the phenylrings is responsible for the twist along the single bond.

Experimentally,trans-stilbene appears to have a larges32°d torsion angle in the gas phase19 and is considerablymore planars5°d in the crystal.20 However, various factors,such as ambiguity in averaging over motions with the largeamplitude in the gas phase, dynamic disorderspedalingmotion21d in crystals, and effects of crystalline environment,complicate the comparison with theoretical values. Indirectexperimental evidence for planarity of the free stilbene mol-ecules in solution is obtained from the fact that simulatedspectrasboth vibrational22 and electronic11d fit much betterthe experimental data with the assumption of molecular pla-narity. Studies of the fine rotational structure of the electronicspectra in a collision-free environment in a molecular beamindicate that stilbene is planar in both the ground and firstexcited states.23

Theoretical results for stilbene geometry optimizationswere recently reviewed in Ref. 14. Almost all semiempiricalandab initio methods predict the planar geometry of stilbenesC2hd to be a saddle point, while the absolute minimum isnonplanarsC2d. In contrast, DFT methods predict the abso-lute minimum to be planar. In general, many gradient-corrected and hybrid density functionals show very good ac-curacy for small molecular systems. It is also known that theperformances of the same functionals are of a lower level inlarge conjugated systems, but this defect becomes evident ingoing towards the limit of an infinite system. Many function-als such as local-density approximationsLDA d and evengradient-corrected versions lead to unphysical behavior ofsome observables. For example, they do not reproduce exci-tonic effects. Mixed hybrid methods partially correct thesedeficiencies.24 Even though, hybrid B3LYP functional is

224505-2 Masunov et al. J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 3: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

known to overdelocalize electrons—a feature inherited fromthe free-electron gas model.25 Specifically, it significantlyoverestimates rotation barriers around single bonds,26 andunderestimates BLA values.27 B3LYP planarity predictions,therefore, should be taken somewhat skeptically. The resultsmay improve when a portion of HF exchange in the hybridfunctional increases. One should note that the energy barrierto planarization of stilbene is predicted to be less than1 kcal/mol at all theory levels. Therefore, the relative stabil-ity of the planar versus nonplanar structures still may bereversed after the zero-point vibration energy is taken intoaccount and/or the basis set is increased. Such a reversal wasfound, for instance, by Masunov and Dannenberg for theurea molecule.28

An extrapolation to the infinite basis set limit was re-cently attempted by Kwasniewskiet al.18 In their focal pointanalysis they used the B3LYP/correlation consistent polar-ized valence triple zetascc-pVTZd optimized geometry forthe minimumsplanar structured, saddle points to the rotationof one and both phenyl groups, and artificially twistedC2

structures as an estimate for the nonplanar geometry. Al-though the planar geometry was found unstable at the HFlimit, account for electron correlation with the Møller–Plesset second-order perturbation theorysMP2d and coupledclusters singles and doubles with triple corrections methodfCCSDsTdg makes it an absolute minimum. This suggeststhat after the large basis set is used and electronic correlationis accounted for, the stilbene molecule is, in fact, planar.Unfortunately, at this level of theory direct geometry optimi-zation is not feasible. Even though the focal point analysisstudy does not provide direct evidence, it presents the bestab initio results to date, and provides three benchmark pointson the trans-stilbene ground-state potential-energy surfacewhich allow us to select appropriate theory level. We testedseveral computationally inexpensive theory levels, includingthose using minimal STO-3G basis set. The minimal basisset sometimes gives surprisingly good results for some mo-lecular properties due to compensation of errors of differenttypes. This compensation, however, does not hold for manyother properties, and STO-3G calculations are not reliablewithout careful comparisons with higher accuracy ap-proaches and experiment. In the present cases the performedchecksspresented in Table Id show that HF/STO-3G resultsto be acceptable for this benchmark set. Incidentally, with the

minimal basis set the planar structure is the most stable evenat the HF level.

Extrapolation of the bond length alternation parameter tothe limit of the infinite basis set is unfeasible, but one canobserve several trends from the calculated values, summa-rized in Table II. At the HF/6-31G and HF/6-31G** levelsboth absolute bond lengths and BLA values are in closestagreement with experiment, and they do not change signifi-cantly when the basis set size is increased. At the MP2/6-31G** level the BLA is decreased by 0.04 Å, but half of thisdecrease is lost when higher-order electron correlations aretaken into account at both CCSD and complete active spaceself-consistent fieldsCASSCFd levels. At the B3LYP/6-31G** level the BLA is very similar to the values obtainedfrom the MP2, and it somewhat increases with the size of thebasis set. One may conclude that both DFT and MP2 over-delocalize double bonds in stilbene and the BLA tends toreturn to its HF value as the basis set size and the order ofelectron correlation are increased. This observation supportsthe use of the HF method for geometry optimizations.

The lowest singlet excited statesS1d of stilbene is ex-

TABLE I. Energiesskcal/mold for the three points ontrans-stilbene poten-tial energy-surface relative to the planar structure: twistedC2 s27°d, transi-tion state to the rotation of one phenyl groupCs, and second-order saddlepoint for the simultaneous rotation of both phenyl groupsC2h.

