Influence of Chemical Structure on the Charge Transfer State

Mar 18, 2014 - *E-mail: [email protected]. .... Antonio Sánchez-Díaz , Xabier Rodríguez-Martínez , Laura Córcoles-Guija , Germán Mora-...
0 downloads 0 Views 3MB Size
Article pubs.acs.org/JPCC

Influence of Chemical Structure on the Charge Transfer State Spectrum of a Polymer:Fullerene Complex Sheridan Few,† Jarvist M. Frost,‡ James Kirkpatrick,† and Jenny Nelson*,† †

Centre for Plastic Electronics, Department of Physics, Imperial College London, London SW7 2AZ, United Kingdom Department of Chemistry, University of Bath, Bath BA2 7AY, United Kingdom



S Supporting Information *

ABSTRACT: Charge transfer (CT) state properties at the donor:acceptor interface in organic photovoltaic materials are widely regarded as crucial in determining the charge separation efficiency of organic photovoltaic devices. In this work, we use time dependent density functional theory (TDDFT) with B3LYP/6-31g* to study the influence of chemical structure on excitation energies, oscillator strengths, and electronic structure. We vertically excite states in a series of polymer:fullerene blends, modeling each blend as an oligomer:fullerene pair in vacuo. Our method reproduces experimentally observed trends in CT state energy with chemical structure as measured by electroluminescence. For oligothiophene:PCBM (PCBM = [6,6]-phenyl C61-butyric acid methyl ester), we find that Coulomb binding tends to reduce in higher excited CT states. In the case of several isoindigo and diketopyrrolopyrrole (DPP) based donors, we find that the first excited state of the pair lies close in energy to the first singlet of the oligomers, and low excited states have hybrid, or incomplete charge transfer, character. The natures and energies of these states are dependent on the fullerene position. We discuss the effect of thiophene substitution in a DPP polymer on charge generation in terms of the calculated CT state properties and rationalize the observed charge separation efficiency of corresponding experimental systems in terms of these calculations.



excitonic states.12,13 Thus the importance of “hot” excitations for photocurrent generation is in dispute. Another possibility is long-range charge separation directly from excitons located deep in the polymer domain, avoiding trapping of charges in the interfacial CT state, although this hypothesis relies on a slow decay of the wave function with distance in organic photovoltaic (OPV) materials.14 Faist et al.15 carried out electroluminescence and device studies on a number of polymers [poly(3-hexylthiophene-2,5diyl) (P3HT), poly(2-methoxy-5-(3′-7′-dimethyloctyloxy)-1,4phenylenevinylene) (MDMO-PPV), poly((9,9-dioctylfluorenyl-2,7-diyl)-alt-5,5-(40,70-di-2-thienyl-20,10,30-benzothiadiazole)) (PFODTBT), poly((4,4′-bis(2- ethylhexyl)dithieno[3,2b:2′,3′-d]silole)-2,6-diyl-alt-(4,7-bis(2-thienyl)-2,1,3- benzothiadiazole)-5,5′-diyl) (Si-PCPDTBT)] blended with fullerene adducts with a range of LUMO levels, but very similar electroluminescence spectra [[6,6]-phenyl C61-butyric acid methyl ester (PCBM), and indeneofullerenes with one (ICMA), two (ICBA), and three (ICTA) adducts. These studies suggested the following empirical condition for efficient charge separation:

