Вы находитесь на странице: 1из 4

Communications

(Alq3) layer.[29] Meanwhile, this has been recognized as a signature of electrode-limited injection into an organic hopping system.[30] Functionally, the process resembles RichardsonSchottky (RS) thermionic emission, except the RS coefficient (b) is 50 % larger than the theoretical value and, most importantly, the limiting current j is of the order of 104 A cm2, i.e., about three orders less than what PooleFrenkel theory would predict.[28] Once the values of b and j are known one can estimate D from the experimental j(F) = jexp[(D bF1/2)/kT] dependence. It yields a value of D = 1.0 0.1 eV for electron injection from Al into PF. This is in excellent agreement with the value 0.9 eV expected from the reduction potential of PF and the work function of Al. Such an agreement is anything but straightforward. It indicates that in this system the injection barrier(s) is (are) solely determined by the ratio Eox/Ered and the work function of the metals. Consequently, modification of the barriers due to interfacial charges, or the formation of dipole layers, is absent. In conclusion, PF (1) may profitably be employed in bi/multilayer LEDs operated with Mg as a cathode. Upon combining with a hole-transporting material with low Eox, efficient interfacial accumulation can be established and, given an electron mobility of ca. 106 cm2 V1 s1 the LED switching time can be as short as 10 ms.[31]
Received: December 23, 1999 Final version: April 5, 2000

[22] F. Uckert, S. Setayesh, K. Mllen, Macromolecules 1999, 32, 4519. [23] J. Pommerehne, H. Vestweber, W. Guss, R. F. Mahrt, H. Bssler, M. Porsch, J. Daub, Adv. Mater. 1995, 7, 551. [24] R. Cervini, X.-C. Li, G. W. C. Spencer, A. B. Holmes, S. C. Moratti, R. H. Friend, Synth. Met. 1997, 84, 359. [25] S. Alvarado, P. Seidler, D. Lidzey, D. Bradley, Phys. Rev. Lett. 1998, 81, 1082. [26] P. W. M. Blom, M. J. M. de Jong, J. J. M. Vleggar, Appl. Phys. Lett. 1996, 68, 3308. [27] Y.-H. Tak, H. Vestweber, H. Bssler, A. Bleyer, R. Stockmann, H. H. Hrhold, Chem. Phys. 1996, 212, 471. [28] U. Wolf, S. Barth, H. Bssler, Appl. Phys. Lett. 1999, 75, 2035. [29] S. Barth, H. Bssler, P. Mller, H. Vestweber, H. Riel, P. F. Seidler, W. Rie, Phys. Rev. B 1999, 60, 8791. [30] U. Wolf, V. I. Arkhipov, H. Bssler, Phys. Rev. B 1999, 59, 7507. [31] V. I. Nikitenko, H. Bssler, J. Appl. Phys. 1999, 85, 6515.

Unexpected Dimerization of Oxidized FullereneOligothiopheneFullerene Triads**


By Joke J. Apperloo, Bea M. W. Langeveld-Voss, Joop Knol, J. C. Hummelen, and Ren A. J. Janssen* Oxidized states of well-defined p-conjugated oligomers have attracted much interest recently. These compounds serve as models for charge carriers in conjugated polymers and have contributed significantly to the understanding of the nature and spectroscopic characteristics of polarons and bipolarons in conducting polymers. An intriguing phenomenon observed for the radical cations of p-conjugated oligomers is the reversible dimerization that occurs in solution at high concentrations or low temperatures.[1] A recent subject of discussion is whether these dimers are composed of a face-to-face interaction of the p-orbitals of the radical cations (p-dimers) or whether a s-bond is formed between two (terminal) carbon atoms (s-dimers).[2,3] X-ray crystallographic data seem to support the face-to-face interaction via the p-orbitals of the constituent radical cations in oligothiophene and oligopyrrole p-dimers.[3,4] In such a parallel complex, the Davydov interaction between the transition dipoles of the radical cations gives rise to a hypsochromic shift of the absorption bands in the p-dimer.[5] Generally, the driving force for the dimerization process is thought to be provided by the splitting of the two singly oc-

