Computational Study of the Structure, Dynamics, and Photophysical

the calculation and often avoid some of the convergence problems in explicit models, where ...... Stuart, S. J.; Dickson, B. M.; Sumpter, B. G.; N...
9 downloads 0 Views 803KB Size
J. Phys. Chem. B 2005, 109, 7671-7685

7671

Computational Study of the Structure, Dynamics, and Photophysical Properties of Conjugated Polymers and Oligomers under Nanoscale Confinement Bobby G. Sumpter,*,† Pradeep Kumar,‡ Adosh Mehta,§ Michael D. Barnes,| William A. Shelton,† and Robert J. Harrison† Oak Ridge National Laboratory, Oak Ridge, Tennessee 37831, and Department of Chemistry, UniVersity of Massachusetts, Amherst, Massachusetts 01003 ReceiVed: NoVember 23, 2004

Computational simulations were used to investigate the dynamics and resulting structures of several paraphenylenevinylene (PPV) based polymers and oligomers (PPV, 2-methoxy-5-(2′-ethyl-hexyloxy)-p-phenylenevinylene f MEH-PPV and 2,5,2′,5′-tetrahexyloxy-7,8′-dicyano-p-phenylenevinylene f CN-PPV). The results show how the morphology and structure are controlled to a large extent by the nature of the solutesolvent interactions in the initial solution-phase preparation. Secondary structural organization is induced by using the solution-phase structures to generate solvent-free single molecule nanoparticles. Isolation of these single molecule nanostructures from microdroplets of dilute solution results in the formation of electrostatically oriented nanostructures at a glass surface. Our structural modeling suggests that these oriented nanostructures consist of folded PPV conjugated segments with folds occurring at tetrahedral defects (sp3 C-C bonds) within the polymer chain. This picture is supported by detailed experimental fluorescence and scanning probe microscopy studies. We also present results from a fully quantum theoretical treatment of these systems which support the general conclusion of structure-mediated photophysical properties.

I. Introduction The confinement of materials at the nanoscale and the resulting alterations of their properties and chemistry have been and continue to be a subject of considerable excitement and interest in numerous science and technology sectors.1 Recent results have shown dramatic effects in this regard including profound changes in kinetics and reaction products of organic molecules in porous media,2 altered structures and enhanced melting points of proteins and other macromolecules,3 new types of fluid dynamics,4 modified fluorescence lifetimes,5 and the modification of the structure of liquids.6 For many macromolecules, in particular polymers, three-dimensional confinement at a nanometer size scale is often comparable to a polymer’s radius of gyration and is of special interest in the context of so-called collapse transitions associated with semiconducting polymers and how the confinement affects intra and interchain organization and associated photophysical properties.7-13 However, commonly used techniques capable of producing these types of structures such as thin-film or self-assembly processes can suffer from substrate interactions which may dominate or obscure the underlying polymer physics. In a departure from conventional solvent-cast film studies, we have recently explored ink-jet printing methods as a means of isolating single chains of conjugated polymers from micronsized droplets of dilute polymer solution. This method is based on using droplet-on-demand generation to create a small drop consisting of a very dilute polymer mixture in a solvent.14,15 In this way, the self-organization of the polymer chain proceeds without interaction of a substrate or host polymer. Applications * Corresponding author. E-mail: [email protected]. † Computer Science and Mathematics Division. ‡ Chemical Sciences Division. § Life Sciences Division, Oak Ridge National Laboratory. | Department of Chemistry, University of Massachusetts.

of these types of particles take advantage of high surface area and confinement effects, leading to interesting nanostructures with different properties that cannot be produced using conventional methods. Clearly, there is extraordinary potential for developing new materials in the form of bulk, composites, and blends that can be used for coatings, optoelectronic components, magnetic media, ceramics and special alloys, micro- or nanomanufacturing, and bioengineering. The key to beneficially exploiting these interesting materials is a detailed understanding of the connection of nanoparticle technology to atomic and molecular origins of the process. The question of morphological control of individual chains of conjugated polymers is important both from the standpoint of fundamental physical understanding of interchain interactions16 as well as from the standpoint of the development of nanoscale polymer-based optoelectronic devices.17 There has been considerable interest in these organic conjugated systems because they provide the basis of novel materials that combine optoelectronic properties of semiconductors with the mechanical properties and processing advantages of plastics. It is relatively easy to functionalize the polymer backbone with a variety of flexible side groups that enhance processability from solution into thin films. However, despite the enormous versatility for practical applications, the fundamental physics underlying the optimization of the optoelectronic properties has remained elusive. Much of the problem in this regard is centered about a poor understanding of the interactions between conjugated polymer chains in solutions and films. It is clear that electronic structure of conjugated polymers depends sensitively on the physical conformation of the polymer chains and the way the chains pack together. Of the large class of conjugated polymers, polyphenylene vinylene (PPV) derivatives have received a great deal of attention in the context of polymer-based optoelectronic devices

10.1021/jp0446534 CCC: $30.25 © 2005 American Chemical Society Published on Web 03/26/2005

7672 J. Phys. Chem. B, Vol. 109, No. 16, 2005 because of its efficient luminescence and charge-transport properties. A single PPV macromolecule may be described structurally as a large number (20-200) of stiff-chain (conjugated) segments with average conjugation length of between 6 and 12 monomer units. The conjugated segments, linked by socalled “tetrahedral defects”, can act as a local optical chromophore within the molecule where the interaction between chromphoric segments depends sensitively on solvent and filmprocessing parameters.18-22 Currently, much insight has been gained through single molecule spectroscopy into exciton dynamics, photochemical stability, and chain organization of PPV-based polymer molecules isolated in dilute thin films.16,23-26 In this paper, we describe our recent work on investigating the effects of three-dimensional confinement on structural and photoluminescence properties of single molecules of conjugated organic polymers. By using a combination of state-of-the-art experimental and computational techniques, we have gained new insight into the organization of stiff-chain polymers confined to nanoscale domains and show that the photophysical properties of these systems are profoundly altered as a result. II. Summary of the Experimental Evidence for Oriented Polymer Nanostructures In recent experiments, we used ink-jet printing methods to isolate single conducting polymer chains in microdroplets in various organic solvents (tetrahydrofuran, toluene, etc.). The droplets generated were typically less than 5-µm initial diameter27,28 and evaporate in route to the coverglass substrate thus allowing self-organization of the polymer chain in the absence of host polymer or substrate interactions. The probe polymer used in our experiments was (poly[2-methoxy-5-(2′-ethylhexyloxy)-1,4-phenylene vinylene]), or MEH-PPV, a polymer commonly used in organic optoelectronic devices. MEH-PPV concentrations typically were in the range of 10-11-10-12 M (with concentrations as low as 10-14 M) ensuring that the nanostructures probed in our experiments are single MEHPPV chains. A combination of dipole emission pattern imaging, polarization-modulated fluorescence,27,28 photon-correlation statistics,29 and atomic force microscopy27 was used to probe structure and morphology of these interesting species. Together, these measurements provide a clear picture of intramolecular organization within individual macromolecules and the modification of photoluminescence properties that derive from the novel orientation. Figure 1 shows a typical high-resolution fluorescence image of single MEH-PPV molecules adsorbed on a glass surface after microdroplet deposition.27 The “donutlike” spatial intensity patterns seen in the fluorescence image are characteristic of single dipole emitters oriented parallel to the optic (Z) axis. These unique spatial intensity patterns derive from the fact that a single quantum system emits light with a sine-squared angular distribution about the dipole axis, µ.30 Since emission is forbidden at angles along µ, dipoles oriented perpendicular to the substrate show donutlike spatial intensity patterns with a characteristic node in the center. This unusual transition dipole orientation is precisely the opposite of that seen for spin-coated films where all molecules appear to show spatial intensity patterns in fluorescence characteristic of in-plane (parallel to the glass substrate) emitters. Depending on the mode of production, highly uniform oriented samples can be prepared with rms deviation in the polar (tilt) angle, q, of a few percent. Recent electric force microscopy measurements of these oriented nanostructures show that the orientation mechanism derives from

Sumpter et al.

Figure 1. High-resolution image of z-oriented MEH-PPV single molecules acquired under 514.5-nm excitation. The image was acquired for 20 s for better signal-to-noise ratio. The size of the emission pattern does not represent the actual size or morphology of the nanoparticle, only the antenna image.

an electrostatic interaction between the charged nanostructure (where the charges are induced upon droplet production) and SiO- groups on the glass surface.31 Observation of z-oriented dipolar emission patterns from these species provides compelling but not definitive evidence of ordered intramolecular structure. If the radiative recombination site within the molecule is localized, some local 3D orientation of the molecule can be perceived experimentally from the spatial intensity pattern (or linear dichroism measurement). However, in a random-coil structure approximation, each local orientation should logically be expected to be different with no net orientation in the sample. The fact that our ensemble sample shows uniform Z orientation is significant in that this particular orientation can only be explained if the molecules possess a high degree of alignment between conjugated segments. This is because the transition moment for the (1Bu) optically pumped exciton is polarized collinearly with the conjugation axis.32 Thus, the only way for the ensemble of polymer chains to show the same transition moment is for all of them to possess the same up-down/left-right structural identity with respect to the surface. Further evidence of a nanocylindrical geometry was seen by scanning the contact with a modified Digital Instruments Bioscope/Dimension scanner with a Nanoscope III controller. All measurements were made in tapping mode. Particle heights ranging from 7 to 12 nm were seen, in good agreement with the persistence length of MEH-PPV (≈10 monomer units), with a small minority of larger (>20 nm) particles, and consistent with a folded molecular structure. As shown by Barbara and co-workers, polarization modulation spectroscopy makes a convenient probe of the intramolecular organization on individual polymer chains. The principle behind the measurement is that the molecule as a whole may act essentially as an antenna array where any chromophore (conjugated) segment within the chain may absorb an excitation photon and generate an electron-hole pair. The excitonic energy may migrate to one or more low-energy radiative traps within the chain, thus luminescence intensity as a function of excitation polarization (angle of E with respect to lab coordinates) provides an estimate of morphology.18 We showed that the distribution of polarization anisotropy parameters obtained from z-oriented polymer nanostructures differed significantly from that simulated for a so-called “rod” structure.27 This difference indicates a structure with more organization than the other models, one that has a structure similar to that which is shown as computed from molecular dynamics and mechanics simulations. This structure has a high degree of chain organization with parallel chain