INTRODUCTION

A high yield of free polarons at the donor:acceptor interface is required for an efficient organic heterojunction photovoltaic device. A simple Coulombic treatment, assuming electrons and holes are point charges, a relative dielectric constant of 3−4, and a separation of 1−2 nm at the interface, results in a Coulombic binding energy of Ebinding = e2/(4πε0εrreh) ≈ 0.1− 0.5 eV, significantly above the average thermal energy at room temperature, Ethermal ≈ 0.025 eV. Nonetheless, devices have been reported with charge separation efficiencies approaching 100%.1 A number of conceptual models exist for how the process of charge separation may take place. Intramolecular2−9 or intermolecular10,11 delocalization of charge may result in a decreased Coulomb binding. Related to this is the involvement of electronically or vibronically hot charge transfer states, whereby the additional electric and/or vibronic energy an exciton possesses upon reaching the interface may be important in bringing about charge separation, possibly by allowing the exciton to access more delocalized higher energy charge transfer states. This concept is supported by IR pump−push experimental results, showing generation of additional charges when a device containing relaxed CT states is excited with an IR pulse (external quantum efficiency (EQE) enhanced by 10− 50% in polymer:polymer and 6% in polymer:fullerene devices). However, other studies have shown that the internal quantum efficiency of photocurrent generation is insensitive to photon energy over the range of photon energies spanning CT and © 2014 American Chemical Society

ΔECS = Eopt,min − ECT,EL ≥ 0.35 eV

(1)

where ΔECS is the energy required for charge separation, ECT,EL is the CT state emission as measured by electroluminescence, Received: December 19, 2013 Revised: March 14, 2014 Published: March 18, 2014 8253

dx.doi.org/10.1021/jp412449n | J. Phys. Chem. C 2014, 118, 8253−8261

The Journal of Physical Chemistry C

Article

and Eopt,min is the lower absorption onset of either blend component. It is suggested that the CT state may be in competition with the exciton on the material with bandgap Eopt,min. In all but one of the blends considered by Faist et al.,15 Eopt,min corresponds to fullerene absorption. A number of systems have subsequently been reported where Eopt,min corresponds to the polymers, which exhibit high charge separation efficiencies despite not meeting condition (1), suggesting that a different rule may be appropriate in these cases.16−18 However, if charge separation efficiency could be correlated to the properties of the charge transfer state, this would be very useful for material design. In this study, we use quantum chemistry to investigate electronically excited charge transfer states, relating these to chemical structure and relative positions of molecules. A number of computational methods have been used to study molecule pairs. Bakulin et al.8 use intermediate neglect of differential overlap (INDO)/singles configuration interaction (SCI) to calculate charge distributions of two donor:acceptor molecule pairs in the lowest CT state and a higher state CT*. They find a more delocalized electron and hole in the latter than the former. INDO neglects differential overlap between electron wave functions localized on different atoms. Because we are interested in states in which charge is likely to delocalize between molecules, we expect these neglected contributions may be important in defining characteristics of CT states. Yi et al.19 use time dependent density functional theory (TDDFT) with B3LYP/6-31g** to study small molecule pairs, calculating electronic couplings between states and resultant time scales for charge transfer and recombination as a function of mutual position of molecules. Difley and Voorhis20 use constrained DFT (CDFT) to consider charge transfer states in small molecule pairs. In this method, a ground state DFT calculation is performed while constraining charge on the donor to +e, on the acceptor to −e, and spin on each to 1/2. This allows CT states to be considered while avoiding problems in reproducing Coulomb interaction between electron and hole exhibited with TDDFT.21 While informative for certain systems, it is not clear that states at the interface in organic photovoltaic materials will always exhibit exactly one electron charge transfer. Some states may bridge both molecules, exhibiting partial electron transfer. This could not be predicted using CDFT. CDFT also offers no ready route for investigation of higher (hot) excited states of the system, which may play an important role in bringing about charge separation. Liu and Troisi7 study the triplet ground state using unrestrained density functional theory as a proxy for the (more difficult to calculate) singlet excited state, arguing that exchange interaction will be negligible for well separated molecules. This method relies upon the CT state energy being below the triplet of either single component, which may not be the case for some studied systems, and again allows us only to access the lowest energy charge transfer state. Tuned range-separated meta generalized gradient approximation (meta-GGA) hybrids may offer accurate charge transfer state descriptions22,23 but require careful tuning to the system in question. The use of many-body Green’s functions theory within Hedin’s GW approximation and the Bethe-Salpeter equation (GW-BSE)24 is expected to give very accurate predictions of CT state energies but is prohibitively computationally expensive for studying systems of the size required to