[1] A. Kraft, A. C. Grimsdale, A. B. Holmes, Angew. Chem. Int. Ed. 1998, 37, 402; Angew. Chem. 1998, 110, 416. [2] Y. Yang, MRS Bull. 1997, 22, 31. [3] M. Deuen, H. Bssler, Chem. Unserer Zeit 1997, 31(2), 76. [4] J. Kido, Trends Polym. Sci. 1994, 2, 350. [5] M. Strukelj, F. Papadimitrakopoulos, T. M. Miller, L. J. Rothberg, Science 1995, 267, 1969. [6] Y. Ohmori, N. Tada, M. Yoshida, A. Fujii, K. Yoshino, J. Phys. D: Appl. Phys. 1996, 29, 2983. [7] V. R. Nikitenko, Y.-H. Tak, H. Bssler, J. Appl. Phys. 1998, 84, 2334. [8] J. Salbeck, Ber. Busenges. Phys. Chem. 1996, 100, 1667. [9] M. Strukelj, T. M. Miller, F. Papadimitrakopoulos, S. Son, J. Am. Chem. Soc. 1995, 117, 11 976. [10] Q. Pei, Y. Yang, Chem. Mater. 1995, 7, 1568. [11] A. Schmidt, M. L. Anderson, N. F. Armstrong, J. Appl. Phys. 1995, 78, 5619. [12] N. Tamoto, C. Adachi, K. Nagai, Chem. Mater. 1997, 9, 1077. [13] J. Pommerehne, A. Selz, K. Book, F. Koch, U. Zimmermann, C. Unterlechner, J. H. Wendorff, W. Heitz, H. Bssler, Macromolecules 1997, 9, 1077. [14] Z. Peng, Z. Bao, M. E. Gavin, Adv. Mater. 1998, 10, 680. [15] P. M. Borsenberger, W. T. Gruenbaum, M. B. O'Regan, L. J. Rossi, J. Polym. Sci., Part B: Polym. Phys. 1995, 33, 2143. [16] H. Tokuhisa, M. Era, T. Tsutsui, S. Saito, Appl. Phys. Lett. 1995, 66, 3433. [17] T. Tsutsui, MRS Bull. 1997, 22, 39. [18] R. Fink, C. Frenz, M. Thelakkat, H.-W. Schmidt, Polym. Prepr. (Am. Chem. Soc., Div. Polym. Chem.), 1997, 38, 323. [19] H. Tokuhisa, M. Era, T. Tsutsui, Adv. Mater. 1998, 10, 404. [20] M. Jandke, P. Strohriel, S. Berleb, E. Werner, W. Brtting, Macromolecules 1998, 31, 6434. [21] H. Antoniadis, M. Inbasekaran, E. P. Woo, Appl. Phys. Lett. 1998, 73, 3055.

[*] Dr. R. A. J. Janssen, J. J. Apperloo, Dr. B. M. W. Langeveld-Voss Laboratory of Macromolecular and Organic Chemistry Eindhoven University of Technology PO Box 513, NL-5600 MB Eindhoven (The Netherlands) Dr. J. Knol, Dr. J. C. Hummelen Stratingh Institute and Materials Science Center University of Groningen Nijenborgh 4, NL-9747 AG Groningen (The Netherlands) [**] This research was supported by the Netherlands Organization for Chemical Research (CW) with financial aid of the Netherlands Organization for Scientific Research (NWO) and the Technische Universiteit Eindhoven through a grant in the PIONIER program.

908

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0935-9648/00/1206-0908 $ 17.50+.50/0

Adv. Mater. 2000, 12, No. 12

Communications

cupied, degenerate p-orbitals into two new energy levels, and the net energy gain when the lower of these two levels becomes doubly occupied.[6] An interesting question is whether the intrinsic tendency for dimerization can be controlled by steric interference. Backbone substituents are commonly introduced for solubility purposes, but modulating either their number or size might regulate the proximity of approach of the two constituent p-conjugated backbones. As an example, full alkyl substitution (two hexyl groups per thiophene) prevents the p-dimerization process in oligothienylenevinylenes.[1f,7] Likewise, oligothiophenes endsubstituted with rather bulky (but flexible) Frchet-type dendrons, show a diminished tendency to dimerize.[8] An unambiguous tool to prevent dimerization might be substitution with end groups exceeding in size the short p-dimer stacking distance. Buckminsterfullerene, C60, with its radius of 5.1 ,[9] was thought to be a suitable candidate for checking this idea. Therefore, two dumbbell-shaped fullereneoligothiophenefullerene triads (C60-6T-C60 and C609T-C60) were synthesized[10] (Scheme 1) and their oxidized states were studied in relation to those of the oligothiophenes, 6T and 9T, which lack any a-substitution.[5] Surprisingly, we found that, despite the anticipated steric inhibition of a face-to-face interaction between the oligothiophene moieties, the radical cations of the fullerene-substituted oligothiophenes show a much stronger tendency to dimerize than the corresponding non-end-capped oligomers.