Conjugated Polymers and Oligomers segments located to within a relatively short distance (3.5 Å) stacked into a cylindrical morphology (full details of the simulations are given below). The size range (5-15 nm) associated with dimensions of a single chain of a conjugated polymer provides an interesting proving ground for theoretical models, as we are now in a position to directly compare experimental measurements with structural calculations. One further important piece of experimental evidence that sheds some light for the driving force behind the highly ordered and cylindrical morphology of these systems comes from fluorescence correlation spectroscopy (FCS) carried out in the solution phase.33 The FCS measurements were carried out using the same types of freshly prepared solutions and data were acquired periodically up to 100 hours to study the polymer chain dynamics. FCS is based on fluctuations in the fluorescence intensity of the molecules diffusing in and out of a laser focal volume. The dilute MEHPPV and CN-PPV solutions were loaded into a sample cuvette and fluorescence from the solutions was collected from which a correlation curve was fit to a single diffusion coefficient model. From extensive FCS studies, we have been able to show a clear correlation of solution-phase morphologies with the luminescence properties (dipole orientation, photostability, and photoluminescence spectra) of single molecules of various PPV systems. The initial nucleation of the compact structured PPV systems is strongly favored by poorer solvents. In fact, a good solvent does not lead to the formation of a structured system as will also be demonstrated below via molecular dynamics methods. It appears that solution-phase molecular organization may be the dominating factor underlying the generation of the highly structured PPV systems in single molecule nanoparticles with some secondary contributions due to the droplet generation, consistent with earlier work by Schwartz and co-workers.34 To more thoroughly examine and understand the dynamical dependence of the structure and morphology on the type of solvent35 and possible secondary structural changes induced from nanoscale confinement of a single PPV molecule, extensive computational experiments are used. In addition, we consider the effects of structural changes, both intramolecular as well as overall morphological shape, on the excited-state electronic structure. III. Computational and Theoretical Methods The interplay between chemical composition, atomic arrangements, microstructures, and macroscopic behavior makes computational modeling and materials design extremely difficult. Even with the fundamental laws of quantum mechanics and statistical physics, the availability of high-performance computers, and the growing range of sophisticated software systems accessible to an increasing number of scientists and engineers, the goal of designing novel materials from first principles continues to elude most attempts. On the other hand, computational experiments have lead to increased understanding of atomistic origins of molecular structure and dynamics. In particular, Monte Carlo and molecular dynamics and mechanics methods have yielded a wealth of knowledge on the structural behavior of various polymeric materials and their dynamic behavior as a function of temperature and pressure. Quantum chemistry methods, although generally difficult to directly apply to large molecular systems, allow ab initio determination of many-body molecular interactions which can be used in the classically based methods. In addition, these methods are now becoming applicable for systems containing several hundreds of atoms, making quantum mechanics based prediction of

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7673 structure and properties feasible. The combination of all of these computational chemistry methods clearly provides a good framework to examine some of the fundamental questions concerned with the role of solvation in controlling molecular self-organization of conjugated macromolecules and the resulting photophysical properties. III. 1. Molecular Dynamics, Monte Carlo, and Molecular Mechanics. The molecular dynamics, Monte Carlo, and molecular mechanics methods are well-reviewed and proven modeling methods and we only discuss our particular implementations. These methods require the specification of molecular potential energy functions to describe the various many-body interactions common in molecular systems. Since our interest here is to understand the structure and morphology in solution and in dry, single molecule nanoparticles, we have elected to use potential functions with a proven record of accurate prediction of structure and vibrational spectra for conjugated organic molecules. These potentials are harmonic or Morse oscillators for the bond-stretching and angle-bending terms (both in- and out-of-plane), truncated Fourier series for the torsion interactions (regular dihedral and improper), and Lennard-Jones 6-12 plus Coulomb potentials for the nonbonded interactions. Several standard force fields fall into this category, such as the MM2, MM3, MM4, Dreiding, UFF, MMFF, CHARMM, AMBER, GROMOS, TRIPOS, and OPLS models.36 In the present study, we have used parameters defined within the MM3 model as this particular parametrization has proven to give very accurate results for structural optimizations of many conjugated organic molecules.37 Of particular importance is the capability of the MM3 model to account for intermolecular interactions of the π-electron densities through the dependence of the stretching and torsion terms on iterative self-consistent field evaluations for the relevant π-conjugated bonds. The overall reliability of this model for structural calculations has continually been demonstrated for numerous aromatic compounds (benzene, biphenyl, annulene)38 and conjugated systems (trans-stilbene and even multiple oligomers of PPV).39 We have also verified many of the MM3 predicted structures by comparing to those obtained from calculations with no assumed potentials, for example, ab initio quantum mechanics. These results indeed suggest that MM3 structural predictions for PPV systems are quite accurate. This is very fortunate from a computational standpoint, as our investigations and those of others clearly point out that extended -π conjugated systems present major challenges to current DFT exchange-correlation functionals (nonlocal parts of exchange are not treated adequately). Thus, structural optimization of these systems often requires using many-body electronic theory (CI, CC, or MBPT), which scale at best as O(N5). Monte Carlo methods used in macromolecular science generally begin by constructing a Markov Chain generated by the Metropolis algorithm (i.e., sampling of states according to their thermal importance: Boltzmann distribution for the ensemble under consideration, usually the canonical ensemble).40 As the chain length of a simulated system becomes longer, it quickly becomes necessary to introduce a series of biased moves in which additional information about the system is incorporated into the Monte Carlo selection process in such a way as to maintain detailed balance. The most commonly used biased sampling techniques are the continuum and concerted rotation moves. These modern algorithms or slightly modified versions can efficiently generate dense fluid polymer systems for chain lengths of 30-100 monomers. Longer polymer chain lengths pose additional convergence problems and often require the use

7674 J. Phys. Chem. B, Vol. 109, No. 16, 2005 of other types of biased moves, in particular the double-bridging moves.41 In the present study, we are primarily interested in understanding the morphology and molecular structure of polymeric molecules composed of PPV-based molecules in dilute solution. The applicability of the Monte Carlo methods to the current situation is somewhat different than in dense fluids but the biased moves developed for that regime are still valid. Addition of solvent via continuum models (discussed below) or explicit atoms can also be easily implemented. One of our primary interests in using the Monte Carlo method is to ensure adequate equilibration of longer chain polymers, that is, those with 100s of monomers. Molecular mechanics and molecular dynamics methods can also be used but generally require considerably long times to equilibrate. For shorter chain lengths, however, we use these methods to obtain data on the dynamical processes of chain self-organization. Molecular mechanics methods use the laws of classical physics to predict structures and properties of molecules by optimizing the positions of atoms based on the energy derived from an empirical force field describing the interactions between all nuclei (electrons are not treated explicitly).42a As such, molecular mechanics can determine the equilibrium geometry in a much more computationally efficient manner than ab initio quantum chemistry methods, yet the results for many systems are often comparable. However, since molecular mechanics treats molecular systems as an array of atoms governed by a set of potential energy functions, the model nature of this approach should always be noted. Molecular dynamics (MD) simulations essentially consist of integrating Hamilton’s equations of motion over small time steps. Although these equations are valid for any set of conjugate positions and momenta, Cartesian coordinates greatly simplifies the kinetic energy term. In our MD simulations, the classical equations of motion are formulated using our geometric statement function approach which reduces the number of mathematical operations required by a factor of ∼60 over many traditional approaches.42b These coupled first-order ordinary differential equations are solved using novel symplectic integrators developed in our laboratory that conserve the volume of phase space and robustly allow integration for virtually any time scale.42c III. 2. Modeling the Solvent. It is well-known that solvation effects are critical to the structural and dynamical properties of macromolecules. In the present case, the polymeric nanoparticles are produced from a very dilute solution and show strong solvent dependences on the formation of oriented nanostructures and thus the observed photophysical properties. Therefore, modeling the structure of these PPV based nanoparticles must include the influence of the solvent. Although a “full” microscopic description of solvation is possible using molecular dynamics or mechanics with explicit solvent molecules, there are still approximations in the many-body electrostatic interactions, and this approach can be very computationally time-consuming. As such, considerable research has previously been devoted toward developing reliable implicit solvent models in which the solvent molecules are generally replaced by a structureless dielectric continuum.43,44 These models greatly increase the speed of the calculation and often avoid some of the convergence problems in explicit models, where longer simulations or different solventstarting geometries can yield different final energies. The continuum models generally divide the solvation effect into nonpolar contributions treated in terms of the amount of solvent-accessible surface area and electrostatic contributions computed on the basis of the Poisson-Boltzmann equation or

Sumpter et al. one of its many simplifications, such as, the generalized Born models.

∆Gsol ) ∆Gnonpolar + ∆Gpolar

(1)

where



∆Gpolar ) 1/2 Φreac(r) F(r) dV

(2)

and

∆Gnonpolar )

∑σi SAi

(3)

In eq 2, Φreac is the reaction field (difference between the solvent and vacuum potentials) and F is the charge distribution. Most schemes for evaluating the nonpolar components of the solvation free energy are ad hoc. As such, a simple model is generally used as shown in eq 3, where the free energy associated with the nonpolar solvation of any atom is assumed to be characteristic for that atom and proportional to its solvent-exposed surface area, where SA is the exposed surface area and σ is the characteristic “surface tension” (this is not the surface tension of the solvent but simply a parameter with units of energy per area) associated with the same atom. The solvent-exposed surface area can be computed by a variety of different procedures but one common method uses a spherical probe molecule rolling over the van der Waals surface of the solute atoms. The atomic surface tension parameters are generally taken from fits to collections of experimental data for the free energy of solvation in a specific solvent minus the electrostatic part computed via the generalized Born (GB) method. These types of data fits are generally available for water, carbon tetrachloride, chloroform, and octanol solvents. Crammer and Truhlar44,45 have developed the SMx models which generalized the computation of the surface tension parameters to any solvent by making these values a function of more quantifiable solvent properties such as macroscopic surface tension, index of refraction, relative percent composition of aromatic carbon atoms and halogen atoms, and hydrogen bonding acidity and basicity. These models also attempt to better account for other contributions to solvation such as the cavitation energy (making a “hole” in the solvent for the solute), attractive dispersion forces between the solute and solvent molecules, and local structural changes in the solvent such as changes in the extent of hydrogen bonding, While molecular modeling and simulation of solvation is still an active area of research and development,43-52 generally satisfactory results can be obtained by using the generalized Born (GB) approximation to the Poisson-Boltzmann equation:

∆Gpolar ) -1/2[1 - 1/]