determine the effect of changes in the chemical structure on CT states. In order to carry out a broad investigation of the influence of chemical structure on CT state properties, including higher excited states, we require a calculation method which is reliable and tractable. For this reason, we use TDDFT with B3LYP/631g* for these studies. Although we may expect errors due to interactions between functional, basis set, and lack of polarizable medium (discussed in detail in subsequent text), we consider that our results are valid provided that the magnitude and sign of these errors remain consistent, since we are considering a set of relatively similar systems (conjugated oligomer:fullerene pairs each at the same separation). B3LYP is a global hybrid functional with a relatively small contribution (20%) from Hartree−Fock (HF) exchange, resulting in an unphysically small Coulomb energy cost associated with charge transfer excitations when using TDDFT.21 This will result in a decrease in CT state excitation energies compared to a functional with more HF, and the appearance of CT states lower in the spectrum. Opposing this effect, the limited basis set prevents electrons from completely relaxing into their lowest energy configurations. This will have a larger impact on excited and charge transfer states than the ground state, due to the larger spatial extent of excited wave functions and the possibility of electron density in regions between the two molecules. The vacuum surrounding our molecules is also likely to increase CT state excitation energies relative to a calculation in which the polarizable surroundings are considered. We optimize geometries of molecules in the ground state in isolation, and calculated energies are compared to CT state emission, as this is much easier to determine experimentally than CT state absorption. The lack of Stokes shift in calculated values will therefore tend to increase calculated values relative to experimental values.25 We seek here to determine whether the cancellation of errors in TD-B3LYP/6-31g* is sufficiently consistent to give an agreement with experimental trends. Our intent is to then investigate the electronic structure of excitations calculated using this method and consider possible implications for design of molecules giving high charge separation efficiencies.



METHOD The geometries of neutral oligomers and fullerenes are energy minimized in isolation using B3LYP/6-31g*. Excited states are calculated using TDDFT. Increasing oligomer length results in lower excitation energies. For consistency between oligomers (some of which have very different monomer length), oligomer lengths for blend studies were selected such that an increase of one repeat unit resulted in a decrease of less than 0.1 eV in the first excitation energy. For studied cases, only small changes in energy levels were predicted for further increases in oligomer length (see Supporting Information (SI) Figures 1−3). In studies of oligomers:fullerene pairs, fullerenes are placed above the center of the selected oligomer moiety, oriented such that the face opposite the fullerene side chain is cofacial with the conjugated plane of the oligomers moiety, at an interatomic closest approach of 3.5 Å. Preliminary studies on dodecothiophene(12T):fullerene pairs show only minimal energy changes for small changes in relative position and orientation (see SI). Excited state properties are calculated using TD-B3LYP/6-31g*. The degree of charge transfer is assessed by summing Mulliken charges 26 on relevant 8254

