Subscriber access provided by University of Florida | Smathers Libraries
Article
Dibenzo[a,e]pentalenes with Low-Lying LUMO Energy Levels as Potential n-Type Materials David Christopher Grenz, Maximilian Schmidt, Daniel Kratzert, and Birgit Esser J. Org. Chem., Just Accepted Manuscript • DOI: 10.1021/acs.joc.7b02250 • Publication Date (Web): 12 Dec 2017 Downloaded from http://pubs.acs.org on December 15, 2017
Just Accepted “Just Accepted” manuscripts have been peer-reviewed and accepted for publication. They are posted online prior to technical editing, formatting for publication and author proofing. The American Chemical Society provides “Just Accepted” as a free service to the research community to expedite the dissemination of scientific material as soon as possible after acceptance. “Just Accepted” manuscripts appear in full in PDF format accompanied by an HTML abstract. “Just Accepted” manuscripts have been fully peer reviewed, but should not be considered the official version of record. They are accessible to all readers and citable by the Digital Object Identifier (DOI®). “Just Accepted” is an optional service offered to authors. Therefore, the “Just Accepted” Web site may not include all articles that will be published in the journal. After a manuscript is technically edited and formatted, it will be removed from the “Just Accepted” Web site and published as an ASAP article. Note that technical editing may introduce minor changes to the manuscript text and/or graphics which could affect content, and all legal disclaimers and ethical guidelines that apply to the journal pertain. ACS cannot be held responsible for errors or consequences arising from the use of information contained in these “Just Accepted” manuscripts.
The Journal of Organic Chemistry is published by the American Chemical Society. 1155 Sixteenth Street N.W., Washington, DC 20036 Published by American Chemical Society. Copyright © American Chemical Society. However, no copyright claim is made to original U.S. Government works, or works produced by employees of any Commonwealth realm Crown government in the course of their duties.
Page 1 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Dibenzo[a,e]pentalenes with Low-Lying LUMO Energy Levels as Potential n-Type Materials David C. Grenz,1 Maximilian Schmidt, 1 Daniel Kratzert,2 and Birgit Esser1,* 1
Institute for Organic Chemistry, University of Freiburg, Albertstraße 21, 79104 Freiburg, Germany 2
Institute for Inorganic and Analytical Chemistry, University of Freiburg, Albertstraße 21, 79104 Freiburg, Germany Email:
[email protected] Abstract Ambipolar organic semiconductors are of high interest for organic field-effect transistors. In order to allow for n-type conduction, low LUMO energies are required. Dibenzo[a,e]pentalenes (DBPs) are promising
ACS Paragon Plus Environment
1
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 2 of 27
compounds, however, few derivatives exist with energetically low-lying LUMO levels. Here we present DBPs derivatives with LUMO energies down to −3.73 eV and small bandgaps down to 1.63 eV, determined through cyclic voltammetry, UV/Vis absorption spectroscopy and TDDFT calculations. Single crystal X-ray diffraction analysis revealed a 1D-π-stacking mode. The addition of arylalkynyl-substituents at the fivemembered rings in a facile and versatile synthetic route allowed for a tuning of the band-gaps and LUMO energies. The synthetic route can easily be modified to access a variety of DBP derivatives. The LUMO energies of the DBP derivatives presented herein make them attractive for an application in n-type or ambipolar field-effect transistors.
Introduction Polycyclic conjugated hydrocarbons containing fused five- and six-membered rings have attracted interest for a long time,1 and many new derivatives have been synthesized in the past few years.2–11 These compounds are so-called non-alternant hydrocarbons.12 They have been discussed in the context of antiaromaticity, since counting π-electrons in such linearly fused systems that contain an even number of five-membered rings results in a multiple of 4n, indicating potentially antiaromatic character.4,5,8,11 True antiaromatic character has been observed for pentalene, an unstable and highly reactive species, which could only be isolated when kinetically stabilized through sterically demanding substituents.13,14 When two benzene rings are fused to the pentalene core, dibenzo[a,e]pentalene (DBP, 1 in Figure 1) is formed, where the aromatic character of the six-membered rings prevails, resulting in a thermodynamically and kinetically stable compound.13 The presence of five-membered rings in polycyclic conjugated hydrocarbons leads to a lowering of the LUMO energy relative to acenes, composed solely of sixmembered rings, and consequently a small band gap, as observed for many compounds such as fullerenes,15 pentalene derivatives,2–5 indenofluorenes,6,16 and others.9 This shift in orbital energies can be explained by the non-alternant nature of the π-systems.12 Due to their low bandgap, DBP derivatives show
ACS Paragon Plus Environment
2
Page 3 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
amphoteric redox behavior. This makes them particularly attractive for applications in organic field-effect transistors (OFETs), as shown for several diaceno[a,e]pentalenes.17–21 When the LUMO energy is lowered further, DBP derivatives have the potential to be used as n-type materials.21 Compared to their p-type analogues, much fewer examples of n-type materials for OFETs exist, and hence the development of new materials is of high interest.22,23 We have recently shown that the HOMO and LUMO energy levels in DBP can be tuned through modification of substituents at the 2,7-positions with the strongest effect on the HOMO levels (2 in Figure 1).24 In the LUMO of DBP, the largest coefficients are found on the C5 and C10positions,25 hence it is expected that substituents here have the highest influence on the LUMO energy levels. We herein present the synthesis, optoelectronic properties and TDDFT calculations on eight DBP derivatives 3, where LUMO energies were varied between −3.36 eV and −3.73 eV. Electronically modulating substituents were introduced at the 5,10-positions via alkynyl bridges to ensure co-planarity and unhindered conjugation. R
5 2
7 10
1 Mes R
Mes
Mes
R
3a–h Mes
3
R
a b c d e f g h
4-NMe2 4-OMe H 4-Cl 3,5-(CF3)2 4-CN 4-NO2 F5
2 R = Electron-donating or -withdrawing group
R
24
Figure 1. Dibenzo[a,e]pentalene (DBP, 1), DBP derivatives 2, previously investigated, and donor/acceptorfunctionalized DBPs 3a–h, presented herein.
Results and Discussion In the past ten years many new synthetic strategies of DBP derivatives have been reported.26–34 In the original route by Brand1 the DBP core is generated by addition of Grignard reagents to diphenyl
ACS Paragon Plus Environment
3
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 4 of 27
succindandione followed by a two-fold elimination of water. This has the advantage of allowing for a latestage functionalization of the DBP core and provides facile access to differently substituted derivatives. The synthesis of 3a–h proceeded through bisethynyl-substituted diol 4, which is a versatile DBP precursor for further functionalization (Scheme 1). 4 was accessed in two steps from dibromodiphenyl succindandione 5, which was first subjected to Suzuki-Miyaura coupling with mesityl boronic acid pinacol ester to introduce solubility-enhancing mesityl substituents at the 2,7-positions, providing diketone 7 in 82% yield. The cerium trichloride-promoted35 nucleophilic addition of ethynyl magnesium bromide to 7 yielded diol 4. Electronically modulating aryl groups were attached to 4 using Sonogashira coupling reactions with the corresponding aryl iodides to furnish diols 8a–h, followed by acid-catalyzed water elimination yielding DBP derivatives 3a–h. Using dehydrated p-TsOH in toluene allowed performing the final water elimination step under mild conditions at room temperature. DBP derivatives 3a–h were dark red to purple-colored solids, indicating their low band gap.
ACS Paragon Plus Environment
4
Page 5 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
(pin)B-Mes (6) Pd(PPh3)4 Na2CO3
O Br
Br O
O
5
Mes
Mes
THF/H2O reflux, 15 h 82%
O
7
R I HO
MgBr CeCl3
Mes
Mes
THF −20 ºC to rt 24 h 88%
Pd(PPh3)2Cl2 CuI NEt3/toluene rt or 60 ºC, o.n.
