Timothy C.
Parker
a,
Dinesh G. (Dan)
Patel
b,
Karttikay
Moudgil
a,
Stephen
Barlow
a,
Chad
Risko†
a,
Jean-Luc
Brédas‡
a,
John R.
Reynolds
ac and
Seth R.
Marder
*a
aGeorgia Institute of Technology, School of Chemistry & Biochemistry, Center for Organic Photonics and Electronics, 901 Atlantic Drive, Atlanta, GA 30332, USA. E-mail: seth.marder@chemistry.gatech.edu
bThe Pennsylvania State University, 76 University Drive, Hazleton, PA 18292, USA. E-mail: dgp15@psu.edu
cGeorgia Institute of Technology, School of Chemistry & Biochemistry, School of Materials Science, Center for Organic Photonics and Electronics, 901 Atlantic Drive, Atlanta, GA 30332, USA. E-mail: reynolds@chemistry.gatech.edu
First published on 6th August 2014
Increasing the acceptor strength of the widely used acceptor benzothiadiazole (BT) by extending the heterocyclic core is a promising strategy for developing new and stronger acceptors for materials in organic electronics and photonics. In recent years, such heteroannulated BT acceptors have been incorporated into a wide variety of materials that have been used in organic electronic and photonic devices. This review critically assesses the properties of these materials. Although heteroannulation to form acceptors, such as benzo[1,2-c:4,5-c′]bis[1,2,5]thiadiazole (BBT), does result in materials with significantly higher electron affinity (EA) relative to BT, in many cases the extended BT systems also exhibit lower ionization energy (IE) than BT. Both the significantly higher EA and lower IE limit the efficacy of these materials in applications such as bulk heterojunction organic photovoltaics (BHJ-OPV) based on C60. Although the relatively high EA may enable some applications such as air stable organic field effect transistors (OFET), more widespread use of heteroannulated BT acceptors will likely require the ability to moderate or retain the high EA while increasing IE.
One approach to tuning the properties of π-conjugated small molecules or polymers6 is the covalent coupling of at least one electron donor (D) to at least one electron acceptor (A), either directly or through a π-conjugated bridge (π). Several different structural motifs have been extensively studied during the past two decades, leading to advances in: (i) second-order nonlinear optical (NLO) chromophores (D–π–A);7 (ii) two-photon absorbing (TPA) chromophores (D–A–D, D–π–A, and A–D–A);8,9 (iii) electrochromics (–Dn–A–);10–13 (iv) chromophores for dye-sensitized solar cells (DSSCs, D–π–A and D2A);14–17 (v) polymers (–D–A–)n and small molecules (D–A) for OLEDs;18 (vi) small-molecule donors (D–A)19 and acceptors20 and polymers (–D–A–)n for OPVs;21–26 (vii) polymers (–D–A–)n for OFETs,27,28 and (viii) polymers (–D–A–)n for electrochromism.29,30
Early D–A polymers that demonstrated the utility of the approach in modifying optical absorption energies typically used discrete electron accepting groups such as cyano, nitro, or sulfonyl groups as substituents on an aryl subunit or a vinylene in the polymer backbone.31,32 More recently, heterocycles with high EAs have gained favor as stronger and more synthetically variable acceptors. Although a variety of heterocyclic acceptors have been studied and are covered in extensive reviews,22,23 some of the more often-used acceptors, shown in Fig. 1, include thieno[3,4-c]pyrrole-4,6-dione (TPD, 1),33,34 esters of 3-fluorothieno[3,4-b]thiophene-2-carboxylic acid35 (2), diketopyrrolo[3,4-c]pyrrole36 (DPP, 3), isoindigo (4),37–39 and 2,1,3-benzothiadiazole (BT, 5).40Fig. 1 also shows the heterocyclic numbering scheme for BT, which will be used extensively in this review. In particular, BT and its 5-monofluoro- (MFBT) and 5,6-difluoro- (DFBT) derivatives have been used in a variety of materials including both polymers41–43 and small molecules44,45 in OPVs, in OFETs,46,47 as electron-deficient π-bridges in DSSC chromophores,48,49 TPA chromophores,50–52 electrochromics,53,54 and as emitters in small-molecule and polymer OLEDs that include white light emitters,55 emitters with colors spanning the visible spectrum,56–59 and near-infrared emitters.60 The 5,6-dinitro BT derivative also has exhibited strong electron withdrawing ability.61
Fig. 1 Heterocyclic acceptors used in organic electronics. Dashed lines indicate bonds to donor groups in materials. |