dx.doi.org/10.1021/jp412449n | J. Phys. Chem. C 2014, 118, 8253−8261

The Journal of Physical Chemistry C

Article

Figure 1. Chemical structures of studied polymers and fullerenes: [6,6]-phenyl C61-butyric acid methyl ester (PCBM), 1′,1″,4′,4″-tetrahydrodi[1,4]methanonaphthaleno[5,6]fullerene-C60 (ICBA), [6,6]-phenyl C71-butyric acid methyl ester (PC71BM), poly(3-hexylthiophene-2,5-diyl) (P3HT), poly(2-methoxy-5-(3′-7′-dimethyloctyloxy)-1,4-phenylenevinylene) (MDMO-PPV), poly((9,9-dioctylfluorenyl-2,7-diyl)-alt-5,5-(40,70-di2-thienyl-20,10,30-benzothiadiazole)) (PFODTBT), poly[[9-(1-octylnonyl)-9H-carbazole-2,7-diyl]-2,5-thiophenediyl-2,1,3-benzothiadiazole-4,7diyl-2,5-thiophenediyl] (PCDTBT), poly[2,1,3-benzothiadiazole-4,7-diyl[4,4-bis(2-ethylhexyl)-4H-cyclopenta[2,1-b:3,4-b′]dithiophene-2,6-diyl]] (PCPDTBT), poly((4,4′-bis(2- ethylhexyl)dithieno[3,2-b:2′,3′-d]silole)-2,6-diyl-alt-(4,7-bis(2-thienyl)-2,1,3- benzothiadiazole)-5,5′-diyl) (SiPCPDTBT), poly(4,8-bis[(2-ethylhexyl)oxy]benzo[1,2-b:4,5-b]dithiophene-2,6-diyl-3-fluoro-2-[(2-ethylhexyl)carbonyl]thieno[3,4-b]thiophenediyl) (PTB7), poly(isoindigothiophene) (P1TI), poly(isoindigoterthiophene) (P3TI), poly(thieno[3,2b]thiophene-diketopyrrolopyrrole-co-thiophene) (PDPP-TT-T), poly(thieno[3,2b]thiophene-diketopyrrolopyrrole-co-selenophene) (PDPP-TT-S), poly(thieno[3,2b]thiophene-diketopyrrolopyrrole-co-thiazole) (PDPP-TT-Tz). Corresponding oligomers used for quantum chemical studies are indicated in parentheses.



molecules/groups (use of Bader population analysis27 for 12T:PCBM was found to give similar results). Atomic partial charges are calculated via fitting to the electrostatic DFT potential (CHELPG).28 To calculate the Coulomb interaction between molecules, the unshielded Coulomb interactions between partial charges assigned to atoms on opposite molecules are summed, as in donor acceptor

UCoulomb = − ∑ i

∑ j

qiqj 4πε0rij

MODELING SINGLE COMPONENT EXCITATIONS AND BLEND CHARGE TRANSFER STATE ENERGIES Chemical structures of studied polymers, along with the oligomer length chosen to represent them in quantum chemical studies, are shown Figure 1. In all cases apart from P3TI, P1TI, and PTB7, the calculated first excited state energy for oligomers lies within 0.2 eV of the electroluminescence emission of the corresponding polymer (Figure 2a). Figure 2b shows calculated first excitation energies for pairs of oligomers with fullerenes, where fullerenes for which electroluminescence results have been reported are selected. The calculated energies are consistently higher than the experimental electroluminescence peak of blend devices. Limitations with the DFT functional and basis set size effects will affect all calculations presented here, and would be expected to lead to higher than experimentally observed energies. The CT state energies will also be more strongly influenced by the lack of polarizable medium and the neglect of

[unshielded] (2)

where qi represents the CHELPG charge on atom i and rij represents the distance between atoms i and j. Sums are over all atoms on the donor and the acceptor. Quantum chemical calculations were carried out using the Gaussian 09 package.29 Isosurfaces were produced using PyMol,30 spectra were produced using gnuplot,31 and chemical structures were produced using ChemDraw. 8255

dx.doi.org/10.1021/jp412449n | J. Phys. Chem. C 2014, 118, 8253−8261

The Journal of Physical Chemistry C

Article

Figure 2. Comparison of calculated and experimental polymer and polymer:fullerene blend excitation properties: comparisons of first calculated (a) oligomer and (b) oligomer:fullerene pair CT or hybrid state excitation energies with experimental (a) polymer and (b) polymer:fullerene blend film electroluminescence emission peaks. Calculated blend values are for fullerene aligned with a central ring. The value presented for SiPCPDTBT:ICBA is the second excitation energy of the molecule pair (first excitation exhibits no CT). For DA copolymers, the mean of values for fullerene aligned with each unit of the central monomer is taken, and an error bar is included indicating the spread between these positions. Experimental results from refs 15,16,32, and 33.