OH 4 R
HO p-TsOH Mes
Mes OH
R
toluene rt, 3-5 h
R = 4-NMe2 (3a) 4-OMe (3b) H (3c) 4-Cl (3d) 3,5-(CF3)2 (3e) 4-CN (3f) 4-NO2 (3g) F5 (3h)
8a–h
15% 48% 68% 38% 77% 50% 64% 30%
Scheme 1. Synthesis of DBP derivatives 3a–h.
5,10-Bis(arylalkynyl)-substituted DBPs have also been reported by Otera and coworkers before.36 Their route, however, started from highly strained 5,6,11,12-tetradehydrodibenzo[a,e]cyclooctene, which is challenging to synthesize.37 Our synthetic route to 3a–h is facile and versatile, since diketone 5 as starting material can be accessed in gram-scale quantities from inexpensive starting materials and reagents.24 Cross-coupling reactions allow for a substitution of the bromine atoms in 5 by any aromatic or other group (in our case mesityl), and the nucleophilic addition of ethynyl magnesium bromide to 7 can also be performed on large scale. In order to allow for an interpretation of the UV/Vis absorption spectra and electrochemical measurements, TDDFT calculations were performed on DBP derivatives 3a–h. For calculations all mesityl groups in 3a–h were replaced by methyl groups, denoted with an asterisk (3a*–h*). In a previous study on 2,7-bisarylalkynyl substituted DBPs 2 the TPSSH38,39 functional best reproduced experimental orbital
ACS Paragon Plus Environment
5
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 6 of 27
energies, while the B3LYP40,41 functional gave the best absorption spectra compared to experimental results (see the Supporting Information for more details).24 Hence, these functionals were used in this study for DBPs 3a*–h* in combination with D342,43 and the def2-TZVP44 basis set.
−1
Figure 2. Cyclic voltammograms of 3a–h (1 mM in CH2Cl2, 0.1 M n-Bu4NPF6, scan rate 0.1 V s , glassy carbon electrode 45
for 3b–f, Pt electrode for 3a, 3g, and 3h).
The amphoteric redox behavior of DBP derivatives 3a–h became apparent in the cyclic voltammograms (Figure 2 and Table 1), where each featured two reduction waves and one oxidation wave. Only for derivatives 3f and 3g with electron-withdrawing substituents, no oxidation was observed in the electrochemical window of the solvent (more CV measurements can be found in the SI). HOMO and LUMO energies were obtained from the onset of the first oxidation/reduction peak vs. Fc/Fc+, assuming an ionization energy of 4.8 eV for ferrocene,46 and compared to calculated orbital energies (TPSSH-D3/def2-
ACS Paragon Plus Environment
6
Page 7 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
TZVP, Figure 3). DBP derivatives 3a–h showed LUMO energies between −3.36 and −3.73 eV, with the latter one being in the range of values reported for typical n-type OFET materials.22
Figure 3. Energy diagram of experimental HOMO/LUMO energies of 3a–h (from CV measurements in solution, +
determined from the onset of the first reduction/oxidation peak vs. Fc/Fc , assuming an ionization energy of 4.8 eV 46
for ferrocene ) and calculated HOMO/LUMO energies of 3a*–h* (TPSSH-D3/def2-TZVP).
The LUMO energy levels of 3a–h lie lower than those of 2,7-substituted DBP derivatives 2, where values between −3.18 and −3.34 eV were measured.24 This is in accordance with our assumption that attaching substituents at the 5,10-positions most strongly affects the LUMO energies. With values between −5.00 and −5.78 eV the HOMO energy levels of 3a–h lie in a range comparable to those measured for 2 (−5.17 to −5.77 eV). The calculated orbital energies of DBP derivatives 3e*–h* with electron-withdrawing substituents were found relatively close in energy to the experimental ones, while those for 3a*–d* with electron-neutral or -rich substituents deviated more strongly. As had been observed for derivatives 2, the
ACS Paragon Plus Environment
7
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 8 of 27
energetic order of the HOMOs and LUMOs of 3a–h corresponds well with the Hammett constants σ of the substituents R: the σpara constants amount to −0.83 for NMe2 (3a), −0.27 for OMe (3b), 0 for H (3c), 0.23 for Cl (3d), 0.43 for CF3 (σmeta, (3e)), 0.66 for CN (3f), and 0.78 for NO2 (3g) (a Hammett substitution constant for five fluorine substituents as in 3h is not available).47 Table 1. Electrochemical and Optical Data for DBPs 3a–h.
a c
E1/2 Red-2 / V
E1/2 Red-1 / V
E1/2 Ox / V
Eg (el. chem.) / eVa
Eg (opt) / eVb
3a
−2.05
−1.52
0.22
1.64
1.63
3b
−1.88
−1.43
0.79c
2.02
1.82
3c
−1.76
−1.34
0.93
2.09
1.88
3d
−1.82
−1.36
0.92
2.09
1.91
3e
−1.60
−1.20
1.05
2.02
1.84
3f
−1.56
−1.21
1.03
2.06
1.90
3g
−1.65
−1.23
–
2.00
1.86
3h
−1.54
−1.14
1.23c
2.07
1.83
b
From the onsets of the reduction/oxidation peaks; from the onset of the longest wavelength absorption band;
anodic peak potential.
The LUMOs in DBPs 3a*–h* are distributed over the DBP core and the arylalkynyl substituents, while the HOMOs are mostly localized on the DBP core (see representative orbital representations in Figure 4). Exceptions are the NMe2- and OMe-substituted derivatives 3a* and 3b*, where the HOMO and HOMO−1 are switched in energetic order compared to the other DBPs. Here the HOMO is spread out over the arylalkynyl substituents, likely due to the electron-donating nature of the NMe2 and OMe substituents. As a result the HOMO has bg symmetry, which leads to an allowed HOMO-LUMO transition (bg à au), visible in the absorption spectrum of 3a and 3b, as will be discussed later.
ACS Paragon Plus Environment
8
Page 9 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Figure 4. Molecular orbitals of 3a* (left) and 3h* (right) (TPSSH-D3/def2-TZVP).
The absorption spectra of DBPs 3a–h demonstrate the low bandgap of these compounds (Figure 5). According to Laporte’s rule the HOMO-LUMO transition is symmetry-forbidden in a centrosymmetric molecule when both orbitals are of au symmetry.48 This is the case for derivatives 3c–h, while in 3a and 3b the HOMO has bg symmetry (see Figure 4). Hence the HOMO à LUMO transitions (S0 to S1) were only visible for 3a and 3b with NMe2 and OMe substituents at 565 and 574 nm, respectively, with extinction
ACS Paragon Plus Environment
9
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 10 of 27
coefficients of log ε = 5.0 and 3.6 (see logarithmic scale in the inset in Figure 5). For 3c–h the lowest visible absorption bands from 494–512 nm correspond to the HOMO−1 à LUMO transitions, according to TDDFT calculations, with log ε values of 4.2–4.7. The other absorption bands represent multiple transitions involving energetically lower or higher lying orbitals (see SI for details on TDDFT calculations). Compared to the 2,7-substituted DBP derivatives 2 described previously, the absorption range in 3a–h is wider, spanning the range from 260 to 790 nm compared to 270–680 nm in 2.24
Figure 5. Absorption spectra of 3a–h in CH2Cl2 solution (Inset: Spectra plotted on a logarithmic scale).