∑qiqj/fGB

(4)

where the most common form of fGB (one that interpolates between an effective Born radius R for small interatomic distances r and r itself at large distances) is taken as

fGB ) [rij2 + RiRj exp(-rij2/4*RiRj]1/2

(5)

The rij are the interatomic distances, and R are the so-called Born radii that are generally computed on the basis of making the Coulomb field approximation (the electric displacement is Coulombic in form). While this approximation may tend to overestimate the effective Born radii by as much as a factor of 2 (the power of the distance dependence is too small, it should be about the sixth power), it in general has proven to do quite well.47

Conjugated Polymers and Oligomers In principle, the atom-centered monopoles used by this GB model generate all of the multipoles required to represent the true electronic distribution. Currently, there are several different GB models, differing mainly in how the Born radii are computed. In the present study, we have implemented the analytical techniques that use a pairwise atomic summation to give the volume integration for the Born radii as described by Still et al.48 and also by Hawkins et al. (the pairwise descreening method).49 We have also used the Eisenberg-McLachlan atomic solvation parameter model,50 the original numerical integration of the solvent-accessible area implemented by Still et al. (ONION),51 and the analytical continuum electrostatics solvation method of Karplus et al.52 In the present case, we are interested in other solvents such as toluene, tetrahydrofurane (THF), and dichloromethane (DCM). These solvents can be considered to span the range of good (good solubility or highly solvated system) to bad (low solubility or low solvation) solvents for the PPV based polymers of this study. Appropriate dielectric constants and surface probe radii were used for these cases. The principle reason for examining the various models was to determine if there were any qualitative changes in the dynamics and resulting structure because of the assumed continuum solvation model. Since we are not directly concerned with quantitatively accurate structures at this point in the polymer particle formation, the particular computational details of the GB model should not cause many large changes. From our studies using the various implementations of continuum solvation, the qualitative solvated morphologies are indeed quite similar; there is a collapse of a PPV-based molecule in a “bad solvent” such as toluene into an organized folded structure as shown in Figure 2. The results we report in the present paper are therefore only given on the basis of those obtained from the GB/SA model as described by Still et al.48 III. 3. Single Molecule Nanoparticle Formation Process. The procedure we used to model the overall experimental process for producing single molecule PPV polymers was to start a GB/SA-MD simulation at a randomly chosen configuration of the PPV polymer, where we used substitution on PPV backbone to give 2-methoxy-5-(2′-ethyl-hexyloxy)-p-phenylenevinylene f (MEH-PPV) and 2,5,2′,5′-tetrahexyloxy-7,8′dicyano-p-phenylenevinylene f (CN-PPV). Molecular dynamics simulations are performed until the geometry of the PPV systems reaches an “equilibrium structure” as measured by the fluctuation in the total nonbonded energy, end-to-end distance, and an orientational autocorrelation function of a unit vector oriented along the main chain backbone. This typically requires a trajectory on the order of nanoseconds, somewhat dependent on the solvation model but mainly on the nature of the substituted PPV polymer. Figure 2 shows the final geometry of an MEH-PPV polymer consisting of 28 monomers and 3 sp3 (tetrahedral) defects located every 7 monomers. In this particular figure, the time evolution of the positions of the chain ends are marked as solid red and blue lines (the side chains are not shown for better clarity) and the green and yellow lines show the progression of the positions for two of the sp3 defects. Following these computations, which are taken as giving the initial qualitative but crucial stages of folding of the polymer in solvent, the polymer structure obtained is used to start a second series of computations which include explicit solvent molecules. This approach is used in an attempt to reduce any “minor” structural dependence on the continuum model as well as to better account for differences between the solvents (THF, DCM, toluene). Molecular dynamics and mechanics simulations are used on the full system to obtain a new structure. These

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7675

Figure 2. A final structure for a 28 monomer (with three tetrahedral defects) MEH-PPV molecule in a bad solvent. The side groups are made transparent and the solid colored lines indicate the progression of the folding dynamics: Red is one of the chain ends and blue is the other. The green and yellow lines mark the progression of two of the tetrahedral defect sites.

new structures are used to make correlations to experimental observations in solvent. To emulate the production of solvent-free single molecule nanoparticles from solution, the MEH-PPV and CN-PPV structures are minimized in a vacuum using a combination of molecular mechanics with simulated annealing. Simulated annealing is used to take the room temperature solvated structures quickly through a temperature decrease as occurs during evaporative cooling in the experimental nanoparticle generation procedure. Combining this with soft boundary conditions to impose an overall cylindrical shape (this was only used for relatively long polymers >70 monomers long and is based on the experimental evidence discussed in section III.1) gives the final PPV polymer structure for a dry single molecule polymer nanoparticle. Figure 3 shows a typical MEH-PPV single molecule nanoparticle structure obtained from the modeling and simulations procedures just described. A rod-shaped, compact structure with conjugated chain stacking (π stacking) is readily apparent. A similar procedure can be used to obtain information on multimolecule (thin films and solutions are composed of many molecules) structural organization in thin films or to probe concentration dependences in solution by using periodic boundary conditions. For thin films, only two dimensions are periodic but with a substrate placed parallel to the periodic imaged axes. To investigate concentration dependencies of the structure and morphologies in solution, 3-D periodic images are used where the size of the imaged box is varied to emulate different concentrated solutions. These results can be compared to the continuum solvation and nonperiodic explicit solvent model results which emulate very dilute solutions thereby giving two concentration extremes, very concentrated and very dilute.

7676 J. Phys. Chem. B, Vol. 109, No. 16, 2005

Sumpter et al.

Figure 3. Final structure of a 35 monomer (four tetrahedral defects) MEH-PPV molecule obtained through a sequence of continuum and explicit solvent molecular dynamics and mechanics simulations combined with simulated annealing.

III. 4. Electronic Excited States: CIS- ZINDO/s. To obtain information pertinent to the electronic excitation such as optical absorption (vertical transitions) and emission (adiabatic transitions), computations that probe the electronic excited states are required. Given a good approximation of the electronic ground state, which clearly requires optimization of the structure within a reasonable level of theory and for a reasonably complete basis set, there are several possible methods which can be used to probe the electronic excited states. However, most of the more rigorous methods such as coupled clusters (CC), configuration interactions (CI), and other multireference methods (CASSCF, etc., generalized valence bond) computationally scale rather poorly with the number of basis functions and therefore cannot be effectively applied to the relatively large systems of the present study. Density functional theory combined with the response due to a linear electric field that is fluctuating (linear response or RPA for HF), referred to as time-dependent density function theory (TDDFT), can in principle overcome the computational scaling problems. However, it has been found that TDDFT is particularly inaccurate for the current systems which tend to exhibit considerable charge-transfer character of the wave function (there is considerable literature on this topic53). While TDHF (time-dependent Hartree-Fock, the RPA) can also be used, the effect of electron correlation is neglected. Another approach is to resort back to the semiempircal methods, in particular those of Zerner et al., which have parametrized the INDO method on the basis of fairly extensive spectroscopic data (sometimes referred to as INDO/s or ZINDO/s).54 This approach, in principle, takes into account some of the effects of electron correlation in that it is fit to experimental data. When combined with configuration interaction singles (CIS), it can be very successful for a variety of transitions, including πfπ* transitions.55 Since this approach is based on a semiempirical model, it is reasonably computationally efficient when compared to the fully ab initio methods. Bredas and co-workers55 have clearly shown how this method, when carefully used, can be very successful for conjugated organic systems. As such, the results discussed for electronic transitions are only given for calculations based on CIS-ZINDO/s. IV. Results and Discussion Molecular modeling can often provide an efficient method for visualizing processes at a sub-macromolecular level that also can be used to connect theory and experiment. Particularly attractive from a computational point of view is that polymer

nanoparticles are very close to the size scale where a complete atomistic model can be studied without using artificial constraints such as periodic boundary conditions, yet these particles are too small for traditional experimental structure/property determination. Polymeric particles in the nano- and micrometer size range have show many new and interesting properties because of size reduction to the point where critical length scales of physical phenomena become comparable to or larger than the size of the structure itself. This size-scale mediation of the properties (mechanical, physical, electrical, etc.) opens a facile avenue for the production of materials with predesigned properties.56 It is therefore extremely important to develop an understanding of these phenomena. It is important to realize that details of the molecular structure for these types of conjugated organic systems are fundamental for determining photophysical properties involving electron excitations since subtle variations of the π-conjugation can strongly influence the outer valence electron levels. To gain more insight into the structural organization of MEH-PPV and CN-PPV, a combination of molecular dynamics and molecular mechanics (using the MM3 potential), simulated annealing (using a linear annealing schedule over a 1-ns trajectory), semiempirical quantum mechanics (AM1 and INDO/s levels),57 and ab initio quantum mechanics (Hartree-Fock, second-order Moller-Plesset perturbation theory, density functional theory, and configuration interaction singles)58 were performed as described above in the methods section. The minimum energy configuration of single molecule systems consisting of 14-112 monomers with tetrahedral defects located every seven monomers, as well as similar computations with more random locations, were determined for both MEH-PPV and CN-PPV. These simulations were performed for isolated single chain systems (no solvent, to make contact to what has typically been done for these types of systems) and with inclusion of solvent via the continuum model of the generalized Born approximation (GB/SA) as well as with explicit solvent molecules (THF, toluene, dichloromethane). Since subtle variation in the π-conjugation is known to strongly influence the outer valence electron levels, the present study is mainly devoted toward describing results obtained from the determination of the structure and morphologies of the various PPV systems produced by our specific experimental-droplet generation technique discussed above. IV. 1. Structure of PPV-Based Systems. On the basis of molecular mechanics energy minimization in a vacuum, a single