Fig. 1. UV/vis absorption spectra of N-methylfulleropyrrolidine (MP-C60) (4.6 104 M, dotted line), C60-nT-C60 (2.3 104 M, solid lines), and nT (2.3 104 M, dashed lines) for n = 6 and n = 9 recorded in o-dichlorobenzene.

Scheme 1.

The absorption spectra of the C60-nT-C60 triads closely correspond to a superposition of the individual spectra of nT oligomers and N-methylfulleropyrrolidine (Fig. 1). Hence, the electronic interaction between the oligothiophenes and fullerene end groups in the ground state is small. Oxidation of 6T with thianthrenium perchlorate in CH2Cl2 at room temperature replaces the 6T absorption at 2.89 eV by two subgap transitions at 0.83 eV (RC1) and . 1.58 eV (RC2) of molecularly dissolved 6T+ radical cat[5] ions (Fig. 2a). The RC1 and RC2 bands exhibit vibronic side bands at 0.98 and 1.80 eV, respectively. At lower tem. peratures (260 and 220 K are shown), the 6T+ radical cat2+ ions are converted into p-dimers, (6T)2 , exhibiting two bands (D1 and D2) at higher energy (1.10 and 1.96 eV) as a result of the Davydov interaction.[5] These spectral changes with temperature are fully reversible. Upon adding the initial quantities of thianthrenium perchlorate to a CH2Cl2 solution of C60-6T-C60, radical cation
Adv. Mater. 2000, 12, No. 12

Fig. 2. a) UV/vis/near-IR spectra of neutral 6T (solid line) at 295 K, 6T+ at 295 K (dashed), 260 K (dotted), and 220 K (dash-dotted) where radical cations (RC1 and RC2) are reversibly converted into p-dimers (D1 and D2). b) Stepwise oxidation of C60-6T-C60 at 295 K to the singly oxidized state.

bands at 0.81 and 1.55 eV, nearly identical to those of 6T+ , are observed (Fig. 2b, solid lines). This supports the conclusion obtained from the spectra of the neutral molecules, that the electronic effect of the fullero-pyrrolidino moieties on the oligothiophene is also small for the radical cation. Unexpectedly, at a certain degree of oxidation (50 %), a remarkable drop in the intensity of the principal transitions of the oxidized species is observed, which does not occur for 6T, and bands at higher energy (1.13, 1.80, and 1.95 eV) become more pronounced (Fig. 2b, dotted lines). As can be seen in Figure 2, the UV/vis/near-IR spectrum of singly oxidized C60-6T-C60 at room temperature (represented by a solid line for clarity) shows a close similarity to the low
0935-9648/00/1206-0909 $ 17.50+.50/0

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

909

Communications

temperature (220 K) spectrum of 6T+ , where the radical cations are largely present as p-dimers. We conclude that, . in remarkable contrast to 6T+ , the oxidized dumbbell C60+. 6T-C60 has already dimerized to a large extent at room temperature. Surprisingly, the dimerization only starts at increased doping levels (50 %). We have used ESR to study the dimerization of these oxidized species as a function of doping level in more de. . . tail. The ESR spectra of 6T+ , 9T+ , C60-6T-C60+ , and C60+. 9T-C60 in dichloromethane all show an ESR transition at . g = 2.00185 0.00005. For 6T+ some hyperfine coupling can be resolved, while the other radical cations exhibit a single structureless transition.[5] Figure 3a shows the ESR signal intensity of equimolar solutions of 6T and C60-6T-C60 in dichloromethane with increasing doping level. By using a combined UV-ESR cell, a direct correlation with the corresponding absorption spectra could be made. Noticeably, the ESR signal (double integral) follows the absorbance of the radical cation (RC2) band at 1.58 eV accurately (Fig. 3, inset), including the characteristic drop at about 50 % doping for C60-6T-C60.

Fig. 4. a) UV/vis/near-IR spectra of 9T (295 K, solid line) and 9T+ (295 K, dashed line; 240 K, dotted line). b) Same for neutral (solid line) and singly oxidized (dashed line) C60-9T-C60 at 295 K.