None of the DBP derivatives 3a–h showed fluorescence, which had also been observed for 2,7-substituted DBP derivatives 2.24 For the structurally related indeno[1,2-b]fluorenes the lack of emissive properties was rationalized through transient absorption spectroscopy measurements that showed that the excited state lifetimes were shorter than the timescale at which fluorescence is usually observed.49 Although challenging to crystallize, single crystals suitable for X-ray diffraction were obtained from chloro- (3d, Figure 6) and methoxy-substituted DBP 3b (see the Supporting Information). With bond lengths between 1.397 and 1.436 Å, the C-C bonds in the six-membered rings of the DBP unit are fairly equally distributed and indicate aromatic conjugation within these rings. In the five-membered rings, on the other hand, significant bond length alternation is observed: With 1.521(3) Å the C9’–C16 bond can be classified
ACS Paragon Plus Environment
10
Page 11 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
as a single bond, while the lengths of the bonds C10–C11 and C10–C10’ (1.496(4) Å, 1.476(5) Å) indicate partial conjugation as does the distance of 1.392(3) Å for the C9–C10 bond. These bond lengths indicate that significant conjugation is present between the double bonds in the pentalene units and the arylalkyne units. As can be seen in Figure 6b the molecules of 3d pack in a 1D-π-stacking pattern, where the distance between the DBP cores in two stacked layers amounts to 4.7 Å (4.8 Å for 3b, see the Supporting Information) with a shift of 7.1 Å parallel to the rings’ mean plane. The packing is stabilized by π-πinteractions between the chlorobenzene units, which align in a parallel fashion, spaced out by 3.51 Å (Figure 6c, 2.39 Å for 3b). The dihedral angle between the mesityl groups and the DBP core amounts to 66°, while that between the chlorobenzene units and the DBP core is 81°.
ACS Paragon Plus Environment
11
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 12 of 27
Figure 6. Molecular structure (a) and packing diagrams (b, c) of 3d (hydrogen atoms are omitted for clarity, thermal 50
ellipsoids (in a) are shown at 50% probability level).
Conclusions In conclusion we have presented facile and versatile access to 5,10-bis(arylalkyne)-substituted dibenzo[a,e]pentalenes 3 with mesityl groups in the 2,7-positions. Electrochemical, optical and theoretical investigations of eight derivatives 3a–h showed that the HOMO/LUMO energy levels can be tuned and low-lying LUMO energies down to −3.73 eV and small bandgaps down to 1.63 eV (optical band gap) can be achieved. X-ray crystallographic analysis of derivative 3d revealed 1D-π-stacking in the solid state. These values render DBP derivatives 3a–h of interest as ambipolar or n-type materials for organic fieldeffect transistors.
ACS Paragon Plus Environment
12
Page 13 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
Experimental Section Chemicals were purchased from Acros-Organics, Alfa-Aesar, Merck, Roth, Sigma-Aldrich, ChemPur TCIEurope or VWR and used directly without further purification unless otherwise noted. Experiments with moisture- or oxygen-sensitive substances were carried out in dried glassware – heated under vacuum (10−3 mbar) – under argon atmosphere using standard Schlenk techniques. Anhydrous solvents (THF, toluene, CH2Cl2) were obtained from an M. BRAUN solvent purification system (MB-SPS-800) and stored over 3 Å molecular sieves for a minimum duration of 48 h before use. Other anhydrous solvents were obtained by drying over activated molecular sieves (3 Å) for several days. Diethyl ether, ethyl acetate and cyclohexane were purchased in technical grade and purified via distillation using a rotary evaporator. Other solvents were purchased and used in analytical or HPLC grade. 1H, 13C{1H} and 19F NMR spectra were recorded at 300 K on the following spectrometers: Bruker Avance III HD [300.1 MHz (1H resonance), 282.4 MHz (19F resonance)], Bruker Avance III HD [500.3 MHz (1H resonance), 125.8 MHz (13C resonance)] and at 303 K on a Bruker Avance II [400.1 MHz (1H resonance), 101.6 MHz (13C resonance)]. 1H NMR spectra were referenced to tetramethylsilane (δ = 0.00 ppm) or the residual proton resonance of the respective solvent: CDCl3: δ = 7.26 ppm; CD2Cl2: δ = 5.32 ppm; DMSO-d6: δ = 2.50 ppm.51 13C NMR spectra were referenced to the carbon resonance of the deuterated solvent: CDCl3: δ = 77.16 ppm; CH2Cl2: δ = 53.84 ppm; DMSO-d6: δ = 39.52 ppm.51
19
F NMR spectra were referenced to tetramethylsilane
following the IUPAC recommendations.52 NMR spectra were analyzed using MestReNova software. Analysis followed first order, and the following abbreviations for multiplets were used: singlet (s), broad singlet (br), doublet (d), triplet (t), quartet (q), multiplet (m) and combinations thereof i.e. doublet of doublets (dd). Coupling constants (J) are given in Hertz [Hz]. Electrospray ionization (ESI) and atmospheric pressure chemical ionization (APCI) mass spectra were recorded with a Thermo Fisher Scientific INC. Exactive or LCQ Advantage with an orbitrab analyzer. Only the structurally significant and the strongest fragmentation peaks are reported. UV/Vis absorption spectra were measured on a PerkinElmer LAMBDA
ACS Paragon Plus Environment
13
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 14 of 27
950 UV/Vis spectrophotometer. Melting points were measured using a MPM-HV2 from SchorppGerätetechnik. Silica gel layered alumina plates (Merck, Silica Gel 60 F254) were used for TLC, and compounds were detected using UV light (λ = 366 and 254 nm) and/or with a KMnO4 staining solution (3.0 g KMnO4, 20 g K2CO3, 5.0 mL 5.0% NaOH, 300 mL H2O). For flash column chromatography, silica gel (Machery-Nagel 60, 40–63 μm) was used as the stationary phase. Eluents are described in the respective synthetic procedures. Synthesis of 2,7-dimesityl-4b,9b-dihydroindeno[2,1-a]indene-5,10-dione (7). Pd(PPh3)4 (118 mg, 102 µmol), diketone 5 (800 mg, 2.04 mmol) and 2-mesityl-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (6 (for synthesis see SI), 1.11 g, 4.49 mmol) were dissolved in degassed THF (40 mL) and degassed aq. Na2CO3 (2 M, 40 mL). The resulting suspension was stirred at 80 °C for 10 h. Additional portions of Pd(PPh3)4 (71.0 mg, 61.4 µmol) and 2-mesityl-4,4,5,5-tetramethyl-1,3,2-dioxaborolane (6, 502 mg, 2.04 mmol) were added, and the reaction mixture was stirred at 80 °C for 15 h. After cooling to rt the reaction mixture was extracted with CH2Cl2 (3 × 50 mL). The combined organic layers were dried (MgSO4), filtered, and the solvent was evaporated under reduced pressure. Column chromatography (silica gel, cyclohexane/EtOAc: 8/1 to 5/1) afforded compound 7 (787 mg, 1.67 mmol, 82%) as a white solid. Rf = 0.60 (cyclohexane/EtOAc: 4/1); m.p. 277 °C; 1H NMR (400 MHz, CDCl3): δ = 7.99 (d, J = 7.8 Hz, 2H), 7.54 (d, J = 1.7 Hz, 2H), 7.49 (dd, J = 7.8, 1.7 Hz, 2H), 6.94 (s, 2H), 6.92 (s, 2H), 4.50 (s, 2H), 2.32 (s, 6H), 1.98 (s, 6H), 1.92 (s, 6H); 13C NMR (101 MHz, CDCl3): δ = 201.9, 148.4, 142.6, 137.7, 137.4, 136.0, 135.8, 135.5, 128.4, 128.4, 126.7, 125.6, 52.9, 27.1, 21.2, 21.0; HRMS (APCI+): m/z calcd. for C34H31O2+ 471.2319 [M+H]+, found 471.2318. 5,10-Diethynyl-2,7-dimesityl-4b,5,9b,10-tetrahydroindeno [2,1-a]indene-5,10-diol (4). Commercially available anhydrous CeCl3 (786 mg, 3.19 mmol) was dried at 130 °C for 4 h in vacuo. After cooling to rt anhydrous THF (20 mL) was added, and the suspension was stirred overnight in a glovebox. The white suspension was cooled to −20 °C, and ethynylmagnesium bromide solution (0.5 M in THF, 1.60 mL,
