Synthesis of Dibenzofurans from Cyclic Diaryliodonium Triflates and

Publication Date (Web): October 10, 2018. Copyright © 2018 American Chemical Society. *E-mail: [email protected]. ACS AuthorChoice - This is an ...
4 downloads 0 Views 836KB Size
This is an open access article published under an ACS AuthorChoice License, which permits copying and redistribution of the article or any adaptations for non-commercial purposes.

Article Cite This: ACS Omega 2018, 3, 12923−12929

http://pubs.acs.org/journal/acsodf

Synthesis of Dibenzofurans from Cyclic Diaryliodonium Triflates and Water via Oxygen−Iodine Exchange Approach Jian Li,*,† Qiu-Neng Xu,† Zheng-Bing Wang,† Yang Li,† and Li Liu‡ †

School of Pharmaceutical Engineering & Life Sciences and ‡School of Petrochemical Engineering, Changzhou University, Changzhou 213164, China

ACS Omega 2018.3:12923-12929. Downloaded from pubs.acs.org by 79.110.18.85 on 10/10/18. For personal use only.

S Supporting Information *

ABSTRACT: An efficient synthesis of a variety of dibenzofuran derivatives via Cu-catalyzed cyclization diaryliodonium salts in water is achieved. Various dibenzofuran derivatives could be obtained in good to excellent yields via this oxygen− iodine exchange approach. A concise synthesis of organic semiconducting material molecule has been achieved using this method.



INTRODUCTION Dibenzofurans can be found in a wide range of natural products; they are used as pharmaceutical candidates1 because of their unique properties, such as anticancer, antibacterial, antiallergy, antimalarial, and anti-inflammatory activities (Scheme 1).2 In addition, fused dibenzofuran skeletons are considered to be potential photoelectronic materials in blue phosphorescent organic light-emitting diodes.3 As a result, many research groups devoted themselves to developing new approaches to construct dibenzofuran motifs,4 which can be divided into intermolecular and intramolecular categories: (A) intramolecular synthesis strategy: cyclization of diaryl ether derivatives,5 formation of C−O bonds using biaryls;6 (B) reaction of ortho-iodophenols with silylaryl triflates in the presence of CsF to afford the O-arylated products, which are subsequently cyclized using a Pd catalyst to dibenzofurans;7 reaction between 6-diazo-2-cyclohexenone and an orthohaloiodobenzene through tandem Pd-catalyzed cross-coupling/aromatization followed by a Cu-catalyzed Ullmann coupling reaction.8 However, these strategies suffer from hash reaction conditions, tedious steps, limited substrate scope, and use of molar excess of catalyst. Recently, cyclic diaryliodoniums have emerged as a new paradigm to synthesis of dibenzothiophenes,9 dibenzopyrroles,10 and dibenzoselenophens,11 which avoids poor atom economy by generating 1 equiv of an iodoarene as waste. Considering our continuous interest in developing application of hypervalent iodine reagent in organic synthesis, we were focused mainly on construction of heterocyclic to exploit diverse diaryliodonium salt transformations.12 Accordingly, intramolecular cyclizations of dibenziodolium triflates were established by our group for the synthesis of dibenzothiophenes13 and fluorenones.14



formation. However, C−O bond formation has received much less attention, partly owing to the large energy gap between M−O HOMO and M−C LUMO frontier orbitals.15 From former investigation and proposal, diaryliodonium salts could be activated by Cu(I) species through oxidative addition.16 On the basis of the above work, we initiated our study by investigating the reaction of cyclic diaryliodonium 1a, a protocol based on CuI; K2CO3 in a mixture solvent (toluene/H2O = 9:1) at 100 °C for 4 h provided the desired compound 2a in 13% yield (Table 1, entry 1). Encouraged by this promising result, catalysts were first investigated. Although various copper catalysts were tested, such as CuCl, CuBr, Cu(OAc)2, Cu(OTf)2, Cu2O, and Cu(MeCN)4PF6, the yield could not be increased. Generally, the ligand backbone has a critical impact on the reactivity and selectivity of transitionmetal catalytic reaction; therefore, ligand examination was carried out. Gratifyingly, the yield was remarkably improved when ligands were used; the reaction with L4 as ligand provided dibenzofuran 2a in 83% yield (Table 1, entry 11). Ligand screening indicated that generally monodentate ligand PPh3 was not efficient and the bidentate nitrogen ligands were found well adapted to this reaction, such as L2, L3, L4, and L5. Next, the impact of base on the reaction was studied (entries 14−18); other inorganic bases, such as Cs2CO3, t-BuOK, NaOH, Na2CO3, led to lower yields; improved yield was not observed when organic base Et3N was employed. A remarkable solvent effect was then observed during solvent investigation. To our delight, the product yield could be further improved to 92% in MeOH, whereas no reaction occurred in 1,4-dioxane and dimethylformamide (DMF). H2O as the source of oxygen is essential for this reaction; we proceed the reaction in pure water in the presence of 5 mol % CuI, 10 mol % of L4, and 2 equiv of K2CO3 at 100 °C to yield the desired product 2a in