C2 Cs C2h

HF/STO-3G 0.85 4.48 8.52HF/6-31G −0.17 3.15 7.71HF/6-31G** −0.36 2.79 6.63B3LYP/STO-31G 1.54 6.07 11.38B3LYP/6-31G 1.10 5.33 9.89Benchmarka 0.24 4.39 9.35

aReference 18.

TABLE II. Bond lengths and bond length alternation parametersBLA, Å d inthe central bridge of the stilbene.

Theory level Symm RsCvCd rsC–Cd BLA

AM1 C2 1.343 1.453 0.110PM3 C2h 1.342 1.457 0.115

HF/STO-3G C2h 1.322 1.494 0.172HF/6-31G C2h 1.332 1.475 0.143HF/6-31G C2 1.332 1.475 0.143HF/VZPa C2h 1.332 1.479 0.147

CASSCFs12:10d/VZPa C2h 1.351 1.479 0.128HF/6-31G** b C2 1.327 1.478 0.151MP2/6-31G** C2 1.351 1.464 0.113CCD/6-31G* C2 1.345 1.476 0.131B3LYP/6-31G C2h 1.357 1.458 0.101

B3LYP/6-31G** b C2h 1.348 1.466 0.118B3LYP/cc-pVTZb C2h 1.342 1.463 0.121B3LYP/cc-pVTZb C2 1.341 1.465 0.124

X-rayc Ci 1.326 1.471 0.145GEDd C2 1.33 1.48 0.15

aReference 29.bReference 18.cReference 21.dReference 19.

TABLE III. Calculated and experimental torsional vibrational frequencyscm−1d.

Method Frequency

HF/STO-3G 15HF/6-31G 43i

B3LYP/STO-3G 21B3LYP/6-31G 12

Experiment,S0a 9

CIS/STO-3G 40CIS/6-31G 19i

Experiment,S1a 35

aReference 31.

224505-3 Solvent effects on electronic structure of paracyclophane J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 4: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

pected to have a quinoidal structure, which leads to pla-narization. There are several spectroscopic evidences con-firming the planar structure ofS1, which is more rigid thanthe ground-statesS0d structure. In particular, while the elec-tronic absorption spectrum of stilbene has a distinguishablevibronic structure at low temperatures only, the emissionspectra retain a vibronic structure even at the roomtemperature.30 The rotational spectrum analysis results in aconclusion of planarity for the excited-stateS1 as well.23 Therelative rigidity of the excited state is reflected in a higherfrequency of the torsional vibrational mode at thetrans-stilbene excited states35 cm−1 vs 9 cm−1 in the groundstated.31 As one can see from Table III, both ground- andexcited-state torsional vibrational frequencies are reproducedwell with the minimal basis set again. BLA parameter in theexcited state changes signs−0.042 Å at the CASSCFs12:10d/VZP leveld. This effect is qualitatively reproduced at the CIS/STO-3G level as wellsBLA=−0.007 Åd.

After we established that the HF/STO-3G and CIS/STO-3G theory levels for ground and excited states, respec-tively, are sufficient to reproduce important geometric pa-rameters in the stilbene, they were applied for the geometryoptimization of the molecules under study in the groundsS0dand first singlet excited statessS1d. The conformation of PCPmolecules was taken from the crystal structures of analogs,although qualitative comparison of bond lengths and BLAvalues are not possible due to static disordersdifferent enan-tiomers, occupying the same positions in the crystald.5sfd Thegeometry was optimized assuming the highest possible sym-metry sC2h for DSB andD2 for PCPd, lower symmetrysC2

and Cs for DSB andC2 for PCPd, and with no symmetryconstraints. Highly symmetric structures were found to bethe energy minima for all ground states and for the excitedstates of DSBn and DSBd.C2 was found more stable for theexcited states of both PCPnsby 18.8 kcal/mold and PCPdsby 6.2 kcal/mold. The BLA was found to have differentvalues in nonequivalentp systems forS1 of PCPds0.173 and0.087 Åd, which are close toS0 of DSBds0.174 Åd andS1 ofDSBd s0.080 Åd, respectively. Similar values were found forS1 of PCPn. This already indicates that the relaxation of theexcited-stateS1 in PCP molecules leads to localization of theexcitation on one of the monomer units. Similar effect wasreported previously.32,33

B. The nature of absorbing and emitting excitedstates

Excited states of DSB and PCP molecules were calcu-lated at the TD-B3LYP/6-31G//HF/STO-3G levelsin con-ventional quantum-chemical notation “single point//optimization level”d. Table IV shows computed excitationenergies of optically active transitions up to four lowest ex-cited statessthese correspond to the vertical absorption val-uesd, and compares them to the available experimental datataken at the absorption maxima. Overall calculations agreewell with experiment, given the uncertainty of the spectral-broadening effects. The respective dominant pairs of NTOssRef. 16d for these states are shown in Fig. 1. As expected,DSBn has the only one excited-stateS1 with appreciable os-

cillator strength. It has essentially a highest occupied mo-lecular orbitalsHOMOd to lowest occupied molecular orbitalsLUMOd character, which means that the particle/hole NTOpair is heavily dominated by HOMO/LUMO pair. In thiscase, HOMO describes the distribution of double bonds in aclassical Lewis structure, and LUMO corresponds to aquinoid distribution of double bonds. The amplitudes ofLUMO are somewhat larger for the central phenyl ring at-oms, than for the terminal ones. This effect can be describedas partial charge transfer to the central phenyl ring uponexcitation.