Generally, BT is an effective electron acceptor; the presence of the imine functionalities with relatively low energy π*-orbitals gives BT a relatively high EA itself. The molecule can be described as a quasi-quinoidal structure (i.e., with localized, relatively short π-bonds in the benzo ring) rather than a 10π-electron heteroaromatic system; this can increase electronic coupling between substituents in the 4- and 7-positions relative to that found across 1,4-substituted aromatic moieties in both small molecules and polymers. Although BT has proven to be useful, there is still a desire to develop stronger electron acceptors, for example, to decrease the optical gap (Eoptg) in D–A polymers in order to increase light absorption in the near infrared (NIR) and, therefore, to utilize the solar spectrum more efficiently in OPVs,62 to provide electron-accepting materials in OPVs that might replace the currently used fullerene derivatives,63 or to provide air-stable electron-transport materials for n-channel OFETs.64 One way to increase the electron-accepting strength of BT is to heteroannulate at the 5- and 6-positions to give acceptors such as [1,2,5]thiadiazolo[3,4-g]quinoxaline (TDQ, 6) and benzo[1,2-c:4,5-c′]bis[1,2,5]thiadiazole (benzobisthiadiazole, BBT, 7). BBT is structurally similar to the known strong acceptor 865 as well as the core heterocyclic framework of both bis([1,2,5]thiadiazolo)tetracyanoquinodimethane (BTDA-TCNQ, 9) (Fig. 1), which is an electron acceptor that forms conductive charge-transfer crystals with organic donors,66–68 and bis[1,2,5]thiadiazolo-p-quinobis(1,3-dithiole) (TDQBT, 10) (Fig. 1), which forms single-component crystals that have high Hall charge-carrier mobility (3 cm2 V−1 s−1).69–74 In fact, the first known isolated derivative of BBT was 4,8-bis(dicyanomethyl) derivative 11, which was obtained from two-electron reduction of BTDA-TCNQ and protonation of the stable, isolable disodium salt.75 Indeed, in recent years the number of research articles on heteroannulated BT-containing materials has increased significantly, allowing an initial critical assessment of their properties and performance, which is the subject of this current review. Herein, we highlight key aspects of the electronic structure of heteroannulated BT derivatives that give rise to controllable optical and electronic properties and describe how materials containing these acceptors have been employed in organic electronic and photonic applications.
In addition to the interest in the effects of BBT on material properties, the bonding in BBT is intriguing itself in terms of the “tetravalent sulfur” (–NSN–) present in the formal representation shown in Fig. 1, which to maintain a formal charge of zero, must expand its octet to a 10 valence-electron configuration. Other compounds with formally tetravalent sulfur have been reported as a means to probe classical structure and bonding theories including thieno[3,4-c]thiophene derivatives 21 and 2282 (Fig. 3) and thiaphenalene 24 that were reported by Cava and coworkers,83 with 24 also independently reported by Schlessinger and Ponticello.84 Although originally shown as having a tetravalent sulfur, these compounds each have the ylidic resonance structure 27 and the 1,3-dipolar resonance structure 28. Indeed, 21, 22, and 24 each were reportedly unstable and their transient existence as intermediates was only inferred via formation of stable adducts with N-phenylmaleimide (29, Fig. 3), which may occur as a result of a 1,3-dipolar addition of the resonance form 28. The following year, isolable versions were reported by Ponticello and Schlessinger (25)85via sterically deactivating the reactive tetravalent sulfur with adjacent, out-of-plane phenyl rings, and were followed by reports of 2386 and 2687 in back-to-back publications. An X-ray structure of 2388 showed that, compared to thiophene, the length of the S–C bonds shortened by 0.008 Å, nearly within experimental error, that the average CC bond lengths adjacent to S increased by 0.037 Å, and that the fused bond increased by 0.029 Å. These bond length increases are consistent with somewhat less intra thiophene localization in favor of a 10 π-electron annulene-like delocalization. Likewise, relevant bond lengths for [1,2,5]thiadiazole 30, which were determined from gas-phase electron-diffraction82 measurements (and independently from microwave spectroscopy81) and from the X-ray crystal structure of 4,7-diphenyl BT 31,89 can be compared to those determined for 4,8-diphenyl BBT 13 (Fig. 3, Table 1) using X-ray crystallography.79 The diphenyl derivatives were chosen for comparison since the Ph groups are twisted out of plane by 35–43° (BT 31, the range of four independent angles for two independent molecules in the crystal) and 44° (BBT 13), thereby reducing conjugation between the substituents and the heteroaromatic rings that might affect the bonding within the heteroaromatic cores. As seen in Table 1, on going from thiadiazole to BT 31, there is a lengthening of both the imine bonds (“B”) and the C–C bond (“C”; 7a–3a in BT; 8a–3a in BBT, (the lettering scheme for the bond is given in Fig. 3)), indicating a shift in BT away from intra thiadiazole delocalization and toward peripheral annulene-type delocalization; however, note that the benzannulated bonds “D” and “F” are longer than “E” by 0.062 Å and 0.052 Å, respectively (the lettering scheme for the bonds are given in Fig. 3). On going from BT to BBT, the decrease in the S–N “A” bond lengths (0.028 Å) is much more pronounced than