RESULTS AND DISCUSSION

Received: September 11, 2018 Accepted: September 25, 2018 Published: October 10, 2018

Inspired by the pioneering achievements, we postulated that the dibenzofurans can be prepared via double C−O bond © 2018 American Chemical Society

12923

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

Scheme 1. Selected Active 3-(2-Oxopropyl)-2-arylisoindolinone and the Synthetic Strategies

high charge carrier mobility and high stability, could be obtained in 72% yield from the double diaryliodonium salts 1v through the oxygen−iodine exchange reaction. Interestingly, when cyclic diaryliodonium salts 1w that contain thiophene rings were treated under the standard conditions, corresponding sulfone product 2w could be afforded in 50% yield, which might be a promising candidate for organic field-effect transistor materials (Scheme 2). Next, we decided to elucidate the reaction mechanism of the formation of dibenzofuran, from the reaction we know wherein the key was the source of oxygen atoms. Accordingly, some control experiments were designed under the standard conditions as shown in Figure 1. Treatment of cyclic diaryliodonium 1a with 5 mol % CuI and 10 mol % L4 in dry MeOH at 100 °C failed to deliver possible dibenzofuran 2a, suggesting that oxygen atoms might not be derived from air. To further explore the source of oxygen atoms, the isotope labeled water H218O was employed in the reaction and 18O labeled product 2a′ was isolated in 81% yield. These results indicated that the oxygen atom was derived from H2O. On the basis of the control experiment results and relevant literature, we proposed the mechanism of this cyclization reaction. A reasonable mechanism suggested that an oxygen− iodine exchange process is involved, as shown in Figure 2. Initially, the oxidative addition of cyclic diaryliodonium salts 1 with CuI and ligand L afforded intermediate A, containing intramolecular iodobenzene. Subsequently, ligand exchange of A with H2O formed hydroiodic acid and complex B. Next, reductive elimination of intermediate B took place to get 2-[2′iodophenyl]phenol C and Cu(I) intermediate; ligand L was released out simultaneously. Then, oxidative addition of 2-[2′iodophenyl]phenol C to Cu(I) would provide a new electrophilic Cu(III) intermediate, which can be attacked by hydroxyl under basic conditions to form intermediate D. Finally, reductive elimination of D gave the desired product 2 and regenerated the catalyst.

only 28% yield at 4 h. However, the results were significantly improved when the reaction time prolonged to 12 and 24 h, as 2a was obtained in up to 96% yield. On the day of submission, we learned Huang and Wen’s group demonstrated a similar idea on the preparation of dibenzofurans.17 Having established the optimal conditions for the dibenzofuran synthesis, we intended to determine its scope and limitation. A series of reactions of cyclic diaryliodonium salts 1 and H2O were carried out under the standard conditions. All reactions of cyclic diaryliodonium salts 1 bearing varied R1 and R2 groups proceeded smoothly to afford the corresponding dibenzofuran 2 in good to excellent yields (Table 2). The transformation was sensitive to the electronic nature of the diaryliodonium salt substituent; generally, reactions of substrates containing electron-donating substituents in 2 position led to higher yields than the reactions of those bearing electron-withdrawing groups (2b−d vs 2e, 2f). The yields were found to be a little lower for the substrates bearing halogen groups at the meta position on the cyclic diaryliodonium salts (2g−i); however, the C4-substituted halogen salts were more favorable than those at meta position (2j, 2k vs 2g, 2h). Subsequently, the C2-substituted groups were also investigated (2l), and the yield was influenced slightly unfavorably by steric effects. Notably, disubstituted diaryliodonium in different positions could be applied in this methodology and afforded the desired products in modest-togood yields (2n−p). Likewise, disubstitution on different aromatic rings was also tested in the reaction and relatively normal yields were obtained (2q−s). In addition, when fused cyclic diaryliodonium salts were employed, desired polycyclic products were detected in modest-to-good yields. Prompted by these results, we sought to design an efficient synthetic route to various ladder-type heteroacenes18 that contain furan rings in the fused-ring framework. Benzo[1,2b:4,5-b′]bisbenzofuran,19 an organic semiconducting material molecule, exhibiting an outstanding electron ability with both 12924

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

Table 1. Optimization of the Reaction Conditionsa,b

entry

catalyst

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 26c 27d 28e

CuI CuCl CuBr Cu(OAc)2·H2O Cu(OTf)2 Cu2O Cu(MeCN)4PF6 CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI CuI

ligand

base

solvent

yield (%)

L1 L2 L3 L4 L5 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4 L4

K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 Cs2CO3 t-BuOK NaOH Na2CO3 Et3N K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3 K2CO3

toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene toluene DCE THF 1,4-dioxane MeCN ether DMF MeOH DMSO H2O H2O H2O