Properties of these excited states are shown in Figure 3, and a full set of isosurfaces is given in SI Figure 7. We calculate five states at energies significantly lower than the first singlet of either 12T or PCBM, all of which exhibit close to complete electron transfer to the fullerene. The calculated states at higher energies include singlet excitations of 12T and PCBM with properties similar to those calculated for the same molecules in isolation. The HOMO and LUMO of C60 are 5- and 3-fold degenerate, respectively, for a perfectly symmetrical molecule.34 The adduction of C60 to form PCBM (ICBA) lifts this degeneracy but leaves the frontier orbitals quasi degenerate. For C60, the first three unoccupied Kohn−Sham orbitals are calculated within 0.002 eV, and the fourth is separated by 1.18 eV; while for PCBM (ICBA) the first three unoccupied Kohn−Sham orbitals are within 0.28 (0.33) eV, and the fourth is separated by 1.19 (1.09) eV. All TDDFT excitations to C60, PCBM, and ICBA, within the spectral range probed, appear to involve transitions into these three fullerene orbitals. We calculate close to identical charge distribution on 12T in the first three excited states of the blend and assign these excitations to charge transfer from the HOMO of the oligomer to LUMO, LUMO+1, and LUMO+2 of the PCBM (Figure 3). Similarly, we attribute the fourth, fifth, and eighth excited states of the blend to transfer from the HOMO-1 of the oligomer to these three PCBM states (sixth and seventh are single molecule excitations). It should be noted that the oligomer states are spatially perturbed by the presence of another molecule. In the three lowest energy CT states, hole density is significantly more concentrated in the region of the oligomer closest to the fullerene than when this state is calculated for the cation of 12T alone. (The hole also follows the fullerene when it is moved along the chain; see SI Figures 8 and 9). In the next three CT states (fourth, fifth, and eighth), hole density is shifted to outer regions of the oligomer.

Stokes shift (resulting from the use of geometries of charge neutral isolated molecules, rather than excited molecules in intimate contact) than the isolated chromophores, due to the large dipole moment and resultant large reorganization energies of the excited CT complex. Nevertheless the experimental trend in emission energy with chemical structure is well reproduced, with the surprising exception of MDMOPPV:PCBM. For MDMO-PPV, we also calculate a significantly higher HOMO energy than that observed experimentally. To quantify the effect of basis set size, we ran a small study on 5T alone, and when paired with PCBM, using B3LYP with 6-31g* and 6-31+g* (with additional diffuse functions). For 5T alone, we calculate a decrease in the first excitation energy of 5T from 2.64 to 2.56 eV when switching to the larger basis set, while, for 5T:PCBM, we calculate a decrease in the first CT excitation energy from 1.60 to 1.44 eV. This suggests that the effect of basis set confinement is significant, particularly for CT excitations, and can help to explain the overestimate in CT energies with respect to experiment. We use 6-31g* because its computational tractability allows us to make predictions for larger systems than would be feasible with 6-31+g*. Also presented in Figure 2b are calculated CT state energies for two systems in which CT state EL emission is not observed, or is too weak to resolve over single component emission. We calculate a CT state for PFODTBT:ICBA, which fails to meet condition (1), and for which no CT state EL is observed. This is consistent with the hypothesis of Faist et al.,15 that CT states exist in blends failing to meet condition (1), but compete unfavorably with emissive singlet states in the bulk material.