that seen going from thiadiazole to BT (0.003 Å), the F bond (equivalent to 8a–3a in BBT (Fig. 3)) lengthens (0.028 Å). Furthermore, in BBT the D bond shortens (0.031 Å), and the E bond lengthens (0.031 Å) such that the bond alternation between D and E is lost and the bonds are equivalent by symmetry at 1.406 Å, similar to the aromatic bonds in benzene. Taken together, such bond distortions on going from BT to BBT are consistent with an increase both in multiple-bond character on the S, which could be expected if there is tetravalent sulfur bonding in BBT, and in annulene-type 14 π-electron delocalization (32) compared to the Kekulé representation (13). To our knowledge, there have been no detailed characterization studies carried out on BBT specifically to test the degree of aromatic delocalization; however, more detailed insight into the bonding in BBT has been explored through computational studies, the results of which are generally useful to rationalize and predict property changes on incorporation of BBT units into small molecules and polymers, as discussed below.
Fig. 3 Non-classical bonding in sulfur heterocycles and bonding considerations in thiadiazole, BT, and BBT. |
Compound | A | B | C | D | E | F |
---|---|---|---|---|---|---|
a Values from gas-phase electron diffraction. b Two complete molecules in the asymmetric unit, each with approximate non-crystallographic Cs symmetry. c Half a molecule in the asymmetric unit; the molecule has crystallographic inversion symmetry and approximate C2h symmetry. d Averages of multiple chemically equivalent but crystallographically independent bond lengths. e Values in parentheses are the difference between the bond length in the compound and the bond length in the compound immediately above. | ||||||
30 (thiadiazole) | 1.632a | 1.329a | 1.413a | — | — | — |
31 DiPh(BT)b | 1.629d (−0.003)e | 1.355d (+0.026)e | 1.452 (+0.039)e | 1.437d | 1.375d | 1.427d |
13 DiPh(BBT)c | 1.601d (−0.028)e | 1.378d (+0.023)e | 1.455d (+0.003)e | 1.406d (−0.031)e | 1.406d (+0.031)e | 1.455 (+0.028)e |
A long-standing question relevant to the proposed tetravalent sulfur bonding is whether the bonding can be better represented as an ylidic structure with an 8-electron sulfur such as in 33 (Fig. 4) rather than as a 10-electron sulfur, as in 34. Strassner and Fabian90 examined a number of structures for various degrees of tetravalent sulfur bonding including acyclic sulfur diimides, thiadiazole, BT, and BBT using density functional theory (DFT, at the B3LYP/6-31G* level). The main finding was that, generally, for compounds that have tetravalent sulfur in Kekulé representations (such as BBT), there was a higher degree of positive charge on the S atoms and a higher degree of negative charge on the N atoms compared to thiadiazole and BT, which is consistent with an increased contribution from the ylidic valence-bond structures such as 33. This is consistent with d-orbitals on S typically having little direct bonding with p-orbitals on N and is also in agreement with a more recent charge density study on acyclic S–N multiple bond containing compounds.91 Another component of the Strassner and Fabian work was the identification of a relatively small singlet–triplet (S0/T1) energy gap of 20.1 kcal mol−1, which, according to the definition proposed by Wirz,92 put BBTs on the borderline of having some degree of diradicaloid character. Similarly, DFT calculations by Bhanuprakash and coworkers93,94 have also shown that BBT and like-molecules, depending on the nature of the chemical modification and the density functional employed (i.e., the amount of [non-local] Hartree–Fock exchange included in hybrid functionals), can have diradicaloid character. These results for BBT were largely confirmed by Shen and coworkers,95 who additionally showed that the Wiberg bond indices of the bonds A, B, and D (Fig. 3) in BBT indicated a “considerable” degree of conjugation around the BBT periphery; this was associated with an aromatic ring current according to their calculations. The natural charges from Shen et al. are also shown in Fig. 4 (34), and they are consistent with Strassner and Fabian's90 calculated ylidic structure. Indeed, this strongly positive sulfur has recently been identified by Reynolds and coworkers as the source of the tendency of BBT and the BT-heteroannulated derivative benzo(triazole-dithiazole) 35 to lower the (B3PW91/6-31**)-calculated LUMO energy of materials relative to BT.96 It should be noted that Yamashita and coworkers reported formation of the Diels–Alder-like adduct 36 in 89% yield by refluxing BBT derivative 13 with N-phenylmaleimide in xylene;79 however, the existence of 36 should not be used to infer a formal 4 + 2 cycloaddition from a Kekulé-type form such as 32 since symmetry allowed cycloadditions may still follow a stepwise “diradical” cyclization mechanism,97 which may be reasonable to expect from diradicaloid structures.