13 1 3 4 7 10 12 5 32 50 83 7 67 27 39 77 78 8 24 0 37 75 0 92 4 28 66 96

spectrometer in CDCl3. High resolution mass spectrometry (HRMS) was performed on an Agilent 6540 Q-TOF mass spectrometer (ESI). Melting points were determined on a SGW X-4B melting point apparatus. Cyclic diaryliodoniums are prepared according to the literature.8 General Procedure for Preparation of Compound 2. In the pressure tube, a solution of diaryliodonium salts 1 (0.1 mmol), K2CO3 (0.2 mmol), CuI (0.05 mmol), and ligand L4 (0.1 equiv) in H2O (2.0 mL) was stirred at 100 °C for 24 h. After completion of the reaction (observed on thin-layer chromatography), the solvent was evaporated under reduced pressure to obtain the crude mixture. The residues were purified by silica-gel column chromatography (ethyl acetate/ petroleum ether = 1/50:1/10) to afford the pure product 2. The obtained product was analyzed by 1H NMR, 13C NMR, and 19F NMR. Dibenzo[b,d]furan (2a). White solid, 96% yield. 1H NMR (400 MHz, CDCl3) δ 7.98 (d, J = 7.7 Hz, 2H), 7.59 (d, J = 8.1 Hz, 2H), 7.48 (t, J = 7.6 Hz, 2H), 7.37 (t, J = 7.5 Hz, 2H). See: J. Am. Chem. Soc. 2009, 131, 4194−4195. 3-Methyldibenzo[b,d]furan (2b). White solid, 92% yield. 1 H NMR (400 MHz, CDCl3) δ 7.95−7.91 (m, 1H), 7.84 (d, J = 7.9 Hz, 1H), 7.57 (d, J = 8.2 Hz, 1H), 7.44 (dt, J = 7.4, 1.3 Hz, 1H), 7.34 (dt, J = 7.6, 0.9 Hz, 1H), 7.19 (d, J = 7.8 Hz, 1H), 2.55 (s, 3H). See: J. Am. Chem. Soc. 2009, 131, 4194− 4195. 3-Methoxydibenzo[b,d]furan (2c). White solid, 87% yield. 1 H NMR (400 MHz, CDCl3) δ 7.88−7.86 (m, 1H), 7.83 (d, J = 8.5 Hz, 1H), 7.54 (d, J = 8.0 Hz, 1H), 7.37 (dt, J = 8.0, 1.3 Hz, 1H), 7.34 (dt, J = 7.5, 0.9 Hz, 1H), 7.11 (d, J = 2.2 Hz, 1H), 6.96 (dd, J = 8.5, 2.3 Hz, 1H), 3.92 (s, 3H). See: J. Am. Chem. Soc. 2011, 133, 9250−9253. 3-Ethyldibenzo[b,d]furan (2d). White solid, 89% yield. Mp: 82−84 °C. 1H NMR (400 MHz, CDCl3) δ 7.93 (d, J = 7.6 Hz, 1H), 7.87 (d, J = 7.8 Hz, 1H), 7.57 (d, J = 8.2 Hz, 1H), 7.46− 7.43 (m, 2H), 7.34 (t, J = 7.6 Hz, 1H), 7.21 (d, J = 8.0 Hz, 1H), 2.84 (q, J = 7.6 Hz, 2H), 1.35 (t, J = 7.6 Hz, 3H). 13C NMR (125 MHz, CDCl3) δ 156.7, 156.3, 144.3, 126.5, 124.4, 122.9, 122.6, 121.8, 120.3, 120.2, 111.6, 110.7, 29.3, 15.9. HRMS (ESI) calcd for C14H13O ([M + H]+): 197.0961; found 197.0958. Ethyl Dibenzo[b,d]furan-3-carboxylate (2e). White solid, 68% yield. 1H NMR (400 MHz, CDCl3) δ 8.27 (s, 1H), 8.10− 8.18 (m, 1H), 8.03−7.99 (m, 2H), 7.63 (d, J = 8.2 Hz, 1H), 7.55 (t, J = 7.5 Hz, 1H), 7.40 (t, J = 7.5 Hz, 1H), 4.46 (q, J = 7.1 Hz, 2H), 1.47 (t, J = 7.1 Hz, 3H). See: Org. Lett. 2011, 13, 5504−5507. 3-(Trifluoromethyl)dibenzo[b,d]furan (2f). White solid, 59% yield. 1H NMR (400 MHz, CDCl3) δ 8.07 (d, J = 8.1 Hz, 1H), 8.03 (d, J = 7.8 Hz, 1H), 7.86 (s, 1H), 7.65−7.63 (m, 2H), 7.56 (t, J = 8.0 Hz, 1H), 7.43 (t, J = 7.4 Hz, 1H). See: J. Am. Chem. Soc. 2011, 133, 9250−9253. 3-Fluorodibenzo[b,d]furan (2g). White solid, 70% yield. 1H NMR (400 MHz, CDCl3) δ 7.93−7.87 (m, 2H), 7.58 (d, J = 8.2 Hz, 1H), 7.45 (dt, J = 7.4, 1.3 Hz, 1H), 7.37 (dt, J = 7.6, 0.9 Hz, 1H), 7.31−7.28 (m, 1H), 7.11 (dt, J = 9.2, 2.3 Hz, 1H). See: J. Am. Chem. Soc. 2009, 131, 4194−4195. 3-Chlorodibenzo[b,d]furan (2h). White solid, 69% yield. 1 H NMR (400 MHz, CDCl3) δ 7.93 (d, J = 7.7 Hz, 1H), 7.87 (d, J = 8.2 Hz, 1H), 7.59−7.57 (m, 2H), 7.49 (t, J = 7.2 Hz, 1H), 7.39−7.33 (m, 2H). See: J. Am. Chem. Soc. 2011, 133, 9250−9253.