Four frontier orbitals of PCPn can be described as sym-metric and antisymmetric linear combinations of thesemonomer HOMOs and LUMOs. They give rise to the firstfour excited states of PCPn; the corresponding NTOs areshown in Fig. 1. StatesS1 and S4 ssymmetric to symmetricand antisymmetric to antisymmetric combinations, respec-tivelyd have the largest oscillator strengths.S2 andS3 statesacquire some oscillator strength as well due to considerablesteric distortions. Two monomer branches in PCP stronglyinteract via both dipolar and through space mechanisms.4–6

We note that the monomer particle orbitals have increasedamplitudes on the central phenyl rings due to the intramono-mer charge transfer upon excitation. This leads to a throughspace delocalization of the particle in the paracyclophanecore sas seen in the symmetric combination of the particleNTOs in Fig. 2d. The observed energetic splitting betweenexcited states of PCPn strongly depends on the state transi-tion dipoles and, thus, oscillator strengths. We observe sig-nificant s,0.6 eVd splitting between strongly allowedS1 andS4 pair and less pronounceds,0.25 eVd in S2 and S3 ofPCPn. Compared toS1 of DSBn, excited states of the dimerPCPn are redshifted which can be attributed to geometricdistortion, breaking symmetric split of the excited states inthe aggregate around midpoint.4 Similar trends are observedin the DSBd/PCPd monomer/dimer pairsselected orbitals areshown in Fig. 2d.

After the geometry of the excited state relaxes, themonomer units are no longer symmetrically equivalent: boththe hole and particle are localized mostly on one of them, inagreement with the BLA values discussed above. Changes inthe electronic structure of the excited state upon localizationare reflected in the transition orbitals shown in Fig. 2. While

TABLE IV. Calculatedsgas phased and experimentalstoluene, Ref. 8d ab-sorption wavelengthl, calculated oscillator strengthsfcalc, and experimentalrelative intensityIexp.

State lcalc, nm seVd lexp, nm seVd fcalc

DSBn S1 351 s3.33d 361 s3.43d 1.87PCPn S1 418 s2.96d 397 s2.97d 0.63

S2 403 s3.08d ¯ 0.04S3 372 s3.33d ¯ 0.26S4 345 s3.58d 343 s3.62d 1.56

DSBd S1 358 s3.46d 361 s3.43d 1.68PCPd S1 426 s2.91d 397 s3.12d 0.54

S2 410 s3.02d ¯ 0.03S3 383 s3.23d ¯ 0.24S4 353 s3.51d 338 s3.67d 1.57

224505-4 Masunov et al. J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 5: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

FIG. 1. sColord Natural transition orbitals of DSBn and PCPn in the ground-sgd state geometry. The percentage indicates a fraction of the NTO paircontribution to a given electronic excitation.

FIG. 2. sColord Natural transition orbitals for the first excited state of DSBd and PCPd in the ground-sgd and excited-sed state geometries.

224505-5 Solvent effects on electronic structure of paracyclophane J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 6: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

neither particle nor hole orbitals of DSB change significantlyupon relaxationsapart from the arbitrary sign of the wavefunctiond, both hole and especially particle of PCP essen-tially localize on the bottom monomer. Such localization ofthe excitation on one of the disjoint-conjugated fragmentswas previously reported in the theoretical studies32 and de-tected experimentally.32 Comparison of the transition orbitalsin Figs. 1 and 2 demonstrates that whilemeta-metoxy sub-stituents somewhat increase involvement of thepara-carbonatom of the terminal phenyl group in both particle and holeorbitalssinductive effectd, lone pairs of the oxygen atoms donot play any appreciable rolesno p-resonance effectd.

The excitation energies and radiative lifetimes calculatedat the optimized geometry of the first excited state are re-ported in Table V. One can see that the calculated and ex-perimental transition energies agree within 0.3 eV for boththe absorption and emission processes, which is typical forTDDFT. The radiative lifetimes for all molecules are in ex-cellent agreement with experiment.

C. Polarizable continuum model and complexeswith explicit solvent molecules

The polarizable continuum modelsPCMd of solvation isan extension of solvent reaction field models, introduced byBorn, Kirkwood, and Onsager for the charge distributions ina spherical cavity surrounded by a dielectric medium.34 ThePCM uses a more realistic shape of the cavity, simulates thedielectric response with discrete charges on the cavity sur-face, and includes nonpolar contributions to the solvationfree energy in addition to the electrostatic component. Vari-ous formalisms of PCM have appeared in the literature. Theydiffer in the implementation details, but typically give simi-lar results for neutral molecules. Most of their differencescan be attributed to the empirical parameters usedsatomicvan der Waals radii, etc.d. We used integral equation formal-ism sIEFd, which was developed within two alternativeformulations.10,35 These formulations were shown to beequivalent,36 and only the first one is implemented in theGAUSSIAN03 code.10 When applied to electronic transitions,the dielectric response is sometimes separated into fastselec-tronicd, and slowsorientationald components with differentdielectric constantssnonequilibrium PCMd. Here we usenonequilibrium PCM which is introduced10 in GAUSSIAN03

sRef. 9d and was reported to increase the solvent effect on theabsorption spectra by 8%–10%35sbd compared to the single-component model.