Fig. 3. a) ESR signal intensity versus doping level for equimolar solutions of 6T (n) and C60-6T-C60 (*) at 295 K. The inset shows the concomitant evolution of the RC2 band (1.58 eV). b) ESR signal for 9T (~) and C60-9T-C60 (^). In all graphs the dotted line represents a doping level of one.

ibly place for a broad red-shifted band (lmax at 1.9 eV for 6T and 2.36 eV for 9T) as a result of the formation of nanoaggregates (Fig. 5). In good solvents, such as chloroform, the neutral dumbbells do not show this aggregation behavior. This observation might lead to a tentative explanation of the surprisingly high extent of dimerization of . . C60-nT-C60+ in CH2Cl2 compared to nT+ and for the un. usual drop in the concentration of C60-nT-C60+ at intermediate doping levels as inferred from both UV/vis/nearIR and ESR data. The presence of small neutral aggregates (dimers, trimers), in which the first oligomer is more readily oxidized than the second one, would, at the first stages of oxidation, give rise to an increase in the number of unpaired electrons (and hence in ESR signal intensity) and radical cation-like absorption spectra. When all nanoaggre-

For 9T and C60-9T-C60, using the same techniques and procedures, we observe the same phenomena. The spectra of 9T (Fig. 4a) reveal that the radical cations are present as . monomers (9T+ ) at room temperature and as p-dimers 2+ ((9T)2 ) at low temperature.[5] Again there is a close similarity between the room-temperature spectra of oxidized . . C60-9T-C60+ and the low-temperature spectra of 9T+ . ESR intensities as a function of the doping level at room temperature (Fig. 3b) confirm the high degree of dimeriza. tion of oxidized dumbbell C60-9T-C60+ . In order to rationalize the remarkable behavior of the dumbbell radical cations, we investigated the spectral changes of the dumbbells as a function of temperature. While the oxidized dumbbells precipitate irreversibly in dichloromethane during cooling, lowering the temperature of neutral dumbbell solutions causes the original transition (lmax at 2.86 eV for 6T and 2.73 eV for 9T) to make revers910
WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

Fig. 5. UV/vis/near-IR spectral changes upon cooling neutral solutions of: a) C60-6T-C60 and b) C60-9T-C60 in dichloromethane.

0935-9648/00/1206-0910 $ 17.50+.50/0

Adv. Mater. 2000, 12, No. 12

Communications

gates contain a single charge, continued oxidation will produce a second cation radical in the aggregate which, via dimerization, results in a decrease of ESR signal intensity and a shift of the electronic absorptions to lower energy. Any subsequent increase in the number of unpaired electrons could originate either from oxidizing the next oligomer in a small aggregate (doping levels < 1), or from further oxidation up to the dicationic state of one of the two constituent singly oxidized oligomers of a dimer or other closed-shell aggregate (doping levels > 1). This explanation is reminiscent of the formation of a . charge-transfer complex of a radical cations (M+ ) with a neutral molecule (M) to form a dimer cation radical . (M)2+ , which is a well-established phenomenon for simple arenes.[11] The formation of intermediate dimer cation radi. cals (M)2+ before they are converted into cation radical dimers (M)22+ for C60-nT-C60 triads would explain the 50 % doping level where the changes occur. However, the dramatic differences in dimerization between the simple oligothiophenes and the triads implies that the chargetransfer interaction between the oxidized and neutral p-sys. tems, which is responsible for the formation of a (M)2+ species, would be strongly different in the two systems. In our opinion the presence of fullerene end groups does not offer a simple rationale for such a difference. In conclusion, we have shown that, despite the anticipated steric inhibition of p-dimerization of dumbbell radical cations, these molecules dimerize to a large extent at room temperature, in remarkable contrast to the unsubstituted oligomers. We attribute this behavior to the limited solubility of these C60-dumbbells, and regard the dimerization, at least in part, as a solvophobic process.