a

Reaction conditions: 1 (1.0 equiv), catalyst (0.05 equiv), ligand (0.1 equiv), base (2 equiv) in {solvents: H2O [9:1 (v/v), 2 mL]}, 100 °C, 4 h. bGas chromatography yields with n-decane as an internal standard. cReaction conditions: 1 (1.0 equiv), catalyst (0.05 equiv), ligand (0.1 equiv), base (2 equiv) in 2 mL of H2O, 100 °C. d12 h. e24 h.



CONCLUSIONS In summary, a Cu-catalyzed cyclization of cyclic diaryliodonium salts in water has been developed via double C− O bond formation. The developed oxygen−iodine exchange methodology provides straightforward access to dibenzofuran derivatives in up to 96% yield. Applying this strategy, a concise synthesis of organic semiconducting material molecule has been accomplished in moderate yields. Further studies to explore the diaryliodonium salts are currently underway in our laboratory.



EXPERIMENTAL SECTION General Information. All of the reactions was carried out under an air atmosphere condition. Various reagents were purchased from Aldrich, Acros, or Alfa. For column chromatography, 200−300 mesh silica gel was used. 1H, 13C and 19F NMR were recorded on Bruker 400 or 500 MHz 12925

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

Table 2. Cyclic Diaryliodonium Salts Scopea,b

Reaction condition: 1 (0.1 mmol), CuI (0.05 equiv), L4 (0.1 equiv), K2CO3 (2.0 equiv), H2O (2.0 mL), 100 °C, 24 h. bIsolated yield.

a

Scheme 2. Syntheses of Ladder-Type Heteroacenes

Figure 1. Control experiments. Figure 2. Proposed reaction mechanism.

3-Bromodibenzo[b,d]furan (2i). White solid, 75% yield. 1H NMR (400 MHz, CDCl3) δ 7.94 (d, J = 7.6 Hz, 1H), 7.83 (d, J = 8.2 Hz, 1H), 7.76 (d, J = 1.5 Hz, 1H), 7.59−7.57 (m, 1H), 12926

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

MHz, CDCl3) δ −119.7. HRMS (ESI) calcd for C12H7ClFO ([M + H]+): 221.0164 found 221.0160. Naphtho[2,3-b]benzofuran (2t). White solid, 81% yield. 1H NMR (400 MHz, CDCl3) δ 8.48 (d, J = 8.1 Hz, 1H), 8.04− 8.01 (m, 3H), 7.81 (d, J = 8.5 Hz, 1H), 7.74 (d, J = 8.2 Hz, 1H), 7.67 (t, J = 7.1 Hz, 1H), 7.59 (t, J = 7.1 Hz, 1H), 7.51 (t, J = 7.2 Hz, 1H), 7.43 (t, J = 7.3 Hz, 1H). See: Tetrahedron 2005, 61, 1353−1362. Phenanthro[9,10-b]benzofuran (2u). White solid, 73% yield. 1H NMR (400 MHz, CDCl3) δ 8.83−8.76 (m, 2H), 8.68 (d, J = 7.7 Hz, 1H), 8.56−8.52 (m, 1H), 8.43−8.40 (m, 1H), 7.82−7.84 (m, 4H), 7.72−7.70 (m, 1H), 7.56−7.50 (m, 2H). See: J. Org. Chem. 2017, 82, 9133−9143. Benzo[1,2-b:4,5-b′]bisbenzofuran (2v). White solid, 67% yield. 1H NMR (400 MHz, CDCl3) δ 8.37 (d, J = 7.4 Hz, 1H), 8.04−8.01 (m, 2H), 7.74 (d, J = 8.1 Hz, 1H), 7.67 (d, J = 8.0 Hz, 1H), 7.62 (d, J = 8.4 Hz, 1H), 7.54−7.50 (m, 3H), 7.42 (t, J = 7.4 Hz, 1H). See: J. Am. Chem. Soc. 2011, 133, 9250−9253. Benzo[b]benzo[4,5]thieno[3,2-g]benzofuran 12,12-Dioxide (2w). White solid, 50% yield. Mp: 172−174 °C. 1H NMR (400 MHz, CDCl3) δ 8.47 (d, J = 7.5 Hz, 1H), 7.91 (d, J = 7.5 Hz, 1H), 7.86−7.82 (m, 2H), 7.81−7.79 (m, 1H), 7.69− 7.62 (m, 3H), 7.57−7.49 (m, 2H). 13C NMR (125 MHz, CDCl3) δ 157.4, 137.7, 134.1, 132.5, 130.2, 129.7, 129.6, 126.8, 124.5, 124.2, 122.4, 121.3, 120.6, 120.4, 119.4, 116.6, 111.9. HRMS (ESI) calcd for C18H11O3S ([M + H]+): 307.0423 found 307.0418.