Even though the quality of PCM for simulation of sol-vent effects on NMR, IR, and Raman properties has beenestablished, this method has demonstrated only moderately

encouraging results when describing solvent effects on theabsorption spectra of small polar moleculesspyridazine, py-rimidine, pyrasine,35sbd s-tetrazine,37 and formaldehyde38d. Inaprotic sespecially, weakly polard solvents the PCM overes-timates solvatochromic shifts from 30% to 300%. In waterthe PCM alone recovers 30%–50% of the experimental gas/water shifts, and additional explicit water molecules formingH bonds with the solutesat least one for each lone paird werefound necessary to recover the rest.39 In contrast, for acroleinthe PCM quantitatively describes the blue and redshifts inn-p* andp-p* excitations, but the addition of explicit watermolecules leads to overestimation of these shifts.40 For non-polar solutes there is also only qualitative agreements0.13 eV versus an experimental gas/heptane shit of 0.19 eVfor stilbene,41 and 0.18 versus an experimental penthane/methanol shift of 0.48 eV for methylcyclopropene42d. Simu-lations of solvent effects on the fluorescence spectra arescarce, but the available data indicate that PCM is qualita-tively correct.43,44

Dipolar molecules are known to generate a strong reac-tion field, and thus the solvent strongly affects their geometryand BLA values.45 The nonpolar molecules studied here havevanishing dipole moment in the ground statesdue to symme-tryd and, therefore, they are neither expected to generate astrong solvent response, nor demonstrate large geometric re-laxation in a polar solution. Indeed, ground state of PCPd hasa vanishing dipole moment by symmetry, and optimization inwater using the PCM model yields a structure that is only0.4 kcal/mol lower in energy, and 0.001 Å different in BLAvalue. In the excited-state geometry the symmetry is lost, butthe dipole moments are rather smalls1.2 D in the excitedstate, and 0.8 D in the ground stated and directed perpendicu-lar to the monomer planes. The change in the BLA value forthe excited monomer is somewhat largers0.005 Åd, but therelaxation in the total energy remains almost the same. Notsurprisingly, the absorption spectra calculated for the DSBdand PCPd molecules with the PCM model demonstrate neg-ligible shifts relative to the vacuum values. The emissionspectra were calculated in the relaxed geometries of the firstexcited state with the equilibrium PCMswhere both compo-

FIG. 3. sColord Stabilization of PCPd in the twisted conformation by twoexplicit water moleculessH bonds are shown in blued.

TABLE V. Calculatedsgas phased and experimentalstoluene, Ref. 8d fluo-rescence wavelengthsl and radiative lifetimest.

lcalc snmd seVd lexp snmd seVd tcalc snsd texp snsd

DSBn 428s2.90d 390 s3.18d 1.0 1.3PCPn 494s2.51d 454 s2.73d 3.3 2.9DSBd 434s2.86d 391 s3.17d 1.1 1.1PCPd 493s2.52d 454 s2.73d 3.7 3.9

224505-6 Masunov et al. J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 7: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

nents of the reaction field are generated by the charge distri-bution in this excited stated. They demonstrate larger toluene/water shiftss,35 nmd, closer to the experiment. However, a20% decrease in the lifetime is in clear discrepancy withexperiment. Similar trendsssmaller in valuesd were obtainedfor the PCPn molecule.

In order to overcome the limitations of the polarizablecontinuum model, we built several complexes of PCPd withexplicit solvent molecules. Out of several complexes studied,only one of themsFig. 3d had a significantly longer lifetime

for the first excited state. In this complex two water mol-ecules each form two H-bonded bridges between the meth-oxy sidechains of the different monomers. These bridges arestrong enough to stabilize the monomers in the twisted con-formation. In the ground state the H-bonding stabilization is5 kcal/mol at the HF/STO-3G levelsvs. 8 kcal/mol for thesimilar bridges linking methoxy side chains in the planarconformationd. The twist of the phenyl groups disruptspconjugation and leads to the decrease of the transition dipolemoment and, subsequently, considerably increases the radia-

FIG. 4. sColord Natural transition orbitals for the first excited state of DSBd and PCPd in the ground-sgd and excited-sed state geometries in vacuum and withcounterion at the equilibrium distance.

224505-7 Solvent effects on electronic structure of paracyclophane J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 8: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

tive lifetime. A similar mechanism is even less possible forPCPn, however, as it has no heteroatoms. Initial geometriesof its complexes with water were built in such a way thatwater molecules formed H bondssCH¯O and OH̄ pd,and n¯p donor-acceptor interactions. The excitation ener-gies of these complexessnot shownd were found to be shiftedby less than 3 nm, an insignificant shift compared to thetoluene/water emission wavelength shifts of,50 nm, ob-served in experiment. Nevertheless, such observation of sol-vent stabilization of conformationally distorted PCP mol-ecules indicates possible existence of an ensemble of soluteconformers at the room temperature.

Overall we conclude that both implicit and explicit sol-vent models fail to reproduce the correct trends, observed inthe photoluminescence spectra of PCP chromophores.