Although the nature of bonding in BBT is interesting from a theoretical standpoint, a more practical question is to what extent material properties are affected upon extension of BT to afford TDQ, BBT, and other possible BT heteroannulation derivatives. DFT studies on small molecules96,98–100 and oligomers101–108 (to represent polymers) have been carried out to explore the electronic, redox, and optical properties of these moieties in a number of donor–acceptor architectures. These investigations show that a subtle interplay between steric and electronic effects (e.g., relative co-planarity of the subunits) on the electronic coupling between the electron donor and acceptor moieties affects the key energy levels of the materials; however, these studies also indicate that replacement of BT with TDQ and, in particular, with BBT, tends to lead to lower ELUMO and similar or slightly higher EHOMO. These tendencies can be explained using a simple perturbational illustration as shown in Fig. 5, which is similar to the analysis used in a computational approach taken by Pandey et al. on a range of acceptors.109 In Fig. 5, the local B3LYP/6-31G* calculated EHOMO and ELUMO of a “donor” (bithiophene, 37), represented by dashed lines across the graph, along with the EHOMO and ELUMO values of BT, TDQ, and BBT are shown. In D–A–D molecules, mixing of the much lower energy LUMOs of TDQ and BBT with the donor LUMO would be expected to significantly decrease ELUMO of the D–A systems compared to their BT analogue (ΔELUMO); on the other hand, the slightly higher local EHOMO of both TDQ and BBT could be expected to raise EHOMO of the D–A system somewhat relative to BT (ΔEHOMO), and such trends are consistent with the calculated EHOMO and ELUMO values for the D–A–D molecules. This would result in a significant narrowing of the HOMO–LUMO gap (green arrows) across the series BT → TDQ → BBT, and thus, assuming that the lowest lying transition is well-described as a HOMO–LUMO transition, lead to a red shift in the absorption band. Another potential consequence of the relatively low ELUMO for TDQ and BBT is that one might expect the LUMO of a D–A–D system would have higher coefficients in the acceptor portion of the molecule for A = TDQ and BBT than for A = BT. On the other hand, there should be relatively little change in the acceptor contributions to the HOMOs of D–A derivatives, since the change in ELUMO going from BT to BBT is significantly larger than the change in EHOMO. This may mean that, for a given donor, there would be less overlap between the HOMO and LUMO in the TDQ and BBT systems compared to their BT analogues, which can in turn decrease the oscillator strength of the optical transitions, such as was computed and discussed in the work by Köse for a range of acceptors;99 however, as is often the case, these effects will be subject to the modulations of the D–A electronic couplings, and may not be manifested if steric interactions force the D–A moieties significantly out of coplanarity in TDQ- or BBT-containing materials.