7.50 (dt, J = 7.4, 1.1 Hz, 2H), 7.38 (t, J = 7.6 Hz, 1H). See: Org. Lett. 2011, 13, 5504−5507. 2-Fluorodibenzo[b,d]furan (2j). White solid, 80% yield. 1H NMR (400 MHz, CDCl3) δ 7.93 (d, J = 7.4 Hz, 1H), 7.63 (dd, J = 8.1, 2.6 Hz, 1H), 7.59−7.57 (m, 1H), 7.53−7.50 (m, 2H), 7.37 (t, J = 7.8 Hz, 1H), 7.19 (dt, J = 9.0, 2.7 Hz, 1H). See: J. Am. Chem. Soc. 2009, 131, 4194−4195. 2-Chlorodibenzo[b,d]furan (2k). White solid, 73% yield. 1H NMR (400 MHz, CDCl3) δ 8.00−7.89 (m, 2H), 7.59 (d, J = 8.3 Hz, 1H), 7.51 (dd, J = 7.9, 4.9 Hz, 2H), 7.46−7.34 (m, 2H). See: Org. Lett. 2011, 13, 5504−5507. 4-Methyldibenzo[b,d]furan (2l). White solid, 85% yield. 1H NMR (400 MHz, CDCl3) δ 7.94 (d, J = 7.6 Hz, 1H), 7.77 (s, 1H), 7.57 (d, J = 8.2 Hz, 1H), 7.46 (t, J = 8.6 Hz, 2H), 7.34 (t, J = 7.5 Hz, 1H), 7.29−7.28 (m, 1H), 2.54 (s, 3H). See: Org. Lett. 2012, 14, 4838−4841. 4-Propyldibenzo[b,d]furan (2m). Colorless oil, 84% yield. 1 H NMR (400 MHz, CDCl3) δ 7.94 (d, J = 7.6 Hz, 1H), 7.86 (d, J = 7.8 Hz, 1H), 7.57 (d, J = 8.2 Hz, 1H), 7.46−7.41 (m, 2H), 7.34 (t, J = 7.5 Hz, 1H), 7.19 (d, J = 7.8 Hz, 1H), 2.78 (t, J = 7.4 Hz, 2H), 1.80−1.72 (m, 2H), 1.00 (t, J = 7.3 Hz, 3H). 13 C NMR (125 MHz, CDCl3) δ 156.6, 156.2, 142.7, 126.6, 124.4, 123.5, 122.6, 121.9, 120.4, 120.2, 111.6, 111.4, 38.5, 24.9, 13.9. HRMS (ESI) calcd for C15H15O ([M + H]+): 211.1117; found 211.1112. 2,4-Dimethyldibenzo[b,d]furan (2n). White solid, 76% yield. 1H NMR (400 MHz, CDCl3) δ 7.94 (d, J = 7.6 Hz, 1H), 7.62−7.60 (m, 2H), 7.49−7.45 (m, 1H), 7.36 (t, J = 7.4 Hz, 1H), 7.12 (s, 1H), 2.61 (s, 3H), 2.52 (s, 3H). See: Org. Lett. 2013, 15, 2754−2757. 3,4-Dimethoxydibenzo[b,d]furan (2o). White solid, 82% yield. 1H NMR (400 MHz, CDCl3) δ 7.85 (d, J = 7.5 Hz, 1H), 7.54 (d, J = 8.1 Hz, 1H), 7.38 (t, J = 6.3 Hz, 2H), 7.34−7.30 (m, 1H), 7.14 (s, 1H), 4.01 (s, 3H), 3.99 (s, 3H). See: J. Org. Chem. 2006, 71, 9104−9113. Benzo[b][1,3]dioxolo[4,5-f ]benzofuran (2p). White solid, 69% yield. 1H NMR (400 MHz, CDCl3) δ 7.82 (d, J = 7.4 Hz, 1H), 7.52 (d, J = 8.1 Hz, 1H), 7.39−7.35 (m, 1H), 7.33−7.29 (m, 2H), 7.08 (s, 1H), 6.07 (s, 2H). See: Aust. J. Chem. 2013, 66, 1334−1341. 2-Fluoro-7-methyldibenzo[b,d]furan (2q). White solid, 80% yield. Mp: 75−76 °C. 1H NMR (400 MHz, CDCl3) δ 7.79 (d, J = 7.9 Hz, 1H), 7.57 (dd, J = 8.2, 2.6 Hz, 1H), 7.48 (dd, J = 8.9, 4.1 Hz, 1H), 7.38 (s, 1H), 7.21−7.11 (m, 2H), 2.55 (s, 3H). 13C NMR (125 MHz, CDCl3) δ 158.9 (d, J = 237.4 Hz), 157.6, 152.2, 138.5, 125.3 (d, J = 10.2 Hz), 124.1, 121.4, 120.4, 113.7 (d, J = 25.7 Hz), 112.2, 106.4 (d, J = 25.1 Hz), 22.0. 19F NMR (376 MHz, CDCl3) δ −120.8. HRMS (ESI) calcd for C13H10FO ([M + H]+): 201.0710 found 201.0705. 2-Fluoro-7-methoxydibenzo[b,d]furan (2r). White solid, 85% yield. 1H NMR (400 MHz, CDCl3) δ 7.98 (d, J = 8.6 Hz, 1H), 7.52 (dd, J = 8.3, 2.6 Hz, 1H), 7.46 (dd, J = 8.9, 4.1 Hz, 1H), 7.08 (dt, J = 11.4, 2.7 Hz, 2H), 6.95 (dd, J = 8.6, 2.2 Hz, 1H), 3.92 (s, 3H). See: J. Org. Chem. 2018, 83, 805−811. 7-Chloro-2-fluorodibenzo[b,d]furan (2s). White solid, 55% yield. Mp: 77−79 °C. 1H NMR (400 MHz, CDCl3) δ 7.83 (d, J = 8.3 Hz, 1H), 7.60−7.57 (m, 2H), 7.52 (dd, J = 8.9, 4.0 Hz, 1H), 7.35 (dd, J = 8.3, 1.7 Hz, 1H), 7.20 (dt, J = 8.9, 1.9 Hz, 1H). 13C NMR (125 MHz, CDCl3) δ 159.3 (d, J = 238.6 Hz), 152.6, 133.5, 124.5 (d, J = 10.3 Hz), 123.6, 122.8, 121.5, 114.9 (d, J = 25.8 Hz), 112.6, 106.8 (d, J = 25.1 Hz). 19F NMR (376