ACS Paragon Plus Environment
14
Page 15 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
3.19 mmol) was added. The suspension was stirred at −20 °C for 4 h. Diketone 7 (485 mg, 1.03 mmol) was added, and the resulting mixture was stirred at −20 °C for 6 h. The reaction mixture was allowed to warm to rt and stirred for another 18 h. Aq. AcOH (10% v/v, 20 mL) was added, and the aqueous layer was extracted with Et2O (3 × 40 mL). The combined organic layers were dried (MgSO4), filtered, and the solvent was removed under reduced pressure. Column chromatography (silica gel, cyclohexane/EtOAc: 10/1 to 4/1) afforded diol 4 (424 mg, 811 µmol, 79%) as a white solid. Rf = 0.45 (cyclohexane/EtOAc: 4/1); m.p. 274 °C (decomposition); 1H NMR (300 MHz, CDCl3): δ = 7.54 (d, J = 7.7 Hz, 2H), 7.36 (d, J = 1.6 Hz, 2H), 7.17 (dd, J = 7.7, 1.6 Hz, 2H), 6.88 (s, 4H), 4.40 (s, 2H), 2.78 (s, 2H), 2.63 (s, 2H), 2.26 (s, 6H), 1.97 (s, 6H), 1.95 (s, 6H); 13C NMR (101 MHz, CDCl3): δ = 147.7, 142.3, 138.5, 136.8, 136.4, 136.1, 135.7, 131.3, 128.2, 128.1, 125.9, 125.6, 84.9, 75.0, 73.4, 62.0, 21.0, 20.9, 20.8; HRMS (ESI+): m/z calcd. for C38H34O2Na+ 545.2451 [M+Na]+, found 545.2449. General procedure A for Sonogashira coupling reactions of 4 with solid iodobenzene derivatives. Diol 4, Pd(PPh3)Cl2, CuI and the respective iodobenzene derivative were dried under vacuum for 1 h. A degassed and anhydrous mixture of toluene and NEt3 was added, and the reaction mixture was stirred at rt for the indicated time. Sat. aq. NH4Cl was added, and the aqueous layer was extracted with Et2O. The combined organic layers were dried (MgSO4), filtered and the solvent removed under reduced pressure. The residue was purified by column chromatography to afford the respective product. General procedure B for Sonogashira coupling reactions of 4 with liquid iodobenzene derivatives. Diol 4 and Pd(PPh3)Cl2 were dried under vacuum for 15 h. CuI and a degassed and anhydrous mixture of toluene and NEt3 were added, followed by the respective iodobenzene derivative. The mixture was stirred at rt for the indicated time. Sat. aq. NH4Cl was added, and the aqueous layer was extracted with Et2O. The combined organic layers were dried (MgSO4), filtered and the solvent removed under reduced pressure. The residue was purified by column chromatography to afford the respective product.
ACS Paragon Plus Environment
15
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 16 of 27
General procedure C for Sonogashira coupling and water elimination to DBP derivatives 3. Diol 4, Pd(PPh3)Cl2, CuI and the respective iodobenzene derivative were dried under vacuum for 1 h. A degassed and anhydrous mixture of toluene and NEt3 was added, and the mixture was stirred at rt for the indicated time. Sat. aq. NH4Cl was added, and the aqueous layer was extracted with Et2O. The combined organic layers were dried (MgSO4), filtered and the solvent removed under reduced pressure. The crude product was dissolved in anhydrous and degassed toluene and anhydrous p-TsOH was added. The mixture was stirred for the indicated time in a glovebox. H2O was added, and the aqueous layer was extracted with CH2Cl2. The combined organic layers were washed with brine, dried (MgSO4), filtered, and the solvent was removed under reduced pressure. The residue was purified by column chromatography. General procedure D for water elimination from diols 8. The respective diol was dissolved in anhydrous toluene and anhydrous p-TsOH was added. The mixture was stirred for the indicated time. H2O was added and the aqueous layer was extracted with CH2Cl2. The combined organic layers were washed with brine, dried (MgSO4), filtered and the solvent was removed under reduced pressure. The residue was purified by flash column chromatography. 5,10-Bis((4-(dimethylamino)phenyl)ethynyl)-2,7-dimesityl-4b,5,9b,10-tetrahydroindeno[2,1-a]indene5,10-diol (8a). General procedure A was followed using diol 4 (60.0 mg, 115 µmol), PdCl2(PPh3)2 (3.20 mg, 4.60 µmol), CuI (0.09 mg, 4.60 µmol), 4-iodo-N,N-dimethylaniline (for synthesis see SI, 85.0 mg, 345 µmol), toluene (4 mL) and NEt3 (4 mL). The reaction mixture was stirred for 15 h at rt. Column chromatography (silica gel, cyclohexane/EtOAc: 20/1 to 1/1) afforded compound 8a (31.0 mg, 40.7 µmol, 35%) as a beige solid. Rf 0.42 (cyclohexane/EtOAc: 4/1); 1H NMR (300 MHz, CDCl3): δ = 7.60 (d, J = 7.7 Hz, 2H), 7.44 (d, J = 1.6 Hz, 2H), 7.33–7.28 (m, 4H), 7.15 (dd, J = 7.7, 1.6 Hz, 2H), 6.88 (s, 4H), 6.58–6.53 (m, 4H), 4.50 (s, 2H), 2.94 (s, 2H), 2.90 (s, 12H), 2.26 (s, 6H), 1.99 (s, 6H), 1.98 (s, 6H). 4,4'-((2,7-Dimesitylindeno[2,1-a]indene-5,10-diyl)bis(ethyne-2,1-diyl))bis(N,N-dimethylaniline)
(3a).
General procedure D was followed using diol 8a (38.6 mg, 48.8 µmol), anhydrous p-TsOH (12.1 mg,
ACS Paragon Plus Environment
16
Page 17 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
70.3 µmol) and toluene (7 mL). The mixture was stirred at rt for 4 h. Column chromatography (silica gel, cyclohexane/CH2Cl2: 3/2 to 0/1) afforded 3a (14.8 mg, 20.4 µmol, 42%) as a dark purple solid. Rf 0.76 (cyclohexane/CH2Cl2: 1/1); m.p. >350 °C; 1H NMR (300 MHz, CDCl3): δ = 7.48–7.43 (m, 6H), 6.98 (s, 4H), 6.95 (d, J = 1.5 Hz, 2H), 6.77 (dd, J = 7.4, 1.5 Hz, 2H), 6.68–6.63 (m, 4H), 3.02 (s, 12H), 2.36 (s, 6H), 2.14 (s, 12H); HRMS (APCI+): m/z calcd. for C54H49N2+ 725.3890 [M+H]+, found 725.3892; UV/Vis (CH2Cl2): λabs (ε) = 295 (176,024), 365 (24,525), 420 (27,809), 565 (103,669) nm. 2,7-Dimesityl-5,10-bis((4-methoxyphenyl)ethynyl)indeno[2,1-a]indene (3b). General procedure C was followed using diol 4 (50.0 mg, 96 µmol), 4-iodoanisole (67.0 mg, 287 µmol), PdCl2(PPh3)2 (3.40 mg, 4.80 µmol), CuI (0.90 mg, 4.80 µmol), NEt3 (0.5 mL) and toluene (2.0 mL). The reaction mixture was stirred for 18 h. After work-up the crude product was dissolved in toluene (7.0 mL), and anhydrous p-TsOH (0.5 M in toluene, 1.0 mL) was added. The dark red solution was stirred for 3 h. Column chromatography (silica gel, cyclohexane/EtOAc/NEt3: 100/0/3 to 100/25/3) afforded 3b (32.3 mg, 46.0 µmol, 48%) as a red solid. Rf 0.51 (cyclohexane/EtOAc: 4/1); 1H NMR (400 MHz, CDCl3): δ = 7.53–7.49 (m, 4H), 7.44 (d, J = 7.5 Hz, 2H), 6.97 (s, 4H), 6.93 (d, J = 1.5 Hz, 2H), 6.92–6.87 (m, 4H), 6.78 (dd, J = 7.5, 1.5 Hz, 2H), 3.84 (s, 6H), 2.35 (s, 6H), 2.13 (s, 12H); 13C NMR (101 MHz, CDCl3): δ = 160.5, 149.0, 148.7, 141.6, 139.0, 136.8, 136.2, 133.8, 132.0, 128.7, 128.2, 123.2, 122.5, 120.1, 115.0, 114.3, 103.6, 83.7, 55.5, 21.1, 20.8; HRMS (APCI+): m/z calcd. for C52H43O2+ 699.3258 [M+H]+, found 699.3259; UV/Vis (CH2Cl2): λabs (ε) = 285 (78,630), 377 (18,555), 480 (33,761), 512 (39,546) nm. 2,7-Dimesityl-5,10-bis(phenylethynyl)-4b,5,9b,10-tetrahydroindeno[2,1-a]indene-5,10-diol
(8c).