To examine the extent to which the above-mentioned trends are supported by experiment, examples of small molecules with a D–A–D general structure are compared in Fig. 6. Care should be taken in comparing data between the six groups of compounds in Fig. 6 since different measurements conditions and assumptions were used by the various authors (see figure notes). The porphyrin derivatives 38–40 reported by Therien and coworkers,100 the thiophene donor molecules 41 and 42 reported by Yamashita and coworkers110,111 and 43 and 44 reported by Wang and coworkers112 all largely show the general trends discussed above in that EA is substantially increased while IE either decreases or is approximately the same. Other D–A–D derivatives of TDQ such as 45 and 46 reported by Reynolds and coworkers113–116 also follow these trends; however, other compounds such 47 and 48 that are substituted with bulkier 3,4-ethylenedioxythiophene (EDOT),117 display higher IE than the BT counterpart (ΔIE = +0.13 eV from BT → TDQ). This is presumably due to a greater distortion from planarity in the case of the TDQ–EDOT system than its BT–EDOT analogue, which was seen in the EDOT derivative 49, where a torsion between the plane of the BBT and the EDOT thiophenes (53° in the X-ray crystal structure)113 results in a blue shift (λmax 650 nm) compared to the weaker donor thiophene 17 (Fig. 1, λmax = 702 nm in CH2Cl2).77
Fig. 6 Small molecule D–A–D compounds for A = BT, TDQ, or BBT. For all compounds: (BT, TDQ, or BBT) refer to the parent compounds and R1; R2 refer to parent substituents in Fig. 2. Data are [λmax (nm); IE (eV); EA* (eV)]. IE and EA given are: a reported E1/2ox or E1/2red + 4.8 eV. beEox onset or eEred onset +4.8 eV. ceE1/2ox or eE1/2red + 4.34 eV. deE1/2ox or eE1/2red + 5.1 eV. |
The importance of coplanarity is also demonstrated in the series of D–A–D compounds 50–52,111,112,118–121 where there is a small blue shift on going from thiophene donor 50 to the more electron-donating pyrrole donor in 51, but the expected large red shift when the t-Boc group is removed (52), allowing for planarization. Thus, it is clear from the available small-molecule data and computations that, in the absence of large donor–acceptor torsions in TDQ and BBT derivatives, substituting BT with TDQ and BBT results in a red-shift of λmax arising mostly from a relatively large decrease in ELUMO (increase in EA) and, in many cases, a relatively small increase in EHOMO (decrease in IE); however, the effect on IE is subtle and may differ depending on the donors and the specifics of bonding.
Although many polymers containing BT, TDQ, and BBT have been prepared, relatively few studies have directly compared polymers containing each of the three acceptors. Again, comparisons among studies are difficult due to differences in conditions and methods for measuring properties such as λmax, Eg from the optical absorption onset (Eoptg), IE, and EA. Because of these difficulties, we restrict discussion of acceptor trends in polymers to a few studies that have incorporated each of BT, TDQ, and BBT units into comparable polymers, and other TDQ and BBT containing polymers will be discussed below in the sections on device properties. Early on, Yamashita and coworkers electropolymerized monomers 15, 16, and 17 (Fig. 2) using an ITO electrode to give corresponding polymers 53–55 (Fig. 7), respectively.77,78 Polymers 53–55 were intractable solids with Eoptg of 1.2 eV, 0.7 eV, and 0.5 eV, respectively, in the solid state, and with EA increased markedly by TDQ and BBT, both of which are roughly in line with what is seen in BT, TDQ, and BBT small molecules and this trend is consistent with calculations on oligomeric systems. Although the 0.5 eV Eoptg was among the narrowest optical gaps reported at the time, the extremely poor solubility of 53–55 limited more complete characterization. More recently, Marder, Reynolds, and coworkers reported soluble polymers 56–58 (Fig. 7) with the strong DTP donor, which were prepared via Stille cross coupling reactions.107,108 The solubility of these polymers allowed more thorough characterization, and the data in Fig. 7 show both increasing λmax and decreasing Eoptg across the series BT → TDQ → BBT while the electrochemically estimated IE decreases between BT polymer 56 and BBT polymer 58 (4.9 eV to 4.7 eV). The EA increase was more pronounced across the series BT → TDQ → BBT (3.2 eV → 3.5 eV → 4.0 eV). Again, the properties of polymers in Fig. 7 were generally consistent with those discussed above for small molecules and for computations. The implications of these property trends for materials and device performance in OPVs, NIR OLEDs, and OFETs are discussed below.