D. Complexes with counterions

To understand the nature of the observed experimentallylarge solvatochromic shifts, one should keep in mind that inorder to make the chromophores soluble in a polar solvent,they had been derivatized with ionogenic side chains. Eventhough the ionic groups are far removed from thep systemsand their charge is largely screened out by the solvent, coun-terions are present in the solvation shells and may form bothcontact- and solvent-separated complexes with chro-mophores.

To simulate this possibility, one counterionsNMe4+ in the

case of DSBd and PCPd, and BrsH2Od6− in the case of DSBn

and PCPnd was placed on the twofold axis of the moleculeapproaching the central ring cofacially. This lowered thesymmetry fromD2 to C2 in the ground state of PCP mol-ecules and made the monomers nonequivalent, similar to theexcited-state geometry. The complex of the solute with twoions on the opposite sides of molecule, which would retainD2 symmetry, seems to be improbable and its simulation wasnot attempted. Such a complex is presumed to be unstable,unlike solvent-separated complex between ions in aqueoussolutions.46 The main stabilizing factor in the solvent-separated like-charged ion pair is the existence of the bridg-ing water molecules. Such bridges cannot form when thespace between ions is occupied by the hydrophobic solute.Therefore, repulsion between the like-charged ions is ex-pected to be strong in aqueous solution.

While optimization in vacuum may describe the solutemolecule relatively well, the complex between the chro-mophore and the ion would require molecular-dynamics

simulations with explicit water molecules,46 which is muchmore computationally expensive. Instead, we optimized thechromophore-counterion pairsdenoted DSB·ci and PCP·cid,and then increased the ion-solute separation by 2 ÅsdenotedDSB¯ci and PCP̄ cid to simulate the separation and par-tial screening effect by the solvent.

The effect of the counterion on electronic structure isillustrated in Fig. 4. In the ground-state geometry ofDSBd·ci, the cation decreases the participation of the atomicp orbitals of the central phenyl ring in the hole orbitalsre-pulsiond, and increases their participation in the particle or-bital sattractiond, so that the excitation acquires a character ofcharge transfer from the terminal to the central phenyl ring.Similar trend is observed in the excited-state geometrysnotshownd. For PCPd·ci the effect is much more dramatic. Cat-ion completely localizes the electron on the nearest monomerbranch, and the hole on the remote branch. Instead of beingdelocalized between the monomerssin the ground-state ge-ometryd, or localized on one monomersin the excited-stategeometryd, the excitation almost entirely represents chargetransfer from one monomer to another.

The calculated electronic properties of complexes at sev-eral geometries are summarized in Table VI. As one can see,both solvatochromic shifts of photoluminescence and in-crease in the excited-state lifetimes can be well described bythe chromophore-ion pair. Addition of BrsH2Od6

− ion toDSBn nearly doubles the radiative lifetime and redshifts thephotoluminescence maximum by 10 nm in excellent agree-ment with experiment. Blueshift of the absorption maximumis considerably overestimated. This may indicate that in theground-state DSBn, ion pair is more spatially separated, sothat the electrostatic field of the ion is largely screened off bythe solvent environment. The approach of the ion to the chro-mophore molecule in the excited state is likely to be due tothe relatively high polarizability of the chromophore mol-ecule in the excited state, which increases charge-induceddipole attraction. In the case of DSBd·NMe4

+ complex, theopposite charge of the ion increases the partial charge trans-fer to the central phenyl ring in the excited state. As a result,the redshift instead of the blueshift of the absorption maxi-mum is predicted.

Similar redshifts of the photoluminescence maxima andincrease in the radiative lifetimes are observed in the equi-librium structure of PCPn·BrsH2Od6

− complex, but the mag-nitude of both effects is 5–20 times stronger. Such a dramaticeffect of the counterion on the radiative lifetime of PCP chro-mophores compared to that of DSB monomers can be under-

TABLE VI. Calculatedscounterion complex vs gas phased and experimentalswater vs toluene, Ref. 8d shifts inabsorptionsAbs.d and photolumenescenesPLd wavelengthsl and respective differences in radiative lifetimest.

PLDlcalc snmd

PLDlexp snmd Dtcalc, ns Dtexp, ns

Abs.Dlcalc snmd

Abs.Dlexp snmd

DSBn·BrsH2d6− 11 18 1.0 1.0 −20 −5

DSBd·NMe4+ 20 24 0.2 0.3 22 −5

PCPn·BrsH2Od6− 56 42 12.7 20.6 −7 −5

PCPn̄ BrsH2Od6− 41 42 7.9 20.6 −7 −5

PCPd·NMe4+ 128 57 64.5 29.0 −7 −2

PCPd̄ NMe4+ 70 57 25.1 29.0 −8 −2

224505-8 Masunov et al. J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 9: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

stood using transition orbitals. The approaching ion signifi-cantly increases the charge-transfer character of theexcitation sfrom one monomer unit to anotherd. In thecharge-transfer state overlap between hole and particle orbit-als is small, so that the transition dipole decreases and thelifetime dramatically increases in agreement with experi-ment. In the case of the PCPd·NMe4

+ complex the magnitudeof the calculated counterion effects significantly increases,while experimental values increase only slightly. This islikely due to the truncation of the aliphatic chains when theNBu4

+ ions present in the experiment are replaced with themodel NMe4

+. As a result, optimal chromophore-ion distanceis artificially shortened, and enlarged chromophore-ion sepa-ration s2 Å versus equilibriumd was found to be a sufficientcompensation.