Conjugated Polymers and Oligomers MEH-PPV or CN-PPV molecule containing tetrahedral defects has a lower total potential energy when folded into cofacial stacked chain segments (see Figure 2) than in an extended chain configuration: about 10s of kcal/mol determined from the MM3 model for the all trans-anti configuration. For MEH-PPV, an all trans-syn configuration lies at ∼3 kcal/mol higher in potential energy than the anti configuration because of steric interactions of the side groups. For CN-PPV, the syn configuration is ∼4 kcal/mol lower in potential energy mainly because of the strong intermolecular interactions of the CN groups. The syn configuration is also lower in energy for PPV, ∼2 kcal/mol relative to anti. Similar results can be found for various isomeric forms, ranging from having the vinyl linkages all cis to random amounts of cis/trans. For molecules with shorter oligomer segments between tetrahedral defects, steric repulsion can become significant and folding is not energetically favorable for oligomers with less than four monomers. Calculations were also performed for these systems using quantum chemistry techniques, in particular, semiempirical calculations using the AM1 and PM3 models. The AM1 computed heats of formation (energy convergence to 10-5) in a vacuum for an all trans-anti MEH-PPV molecule containing 14 monomers linked by one tetrahedral defect are ∆Hf (linear)) -1070.20 kcal/mol and ∆Hf (folded) ) -1059.74 kcal/mol. For a PPV molecule (no side chains) of the same size, ∆Hf (linear) ) 499.37 kcal/mol and ∆Hf (folded) ) 508.92 kcal/ mol. Both of these isolated systems have heats of formation which indicate a greater stability for the linear form in a vacuum. On the other hand, in an explicit solvent model using QM/MM methods, where the MEH-PPV molecule is modeled with AM1 and the solvent molecules are modeled with MM3, the computed heats of formation are in THF ∆Hfsolv (linear) ) -1081.1 kcal/ mol and ∆Hfsolv (folded) ) -1373.9 kcal/mol, in toluene ∆Hfsolv (folded) ) -1390.1 kcal/mol, and in dichloromethane ∆Hfsolv(linear) ) -1288.6 kcal/mol and ∆Hfsolv (folded) ) -1068 kcal/mol. The folded MEH-PPV molecule in a THF or toluene solution at room temperature is more favorable on the basis of heat of formation than the corresponding linear form, while for DCM it points toward more stability for extended chain conformations. For PPV, the computed heats of formation are positive and smaller for the linear form, both in the gas phase and in solvent (PPV in THF: ∆Hfsolv (linear) ) 487.6 kcal/mol, ∆Hfsolv (folded) ) 496.9 kcal/mol). The influence of solvent (using the continuum GB/SA model) on the total potential energy as computed from MM3 gives a greater differential between the linear and folded systems, [Esolv (linear) - Esolv (folded)] - [Evac (linear) - Evac (folded)] ≈ 11 kcal/ mol, also indicating that MEH-PPV folded structures in a THF or toluene solvent are preferential. In addition to these computations on the total classical potential energy and semiempirical heats of formation, we also determined the approximate intermolecular binding energies, Eb, between two cofacial PPV, MEH-PPV, and CN-PPV oligomers with seven monomers. These calculations were performed using the classical MM3 and the semiempirical quantum AM1 models: for MM3, Eb(syn-PPV) ) 42.1 kcal/ mol, Eb(anti-MEH-PPV) ) 84.7 kcal/mol, and Eb(syn-CNPPV) ) 133.7 kcal/mol; for AM1, Eb(syn-PPV) ) 1.56 kcal/ mol, Eb(anti-MEH-PPV) ) 6.82 kcal/mol, and Eb(syn-CNPPV) ) 35.7 kcal/mol. The main information that can be extracted from the binding energies (for either the MM3 or the AM1 models) is the trend toward increasing intermolecular interactions between the two oligomers going from PPV to MEH-PPV to CN-PPV (the values of the binding energies

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7677 should not be taken to be quantitative, full quantum calculations at a high level of theory would be required for this). There is a clear increase in the binding energy as the PPV backbone has more substitution which presumably contributes to the heats of formation as was indicated from the calculations discussed earlier in this section. The chemical substitution on the vinyl group of the PPV backbone to produce CN-PPV has the most dramatic effect, decreasing the interchain separation, significantly increasing the binding energy, and changing the preferred configuration back to syn. Higher level quantum calculations using BSSE corrected MP2/aug-cc-pvdz calculations indicate that a CN group increases the binding energy ∼2.5 kcal/mol and decreases the corresponding interdimer distance by 0.1 Å for a dimer of 2-cyano-vinylbenzene (a small fragment of CNPPV). These changes originate mainly from the increased intermolecular interactions induced by the CN groups which have a very high electron affinity that strongly affects the π electron density of the phenyl rings. The resulting change in binding is elegantly discussed by the recent work of Sinnokrot and Sherrill for substituted benzene dimmers.59 In the present case, the difference is the substitution of CN occurs on the vinyl linkages between the phenyl rings instead of directly on the aromatic ring itself. The general effect appears to be similar, and we have carried out a series of MP2 and CC electronic structure calculations to verify these effects.60 The intermolecular interactions in all of the PPV-based systems are strong enough to lead to efficient nucleation of multiple oligomer chains into aggregates with varying degrees of crystalline order. For a PPV system consisting of seven chains, each with eight monomers, on the basis of molecular mechanics optimization using the MM3 model (converged to a root-mean-square gradient of 10-5), a herringbone-type packing arrangement is found with identifiable crystallographic parameters of a ) 5.2 Å, b ) 8.0 Å, and c ) 6.4 Å and a setting angle of 58°. This is in good agreement with experimental determination and with recent molecular mechanics simulations.61 For a MEH-PPV system of the same size, aggregate formation occurs but the packing does not conform to a herringbone-type arrangement but instead more to a cofacial staking of the oligomers with an average interchain separation of ∼3.7 Å. CN-PPV behaves quite similar to MEH-PPV except with a closer interchain separation of 3.4 Å and a greater degree of helical twisting (the phenyl rings remain cofacial). This type of π stacking of the conjugated oligomers leads to increased charge carrier mobility by generating large valence or conduction bandwidths proportional to the orbital overlap of adjacent oligomers. Quantum chemistry computations show increased transport properties and indicate decreased luminescence quenching for π staked structures62 as well as substantial self-solvation affects that lead to greater enhancement of orbital overlap (see below). A herringbone packing arrangement destroys the orbital overlap between oligomers (see discussion below on electronic spectra), and because of the two symmetryinequivalent oligomers in a unit cell, leads to a splitting of the bands (Davydov splitting) which subsequently can cause luminescent quenching. Thus, it is desirable to generate PPVbased systems with a π stacked arrangement.63 The single molecule nanoparticles generated by our experimental procedure provide this unique capability by using a combination of 3-D confinement and solvent-induced morphologies in the absence of an interacting substrate. We have also carried out extensive semiempirical quantum mechanics calculations to determine single molecule structures. For these calculations, the initial geometry was taken from an

7678 J. Phys. Chem. B, Vol. 109, No. 16, 2005

Sumpter et al.

Figure 4. Structure of two, three monomer PPV oligomers (all trans-syn) determined from MP2/3-21++G geometry optimization.

MM3 optimized structure. The results obtained from the AM1 calculations (PM3 results were very similar) show larger interchain separation for both MEH-PPV and CN-PPV molecules with a fold (as much as 0.5 Å). The optimized AM1 structures for folded 14 monomer MEH-PPV and CN-PPV molecules have larger torsional rotations about the bonds adjacent to the vinyl CdC which tends to force the oligomer chains farther apart. On the other hand, for just PPV (no side groups), the structures for small oligomers (up to four monomers) and even multiple cofacial oligomers are reasonably similar to the MM3 results. However, as the number of monomers increase, deviations from planarity occur even for PPV. Again, the result is a significantly different structure than that obtained from classical MM3 molecular mechanics. Either the enhanced intermolecular interactions caused by the cofacial arrangements of the substituted PPV oligomer chains (dispersive van der Waals Forces) cause considerable changes that are not accounted for in the MM3 model or the semiempirical AM1 parametrization or model is not appropriate for this type of conjugated system. Since the MM3 generated multioligomer aggregate structure is particularly accurate compared to experimental determinations for PPV, the source of the difference in structures would appear to be in the semiempirical models capability for describing dispersion forces. In addition, the AM1 optimized geometric structures for MEH-PPV do not give electronic structure results for the vertical transitions that agree with experiment, while the vertical transitions computed on the basis of the MM3 optimized structures conform quite closely to experimental results. Full quantum calculations using wave function or density functional theory also give approximately planar structures with interchain distance about 3.5 Å for both PPV and MEH-PPV (see below). As such, one should be very careful with geometry optimizations of multioligomer PPVbased systems using the semiempirical AM1 and PM3 models. We recommend using the much faster, and for this particular case, accurate, classical molecular mechanics minimization of the MM3 force field for the PPV-based systems. In addition, structural differences between semiempirical AM1, DFT, and HF calculations have also been previously observed for small oligomers of PPV.64 One consistent structural observation obtained from the various optimizations using the MM3 and AM1 models is that while small oligomers of PPV with no substitution tends to form cofacial geometries with a shift along the chain axis, MEHPPV and CN-PPV tend to have some “helical” twisting of the PPV backbone, with a larger angle for CN-PPV. The phenyl rings in these structures are rotated about the backbone by about (6° for MEH-PPV and (10° for CN-PPV but maintain an approximately cofacial orientation with respect to the phenyl rings on the neighboring chain. Addition of a fold and larger oligomers causes the twist angle to decrease (depending on the oligomer length between the folds). For example, a twist is present in structures determined from MM3 calculations with

folds and longer oligomer segments but not for those with less than four monomers. Quantum calculations based on HartreeFock SCF theory using a modest basis set (6-31G*) give a structure for MEH-PPV consisting of eight monomers and one tetrahedral defect that has an interchain separation of dIC ) 3.53 Å and has very little helical twisting. The intermolecular interactions of a multiply folded MEH-PPV molecule will tend to decrease any backbone twisting as well as decrease the interchain separation. This influence, often referred to as selfsolvation, is discussed in more detail below. The values for interchain separations are probably not as accurate as the general trend of decreasing interchain separation for CN-PPV as compared to MEH-PPV. The interchain distance reported from X-ray diffraction studies65 of MEHPPV thin films is dIC ) 3.56 Å and is in very good accord with our results from MM3 (discussed above) and full ab initio quantum calculations carried out for short oligomer chains of PPV (no side chains) as well as folded MEH-PPV molecules at the MP2/3-21++G level of theory. Figure 4 shows the minimum energy geometry for PPV consisting of two oligomers with three monomers each. For these types of short oligomer dimer systems, the syn conformation is of lower energy than the anti conformation by ∼0.96 kcal/mol (side chains change this result as noted earlier). For the optimized structure shown in Figure 4, the interchain distance is dIC ) 3.5 Å and there is a shift of 0.98 Å along the chain backbone axis and one of 0.7 Å along the remaining axis. Addition of side groups to produce MEH- or CN-PPV leads to geometries that do not show the shifts and gives dIC ) 3.53 Å for MEH-PPV and dIC ) 3.4 Å for CN-PPV. Whether the interchain distance and the shifts (for PPV) about the remaining axes depend on the number of monomers in the oligomer segments or to chain folds is clearly of interest. Unfortunately, the only way the shifts can be quantified is through full quantum calculations using manybody perturbation theory which computationally scales as N5. The largest system that we have been able to treat at this level of theory and get converged results is for two oligomers consisting of four monomers each, which does maintain the shifts for PPV. While we believe the shift gives a true structural minimum, the values we obtain are significantly smaller than those typically used for stacked PPV molecules (generally half a unit cell length, ∼ 3.3 Å). As will be shown below, the degree of orbital overlap is strongly dependent on both the interchain distance and the shifts along the other axes and it is therefore important to obtain an accurate initial structure. The degree of the shifts along the two axes changes as the system is taken from the very small isolated cluster to the bulk and can be examined by using periodic DFT-LDA calculations. Here, we have carefully calibrated the particular implementation (plane wave norm-conserving pseudopotentials) of the DFTLDA method to the geometry of the MP2 study to ensure consistency. We emphasize the importance of doing this calibration as many geometry optimizations for short oligomers

Conjugated Polymers and Oligomers

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7679

Figure 5. Contours of the one electron frontier LUMO orbital determined from HF/3-21++G calculations.