Fig. 7 Polymers comparing A = BT, TDQ, and BBT. For all compounds: (BT, TDQ, or BBT) refer to the parent compounds and R2 refers to parent substituents in Fig. 2. Data are [λmax (nm); (Eoptg (eV)); IE (eV); EA* (eV)]. a EA = reported eEpc + 4.4, IE = EA + Eoptg. b IE = eEonsetox + 4.4; EA = eEonsetred + 4.4. |
The low-lying LUMOs of the TDQ and BBT acceptors lead to lower energy optical absorption in donor–acceptor polymers incorporating these moieties than in analogous BT systems (see Table 2). Examples examined in OPVs as blends with fullerenes include (Fig. 8): oligothiophene donor polymers 59, 60,127,128 and 61;129 the rigid thiophene–phenylene–thiophene (TPT) donor polymers 62 and 63;130 the dithiophene–pyrrole donor polymer 64;131 the thiophene–fluorene–thiophene donor polymers 65132 and 66;133 and the DTP donor polymers 56–58108 (Fig. 7). As well as optical and OPV data, the table also lists IE and EA values, along with driving forces for formation of a well-separated polymer/fullerene ion pair from the polymer excited state (−ΔGCS). These values are estimates and comparison between studies is complicated by the use of different assumptions and approximations; however, we have attempted to standardize the methods used for estimation for all compounds; moreover, the trends, if not the absolute numbers, within a given study can provide useful insight.
X | A | λ max/nm | E optg /eV | IE/eV | EA/eV | −ΔGCSb/eV | J SC/mA cm−2 | V OC/V | PCE/% |
---|---|---|---|---|---|---|---|---|---|
a From onset of absorption. b Driving force for charge separation from X to [60]PCBM estimated by ΔG = IE(X) − Eoptg(X) − EAPCBM, using an EA[60]PCBM value of 3.8 eV from IPES.144 A similar value of 3.7 eV is also obtained for both [60]- and [70]PCBM using electrochemical data145 and assuming EA = eE1/2(PCBM) + 4.8 eV (potential vs. FeCp2+/0). c From UPS. d Estimated from electrochemical onset oxidation potentials using the assumption that IE = eEox + 4.8 eV where the potential is quoted vs. FeCp2+/0 (equivalent to offsets of 4.4 eV for SCE or AgCl/Ag reference or 4.7 eV for Ag+/Ag reference). e Estimated from electrochemical onset reduction potential using EA = eEred + 4.8 eV (potential vs. FeCp2+/0) or an equivalent relation. f Driving force given applies to PCBM; a somewhat larger driving force is expected for the fullerene employed in the OPV work, the EA of which was estimated to be 4.2 eV. g Estimated from electrochemical data, but assumptions used not specified. h Reduction potentials vs. AgCl/Ag used to estimate EAs using an offset of 4.4 eV, rather than the 4.72 eV in the original publication, for consistency with other data here. No oxidation potentials given; lower limit for IEs estimated from values of EA and Eoptg. i Values refer to devices with [60]PCBM. j Values refer to devices with [70]PCBM. k higher VOC, JSC, and PCE (up to 2.02%) were obtained with a related acceptor that presumably exhibits a similar EA, but contributes increased light absorption. | |||||||||
59 | BT | 597 | 1.7 | 4.7c | — | 0.8 | 2.87i | 0.60 | 0.55 |
60 | BBT | 902 | 0.7 | 4.3c | — | 0.1 | 0.00719i | 0.04 | — |
62 | BT | 514 | 1.7 | 5.3d | 3.5e | 0.2 | 10.1j | 0.8 | 4.3 |
63 | TDQ | ∼875 | 1.0 | 5.2d | 3.7e | −0.4 | 3.6j | 0.54 | 0.84 |
56 | BT | 674 | 1.4 | 4.9d | 3.2e | 0.3 | 3.9i | 0.510 | 1.3 |
57 | TDQ | 931 | 0.9 | 4.9d | 3.7e | −0.2 | 1.7i | 0.268 | 0.22 |
58 | BBT | 1154 | 0.6 | 4.7d | 4.0e | −0.3 | 0.20i | 0.109 | 0.001 |
65 | TDQ | 815 | 1.2 | 5.1d | 3.9e | −0.1f | 3.4 | 0.58 | 0.70 |
66 | TDQ | 788 | 1.3 | 5.2g | 3.9g | −0.1g | 7.35j | 0.82 | 2.36 |
61 | TDQ | 703 | 1.4 | 5.0d | 3.8e | 0.2 | 1.58i | 0.58 | 0.48 |
64 | TDQ | 756 | 1.1 | 4.8d | 3.7e | 0.1 | 3.41j | 0.39 | 0.43 |
67 | TDQ | 820 | 1.5 | 5.5d | 3.6e | −0.2 | 5.75i | 0.77 | 2.44 |
68 | TDQ | 833 | 1.4 | 5.4d | 3.6e | −0.1 | 3.50i | 0.72 | 1.32 |
69 | TDQ | 872 | 1.4 | 5.3d | 3.5e | −0.1 | 4.25i | 0.65 | 1.42 |
70 | TDQ | 776 | 1.7 | >5.2h | 3.5h | <0.3h | ∼3.04j | ∼0.72 | — |
71 | BBT | 810 | 1.5 | >5.2h | 3.6h | <0.2h | 4.95j | 0.66 | 1.02 |