When the PCM was applied to the counterion com-plexes, the excitation energies and oscillator strengths of thePCP molecules and the corresponding PCP·ci complexesshowed very little differencessnot shownd. This may be dueto the fact that the solvent around the ion is much morestructured than the bulk solvent, and the polarizable con-tinuum model using the bulk dielectric constant fails to re-produce this effect.

IV. CONCLUSIONS

The present theoretical study investigates the absorptionand fluorescent properties of PCP-based chromophoresshown in scheme 1 in the solvent environment. We use com-putationally inexpensiveab initio approaches to obtain thegeometry of the ground and excited statessHF/STO-3G andCIS/STO-3G theory levels, respectivelyd and show that thesemethods reproduce important geometric features of similarp-conjugated molecular systems with acceptable accuracy.The time-dependent density-functional theorysat TD-B3LYP/6-31G leveld is utilized next to describe excitationenergies and radiative lifetimes for transition states. Com-parison of the ground-sabsorptiond and excited-semissiondstate geometries and correspondent natural transition orbitalsof isolated moleculesin vacuoreveals that the emitting stateis essentially localized on a single monomer branch, whereasthe absorbing state remains always delocalized.

To understand possible mechanisms for the strong shiftsin the experimentally observed emission maxima and radia-tive lifetimes in polar versus nonpolar solvents, several sol-vent models have been tested. Polarizable continuum modelis able to reproduce both small solvatochromic wavelengthshifts in the absorption spectra, and larger shifts in the fluo-rescence spectra. However, decrease in the radiative lifetimepredicted by PCM disagrees with experiment. The theoreticalmodeling with explicit solvent shows that water moleculescan stabilize the terminal phenyl groups of PCPd in thetwisted conformation Fig. 3 by forming hydrogen-bondedbridges between polar substituents on the different monomerbranches. However, it leads to the blue solvatochromic shiftfor PCPd molecules, in contrast to experiment findings.Moreover, this mechanism cannot be applied to PCPn mol-ecules, where absence of oxygen atoms excludes the forma-

tion of hydrogen bonds. Thus, neither polarizable continuummodel nor explicit water molecules are able to account forexperimentally observed behavior.

A possible scenario, which enables us to reproduceclosely the experimental data, is the formation of a complexbetween the chromophore and counterion in a polar solvent.Excited state in these complexes can be described as ion-induced charge transfer. Based on our simulations, themechanism of the increase in radiative lifetimes of the emit-ting excited state can be described as follows. In nonpolarsolvent, the excitationsboth the hole and the particled is lo-calized on the same branch of PCP chromophore. Increasedpolarizability of the chromophore in the excited state attractsthe nearby ion by charge-induced dipole interaction. The ap-proaching ion then localizes the hole and the particle on thedifferent branches of PCP chromophore, thus decreasing theoverlap between them, and, therefore, the transition dipole.In turn this leads to a dramatic increase of the fluorescencelifetimes.8 A similar unusually large Stokes shift and fivefoldincrease in the lifetime were recently reported for tetracy-cline forming complexes with divalent cations.47

In the solvents of weaker polarityfsuch as dimethyl for-mamidesDMFdg, counterions are expected to form contaction pairs with the charged side chains, rather than with po-larizable yet neutral PCP core. The contribution of the ion-induced charge-transfer state will consequently diminish.Therefore, excited-state lifetime in less polar solvents will besimilar to that of the neutral derivative, which was indeedobserved in experiment.8 An increase in the radius of coun-terion by exchange or complexationsfor instance, with thecrown etherd should lead to a similar result. This can be usedas an experimental verification of the proposed mechanism.

The mechanisms described above may have implicationson the photophysics of organic molecules and polymers usedin organic light-emitting devices and nanoassemblies, as wellas soluble conducting polymers used as biosensors. Based onthese mechanisms, fluorescent signal from long-lived inter-molecular charge-transfer state should decrease if counteri-ons of larger size are used, and increase with ionic strengthin the solution. To reduce unspecific wavelength dependenceof the biosensor on the polarity of environment, the face ofp-electron system should be stericaly protected from the ap-proaching ions.

In organic light-emitting devicessOLEDsd, for example,a thin LiF layer was found to improve the efficiency of elec-tron injection from aluminum catode into organic layer oftris-s8-hydroxyquinolined aluminumsAlqd.48 Several mecha-nisms of this improvement have been proposed, includingformation of ion pairs between Li+ and Alq anion radicals,48

and charge-transfer complexes between fluoride anion andAlq.49 The results of our work suggest that either ion caninduce intermolecular charge-transfer complex in a stack oftwo or morep-conjugated molecules. This indicates the im-portance of molecular packing and ion intercalation inOLED design.