ASSOCIATED CONTENT

S Supporting Information *

The Supporting Information is available free of charge on the ACS Publications website at DOI: 10.1021/acsomega.8b02345. Spectral data for all new compounds (PDF)



AUTHOR INFORMATION

Corresponding Author

*E-mail: [email protected]. ORCID

Jian Li: 0000-0001-6713-1302 Notes

The authors declare no competing financial interest.



ACKNOWLEDGMENTS We are grateful to National Natural Science Foundation of China (21402013) and the Natural Science Foundation of Jiangsu Province (BK20140259), Advanced Catalysis and Green Manufacturing Collaborative Innovation Center in Changzhou University.



REFERENCES

(1) (a) Kokubun, T.; Harborne, J. B.; Eagles, J.; Waterman, P. G. Dibenzofuran phytoalexins from the sapwood tissue of Photinia, Pyracantha and Crataegus species. Phytochemistry 1995, 39, 1033− 1037. (b) Collin, G.; Höke, H. Benzofurans. In Ullmann’s Encyclopedia of Industrial Chemistry; Wiley-VCH Verlag GmbH & Co.: Dordrecht, The Netherlands, 2000; Vol. 5, pp 325−338. (2) (a) Martel, R. R.; Rochefort, J. G.; Klicius, J.; Dobson, T. A. Anti-inflammatory properties of furobufen. Can. J. Physiol. Pharmacol. 1974, 52, 669−673. (b) Carney, J. R.; Krenisky, J. M.; Williamson, R. T.; Luo, J. Achyrofuran, a new antihyperglycemic dibenzofuran from the south american medicinal plant Achyrocline satureioides. J. Nat.