General procedure A was followed using diol 4 (60.0 mg, 115 µmol), PdCl2(PPh3)2 (8.10 mg, 11.5 µmol), CuI (2.20 mg, 11.5 µmol), iodobenzene (110 mg, 544 µmol), toluene (5 mL) and NEt3 (1 mL). The mixture was stirred for 17 h at rt. Column chromatography (silica gel, cyclohexane/EtOAc: 5/1 to 4/1) afforded 8c (61.6 mg, 91.3 mmol, 79%) as a yellow solid. Rf 0.54 (cyclohexane/EtOAc: 4/1); 1H NMR (300 MHz, CDCl3): δ = 7.68 (d, J = 7.6 Hz, 2H), 7.53–7.49 (m, 6H), 7.35–7.33 (m, 6H), 7.36 (dd, J = 7.6, 1.6 Hz, 2H), 6.95 (s, 4H),
ACS Paragon Plus Environment
17
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 18 of 27
4.60 (s, 2H), 3.01 (s, 2H), 2.33 (s, 6H), 2.06 (s, 12H); HRMS (APCI+): m/z calcd. for C50H39+ 639.3046 [M+H]+, found 639.3042. 2,7-Dimesityl-5,10-bis(phenylethynyl)indeno[2,1-a]indene (3c). General procedure D was followed using crude diol 8c (61.6 mg, 91.3 µmol), anhydrous p-TsOH (20.0 mg, 116 µmol) and toluene (10 mL). The mixture was stirred at rt for 5 h. Recrystallization from MeCN afforded 3c (50.0 mg, 78.3 µmol, 86%) as a dark red solid. Rf 0.72 (cyclohexane/EtOAc: 5/1); m.p. >350 °C; 1H NMR (300 MHz, CDCl3): δ = 7.52–7.48 (m, 4H), 7.41 (d, J = 7.4 Hz, 2H), 7.37–7.32 (m, 4H), 7.26 (s, 2H), 6.97 (s, 4H), 6.90 (d, J = 1.4 Hz, 2H), 6.78 (dd, J = 7.4, 1.4 Hz, 2H), 2.34 (s, 6H), 2.12 (s, 12H); 13C NMR (126 MHz, CD2Cl2) δ = 150.2, 149.2, 143.0, 139.1, 137.3, 136.4, 136.0, 133.9, 131.9, 129.7, 129.5, 128.6, 123.9, 123.2, 121.7, 120.3, 102.6, 85.5, 21.3, 21.0; HRMS (APCI+): m/z calcd. for C50H39+ 639.3046 [M+H]+, found 639.3043; UV/Vis (CH2Cl2): λabs (ε) = 284 (70,087), 302 (63,764), 468 (16,213), 498 (18,225) nm. 5,10-Bis((4-chlorophenyl)ethynyl)-2,7-dimesityl-4b,5,9b,10-tetrahydroindeno[2,1-a]indene-5,10-diol (8d). General procedure A was followed using diol 4 (93.0 mg, 178 µmol), PdCl2(PPh3)2 (10.0 mg, 14.3 µmol), CuI (4.70 mg, 24.7 µmol), 1-chloro-4-iodobenzene (130 mg, 543 µmol), toluene (10 mL) and NEt3 (3 mL). The reaction mixture was stirred at rt for 14 h and at 60 °C for 7 h. Column chromatography (silica gel, cyclohexane/EtOAc: 10/1 to 4/1) afforded 8d (65.9 mg, 88.6 mmol, 50%) as a white solid. Rf 0.70 (cyclohexane/EtOAc: 4/1); m.p. 218 °C (decomposition); 1H NMR (300 MHz, CDCl3): δ = 7.66 (d, J = 7.7 Hz, 2H), 7.50 (d, J = 1.5 Hz, 2H), 7.45–7.42 (m, 4H), 7.32–7.30 (m, 4H), 7.26 (dd, J = 7.7, 1.5 Hz, 2H), 6.96 (s, 4H), 4.58 (s, 2H), 2.99 (s, 2H), 2.33 (s, 6H), 2.05 (s, 6H), 2.05 (s, 6H); 13C NMR (126 MHz, CDCl3): δ = 148.2, 142.4, 138.6, 137.0, 136.6, 136.4, 135.9, 134.9, 133.2, 131.4, 128.9, 128.4, 128.3, 126.2, 125.8, 121.0, 91.2, 84.4, 75.8, 62.5, 27.1, 21.2, 21.1. 5,10-Bis((4-chlorophenyl)ethynyl)-2,7-dimesitylindeno[2,1-a]indene (3d). General procedure D was followed using diol 8d (65.9 mg, 88.6 µmol), anhydrous p-TsOH (20.0 mg, 116 µmol) and toluene (10 mL).
ACS Paragon Plus Environment
18
Page 19 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
The reaction mixture was stirred at rt for 1 h. Column chromatography (silica gel, cyclohexane/CH2Cl2: 2/1 to 0/1) afforded 3d (48.4 mg, 68.4 µmol, 76%) as a dark red solid. Rf 0.72 (cyclohexane/EtOAc: 5/1); m.p. >350 °C; 1H NMR (300 MHz, CD2Cl2): δ = 7.62–7.58 (m, 4H), 7.47 (dd, J = 7.5, 0.5 Hz, 2H), 7.41–7.37 (m, 6H), 6.96–6.95 (m, 4H), 6.80 (dd, J = 7.5, 1.5 Hz, 2H), 2.32 (s, 6H), 2.12 (s, 12H); 13C NMR (101 MHz, CD2Cl2): δ = 150.0, 149.4, 142.8, 139.2, 137.3, 133.9, 132.7, 132.1, 130.0, 129.6, 129.5, 129.2, 128.6, 123.9, 123.2, 120.6, 103.9, 84.7, 21.3, 21.0; HRMS (APCI+): m/z calcd. for C50H37Cl2+ 707.2267 [M+H]+, found 707.2263; UV/Vis (CH2Cl2): λabs (ε) = 286 (56,117), 320 (26,720), 470 (25,946), 503 (28,790) nm. 5,10-Bis((3,5-bis(trifluoromethyl)phenyl)ethynyl)-2,7-dimesityl-4b,5,9b,10-tetrahydroindeno[2,1a]indene-5,10-diol (8e). General procedure A was followed using diol 4 (50.0 mg, 96.0 µmol), PdCl2(PPh3)2 (5.60 mg, 8.00 µmol), CuI (2.28 mg, 8.00 µmol), 1-iodo-3,5-bis(trifluoromethyl)benzene (130 mg, 384 µmol), toluene (3 mL) and NEt3 (2 mL). The reaction mixture was stirred at rt for 18 h. Column chromatography (silica gel, cyclohexane/EtOAc: 10/1 to 2/1) afforded 8e (77.2 mg, 81.5 µmol, 85%) as a colorless solid. Rf 0.48 (cyclohexane/EtOAc: 5/1); 1H NMR (500 MHz, CDCl3): δ = 7.95 (s, 4H), 7.85 (s, 2H), 7.67 (d, J = 7.7 Hz, 2H), 7.50 (d, J = 1.5 Hz, 2H), 7.31 (dd, J = 7.7, 1.5 Hz, 2H), 6.97 (s, 4H), 4.63 (s, 2H), 2.99 (s, 2H), 2.33 (s, 6H), 2.06 (s, 6H), 2.05 (s, 6H); 13C NMR (126 MHz, CDCl3): δ = 147.5, 142.6, 138.2, 137.0, 136.1, 135.7, 132.1 (q, J = 33.8 Hz), 131.7, 128.3, 128.2, 126.1, 125.7, 124.6, 122.9 (q, J = 273.2 Hz), 93.6, 82.4, 75.6, 62.5, 21.0, 20.9, 20.9; 19F NMR (282 MHz, CDCl3): δ = −63.1; HRMS (ESI+): m/z calcd. for C54H38O2F12Na+ 969.2572 [M+Na]+, found 969.2571. 5,10-Bis((3,5-bis(trifluoromethyl)phenyl)ethynyl)-2,7-dimesitylindeno[2,1-a]indene
(3e).