(from three to four monomers per oligomer) on the basis of either DFT or SCF-HF theory can generate optimal geometries that are “T shaped”, clearly not a possibility for a folded system. Inclusion of electron correlation is crucial in obtaining the cofacial geometries for the given basis set ranging from STO3G* to cc-aug-pVTZ. Errors caused from the incompleteness of the basis sets and from approximation to electron correlation (contributions from triple and quadruple excitations are important for the benzene dimer,66 for example, a simple prototype for π-π interactions) can cause significant variation in the optimized geometries and likewise for binding energies and vertical transitions. On the other hand, a plane-wave basis function representation is a “complete” one (no basis set superposition error) and with careful selection of the representation of the core via pseudopotentials, these calculations are considerably quicker yet often provide very accurate results. From the periodic DFT-LDA calculations of four monomer PPV oligomers, we find that interchain separation decreases somewhat but the shifts about the two other axes are not significantly altered. The decreased interchain distance going toward a more bulklike phase is assignable to a self-solvation affect that we discuss in more detail below and the persistence of the shift about the other axes means these structural details should be noted in any excited-state calculation for these types of stacked PPV oligomers. A contour plot of the LUMO for the PPV structure is shown in Figure 5. There is clearly significant orbital overlap between the two PPV oligomers. The transition dipole moment for electronic 1Bu excitation to the LUMO is polarized collinear with the conjugation axis. The π bonding and π* antibonding orbitals form the delocalized valence and conduction wave functions which support mobile charge carriers. IV. 2. The Effects of Solvent on Structure and Morphology. On the basis of the above results, it would not appear entirely unreasonable to begin geometry optimization with an organized (stacked oligomer segments) folded or a cofacial oligomer structure in a vacuum which is indeed the most common starting point assumed in previously reported calculations for PPV.39,67 On the other hand, previous Monte Carlo simulations for simplified models consisting of beads on a chain (no atomic structure or interactions were present in the models) suggest random coil-like geometries (defect-coil structures since little to no rotation happens about the unsaturated CdC bonds) would be preferential.18 Indeed, these types of results appear more in accord with the standard interpretation within polymer science for the structure of macromolecules in dense fluids.68 However, in the past, molecular mechanics models for the determination of the minimum energy configuration for PPVbased oligomers have been performed assuming a cofacial arrangement of very short oligomers without folds. This implies

highly organized initial structures. While either modeling approach is certainly not unreasonable, neither provides information directly related to the solution-phase morphologies which are clearly important as indicated by the QM/MM results discussed above, nor do they provide any details on the dynamical processes leading to such structures. Since there is fairly substantial evidence from experiment as discussed in section II that the solution-phase morphologies are crucial to those of thin films and in particular to those of single molecule nanoparticles, it is clear one needs to directly take into account solvent effects. In the computations discussed below, it is shown that the solvent indeed provides the key to producing selforganization into compact and structured morphologies. Figure 6 (also see Figure 2) shows the progression of a typical GB/SA molecular dynamics simulation over a 1-ns trajectory (the total trajectory time to reach an “equilibrated” conformation depends on the number of monomers in the molecule and on the nature of the side groups). The initial configuration (6a) was obtained by propagating an MD simulation of a linear chain of MEH-PPV consisting of 28 monomers and 3 tetrahedral defects at an elevated temperature (800 K) for 10 ps. This allows the system to sample some of the possible phase space available from which an individual geometry is randomly selected. The next set of snapshots (Figure 6b-6d) shows how the MEHPPV chain folds in a bad solvent (treated as a continuum dielectric) at the tetrahedral defects during a 1-ns MD simulation. The structure and morphology obtained from this simulation (Figure 6d) is one that shows considerable folding into a reasonably compact structure with a rod-shaped morphology. While the oligomer chains between the tetrahedral defects do not stack perfectly cofacial (interchain distances range from 3.5 to 4 Å and there are shifts of about one phenyl ring along the chain axis of one oligomer with respect to another as well as some backbone twisting along the chains), there is a definite preference for this organization. Determination of the final or lowest energy configuration (Figure 6e) was obtained by using a combination of simulated annealing and molecular mechanics without any solvent interactions (an attempt to emulate a dry particle as is obtained from the experimental generation) starting with the structure obtained from the MD simulation with solvent (Figure 6d). Some secondary organization is notable, in particular the preponderance toward cofacial oligomer chain stacking (the interchain distance becomes more uniform) and reduced backbone twisting (see Figures 6e and 3). Similar simulations with inclusion of shorter oligomer segments (three to six monomer defects) only show folding at the defects sites separated by at least four monomers. The single molecule

7680 J. Phys. Chem. B, Vol. 109, No. 16, 2005

Sumpter et al.

Figure 6. MD Snapshots taken during a 1-ns GB/SA-MD simulation for a model of a bad solvent.

systems appear to nucleate at a particular oligomer length and the remaining oligomer segments conform very closely to this nucleated size even if there are oligomer segments composed of fewer monomers. This structural arrangement tends to maximize the intermolecular interactions and still leads to a rodshaped structure of near uniform length. While this type of a single molecule nanoparticle is composed of multiple conjugated oligomer segments, the shape and length conform to one particular size and because of extensive orbital overlap induced by this structure, facile nonradiative resonant energy transfer (a Fo¨rster E-transfer process)69 can occur thereby leading to luminescence properties of a single chromophore of fixed chromophore length. Longer MEH-PPV molecules, those with more than 100 monomers, with regularly spaced tetrahedral defects tend to have a much slower equilibration time and go through several complicated geometric and structural changes on a time scale of 10 nanoseconds. Multiple nucleation sites can occur leading to self-organization into several different regions of structural regularity as is shown in Figure 7. This results in a single molecule system that is composed of several regularly stacked oligomer segments separated by several angstroms. This type of solution-phase system would not exhibit the luminescent properties we have measured experimentally for the PPV-based nanoparticles, and the secondary structural changes induced by nanoparticle formation process become an essential component of driving these conjugated systems to the required rodlike structures. Finally, there is a third component to generating the high amount of structural regularity unique to the single molecule nanoparticles. This comes from the excess electronic charge on the nanoparticles, which is also critical for the z-orientation on the deposition substrate. On the basis of semiempirical AM1 quantum calculations for an MEH-PPV structure (14 monomers with one fold) with one excess electron, we note a clear preference for the z-orientation on a substrate with surface Si-O- groups as well as some enhancement in the structural organization toward π-stacks with reduced interchain separation.

Figure 7. Self-organization of a 112 monomer MEH-PPV molecule (16 tetrahedral defects every 7 monomers) during a 10-ns trajectory showing the formation of multiple nucleation sites. The hydrogen atoms are not shown but were explicitly included in the calculations.

The same type of MD simulations but in a continuum model of a good solvent such as DCM does not lead to folded or compact structures but to extended chain conformations with random amounts of folding at the tetrahedral defects as shown in Figure 8. The final structure obtained from this particular type of simulation does not have a compact morphology and is more accurately described as a defect-extended chain. There is clearly a substantial difference in the dynamics and resulting structure and morphology of an MEH-PPV molecule in the different solvents. In a later part of this section, we discuss results obtained by using an explicit solvent model which allows us to account more accurately for the differences between three

Conjugated Polymers and Oligomers

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7681

Figure 8. MD snapshots taken during a 1-ns GB/SA-MD simulation for a model of a good solvent.

solvents, toluene, THF, and DCM. The solution-phase structural differences are small enough not to induce any large alterations in the final dry single molecule nanoparticle structures or morphologies, although there are some details that may be important to the solution-phase spectral measurements. The effects of explicit solvent (full inclusion of all of the atoms) on the structure of MEH-PPV and CN-PPV were also examined. In these calculations, explicit solvent molecules were added to an MD box to achieve the appropriate density followed by aggressive energy minimization using molecular mechanics. To reduce the number of required solvent molecules, we started all calculations with a polymer molecule that had already been equilibrated in a continuum solvent simulation (such as that shown in Figure 6d). This approach also allows us to examine solvents that have very similar dielectric constants. It should be stressed that these simulations began with an initial structure determined from a continuum model for toluene. The structure is already folded. A continuum model for DCM does not lead to a folded structure but one that has very little folding and is more accurately described as a defect-extended chain (Figure 8d). The final structures obtained were qualitatively similar to those obtained for the continuum solvent systems but with some notable quantitative differences. For MEH-PPV, the interchain distance increased from dIC ) 3.7 to 4 Å in THF and to dIC ) 4.1 Å for DCM but decreased to dIC ) 3.4 Å in toluene. The interchain distance for CN-PPV did not seem to be as strongly dependent on the solvent but did show an increased helical backbone structure over the continuum model. The PPV structures in DCM were clearly much more disorganized compared to the other solvents, with very little alignment of the chain segments, even though the simulation started from a prefolded and compact structure. Interestingly, there were no strong transitions to the first excited state observed using the CIS-ZINDO/s for the PPV systems in DCM. The disorganization of the chain segments coupled with the increased separation clearly has dramatic effects on the vertical transitions. The average interchain distance, dIC, obtained for the optimized single molecule MEH-PPV nanoparticle created from the solution-phase bad solvent morphology (Figure 6d) is dIC ) 3.7 Å (determined between the two center chains). For

CN-PPV of the same backbone length, very similar chain dynamics (although the rate of folding was ∼2 times slower mainly because of the increased steric hindrance about the sp3 C-C bond because of the relatively large CN groups) was found, and the minimum energy configuration had an interchain distance of dIC ) 3.5 Å, a relatively large decrease in the separation These values are somewhat different than that determined by Conwell et al. using the MM2 model and assuming cofacial initial geometries. Here, we are using slightly different potential energy functions, MM3, and have explicitly included chain folds and all of the atoms of the system for longer oligomer segments and larger number of “stacked oligomers” (some self-solvation effects are thus possible), and we have accounted for the influence of solvent on the folding process and resulting geometry as well as emulated the process of single molecule nanoparticle production from the solution phase. The “secondary” role of nanoparticle formation from the dilute solution-phase structures induces some notable changes as mentioned above. In particular, the 3-D confinement coupled with the extremely rapid evaporation of the solvent tends to cause much more compact and regular stacks of oligomer segments to form. The effects of solution concentration on the structure of MEH-PPV was also investigated by using 3-D periodic images of the explicit atom solvent model. This approach was used to emulate a concentrated solution (since the size of the imaged box is constrained to be small) and can be compared to the dilute solution structure obtained from the results discussed above (which are for extremely dilute solutions as no boundary conditions were used). Figure 9 shows how the interchain structure becomes disrupted but maintains some cofacial ordering between the oligomer segments. The degree of the interchain disorganization is dependent on the size of the MD box, with smaller or higher concentrated MEH-PPV solutions having less structural organization. These types of more disorganized folded structures do not have a transition dipole oriented along the chain axis, and as noted in the experiments, do not exhibit the donutlike emission patterns. IV. 3. Geometric and Electronic Structural Effects Induced by Interchain Interactions. The general trend observed

7682 J. Phys. Chem. B, Vol. 109, No. 16, 2005

Sumpter et al.