ACKNOWLEDGMENTS

We wish to thank Dr. Glenn P. Bartholomew for usefuldiscussions, and Dr. Antonio Redondo for his help with the

224505-9 Solvent effects on electronic structure of paracyclophane J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp

Page 10: Theoretical study of the effects of solvent environment on ...cnls.lanl.gov/~serg/postscript/JChemPhys_122_224505.pdf · The PCP core holds two p-conjugated systems cofacially sscheme

manuscript. The research at UCSB is supported by Mitsub-ishi Chemical Center for Advanced Materials, and the re-search at LANL is supported by the Center for NonlinearStudies, and the LDRD program of the US Department ofEnergy. This support is gratefully acknowledgedsLA-UR-04-7373d.

1sad M. A. Fox, Acc. Chem. Res.32, 201 s1999d; sbd G. D. Scholes, J.Phys. Chem.100, 18731s1996d.

2D. T. McQuade, A. E. Pullen, and T. M. Swagert, Chem. Rev.sWashing-ton, D.C.d 100, 2537s2000d.

3sad S. Canuto and M. C. Zerner, Chem. Phys. Lett.157, 353 s1989d; sbdS. Canuto and M. Zerner, J. Am. Chem. Soc.112, 2114s1990d.

4sad G. C. Bazan, W. J. Oldham, Jr., R. J. Lachicotte, S. Tretiak, V.Chernyak, and S. Mukamel, J. Am. Chem. Soc.120, 9188s1998d; sbd S.Wang, G. C. Bazan, S. Tretiak, and S. Mukamel,ibid. 122, 1289s2000d;scd S. Tretiak and S. Mukamel, Chem. Rev.sWashington, D.C.d 102,3171 s2002d.

5sad Y. Morisaki and Y. Chujo, Macromolecules37, 4099s2004d; sbd D. S.Seferos, D. A. Banach, N. A. Alcantar, J. N. Israelachvili, and G. C.Bazan, J. Org. Chem.69, 1110s2004d; scd J. Gierschner, H. G. Mack, D.Oelkrug, I. Waldner, and H. Rau, J. Phys. Chem.108, 257s2004d; sdd R.Salcedo, N. Mireles, and L. E. Sansores, J. Theor. Comput. Chem.2, 171s2003d; sed L. Guyard, C. Dumas, F. Miomandre, R. Pansu, R. Renault-Meallet, and P. Audebert, New J. Chem.27, 1000 s2003d; sfd G. P.Bartholomew and G. C. Bazan, J. Am. Chem. Soc.124, 5183s2001d; sgdG. P. Bartholomew and G. C. Bazan, Acc. Chem. Res.34, 30 s2001d; shdN. Verdal, J. T. Godbout, T. L. Perkins, G. P. Bartholomew, G. C. Bazan,and A. M. Kelley, Chem. Phys. Lett.320, 95 s2000d; sid W. J. Warren, Jr.,Y.-J. Miao, R. J. Lachicotte, and G. C. Bazan, J. Am. Chem. Soc.120,419 s1998d.

6sad X. Zhou, A.-M. Ren, J.-K. Feng, and X.-J. Liu, J. Phys. Chem. A107,1850 s2003d; sbd G. P. Bartholomew, I. Ledoux, S. Mukamel, G. C.Bazan, and J. Zyss, J. Am. Chem. Soc.124, 13480s2002d; scd J. Zyss, I.Ledoux, S. Volkov, V. Chernyak, S. Mukamel, G. P. Bartholomew, andG. C. Bazan,ibid. 122, 11956s2000d.

7G. P. Bartholomew, M. Rumi, S. J. K. Pond, J. W. Perry, S. Tretiak, andG. C. Bazan, J. Am. Chem. Soc.126, 11529s2004d.

8sad J. W. Hong, B. S. Gaylord, and G. C. Bazan, J. Am. Chem. Soc.124,11868s2002d; sbd J. W. Hong, H. Benmansour, and G. C. Bazan, Chem.-Eur. J. 9, 3186 s2003d; scd J. W. Hong, H. Y. Woo, B. Liu, and G. C.Bazan, J. Am. Chem. Soc.127, 7435s 2005d.

9M. J. Frisch, G. W. Trucks, H. B. Schlegelet al., GAUSSIAN 03, RevisionC.02, Gaussian, Inc., Wallingford, CT, 2004.

10sad E. Cances, B. Mennucci, and J. Tomasi, J. Chem. Phys.107, 3032s1997d; sbd B. Mennucci, R. Cammi, and J. Tomasi,ibid. 109, 2798s1998d.

11A. Masunov and S. Tretiak, J. Phys. Chem. B108, 899 s2004d.12C. Katan, F. Terenziani, O. Mongin, M. H. V. Werts, L. Porrès, T. Pons, J.

Mertz, S. Tretiak, M. Blanchard-Desce, J. Phys. Chem. A109, 3024s2005d.

13sad E. Runge and E. K. U. Gross, Phys. Rev. Lett.52, 997s1984d; sbd M.E. Casida, inRecent Advances in Density-Functional Methods, edited byD. A. ChongsWorld Scientific, Singapore, 1995d, Vol. 3.

14J. Fabian, L. A. Diaz, G. Seifert, and T. Niehaus, J. Mol. Struct.:THEOCHEM 594, 41 s2002d.

15D. A. McQuarrie and J. D. Simon,Physical Chemistry: A MolecularApproachsUniversity Science Books, Sausalito, CA, 1997d.

16R. L. Martin, J. Chem. Phys.118, 4775s2003d.17A. Kokalj, J. Mol. Graphics Modell.17, 176s1999d; Code available from

http://www.xcrysden.org.