General
procedure D was followed using crude diol 8e (58.3 mg, 62.0 µmol), anhydrous p-TsOH (23.1 mg, 134 µmol) and toluene (10 mL). The reaction mixture was stirred at rt for 2.5 h. Column chromatography (silica gel, cyclohexane/CH2Cl2: 25/1 to 5/1) afforded 3e (50.7 mg, 55.7 µmol, 90%) as a red solid. Rf 0.60 (cyclohexane/CH2Cl2: 1/1); m.p. >350 °C; 1H NMR (500 MHz, CD2Cl2): δ = 7.98 (s, 4H), 7.86 (s, 2H), 7.41 (d, J = 7.4 Hz, 2H), 6.98 (s, 4H), 6.90 (d, J = 1.4 Hz, 2H), 6.84 (dd, J = 7.4, 1.4 Hz, 2H), 2.35 (s, 6H), 2.13 (s, 12H);
ACS Paragon Plus Environment
19
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
13
Page 20 of 27
C NMR (126 MHz, CDCl3): δ = 151.1, 148.5, 142.9, 138.3, 137.1, 135.9, 132.4 (q, J = 34.1 Hz), 131.7, 131.0,
129.3, 128.2, 123.9, 123.4, 123.2, 122.8 (q, J = 273.0 Hz), 119.3, 99.0, 86.7, 21.0, 20.7; 19F NMR (282 MHz, CDCl3): δ = −63.1; HRMS (APCI+): m/z calcd. for C54H34F12+ 910.2474 [M]+, found 910.2474; UV/Vis (CH2Cl2): λabs (ε) = 286 (112,007), 338 (61,768), 468 (43,308), 494 (45,733) nm. 4,4'-((2,7-Dimesitylindeno[2,1-a]indene-5,10-diyl)bis(ethyne-2,1-diyl))dibenzonitrile
(3f).
General
procedure C was followed using diol 4 (82.0 mg, 157 µmol), 4-iodobenzonitrile (180 mg, 784 µmol), PdCl2(PPh3)2 (8.80 mg, 12.6 µmol), CuI (2.40 mg, 12.6 µmol), NEt3 (0.3 mL) and toluene (7.0 mL). The reaction mixture was stirred for 2 h. After work-up and filtration through a silica plug the crude product was dissolved in toluene (5.0 mL) and p-TsOH (34.0 mg, 197 µmol) was added. The dark red solution was stirred at rt for 3 h. Two consecutive column chromatography steps (silica gel, cyclohexane/EtOAc: 1/0 to 0/1 and cyclohexane/CH2Cl2: 1/2 to 0/1) afforded 3f (54.4 mg, 79.0 µmol, 50%) as a red solid. Rf 0.51 (cyclohexane/EtOAc: 4/1); m.p. >350 °C; 1H NMR (300 MHz, CDCl3): δ = 7.69–7.63 (m, 8H), 7.40 (d, J = 7.4 Hz, 2H), 6.98 (s, 4H), 6.89 (d, J = 1.4 Hz, 2H), 6.80 (dd, J = 7.4, 1.4 Hz, 2H), 2.35 (s, 6H), 2.13 (s, 12H); 13
C NMR (126 MHz, CDCl3): δ = 151.0, 148.7, 142.8, 138.5, 137.2, 136.1, 132.6, 132.3, 131.3, 129.3, 128.3,
127.4, 123.5, 123.2, 119.7, 118.4, 112.6, 101.0, 87.9, 21.2, 20.8; HRMS (APCI+): m/z calcd. for C52H40N3+ 706.3217 [M+NH4]+, found 706.3215; UV/Vis (CH2Cl2): λabs (ε) = 292 (71,386), 347 (29,580), 474 (27,434), 506 (28,588) nm. 2,7-Dimesityl-5,10-bis((4-nitrophenyl)ethynyl)-4b,5,9b,10-tetrahydroindeno[2,1-a]indene-5,10-diol (8g). General procedure A was followed using diol 4 (81.2 mg, 155 µmol), PdCl2(PPh3)2 (9.20 mg, 13.1 µmol), CuI (2.00 mg, 15.8 µmol), 1-iodo-4-nitrobenzene (155 mg, 622 µmol), toluene (12.0 mL) and NEt3 (3.0 mL). The reaction mixture was stirred at rt for 40 h. Column chromatography (silica gel, cyclohexane/EtOAc: 8/1) afforded 8g (83.0 mg, 109 µmol, 70%) as a brown solid. Rf 0.51 (cyclohexane/EtOAc: 3/1); m.p. 189 °C (decomposition); 1H NMR (300 MHz, CDCl3): δ = 8.25–8.21 (m, 4H), 7.69–7.65 (m, 6H), 7.51 (d, J = 1.6 Hz, 2H), 7.31 (dd, J = 7.7, 1.6 Hz, 2H), 6.97 (s, 4H), 4.64 (s, 2H), 3.07 (s,
ACS Paragon Plus Environment
20
Page 21 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
2H), 2.34 (s, 6H), 2.06 (s, 12H); 13C NMR (101 MHz, CDCl3): δ = 147.7, 147.5, 142.6, 138.3, 137.0, 136.2, 136.1, 135.7, 132.6, 131.6, 129.2, 128.3, 128.2, 126.1, 125.7, 123.6, 95.2, 83.5, 75.7, 62.3, 21.0, 20.9, 20.9; HRMS (ESI+): m/z calcd. for C50H40O6N2Na+ 787.2779 [M+Na]+, found 787.2777. 2,7-Dimesityl-5,10-bis((4-nitrophenyl)ethynyl)indeno[2,1-a]indene (3g). General procedure D was followed using crude diol 8g (48.4 mg, 62.3 µmol), anhydrous p-TsOH (42.2 mg, 245 µmol) and toluene (13 mL). The reaction mixture was stirred at rt for 3.5 h. Column chromatography (silica gel, cyclohexane/CH2Cl2: 1/1 to 0/1) afforded 3g (41.3 mg, 56.7 µmol, 91%) as a red solid. Rf 0.40 (cyclohexane/CH2Cl2: 1/1); 1H NMR (300 MHz, CDCl3): δ = 8.27–8.22 (m, 4H), 7.73–7.69 (m, 4H), 7.50 (d, J = 7.4 Hz, 2H), 6.98 (s, 4H), 6.90 (d, J = 1.4 Hz, 2H), 6.82 (dd, J = 7.4, 1.4 Hz, 2H), 2.35 (s, 6H), 2.13 (s, 12H); HRMS (APCI+): m/z calcd. for C50H40O4N3+ 746.3013 [M+NH4]+, found 746.3008; UV/Vis (CH2Cl2): λabs (ε) = 300 (42,251), 365 (21,941), 483 (17,505), 512 (17,907) nm. 2,7-Dimesityl-5,10-bis((perfluorophenyl)ethynyl)-4b,5,9b,10-tetrahydroindeno[2,1-a]indene-5,10-diol (8h). General procedure A was followed using diol 4 (50.0 mg, 96.0 µmol), PdCl2(PPh3)2 (5.60 mg, 8.00 µmol), CuI (2.28 mg, 8.00 µmol), pentafluoroiodobenzene (113 mg, 384 µmol), toluene (3 mL) and NEt3 (2 mL). The reaction mixture was stirred at rt for 18 h. Column chromatography (silica gel, cyclohexane/EtOAc: 10/1 to 6/1) afforded 8h (26.7 mg, 31.2 µmol, 33%) as a colorless solid. Rf 0.40 (cyclohexane/EtOAc: 5/1); 1H NMR (300 MHz, CDCl3): δ = 7.65 (d, J = 7.7 Hz, 2H), 7.50 (d, J = 1.5 Hz, 2H), 7.30 (dd, J = 7.7, 1.5 Hz, 2H), 6.97 (s, 4H), 4.63 (s, 2H), 3.02 (s, 2H), 2.34 (s, 6H), 2.05 (s, 12H); 13C NMR (126 MHz, CDCl3): δ = 148.6–148.4 (m), 147.2, 146.6–146.4 (m), 142.7, 141.1–140.9 (m), 138.8–138.5 (m), 138.2, 136.9, 136.7–136.5 (m), 136.1, 136.0, 135.8, 131.7, 128.2, 128.2, 126.1, 125.7, 102.6, 75.7, 69.5, 62.3, 21.0, 20.9, 20.8; 19F NMR (282 MHz, THF-d8): δ = −(138.1–138.2) (m), −(154.7–154.8) (m), −(164.0– 164.2) (m); HRMS (APCI+): m/z calcd. for C50H32O2F10Na+ 877.2135 [M+Na]+, found 877.2133.