72 | BBT | 636 | 1.6 | 5.3d | 3.5e | 0.2 | 3.50i | 0.72 | 1.05k |
Fig. 8 Polymers containing A = BT, TDQ, and BBT used to fabricate BHJ OPVs. For all compounds: (BT, TDQ, or BBT) refer to the parent compounds and R2 refers to parent substituents in Fig. 2. |
Although the low-energy absorptions possible in donor–acceptor polymers incorporating BBT and TDQ acceptors are attractive from the point of view of extending solar harvesting into the near-IR, in many cases, especially those where these acceptors are coupled with electron-donating quaterthiophene (60), dithiophene-pyrrole (64), or dithiophene DTP donors (57, 58), these absorptions extend to longer wavelengths than the optimal 830–950 nm referenced above. In several studies, this results in BBT and TDQ derivatives performing much more poorly than the corresponding BT derivative (60vs.59; 63vs.62, 57 and 58vs.56). Multiple factors may contribute to this; for example, a significantly lower absorptivity is found for 60vs.59, and open-circuit voltages are often smaller for the BBT and TDQ derivatives (this effect being much more significant than the effect attributable to the slightly lower estimated IEs relative to BT analogues). However, an important role is presumably played by the lower, in some cases negative, driving forces estimated for charge-separation from the excited states of the BBT and TDQ materials. Indeed, for some compounds (e.g., 58) the electrochemically estimated EAs exceed that of [60]PCBM. Reasonably efficient devices with large open-circuit voltages are obtained for the TDQ derivative, 66; however, in this particular case the assumptions used to estimate the IE (and, therefore, −ΔGCS) and EA were not described, meaning these values are difficult to reliably compare to those for other materials.
The remainder of the table summarizes the properties of materials with less straightforward alternating donor–acceptor structures: copolymers containing both TDQ- and BT-containing monomers (67–69),134 polymers containing platinum in the main chain (70, 71),135 and a small-molecules with electron-withdrawing substituents on the periphery (72).136 Other more recent polymers incorporating dithienosilole137 and benzodithiophene138 donors showed similar results. Although some of these materials lead to moderate efficiencies, charge separation is again estimated to be of marginal thermodynamic feasibility.
Overall, the relatively low fundamental HOMO–LUMO gaps of the TDQ and, especially, BBT building blocks create a challenge in achieving both efficient charge separation and a moderate open-circuit voltage in simple single-layer BHJ devices. The best prospects are for materials with less electron-donating co-monomers or substituents in the case of small molecules. However, it is worth noting that, given the very low-energy absorptions seen for some of these materials, TDQ and BBT materials may be useful in tandem cells,139–141 in which a low-voltage cell that harvests longer wavelength light is sandwiched with, and connected in series to, a higher voltage cell that harvests shorter wavelengths. TDQ and BBT materials with near-IR absorptions could potentially be used in the low-voltage portion either as hole-transport materials, perhaps in conjunction with higher EA electron-transport materials than [60]- or [70]PCBM to ensure efficient charge separation, or, given that electron-transport has been observed in some BBT derivatives (see below), as a light-harvesting electron-transporting component in conjunction with a more electron-donating hole-transporting polymer. Indeed, Wong and coworkers recently reported a dual acceptor polymer 73 containing both BBT and benzotriazole that was used as a PCBM replacement in conjunction with P3HT as a donor, albeit with low reported efficiencies so far (PCE = 0.4%).142 Moreover, some of these materials, in conjunction with appropriate partner materials, may be useful in near-IR photodetectors,143 where a measureable efficient photocurrent generation is required, but a photovoltage is not.