thesis was performed via the Prato reaction of OHC-nT-CHO with N-methylglycine (8 equiv.) and C60 (4 equiv.) in chlorobenzene during 18 h reflux. The structure and purity of the C60-nT-C60 triads was confirmed with 1 H NMR, 13C NMR, UV/vis, MALDI-TOF mass spectrometry, and HPLC. Full details will be published elsewhere [10]. Since the Prato reaction creates a stereocenter at each pyrrolidine ring, the triads are obtained as mixture of stereo- and regioisomers. Solvents were distilled, dried, and deoxygenated before use. Oxidation of the neutral oligomers was accomplished by adding solutions of thianthrenium perchlorate (THIClO4) [12] from a gas-tight syringe. UV/vis/near-IR spectra were recorded using a Perkin Elmer Lambda 900 spectrometer equipped with an Oxford Optistat cryostat for variable temperature experiments. ESR experiments were carried out with an X-band Bruker ESP 300E spectrometer, operating with a standard cavity, an ER 035 M NMR Gauss meter, and a HP 5350B frequency counter. Received: March 7, 2000 Final version: April 6, 2000 [1] a) M. G. Hill, K. R. Mann, L. L. Miller, J.-F. Penneau, J. Am. Chem. Soc. 1992, 114, 2728. b) P. Audebert, P. Hapiot, J.-M Pernaut, P. Garcia J. Electroanal. Chem. 1993, 361, 283. c) P. Buerle, U. Segelbacher, A. Maier, M. Mehring, J. Am. Chem. Soc. 1993, 115, 10 217. d) J. A. E. H. Van Haare, L. Groenendaal, E. E. Havinga, R. A. J. Janssen, E. W. Meijer, Angew. Chem. Int. Ed. Engl. 1996, 35, 638. e) A. Sakamoto, Y. Furukawa, M. Tasumi, J. Phys. Chem. B 1997, 101, 1726. f) E. Levillain, J. Roncali, J. Am. Chem. Soc. 1999, 121, 8760. [2] a) A. Smie, J. Heinze, Angew. Chem. Int. Ed. Engl. 1997, 36, 363. b) P. Tschuncky, J. Heinze, A. Smie, G. Engelmann, G. Kossmehl, J. Electroanal. Chem. 1997, 433, 223. [3] A. Merz, J. Kronberger, L. Dunsch, A. Neudeck, A. Petr, L. Parkanyi, Angew. Chem. Int. Ed. 1999, 38, 1442. [4] D. D. Graf, R. G. Duan, J. P. Campbell, L. L. Miller, K. R. Mann, J. Am. Chem. Soc. 1997, 119, 5888. [5] J. A. E. H. Van Haare, E. E. Havinga, J. L. J. Van Dongen, R. A. J. Janssen, J. Cornil, J.-L. Brdas, Chem. Eur. J. 1998, 4, 1509. [6] p-Dimers and p-stacks in solution were last reviewed in 1996, see: L. L. Miller, K. R. Mann, Acc. Chem. Res. 1996, 29, 417. [7] J. J. Apperloo, J.-M. Raimundo, P. Frre, J. Roncali, R. A. J. Janssen, Chem. Eur. J. 2000, 6, 1698. [8] J. J. Apperloo, R. A. J. Janssen, P. R. L. Malenfant, L. Groenendaal, J. M. J. Frchet, unpublished. [9] M. S. Dresselhaus, G. Dresselhaus, P. C. Eklund, Science of Fullerenes and Carbon Nanotubes, Academic, London 1996. [10] P. A. van Hal, J. Knol, B. M. W. Langeveld-Voss, S. C. J. Meskers, J. C. Hummelen, R. A. J. Janssen, J. Phys. Chem. A, in press. [11] a) B. Badger, B. Brocklehurst, Trans. Faraday Soc. 1969, 65, 2576, 2582, 2588. b) A. Terahara, H. Ohya-Nishiguchi, N. Hirota, A. Oku, J. Phys.Chem. 1986, 90, 1564. c) H. Yokoi, A. Hatta, K. Ishiguro, Y. Sawaki, J. Am. Chem. Soc. 1998, 120, 12 728. [12] H. J. Shine, C. F. Dais, R. Small, J. Org. Chem. 1964, 29, 21.

Experimental
The C60-nT-C60 triads were synthesized from the corresponding a,w-diformyl-nT oligomers that carry a dodecyl substituent at one of the b-positions of every third thiophene ring, starting at the second ring [5]. The syn-

_______________________

Adv. Mater. 2000, 12, No. 12

WILEY-VCH Verlag GmbH, D-69469 Weinheim, 2000

0935-9648/00/1206-0911 $ 17.50+.50/0

911

Вам также может понравиться