HIGHER EXCITED STATES IN P3HT:PCBM Recent studies8,11 have suggested a role for higher energy CT states in facilitating the separation of charge pairs. In order to probe the energy spectrum and characteristics of higher energy CT states, we focus on the well studied material system P3HT:PCBM, modeled as a 12T:PCBM molecule pair. 8256

dx.doi.org/10.1021/jp412449n | J. Phys. Chem. C 2014, 118, 8253−8261

The Journal of Physical Chemistry C

Article

driving charge separation. Note that these Coulomb binding energies do not take into account the polarization of the surrounding medium, and as such are included only for comparison between binding in different states, rather than as absolute values. We calculate a hybrid excitonic/charge transfer character in the ninth and 10th (Figure 3e) excited states, which we associate with the close energetic proximity of the charge transfer excitations to the first singlets of lone donor and acceptor molecules. Charge separation from such a state would necessarily occur via a state which exhibits complete charge transfer (though possibly one involving molecules further from the interface, which cannot be reliably computed using our current method). Excitation spectra for 12T:PC71BM and 12T:ICBA (see SI Figures 12−15) are qualitatively similar to those of 12T:PCBM, with the first three excitations dominated by transitions from a HOMO localized on regions of the oligomer oriented with the fullerene to the near degenerate LUMO, LUMO+1, and LUMO+2 of the fullerene, and transitions from an oligomer HOMO-1 localized on units far from the fullerene higher in the spectrum. For ICBA, CT state energies are blue-shifted (the lowest by 0.14 eV) in accordance with its higher LUMO level. Our calculataion of the CT state in 12T:PCBM (12T:ICBA) being 0.6 (0.4) eV below S1 of the oligomer implies energetically favorable charge separation, consistent with P3HT:PCBM (P3HT:ICBA) device measurements.35,36



DONOR:ACCEPTOR COPOLYMERS We next calculate excited state spectra for a series of push−pull donor polymers, containing submonomer electron rich and electron poor units, such that electron density is transferred within the monomer upon excitation. We investigate the effects of the position of the fullerene, relative to electron rich and electron poor units, on the charge distributions, energies, and oscillator strengths of the first few excited states for a range of such polymer:fullerene blends. We consider separately the case of material combinations for which the first blend excited state ECT lies well below the optical gap Eg, and those for which ECT and Eg are similar. Cases Where ECT ≪ Eg: PFODTBT:PCBM and PFODTBT:ICBA. For systems in which the first excited state of the blend lies significantly below the excitation energy of individual molecules, close to complete charge transfer is observed in low energy states (see SI Figures 17−19). This is the case in our calculations for PFODTBT (modeled as 3FDTBT) with both PCBM and ICBA. Small shifts in excitation energies (of order 0.01 eV) are calculated as the fullerene aligns with different submonomer units. Cases Where ECT ≈ Eg: P3TI:PC71BM and P1TI:PCBM. Cases of blends where ECT lies close to Eg are particularly interesting. Faist et al.15 found that, for a range of material systems, photocurrent in solar cells appears to switch off when ECT comes within 0.3−0.4 eV of Eg (see condition 1). However, other materials exist for which efficient photocurrent is observed even though CT state emission lies within around 0.1 eV of the polymer emission.16,37,38 Since systems which require a smaller driving force for current generation, quantified by a small (Eg − ECT), offer a higher theoretical VOC and power conversion efficiency, it is of great interest to understand how the excited states of such systems differ from those of others. For both P3TI and P1TI, we calculate a significant degree of electron transfer from the central thiophene block to the central

Figure 3. 12T:PCBM excited states (a) Calculated excitation spectrum for 12T and PCBM, each alone and as a molecule pair. Excitations are colored according to the degree of charge transfer exhibited. Single molecule excitations (CT ≤ 0.10 e) are blue, complete charge transfer excitations (CT ≥ 0.90 e) are red, and hybrid states (0.10 e < CT < 0.90 e) are green. For hybrid states, the degree of charge transfer is indicated. This scheme is used in spectra throughout this work. (b) Isosurfaces showing change in charge distribution between ground and first, second, fourth, and tenth excited states (electron density moves from red to blue regions). Assignment of orbitals involved in transition, and Coulomb binding energy EB, as calculated using eq 2, also indicated (with the exception of 0 → 10, in which this quantity is not relevant as only 0.74 electrons are transferred to the acceptor). Forbidden transitions with negligible oscillator strength (