18S. P. Kwasniewski, L. Claes, J. P. Francois, and M. S. Deleuze, J. Chem.Phys. 118, 7823s2003d.

19M. Traetteberg, E. B. Frantsen, F. C. Mijlhoff, and A. Hoekstra, J. Mol.Struct. 26, 57 s1975d.

20J. A. Bouwstra, A. Schouten, and J. Kroon, Acta Crystallogr., Sect. C:Cryst. Struct. Commun.40, 428 s1984d.

21J. Harada and K. Ogawa, J. Am. Chem. Soc.123, 10884s2001d.22C. H. Choi and M. Kertesz, J. Phys. Chem. A101, 3823s1997d.23B. B. Champagne, J. F. Pfanstiel, D. F. Plusquellic, D. W. Pratt, W. M.

Van Herpen, and W. L. Meerts, J. Phys. Chem.94, 6 s1990d.24S. Tretiak, K. Igumenshchev, and V. Chernyak, Phys. Rev. B71, 033201

s2005d.25N. Kobko, A. Masunov, and S. Tretiak, Chem. Phys. Lett.392, 444

s2004d.26sad A. Karpfen, C. H. Choi, and M. Kertesz, J. Phys. Chem. A101, 7426

s1997d; sbd J. C. Sancho-Garcia, J. L. Bredas, and J. Cornil, Chem. Phys.Lett. 377, 63 s2003d; scd J. C. Sancho-Garcia and J. Cornil, J. Chem.Phys. 121, 3096s2004d.

27C. H. Choi, M. Kertesz, and A. Karpfen, J. Chem. Phys.107, 6712s1997d.

28A. Masunov and J. J. Dannenberg, J. Phys. Chem. A103, 178 s1999d.29V. Molina, M. Merchán, and B. O. Roos, J. Phys. Chem.101, 3478

s1997d.30J. Gierschner, H. G. Mack, L. Luer, and D. Oelkrug, J. Chem. Phys.116,

8596 s2002d.31W.-Y. Chiang and J. Laane, J. Chem. Phys.100, 8755s1994d.32sad E. Zojer, P. Buchacher, F. Wudl, J. Cornil, J. P. Calbert, J. L. Bredas,

and G. Leising, J. Chem. Phys.113, 10002s2000d; sbd I. Franco and S.Tretiak, J. Am. Chem. Soc.126, 12130s2004d.

33D. A. Bussian, M. A. Summers, B. Liu, G. C. Bazan, and S. K. Buratto,Chem. Phys. Lett.388, 181 s2004d.

34sad M. Born, Z. Phys.1, 45 s1920d; sbd J. G. Kirkwood, J. Chem. Phys.2, 351 s1934d; scd L. Onsager, J. Am. Chem. Soc.58, 1486s1936d.

35sad D. M. Chipman, J. Chem. Phys.112, 558s2000d; sbd M. Cossi and V.Barone,ibid. 115, 4708s2001d.

36M. A. Aguilar, J. Phys. Chem. A105, 10393s2001d.37C. Adamo and V. Barone, Chem. Phys. Lett.330, 152 s2000d.38C. Zazzaa, L. Bencivenni, A. Grandia, and A. Aschi, J. Mol. Struct.:

THEOCHEM 680, 117 s2004d.39F. Aquilante, M. Cossi, O. Crescenzi, G. Scalmani, and V. Barone, Mol.

Phys. 101, 1945s2003d.40F. Aquilante, V. Barone, and B. O. Roos, J. Chem. Phys.119, 12323

s2003d.41R. Improta, F. Santoro, C. Dietl, E. Papastathopoulos, and G. Gerber,

Chem. Phys. Lett.387, 509 s2004d.42B. Mennucci, J. Tomasi, K. Ruud, and K. V. Mikkelsen, J. Chem. Phys.

117, 13 s2002d.43W. G. Han, T. Q. Liu, F. Himo, A. Toutchkine, D. Bashford, K. M. Hahn,

and L. Noodleman, ChemPhysChem4, 1084s2003d.44T. Q. Liu, W. G. Han, F. Himo, G. M. Ullmann, D. Bashford, A. Toutch-

kine, K. M. Hahn, and L. Noodleman, J. Phys. Chem. A108, 3545s2004d.

45S. Corni, C. Cappelli, M. Del Zoppo, and J. Tomasi, J. Phys. Chem. A107, 10261s2003d.

46A. Masunov and T. Lazaridis, J. Am. Chem. Soc.125, 1722s2003d.47S. Schneider, M. O. Schmitt, G. Brehm, M. Reiher, P. Matousek, and M.

Towrie, Photochem. Photobiol. Sci.2, 1107s2003d.48M. G. Mason, C. W. Tang, L. S. Hunget al., J. Appl. Phys.89, 2756

s2001d.49Y. Yuan, D. Grozea, S. Han, and Z. H. Lu, Appl. Phys. Lett.85, 4959

s2004d.

224505-10 Masunov et al. J. Chem. Phys. 122, 224505 ~2005!

Downloaded 30 Dec 2005 to 128.165.22.22. Redistribution subject to AIP license or copyright, see http://jcp.aip.org/jcp/copyright.jsp