ACS Paragon Plus Environment
21
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 22 of 27
2,7-Dimesityl-5,10-bis((perfluorophenyl)ethynyl)indeno[2,1-a]indene (3h). General procedure D was followed using crude diol 8h (21.2 mg, 25.0 µmol), anhydrous p-TsOH (23.1 mg, 134 µmol) and toluene (20 mL). The reaction mixture was stirred at rt for 2.5 h. Column chromatography (silica gel, cyclohexane/CH2Cl2: 8/1 to 0/1) afforded 3h (18.4 mg, 22.4 µmol, 90%) as a red solid. Rf 0.55 (cyclohexane/CH2Cl2: 4/1); m.p. >350 °C; 1H NMR (500 MHz, CD2Cl2): δ = 7.41 (d, J = 7.4 Hz, 2H), 6.96 (s, 4H), 6.87 (d, J = 1.4 Hz, 2H), 6.79 (dd, J = 7.4, 1.4 Hz, 2H), 2.34 (s, 6H), 2.12 (s, 12H); 19F NMR (282 MHz, THF-d8): δ = −(134.6–134.7) (m), −(150.2–150.4) (m), −(160.9–161.1) (m); HRMS (APCI+): m/z calcd. for C50H28F10+ 818.2037 [M]+, found 818.2040; UV/Vis (CH2Cl2): λabs (ε) = 284 (44,813), 468 (15,574), 500 (16,531) nm. Electrochemical measurements: Cyclic voltammograms (CVs) were measured inside a glovebox using a PGSTAT128N by Metrohm Autolab. As working electrode a glassy carbon disc electrode (2 mm diameter) or a platinum disc electrode (2 mm diameter) were used, as counter electrode a platinum rod and as reference electrode a Ag/AgNO3-electrode containing a silver wire immersed in an inner chamber filled with 1 M AgNO3 and 0.1 M n-Bu4NPF6 in anhydrous CH3CN. The analyte solution contained 10 mL of solvent (anhydrous CH2Cl2) with 0.1 M n-Bu4NPF6 and the specified analyte concentration. The ferrocene/ferrocenium redox couple was used as internal reference. HOMO and LUMO levels were calculated using the following equations: ELUMO (eV) = − (Ei,Fc + xRed) (with Ei,Fc = ionization energy of ferrocene = 4.8 eV46; xRed = onset of the first reduction peak, calibrated vs. Fc/Fc+ in eV), EHOMO (eV) = − (Ei,Fc + xOx) (with xOx = onset of the first oxidation peak, calibrated vs. Fc/Fc+ in eV). DFT calculations: TDDFT calculations were performed with the TURBOMOLE v7.0 program package.53 The resolution-of-identity (RI, RIJDX for SP)54,55 approximation for the Coulomb integrals was used in all DFT calculations employing matching auxiliary basis set def2-XVP/J.56 Further, the D3 dispersion correction scheme42,43 with the Becke-Johnson damping function was applied.57–59 The geometries of 3a*–h* were optimized without symmetry restrictions at the PBE60-D3/def2-SVP61 level followed by harmonic
ACS Paragon Plus Environment
22
Page 23 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
vibrational frequency analyses to confirm minima as stationary points. The structures were then reoptimized using TPSS38-D3/def2-TZVP.44 Vertical excitation energies were calculated using TDDFT using the TPSSH38,39 and B3LYP40,41 functionals with the def2-TZVP basis set. For TDDFT calculations geometries were optimized with symmetry restriction within the C2h point group at the functional-D3/def2-TZVP level of theory. Vertical excitation energies were then calculated using TDDFT at the same level of theory.
Associated Content The Supporting Information contains additional synthetic procedures, CVs, NMR spectra, energies and Cartesian coordinates of calculated structures, results of TDDFT calculations, and X-ray crystallographic data.
Acknowledgments Generous support by the German Research Foundation (Emmy Noether grant ES 361/2-1 and grant No. INST 40/467-1 FUGG) and the state of Baden-Württemberg through bwHPC is gratefully acknowledged.
References (1)
Brand, K. Chem. Ber. 1912, 45, 3071–3077.
(2)
Dai, G.; Chang, J.; Luo, J.; Dong, S.; Aratani, N.; Zheng, B.; Huang, K. W.; Yamada, H.; Chi, C. Angew. Chem. Int. Ed. 2016, 55, 2693–2696.
(3)
Yuan, B.; Zhuang, J.; Kirmess, K. M.; Bridgmohan, C. N.; Whalley, A. C.; Wang, L.; Plunkett, K. N. J. Org. Chem. 2016, 81, 8312–8318.
(4)
Kato, S.; Kuwako, S.; Takahashi, N.; Kijima, T.; Nakamura, Y. J. Org. Chem. 2016, 81, 7700–7710.
(5)
Oshima, H.; Fukazawa, A.; Yamaguchi, S. Angew. Chem. Int. Ed. 2017, 56, 3270–3274.
(6)
Chase, D. T.; Fix, A. G.; Rose, B. D.; Weber, C. D.; Nobusue, S.; Stockwell, C. E.; Zakharov, L. N.;
ACS Paragon Plus Environment
23
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 24 of 27
Lonergan, M. C.; Haley, M. M. Angew. Chem. Int. Ed. 2011, 50, 11103–11106. (7)
Chase, D. T.; Rose, B. D.; McClintock, S. P.; Zakharov, L. N.; Haley, M. M. Angew. Chem. Int. Ed. 2011, 50, 1127–1130.
(8)
Cao, J.; London, G.; Dumele, O.; von Wantoch Rekowski, M.; Trapp, N.; Ruhlmann, L.; Boudon, C.; Stanger, A.; Diederich, F. J. Am. Chem. Soc. 2015, 137, 7178–7188.
(9)
Wood, J. D.; Jellison, J. L.; Finke, A. D.; Wang, L.; Plunkett, K. N. J. Am. Chem. Soc. 2012, 134, 15783– 15789.