12927

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

Prod. 2002, 65, 203−205. (c) Kaul, R.; Deechongkit, S.; Kelly, J. W. Synthesis of a negatively charged dibenzofuran-based B-turn mimetic and its incorporation into the WW miniprotein-enhanced solubility without a loss of thermodynamic stability. J. Am. Chem. Soc. 2002, 124, 11900−11907. (d) Petrassi, H. M.; Johnson, S. M.; Purkey, H. E.; Chiang, K. P.; Walkup, T.; Jiang, X.; Powers, E. T.; Kelly, J. W. Potent and selective structure-based dibenzofuran inhibitors of transthyretin amyloidogenesis: kinetic stabilization of the Native State. J. Am. Chem. Soc. 2005, 127, 6662−6671. (e) Momotake, A.; Lindegger, N.; Niggli, E.; Barsotti, R. J.; Ellis-Davies, G. C. R. The nitrodibenzofuran chromophore: a new caging group for ultraefficient photolysis in living cells. Nat. Methods 2006, 3, 35−40. (f) Lin, R.-D.; Chen, M.-C.; Liu, Y.-L.; Lin, Y.-T.; Lu, M.-K.; Hsu, FL.; Lee, M.-H. New whitening constituents from Taiwan-Native Pyracantha koidzumii: structures and tyrosinase inhibitory analysis in human epidermal melanocytes. Int. J. Mol. Sci. 2015, 16, 28598− 28613. (g) Love, B. E. Isolation and synthesis of polyoxygenated dibenzofurans possessing biological activity. Eur. J. Med. Chem. 2015, 97, 377−387. (3) (a) Deng, L.; Li, J.; Wang, G.-X.; Wu, L.-Z. Simple bipolar host materials incorporating CN group for highly efficient blue electrophosphorescence with slow efficiency roll-off. J. Mater. Chem. C 2013, 1, 8140−8145. (b) Zhang, L.; Zhang, Y.-X.; Hu, Y.; Shi, X.-B.; Jiang, Z.-Q.; Wang, Z.-K.; Liao, L.-S. Highly efficient blue phosphorescent organic light-Emitting diodes employing a host material with small bandgap. ACS Appl. Mater. Interfaces 2016, 8, 16186−16191. (4) (a) Ames, D. E.; Opalko, A. Synthesis of dibenzofurans by palladium-catalyzed intramolecular dehydrobromination of 2-bromophenyl phenyl ethers. Synthesis 1983, 234−235. (b) Liu, Z.; Larock, R. C. Synthesis of carbazoles and dibenzofurans via cross-coupling of oiodoanilines and o-iodophenols with silylaryl triflates and subsequent Pd-catalyzed cyclization. Tetrahedron 2007, 63, 347−355. (c) Lockner, J. W.; Dixon, D. D.; Risgaard, R.; Baran, P. S. Practical Radical Cyclizations with Arylboronic Acids and Trifluoroborates. Org. Lett. 2011, 13, 5628−5631. (d) Nervig, C. S.; Waller, P. J.; Kalyani, D. Palladium-catalyzed intramolecular C-H arylation of arenes using tosylates and mesylates as electrophiles. Org. Lett. 2012, 14, 4838− 4841. (e) Maetani, S.; Fukuyama, T.; Ryu, I. Rhodium-catalyzed decarbonylative C-H arylation of 2-aryloxybenzoic acids leading to dibenzofuran derivatives. Org. Lett. 2013, 15, 2754−2757. (f) Campeau, L.-C.; Parisien, M.; Jean, A.; Fagnou, K. Catalytic direct arylation with aryl chlorides, bromides, and iodides: intramolecular studies leading to new intermolecular reactions. J. Am. Chem. Soc. 2006, 128, 581−590. (g) Wang, C.; Piel, I.; Glorius, F. Palladiumcatalyzed intramolecular direct arylation of benzoic acids by tandem decarboxylation/C−H activation. J. Am. Chem. Soc. 2009, 131, 4194− 4195. (h) Shen, Z.; Ni, Z.; Mo, S.; Wang, J.; Zhu, Y. Palladiumcatalyzed intramolecular decarboxylative coupling of arene carboxylic acids/esters with aryl bromides. Chem. - Eur. J. 2012, 18, 4859−4865. (i) Ishida, T.; Tsunodaa, R.; Zhanga, Z.; Hamasaki, A.; Honma, T.; Ohashi, H.; Yokoyama, T.; Tokunaga, M. Supported palladium hydroxide-catalyzed intramolecular double CH bond functionalization for synthesis of carbazoles and dibenzofurans. Appl. Catal., B 2014, 150−151, 523−531. (j) Niu, L.; Yang, H.; Jiang, Y.; Fu, H. Efficient synthesis of dibenzoxaborininols from diaryl ethers and their application to dibenzofuran synthesis. Adv. Synth. Catal. 2013, 355, 3625−3632. (k) Ames, D. E.; Opalko, A. Synthesis of dibenzofurans by palladium-catalyzed intramolecular dehydrobromination of 2bromophenyl phenyl ethers. Synthesis 1983, 234−235. (l) Campeau, L.-C.; Thansandote, P.; Fagnou, K. High-yielding intramolecular direct arylation reactions with aryl chlorides. Org. Lett. 2005, 7, 1857− 1860. (m) Parisien, M.; Valette, D.; Fagnou, K. Direct arylation reactions catalyzed by Pd(OH)2/C: evidence for a soluble palladium catalyst. J. Org. Chem. 2005, 70, 7578−7584. (n) Ferguson, D. M.; Rudolph, S. R.; Kalyani, D. Palladium-Catalyzed Intra- and Intermolecular C-H Arylation Using Mesylates: Synthetic Scope and Mechanistic Studies. ACS Catal. 2014, 4, 2395−2401. (o) Du, Z.; Zhou, J.; Si, C.; Ma, W. Synthesis of dibenzofurans by palladium-