Figure 9. Structure of a 14 monomer MEH-PPV molecule with one tetrahedral defect in a high concentration solution (top structure) compared to the same system in a very dilute solution (bottom structure). Note the disruption of the interchain stacking (compare top and bottom structures) which tends to get worse with higher concentrations.

Figure 10. Hybrid molecular mechanics and semiempirical quantum chemistry calculations for the influence of MEH-PPV chains on the interior interchain organization and packing distance.

from the MM3 results is a decrease in the interchain separation as larger numbers of oligomer segments are added. This decrease is about 0.3 Å for oligomer segments in the center of a particle and indicates some type of chain-chain self-solvation effect. To get a better idea of the change in the distance because of self-solvation, we performed limited multiscale modeling (QM/MM) simulations. Figure 10 shows how these simulations were set up by modeling the inner folded MEH-PPV molecule with semiempirical quantum mechanics (AM1 model) and the outer chains with molecular mechanics (MM3 model). The outer chains were fixed at a distance of dIC ) 3.7 Å from the center MEH-PPV molecule in accord with the MM3 results and the MEH-PPV molecule was optimized using AM1. Since we are only comparing differences here instead of absolute structure, the AM1 model for the molecule should be reasonable. The results of these QM/MM calculations show a significant decrease in the interchain separation of about 0.9 Å. The resulting structure of the MEH-PPV molecule was actually very similar to that obtained from the MM3 calculations (interchain separation of dIC ∼ 3.5 Å and much smaller torsional rotations about the bonds adjacent to the vinyl group) but differed substantially from the AM1 vacuum results. This provides some interesting evidence that interchain separation of the rod-shaped morphologies tends to decrease toward the center of the single molecule nanoparticle. As we will show below, the electronic structure depends quite strongly on this interchain distance, with enhanced singlet-to-singlet transition moments for shorter distances. This

might provide some rationalization of the definitive experimental results that show photon antibunching for single molecule z-oriented nanoparticles.29 These results showed that the zoriented single molecule nanoparticles act as single photon emitters but multichromophore absorbers. The self-solvated inner core structure, where the interchain distance becomes closer and there is a much higher degree of structural organization, is probably acting as the primary emission site. From our semiempirical results as well as others, we know that the HOMO-LUMO band gaps in molecular systems are strongly dependent on the degree of confinement, generally decreasing with increasing confinement. For the PPV-based systems, we have observed the following HOMO-LUMO gap dependences: (1) A decreasing HOMO-LUMO gap with increasing numbers of folded oligomer segments for PPV, CN-PPV, and MEH-PPV. (2) For a fixed oligomer segment length between folds, the magnitude of the HOMO-LUMO gap decrease becomes smaller with increasing number of segments. (3) For a fixed number of oligomer segments but a varying number of monomers in each oligomer, there is also a general decrease in the HOMO-LUMO band gap for the PPV systems. (4) The computed HOMO-LUMO gap dependence on selfsolvation as determined for the system shown in Figure 12 is a decrease (red shift) of ∼0.1 eV and an increase in the wavelength for the first excited-state transition of ∼23 nm. The change in the electronic structure comes primarily from a LUMO

Conjugated Polymers and Oligomers lowering of ∼0.09 eV. The HOMO, which often shows greater sensitivity to confinement, does not change as much as the LUMO, increasing by ∼0.02 eV. The self-solvation effects noted in the present simulations lead to reasonably large changes in the electronic structure and related optical transitions. These results suggest that single molecule PPV-based nanoparticles act as dielectric core-shell systems, with the emissive core being composed of the selfsolvated more ordered and tightly packed chains and the shell as less tightly packed chains and larger interchain separation but still close enough to allow orbital overlap that facilitates fluorescence resonant energy-transfer processes. This type of energy transfer is a nonradiative process (Fo¨rster energy transfer69) in which fluorescent donor molecules (the shell) are excited at particular wavelengths and transfer their absorbed photonic energy with an efficiency approximately proportional to 1/r6 (where r is the distance between the donor and acceptor) to a fluorescent acceptor molecule (the core) in close proximity, which then re-emits it at a second wavelength. This series of coupled chromophores provides an energy gradient toward the emission center. An extended energy-transfer system with multiple donor groups that directionally transfer photonic energy to a core acceptor group (energy funnel)70 as occurs for the PPV-based nanoparticles is like a molecular photonic antenna or amplifier structure. This interpretation would also allow direct implementation of classical electrodynamics, which describes the attenuation of the fluorescence lifetimes because of interactions of a dipole with the electromagnetic vacuum field,71 a property which has been observed experimentally.72 IV. 4. Electronic Structure: Vertical Transitions. To investigate the dependences of the vertical transitions on the intrerchain distance, we carried out a study similar to that of Bredas and co-workers73 and more recently Gierschner et al.74 Here, we examine the electronic vertical transitions for a MEHPPV system composed of two oligomer chains of seven monomers each with explicit treatment of the side chains. Optimization of this structure was performed using the MM3 model followed by a restricted CIS-ZINDO/s calculation that included eight HOMOs and eight LUMOs (larger CIS calculations, up to 50 orbitals, did not change the dependences but only the values for the transitions). AM1 optimization was not used since we have found these structures to be very questionable for MEH-PPV and CN-PPV (see the above discussion). The computed electronic spectra for the equilibrium structure has two primary peaks: the most intense peak (oscillator strength of 8.9) occurs at 462 nm and a peak also occurs at 386 nm (oscillator strength of 0.96). These peaks correspond to the third (HfL+2) and seventh (H-3fL) allowed transitions. The first (HfL) allowed vertical transition occurs at 511 nm with very little oscillator strength (0.013). These results are reasonably consistent with those reported by previous studies on stilbene and similar molecules and are in good agreement with the main absorption peak for MEH-PPV of 480 nm determined from experiment. The dependence of the electronic structure on the interchain separation was examined by varying it from its equilibrium value. No optimization of the geometry was performed along this coordinate as we are interested in showing how the separation alters the nature of the vertical transitions. A plot of this dependence is shown in Figure 11 where the transitions with the greatest oscillator strength, corresponding to the third allowed transition (open circles) and the next largest oscillator strength transition, the seventh (squares), are shown along with the first allowed transition (triangles). The transition with the greatest oscillator strength, third allowed transition

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7683

Figure 11. Plot showing the dependence of the vertical transitions on interchain displacement (in Å) from equilibrium (HfL+2 is O symbol, H-3fL is 0, HfL is 4). The transition with the most oscillator strength is HfL+2 and is the third allowed transition while the first allowed transition has the smallest oscillator strength and corresponds to the HfL transition.

(circles), has a dependence on the interchain separation that causes a variation in energy over 14 nm, decreasing initially followed by a slow increase that asymptotically approaches 465 nm for large displacements. The first allowed transition which has the smallest oscillator strength of those shown in the plot (triangles) varies over 112 nm, decreasing rather quickly for displacements up to 1 Å followed by a slow decrease which asymptotically approaches 470 nm. The seventh (second largest oscillator strength) transition (squares) has a rather weak initial dependence on the interchain displacement up to about 1 Å where there is a sudden decrease in energy by 49 nm followed by a slow increase asymptotically approaching 351 nm. There are clearly multiple regimes of influence because of intermolecular coupling. By examining the molecular orbitals, we find a weak coupling regime for an oligomer separation from its equilibrium value of greater than 4 Å where the LCAO coefficients of a given orbital become localized on a single oligomer chain. The optical properties of the MEH-PPV system in this region might be understood in terms of constructive and destructive interactions of the oligomer transition dipole moments which agree reasonably well with molecular exciton models.75 The energy splitting between these two electronic states increases with decreasing separation fairly monotonically until ∼1 Å. For separations between 1 and 4 Å from equilibrium, the LCAO coefficients of the MOs show some delocalization over both MEH-PPV oligomer chains, and for separations less than 1 Å the LCAO coefficients of the MOs show significant delocalization over both MEH-PPV oligomer chains. Since the full quantum computations indicate a shift along the chain axis of a PPV oligomer chain with respect to the other chain (see Figure 4) as well as along the other axis in the same plane, we also examined these influences on the vertical transitions. In these cases, displacement from equilibrium cannot be very large since the side chains begin interfering with each other and the PPV backbone (only about a 1.5 Å shift is allowed along the chain axis without performing any relaxation). Since the interchain distance is kept constant, it is reasonable to expect the intermolecular interactions will not contribute in the same fashion as for Figure 11. The dependence of the vertical transition energy on the shift about the backbone axis shows both a rapid increase and decrease, with a maximum in wavelength (or minimum in energy) occurring at 0.75 Å. The absorption wavelength obtained