(10) Maekawa, T.; Ueno, H.; Segawa, Y.; Haley, M. M.; Itami, K. Chem. Sci. 2016, 7, 650–654. (11) Frederickson, C. K.; Zakharov, L. N.; Haley, M. M. J. Am. Chem. Soc. 2016, 138, 16827–16838. (12) Gleiter, R.; Haberhauer, G. Aromaticity and Other Conjugation Effects; Wiley-VCH: Weinheim, Germany, 2012. (13) Hopf, H. Angew. Chem. Int. Ed. 2013, 52, 12224–12226. (14) Hafner, K.; Süss, H. U. Angew. Chem. Int. Ed. Engl. 1973, 12, 575–577. (15) Xie, Q.; Perez-Cordero, E.; Echegoyen, L. J. Am. Chem. Soc. 1992, 114, 3978–3980. (16) Chase, D. T.; Fix, A. G.; Kang, S. J.; Rose, B. D.; Weber, C. D.; Zhong, Y.; Zakharov, L. N.; Lonergan, M. C.; Nuckolls, C.; Haley, M. M. J. Am. Chem. Soc. 2012, 134, 10349–10352. (17) Liu, C.; Xu, S.; Zhu, W.; Zhu, X.; Hu, W.; Li, Z.; Wang, Z. Chem. Eur. J. 2015, 21, 17016–17022. (18) Li, C.; Liu, C.; Li, Y.; Zhu, X.; Wang, Z. Chem. Commun. 2015, 51, 693–696. (19) Kawase, T.; Fujiwara, T.; Kitamura, C.; Konishi, A.; Hirao, Y.; Matsumoto, K.; Kurata, H.; Kubo, T.; Shinamura, S.; Mori, H.; Miyazaki, E.; Takimiya, K. Angew. Chem. Int. Ed. 2010, 49, 7728–7732. (20) Dai, G.; Chang, J.; Zhang, W.; Bai, S.; Huang, K.-W.; Xu, J.; Chi, C. Chem. Commun. 2015, 51, 503– 506. (21) Dai, G.; Chang, J.; Jing, L.; Chi, C. J. Mater. Chem. C 2016, 4, 8758–8764. (22) Gao, X.; Hu, Y. J. Mater. Chem. C 2014, 2, 3099–3117.
ACS Paragon Plus Environment
24
Page 25 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
(23) Gao, X.; Zhao, Z. Sci. China Chem. 2015, 58, 947–968. (24) Wilbuer, J.; Grenz, D. C.; Schnakenburg, G.; Esser, B. Org. Chem. Front. 2017, 4, 658–663. (25) Esser, B. Phys. Chem. Chem. Phys. 2015, 17, 7366–7372. (26) Saito, M. Symmetry 2010, 2, 950–969. (27) Saito, M.; Nakamura, M.; Tajima, T.; Yoshioka, M. Angew. Chem. Int. Ed. 2007, 46, 1504–1507. (28) Zhang, H.; Karasawa, T.; Yamada, H.; Wakamiya, A.; Yamaguchi, S. Org. Lett. 2009, 11, 3076–3079. (29) Chen, C.; Harhausen, M.; Liedtke, R. R.; Bussmann, K.; Fukazawa, A.; Yamaguchi, S.; Petersen, J. L.; Daniliuc, C. G.; Fröhlich, R.; Kehr, G.; Erker, G.; Fr??hlich, R.; Kehr, G.; Erker, G. Angew. Chem. Int. Ed. 2013, 52, 5992–5996. (30) Hashmi, A. S. K.; Wieteck, M.; Braun, I.; Nösel, P.; Jongbloed, L.; Rudolph, M.; Rominger, F. Adv. Synth. Catal. 2012, 354, 555–562. (31) Levi, Z. U.; Tilley, T. D. J. Am. Chem. Soc. 2009, 131, 2796–2797. (32) Kawase, T.; Konishi, A.; Hirao, Y.; Matsumoto, K.; Kurata, H.; Kubo, T. Chem. Eur. J. 2009, 15, 2653– 2661. (33) Maekawa, T.; Segawa, Y.; Itami, K. Chem. Sci. 2013, 4, 2369–2373. (34) Zhao, J.; Oniwa, K.; Asao, N.; Yamamoto, Y.; Jin, T. J. Am. Chem. Soc. 2013, 135, 10222–10225. (35) Imamoto, T.; Takiyama, N.; Nakamura, K.; Hatajima, T.; Kamiya, Y. J. Am. Chem. Soc. 1989, 111, 4392–4398. (36) Xu, F.; Peng, L.; Orita, A.; Otera, J. Org. Lett. 2012, 14, 3970–3973. (37) Orita, A.; Hasegawa, D.; Nakano, T.; Otera, J. Chem. Eur. J. 2002, 8, 2000–2004. (38) Tao, J.; Perdew, J. P.; Staroverov, V. N.; Scuseria, G. E. Phys. Rev. Lett. 2003, 91, 146401. (39) Staroverov, V. N.; Scuseria, G. E.; Tao, J.; Perdew, J. P. J. Chem. Phys. 2003, 119, 12129–12137. (40) Stephens, P. J.; Devlin, F. J.; Chabalowski, C. F.; Frisch, M. J. J. Phys. Chem. 1994, 98, 11623–11627. (41) Becke, A. D. J. Chem. Phys. 1993, 98, 5648–5652.
ACS Paragon Plus Environment
25
The Journal of Organic Chemistry 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
Page 26 of 27
(42) Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. J. Chem. Phys. 2010, 132, 154104. (43) Grimme, S.; Ehrlich, S.; Goerigk, L. J. Comput. Chem. 2011, 32, 1456–1465. (44) Schäfer, A.; Huber, C.; Ahlrichs, R. J. Chem. Phys. 1994, 100, 5829–5835. (45) Large peak separations might be due to a temperature effect (CVs were measured in the glove box at 25-30 oC) or the specific cell setup. (46) D’Andrade, B. W.; Datta, S.; Forrest, S. R.; Djurovich, P.; Polikarpov, E.; Thompson, M. E. Org. Electron. 2005, 6, 11–20. (47) Hansch, C.; Leo, A.; Taft, R. W. Chem. Rev. 1991, 91, 165–195. (48) Wermuth, C. G.; Ganellin, C. R.; Lindberg, P.; Mitscher, L. A. Pure Appl. Chem. 1979, 51, 1129–1143. (49) Rose, B. D.; Shoer, L. E.; Wasielewski, M. R.; Haley, M. M. Chem. Phys. Lett. 2014, 616–617, 137– 141. (50) CCDC 1550378 contains the supplementary crystallographic data for this paper. These data can be obtained free of charge from The Cambridge Crystallographic Data Centre via www.ccdc.cam.ac.uk/structures. (51) Fulmer, G. R.; Miller, A. J. M.; Sherden, N. H.; Gottlieb, H. E.; Nudelman, A.; Stoltz, B. M.; Bercaw, J. E.; Goldberg, K. I. Organometallics 2010, 29, 2176–2179. (52) Harris, R. K.; Becker, E. D.; Cabral De Menezes, S. M.; Goodfellow, R.; Granger, P. Concepts Magn. Reson. Part A Bridg. Educ. Res. 2002, 14, 326–346. (53) TURBOMOLE V7.0 2015, a development of the University of Karlsruhe and Forschungszentrum Karlsruhe
GmbH,
1989-2007,
TURBOMOLE
GmbH,
since
2007;
available
from
http://www.turbomole.com. (54) Eichkorn, K.; Treutler, O.; Öhm, H.; Häser, M.; Ahlrichs, R. Chem. Phys. Lett. 1995, 240, 283–290. (55) Eichkorn, K.; Treutler, O.; Öhm, H.; Marco, H.; Ahlrichs, R. Chem. Phys. Lett. 1995, 242, 652–660. (56) Weigend, F. Phys. Chem. Chem. Phys. 2006, 8, 1057–1065.
ACS Paragon Plus Environment
26
Page 27 of 27 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60
The Journal of Organic Chemistry
(57) Becke, A. D.; Johnson, E. R. J. Chem. Phys. 2005, 123, 154101. (58) Johnson, E. R.; Becke, A. D. J. Chem. Phys. 2005, 123, 24101. (59) Johnson, E. R.; Becke, A. D. J. Chem. Phys. 2006, 124, 174104. (60) Perdew, J. P.; Burke, K.; Ernzerhof, M. Phys. Rev. Lett. 1996, 77, 3865–3868. (61) Schäfer, A.; Horn, H.; Ahlrichs, R. J. Chem. Phys. 1992, 97, 2571–2577.
ACS Paragon Plus Environment
27