Fig. 9 A = BT, TDQ, or BBT polymers 74,75 and D–A–D compounds 76–82 studied for NIR electroluminescence and fluorescence. Data are [λmax (nm)]. a In toluene. b In CH2Cl2. For all compounds: (BT, TDQ, or BBT) refer to the parent compounds and R1; R2 refer to parent substituents in Fig. 2. |
X | A | IE/eV | EA/eV | ϕ f/% | λ El/nm | EQE/% |
---|---|---|---|---|---|---|
a Estimated from electrochemical oxidation potential using IE = eEox + 4.8 eV where the potential is quoted vs. FeCp2+/0 (equivalent to offsets of 4.3–4.4 eV for SCE or AgCl/Ag references). b Estimated from EA = eEred + 4.8 eV (potential vs. FeCp2+/0) or equivalent expression. c eE 1/2ox or eE1/2red + 4.34 eV. d Estimated from electrochemical data and optical absorption data, but assumptions used are not specified. | ||||||
74 | TDQ | 5.7a | 4.2b | — | 849 | 0.013 |
75 | TDQ | 5.7a | 4.3b | — | 859 | — |
76 | BBT | 5.2a | 3.8d | 5.8 | 1080 | 0.73 |
77 | BBT | 5.2a | 3.9d | 7.4 | 1050 | 0.05 |
78 | BBT | 5.0a | 4.0c | 4.9 | — | |
43 | TDQ | 5.4a | 3.5c | 10.1 | 706 | 0.89 |
44 | BBT | 5.4a | 3.7c | 13.0 | 802 | 0.43 |
80 | TDQ | 5.6d | 3.7d | 21 | 692 | 1.6 |
49 | BBT | 5.6d | 4.1d | 7.6 | 815 | 0.51 |
Fig. 10 Polymers with A = BT, TDQ, or BBT studied in OFETs. For all compounds: (BT, TDQ, or BBT) refer to the parent compounds and R2 refers to parent substituents in Fig. 2. |
X | A | IE/eV | EA/eV | μ h/cm2 V−1 s−1 | μ e/cm2 V−1 s−1 |
---|---|---|---|---|---|
a Estimated from electrochemical onset oxidation potentials using the assumption that IE = eEox + 4.8 eV where the potential is quoted vs. FeCp2+/0 (equivalent to offsets of 4.4 eV for SCE or AgCl/Ag reference or 4.7 where the reference is Ag+/Ag). b Estimated from electrochemical onset reduction potential using EA = eEred + 4.8 eV (potential vs. FeCp2+/0) or an equivalent relation. c Method for determining IE and EA not specified. d Estimated from onset of eEox or eEred, but assumptions used not specified. e From UPS. f Estimated from EA − Eoptg (determined from the absorption onset). | |||||
41 | BT | 5.50a | 3.30b | — | 0.04 |
42 | BBT | 5.29a | 4.04b | — | 0.40 |
17 | BBT | 5.32a | 3.96b | 3.4 × 10−7 | 1.6 × 10−4 |
56 | BT | 4.9a | 3.2b | 1.2 × 10−4 | — |
57 | TDQ | 4.9a | 3.7b | 2.2 × 10−3 | — |
58 | BBT | 4.7a | 4.0b | 1.6 × 10−3 | 7.9 × 10−4 |
83 | TDQ | 4.84a | 3.63b | 3.8 × 10−3 | — |
84 | BBT | 4.8c | 4.0c | 1.1 × 10−1 | 7.4 × 10−2 |
86 | BBT | 4.8c | 4.1c | 1.9 × 10−3 | 1.1 × 10−2 |
87 | BBT | 5.1c | 3.9c | 5.6 × 10−3 | 7.0 × 10−4 |
85 | BBT | 5.33d | 4.32d | 7.1 × 10−4 | 3.3 × 10−3 |
88 | BBT | 5.12d | 3.92d | 3.1 × 10−4 | — |
89 | BBT | 4.6e | 3.8b | 2.5 | Low |
90 | BBT | 4.36f | 3.8b | 1.0 | 0.7 |
91 | BT | 4.75f | 3.4b | 0.17 | — |
92 | BBT | 4.55f | 3.9b | 0.89 | 0.99 |
Footnotes |
† New permanent address: Department of Chemistry, University of Kentucky, Lexington, KY, 40506-0055, USA. |
‡ New permanent address: Division of Physical Sciences and Engineering, King Abdullah University of Science and Technology, Thuwal, 23955-6900, Saudi Arabia. |
This journal is © The Royal Society of Chemistry 2015 |