7684 J. Phys. Chem. B, Vol. 109, No. 16, 2005 at a 1.5 Å shift along the backbone is the same as that at the equilibrium position. However, while the characteristic π delocalization over both oligomers is maintained up to 0.75 Å, it significantly decreases beyond this point. The electronic absorption spectrum at a 1.5 Å shift has extra peaks compared to that for shifts between 0 and 0.75 Å which apparently correspond to individual oligomer subsegments (a peak for a four monomer segment shows up) induced by the reduced orbital delocalization between the two oligomers. V. Summary and Conclusions By combining experimental observations and developments with extensive computational chemistry studies, we have presented substantial evidence suggesting highly ordered rodshaped structures for single molecule MEH-PPV and CNPPV systems. The chain organization is crucial to the photophysical properties and can be controlled to a large extent by the solvent. A relatively bad solvent leads to a structure that is tightly folded into a rod-shaped morphology while a good solvent produces structures with little or random folding and more of a defect, extended-chain morphology. Secondary structural regularity is also induced by producing single molecule nanoparticles from microdroplets of the solution without the presence of a substrate. For toluene and THF solvent preparations of MEH-PPV and CN-PPV, the resulting solvent-free single molecule nanoparticle structures show a very high level of organization consisting of π-stacked folded chains. Because of this structural organization which imposes a rod-shaped morphology, single molecule nanoparticle orientation on an appropriate surface occurs such that the particle stands on its end with near-perfect z-orientation. This orientational alignment is caused by excess charges that are induced on the particle during its production which tend to localize at the surface (as can be shown through ab initio calculations),31 causing the rodshaped particle to minimize the Coulombic interactions with the unprotonated (Si-O)- surface groups. The excess charge also appears to increase the structural organization into π-stacks. The resulting z-oriented single molecule rod-shaped nanoparticles are organized into something like a core-shell system, where the inner core is a self-solvated PPV system with interchain distances ∼0.3 Å closer than the surrounding chains (see Figure 10). Since there is still orbital overlap throughout the system, facile fluorescence resonant energy transfer without the emission of a photon can occur to the core emission site. In addition, this type of interchain distance anisotropy can give rise to behavior observed for a single quantum emitter in a dielectric medium at the nanometer scale (Raleigh scattering regime) and thus can show strong effects because of classical electromagnetic interactions (vacuum fluctuations are altered from boundary reflections) which would lead to the observed altered fluorescence lifetimes. The overall ramifications of developing this fundamentally new processing technique for generating nanoscale optoelectronic materials are far-reaching. By achieving uniform orientation perpendicular to the substrate with enhanced luminescence lifetimes and photostability under ambient conditions, the door is now open for major developments in molecular photonics, display technology, and bioimaging, as well as new possibilities for optical coupling to molecular nanostructures and for novel nanoscale optoelectronics devices. Acknowledgment. This work was supported in part by the ORNL-LDRD program and the Division of Chemical Sciences, Office of Basic Energy Sciences, U.S. Department of Energy,

Sumpter et al. under contract DE-AC05-00OR22725 with Oak Ridge National Laboratory (ORNL), managed and operated by UT-Battelle, LLC. The extensive computational work was performed on the computers at ORNL’s Center for Computational Science (CCS). References and Notes (1) El-Sayed, M. A. Acc. Chem. Res. 2001, 34, 257. Niemer, C. M. Angew. Chem., Int. Ed. 2001, 40, 4128. Adams, D. M.; Brus, L.; Chidsey, C. E. D.; Creager, S.; Creutz, C.; Kagan, C. R.; Kamat, P. V.; Liberman, M.; Lindsay, S.; Marcus, R. A.; Metzger, R. M.; Michel-Beyerle, M. E.; Miller, J. R.; Newton, M. D.; Rolison, D. R.; Sankey, O.; Schanze, K. S.; Yardley, J.; Zhu, X. J. Phys. Chem. B. 2003, 107, 6668. Antonietti, M.; Landfester, K. ChemPhysChem 2001, 2, 207. (2) Turner, C. H.; Brennan, J. K.; Johnson, J. K.; Gubbins, K. E. J. Chem. Phys. 2002, 116, 2138. Turner, C. H.; Johnson, J. K.; Gubbins, K. E. J. Chem. Phys. 2001, 114, 1841. Stuart, S. J.; Dickson, B. M.; Sumpter, B. G.; Noid, D. W. Proc. MRS 2001, 651, T7.1.1; 2001, T1.8.1. Kidder, M. K.; Britt, P. F.; Zhang, Z.; Dai, S.; Buchannan, A. C. Chem. Commun. 2003, 2804. (3) Eggers, D. K.; Valentine, J. S. Protein Sci. 2001, 10, 250. Zhou, H.-X.; Dill, K. A. Biochemistry 2001, 40, 11289. Rittg, R.; Huwe, A.; Fleischer, G.; Karger, J.; Kremer, F. Phys. Chem. Phys. 1999, 1, 519. (4) Hummer, G.; Rasaiah, J. C.; Noworyta, J. P. Nature 2001, 414, 118. Tuzun, R. E.; Noid, D. W.; Sumpter, B. G.; Merkle, R. C. Nanotechnology 1997, 8, 112; 1996, 7, 241. (5) Barnes, W. L. J. Mod. Opt. 1998, 45, 661. Schniepp, H.; Sandoghdar, V. Phys. ReV. Lett. 2002, 89, 257403. Chew, H. Phys. ReV. A 1988, 38, 3410. (6) Striolo, A.; Chialvo, A. A.; Cummings, P. T.; Gubbins, K. E. Langmuir 2003, 19, 8583. Allen, R.; Hansen, J.-P.; Melchionna, S. J. Chem. Phys. 2003, 119, 3905. Koga, K.; Gao, G. T.; Tanaka, H.; Zeng, X. C. Nature 2001, 412, 802. Toriyama, K.; Okazaki, M. J. Phys. Chem. B 2004, 108, 12917. (7) Eisenberg, P.; Lucas, J. C.; Williams, R. J. J. Macromol. Symp. 2002, 189, 1. (8) Heeger, A. J. J. Phys. Chem. B 2001, 105, 8475. (9) Feller, J. F.; Linossier, I.; Grohens, Y. Mater. Lett. 2002, 57, 64. (10) Gensler, R.; Groppel, P.; Muhrer, V.; Muller, N. Part. Part. Syst. Charact. 2002, 19, 293. (11) Friend, R. H.; Gymer, R. W.; Holmes, A. B.; Burroughes, J. H.; Marks, R. N.; Taliani, C.; Bradley, D. D. C.; dos Santos, D. A.; Salaneck, W. R. Nature 1999, 397, 121. (12) Hide, F.; Diazgarcia, M. A.; Schwartz, B. J.; Heeger, A. J. Acc. Chem Res. 1997, 30, 430. (13) Srinivasarao, M.; Collings, D.; Philips, A.; Patel, S. Science 2001, 292, 79. (14) Kung, C.-Y.; Barnes, M. D.; Lermer, N.; Whitten, W. B.; Ramsey, J. M. Appl. Opt. 1999, 38, 1481. (15) Barnes, M. D.; Ng, K. C.; Fukui, K.; Sumpter, B. G.; Noid, D. W. Macromolecules 1999, 32, 7183. (16) Schwartz, B. J. Annu. ReV. Phys. Chem. 2003, 54, 141. (17) Peyser, L. A.; Vinson, A. E.; Bartko, A. P.; Dickson, R. M. Science 2001, 291, 103-106. (18) Hu, D.; Yu, L.; Barbara, P. F. J. Am. Chem. Soc. 1999, 121, 6936. Hu, D.; Yu, J.; Bagchi, B.; Rossky, P. J.; Barbara, P. F. Nature 2000, 405, 1030. (19) Vandenbout, D. A. et al. Science 1997, 277, 1074. (20) Yu, J.; Hu, D.; Barbara, P. F. Science 2000, 289, 1327. (21) Padmanaban, G.; Ramakrishnan, S. J. Am. Chem. Soc. 2000, 122, 2244. (22) Gettinger, C. L.; Heeger, A. J.; Drake, J. M.; Pine, D. J. J. Chem. Phys. 1994, 101, 1673. (23) Teetsov, J.; Vanden Bout, D. A. J. Phys. Chem. B 2000, 104, 9378. (24) Schaller, R. D.; Lee, L. F.; Johnson, J. C.; Haber, L. H.; Vieceli, J.; Benjamin, H.; Nguyen, T.-Q.; Schwartz, B. J.; Saykally, R. J. J. Phys. Chem. B 2002, 106, 9496. (25) Scheblykin, I.; Zoriniants, G.; Hofkens, J.; De Feyter, S.; Van der Auweraer, M.; De Schryver, F. Chem. Phys. Chem. 2003, 4, 260. Sartori, S. S.; Feyter, S. D.; Hofkens, J.; Auweraer, M. V.; Schryver, F.; Brunner, K.; Hofstraat, J. W. Macromolecules 2003, 36, 500. (26) Nguyen, T. Q.; Kwong, R. C.; Thompson, M. E.; Schwartz, B. J. Appl. Phys. Lett. 2000, 76, 2454. Shi, Y.; Liu, J.; Yan, Y. J. Appl. Phys. 2000, 87, 4254. Liu, J.; Shi, Y.; Ma, L.; Yang, Y. J. Appl. Phys. 2000, 88, 605. Lee, T. W.; Park, O. O. Appl. Phys. Lett. 2000, 77, 3334. Lee, T. W.; Park, O. O. AdV. Mater. 2000, 12, 801. (27) Mehta, A.; Kumar, P.; Dadmun, M.; Zheng, J.; Dickson, R. M.; Thundat, T.; Sumpter, B. G.; Barnes, M. D. Nanoletters 2003, 3, 603. (28) Kumar, P.; Mehta, A.; Dadmun, M.; Zheng, J.; Peyser, L.; Dickson, R. M.; Thundat, T.; Sumpter, B. G.; Barnes, M. D. J. Phys. Chem. B 2003, 107, 6252.

Conjugated Polymers and Oligomers (29) Kumar, P.; Lee, T. H.; Mehta, A.; Sumpter, B. G.; Dickson, R. M.; Barnes, M. D. J Am. Chem. Soc. 2004, 126, 3376. (30) Hellen, E. H.; Axelrod, D. J. Opt. Soc. Am. B 1987, 4, 337. (31) Mehta, A.; Thundat, T.; Kumar, P.; Barnes, M. D. Appl. Phys. Lett. 2004, submitted. (32) Bredas, J. L.; Beljonne, D.; Cornil, J.; Calbert, J. P.; Shuai, Z.; Silbey, R. Synth. Met. 2001, 125, 107. (33) Kumar, P.; Mehta, A.; Mahurin, S. M.; Dai, S.; Dadmun, M. D.; Sumpter, B. G.; Barnes, M. D. Macromolecules 2004, 37, 6132. (34) Nguyen, T. Q.; Yee, R. Y.; Schwartz, B. J. J. Photochem. Photobiol., A. 2001, 144, 21. Nguyen, T. Q.; Doan, V.; Schwartz, B. J. J. Chem. Phys. 1999, 110, 4068. (35) Nguyen, T. Q.; Doan, V.; Schwatz, B. J. J. Chem. Phys. 1999, 110, 4068. Hollars, C. W.; Lane, S. M.; Huser, T. Chem. Phys. Lett. 2003, 370, 393. (36) See, for example, Cramer, C. J. Essentials of Computational Chemistry: Theory and Models; Wiley, 2002. (37) Allinger, N. L.; Yuh, Y. H.; Lii, J.-H. J. Am. Chem. Soc. 1989, 111, 8551. (38) Tai, J. L.; Allinger, N. L. J. Comput. Chem. 1998, 19, 475. Lii, J.-H.; Allinger, N. L. J. Am. Chem. Soc. 1989, 111, 8576. (39) Fratini, A. V.; Baker, K. N.; Resch, T.; Knachel, H. C.; Adams, W. W.; Socci, E. P.; Farmer, B. L. Polymer 1993, 43, 1571. Nevins, N.; Lii, J.-H.; Allinger, N. L. J. Comput. Chem. 1996, 17, 695. Claes, L.; Deleuze, M. S.; Francois, J.-P. J. Mol. Struct. (THEOCHEM) 2001, 549, 63. (40) See, for example, Frenkel, D.; Smit, B. Understanding Molecular Simulations: From Algorithms to Applications; Academic Press: New York, 2002. (41) Karayiannis, N. C.; Giannousaki, A. E.; Mavrantzas, V. G.; Theodorou, D. N. J. Chem. Phys. 2002, 17, 5465. Karayiannis, N. C.; Giannousaki, A. E.; Mavrantzas, V. G. J. Chem. Phys. 2003, 118, 2451. (42) (a) Burkert, U.; Allinger, N. L. Molecular Mechanics; ACS Monograph 177; Amercian Chemical Society: Washington, DC, 1992. (b) Tuzun, R. E.; Noid, D. W.; Sumpter, B. G. Macromol. Theory Simul. 1997, 6, 855. (c) Gray, S. K.; Noid, D. W.; Sumpter, B. G. J. Chem. Phys. 1994, 101, 4062. (43) Bashford, D.; Case, D. A. Annu. ReV. Phys. Chem. 2000, 51, 129. (44) Cramer, C. J.; Truhlar, D. G. Chem. ReV. 1999, 99, 2161. (45) Li., J.; Cramer, C. J.; Truhlar, D. G. Biophys. Chem. 1999, 78, 147. (46) Lee, M. S.; Salsbury, F. R.; Brooks, C. L., III. J. Chem. Phys. 2002, 116, 10606. Onufriev, A.; Case, D. A.; Bashofrd, D. J. Comput. Chem. 2002, 23, 1297. (47) Grycuk, T. J. Chem. Phys. 2003, 199, 4817. (48) Qiu, D.; Shenkin, P. S.; Hollinger, F. P.; Still, W. C. J. Phys. Chem. A 1997, 101, 3005. (49) Hawkins, G. D.; Cramer, C. J.; Truhlar, D. G. Chem. Phys. Lett. 1995, 246, 122; J. Phys. Chem. 1996, 100, 19824. (50) Eisenberg, D.; McLachlan, A. D. Nature 1986, 319, 199. (51) Still, W. C.; Tempczyk, A.; Hawley, R. C.; Hendrickson, T. J. Am. Chem. Soc. 1990, 112, 6127. (52) Schaefer, M.; Karplus, M. J. Phys. Chem. 1996, 100, 1578. Schaefer, M.; Bartels, C.; Karplus, M. J. Mol. Biol. 1998, 284, 835. (53) Dreuw, A.; Weisman, J. L.; Head-Gordon, M. J. Chem. Phys. 2003, 119, 2943. Tozer, D. J. J. Chem. Phys. 2003, 119, 12697. Iikura, H.; Tsuneda, T.; Yanai, T.; Hirao, K. J. Chem. Phys. 2001, 115, 3540. Cai, Z.-L.; Sendt, K.; Reimers, J. R. J. Chem. Phys. 2002, 117, 5543. Grimme, S.; Parac, M. ChemPhysChem 2003, 292, 1439. Van Gisbergen, S. J. A.; Schipper, P. R. T.; Gritsenko, O. V.; Baerends, E. J.; Snijders, J. G.; Champagne, B.; Kirtman, B. Phys. ReV. Lett. 1999, 83, 694. (54) Zerner, M. C. In ReViews in Computational Chemistry; Lipkowitz, K. B., Boyd, D. B., Eds.; VCH: New York, 1991; Vol. II, p 313. Pearl, G. M.; Zerner, M. C.; Broo, A.; McKelvey, J. J. Comput. Chem. 1998, 19, 781.

J. Phys. Chem. B, Vol. 109, No. 16, 2005 7685 (55) Hill, I. G.; Kahn, A.; Cornil, J.; dos Santos, D. A.; Bredas, J. L. Chem. Phys. Lett. 2000, 317, 444. Han, Y.; Lee, S. U. Chem. Phys. Lett. 2002, 366, 9. Cornil, J.; Belojonne, D.; Heller, C. M.; Campbell, I. H.; Laurich, B. K.; Smith, D. L.; Bradley, D. D. C.; Mullen, K.; Bredas, J. L. Chem. Phys. Lett. 1997, 278, 139. Cornil, J.; Heeger, A. J.; Bredas, J. L. Chem. Phys. Lett. 1997, 272, 463. Cornil, J.; Beljonee, D.; Bredas, J. L. J. Chem. Phys. 1995, 103, 834. Cornil, J.; Calbert, J. Ph.; Beljonne, D.; Silbey, R.; Bredas, J. L. Synth. Met. 2001, 119, 1. Pogantsch, A.; Mahler, A. K.; Hayn, G.; Saf, R.; Stelzer, F.; List, E. J. W.; Bredas, J. L.; Zojer, E. Chem. Phys. 2004, 297, 143. (56) Hayashi, C.; Uyeda, R.; Tasaki, A. Ultra Fine Particles Technology; Noyes: NJ, 1997. (57) Semiempirical quantum calculations were performed using GAMESS, Schmidt, M. W.; Baldridge, K. K.; Boatz, J. A.; Elbert, S. T.; Gordon, M. S.; Jensen, J. H.; Koseki, S.; Matsunaga, N.; Nguyen, K. A.; Su, S. J.; Windus, T. L.; Dupuis, M.; Montogemery, J. A. J. Comput. Chem. 1993, 14, 1347-1363, and Hyperchem7.0, Hypercube, Inc.: Gainesville, FL, 2003. (58) Ab initio quantum calculations were performed using NWChem Version 4.5 as developed and distributed by Pacific Northwest National Laboratory, P.O. Box 999, Richland, WA 99352, and funded by the U.S. Department of Energy. (59) Sinnokrot, M. O.; Sherrill, C. D. J. Chem. Phys. 2003, 107, 8377. (60) Sumpter, B. G.; Harrison, R. J.; Shelton, W. A. in preparation. (61) Chen, D.; Winokur, M. J.; Masse, M. A.; Karaz, F. E. Phys. ReV. B. 1990, 41, 6759. Van Hutten, P. F.; Wildeman, J.; Meetsma, A.; Hadziioannou, G. J. Am. Chem. Soc. 1999, 121, 5910. Grainer, T.; Thomas, E. L.; Gagnon, D. R.; Karasz, F. E.; Lenz, R. W. J. Polym. Sci., Part B 1986, 24, 2793. Claes, L.; Francois, J.-P.; Deleuze, M. S. Chem. Phys. Lett. 2001, 339, 216. (62) Ferretti, A.; Ruini, A.; Molinari, E. Phys. ReV. Lett. 2003, 90, 086401-1. Ruini, A.; Caldas, M. J.; Bussi, G.; Molinari, E. Phys. ReV. Lett. 2002, 88, 206403. (63) Curtis, M. D.; Cao, J.; Kampf, J. W. J. Am. Chem. Soc. 2004, 125, 4318. (64) Grozema, F. C.; Candeias, L. P.; Swart, M.; Van Duijnen, P. Th.; Wildeman, J.; Hadziioanou, G.; Siebbeles, L. D. A.; Warman, J. M. J. Chem. Phys. 2002, 117, 11366. (65) Yang, C. Y.; Hide, F.; Diaz-Garcia, M. A.; Heeger, A. J.; Cao, Y. Polymer 1998, 39, 2299-2304. (66) Sinnokrot, M. O.; Valeev, E. F.; Sherrill, C. D. J. Am. Chem. Soc. 2002, 124, 10887. Hopkins, B. W.; Tschumper, G. S. J. Phys. Chem. A 2004, 108, 2941. (67) Conwell, E. M.; Perlstein, J.; Shaik, S. Phys. ReV. B. 1996, 54, R2308. Cornil, J.; Heeger, A. J.; Bredas, J. L. Chem. Phys. Lett. 1997, 272, 463. (68) Ivanov, V. A.; Paul, W.; Binder, K. J. Chem. Phys. 1998, 109, 5659. Schweizer, K. S. J. Chem. Phys. 1986, 85, 4181. Kohler, B. E.; Samuel, I. D. W. J. Chem. Phys. 1995, 103, 6248. (69) Fo¨rster, F. Naturwissenschaften 1946, 33, 16. Canter, C. R.; Pechukas, P. Proc. Natl. Acad. U.S.A. 1971, 68, 2099. Demidov, A. A. Resonance Energy Transfer; Wiley: 1999. (70) Yu, J.; Hu, D.; Barbara, P. F. Science 2000, 289, 1327. (71) Chew, H. Phys. ReV. A. 1988, 38, 3410. (72) Schniepp, H.; Sandoghdar, V. Phys. ReV. Lett. 2002, 89, 2574031. (73) Cornil, J.; Dos Santos, D. A.; Crispin, X.; Silbey, R.; Bredas, J. L. J. Am. Chem. Soc. 1998, 120, 1289. (74) Gierschner, J.; Mack, H.-G.; Oelkrug, D.; Waldner, I.; Rau, H. J. Phys. Chem. A 2004, 208, 257. (75) Hochstrasser, R. M.; Kasha, M. Photochem. Photobiol. 1964, 3, 317. Kasha, M.; Rawls, R. H.; El- Bayoumi, M. A. Pure Appl. Chem. 1965, 11, 371.