catalysed tandem denitrification/C-H activation. Synlett 2011, 3023− 3025. (5) (a) Maevskii, Y. V.; Migachev, G. I. Synthesis of dibenzofuran and its nitro-substituted derivatives. Chem. Heterocycl. Compd. 1986, 22, 676. (b) Liu, J.; Fitzgerald, A. E.; Mani, N. S. Facile Assembly of Fused Benzo[4,5]furo Heterocycles. J. Org. Chem. 2008, 73, 2951− 2954. (c) Panda, N.; Mattan, I.; Nayak, D. K. Synthesis of dibenzofurans via C−H activation of o-Iodo diaryl ethers. J. Org. Chem. 2015, 80, 6590−6597. (d) Wei, Y.; Yoshikai, N. Oxidative cyclization of 2-arylphenols to dibenzofurans under Pd(II)/peroxybenzoate catalysis. Org. Lett. 2011, 13, 5504−5507. (e) Zhao, J.; Zhang, Q.; Liu, L.; He, Y.; Li, J.; Li, J.; Zhu, Q. CuI-Mediated sequential iodination/cycloetherification of o-Arylphenols: synthesis of 2- or 4-iododibenzofurans and mechanistic Studies. Org. Lett. 2012, 14, 5362−5365. (f) Zhao, J.; Wang, Y.; He, Y.; Liu, Y.; Zhu, Q. Cucatalyzed oxidative C(sp2)−H cycloetherification of o-arylphenols for the preparation of dibenzofurans. Org. Lett. 2012, 14, 1078−1081. (g) Yang, W.; Zhou, J.; Wang, B.; Ren, H. Lewis Acid-promoted synthesis of Unsymmetrical and highly functionalized carbazoles and dibenzofurans from biaryl triazenes: application for the total synthesis of Clausine C, Clausine R, and Clauraila A. Chem. - Eur. J. 2011, 17, 13665−13669. (h) Cho, J. Y.; Roh, G.-B.; Cho, E. J. Visible-lightpromoted synthesis of dibenzofuran derivatives. J. Org. Chem. 2018, 83, 805−811. (i) Xiao, B.; Gong, T.; Liu, Z.; Liu, J.; Luo, D.; Xu, J.; Liu, L. Synthesis of dibenzofurans via Palladium-catalyzed phenoldirected C-H Activation/C-O cyclization. J. Am. Chem. Soc. 2011, 133, 9250−9253. (6) Liu, Z.; Larock, R. C. Synthesis of carbazoles and dibenzofurans via cross-coupling of o-iodoanilines and o-iodophenols with silylaryl triflates. Org. Lett. 2004, 6, 3739−3741. (7) Zhao, H.; Yang, K.; Zheng, H.; Ding, R.; Yin, F.; Wang, N.; Li, Y.; Cheng, B.; Wang, H.; Zhai, H. A one-Pot synthesis of dibenzofurans from 6-diazo-2-cyclohexenones. Org. Lett. 2015, 17, 5744−5747. (8) Wang, M.; Wei, J.; Fan, Q.; Jiang, X. Cu(II)-catalyzed sulfide construction: both aryl groups utilization of intermolecular and intramolecular diaryliodonium salt. Chem. Commun. 2017, 53, 2918− 2921. (9) (a) Wang, M.; Fan, Q.; Jiang, X. Nitrogen−Iodine Exchange of Diaryliodonium Salts: Access to Acridine and Carbazole. Org. Lett. 2018, 20, 216−219. (b) Riedmüller, S.; Nachtsheim, B. J. Palladiumcatalyzed synthesis of N-arylated carbazoles using anilines and cyclic diaryliodonium salts. Beilstein J. Org. Chem. 2013, 9, 1202−1209. (10) Tan, Q.; Zhou, D.; Zhang, T.; Liu, B.; Xu, B. Iodine-doped sumanene and its application for the synthesis of chalcogenasumanenes and silasumanenes. Chem. Commun. 2017, 53, 10279−10282. (11) (a) Liu, L.; Qiang, J.; Bai, S.; Sung, H.-L.; Miao, C.; Li, J. Direct access to isoindolin-1-one scaffolds by copper-catalyzed divergent Cyclizations of 2-formylbenzonitrile and diaryliodonium salts. Adv. Synth. Catal. 2017, 359, 1283−1289. (b) Liu, L.; Bai, S.; Li, Y.; Wang, L.-X.; Hu, Y.; Sung, H.-L.; Li, J. Synthesis of 2,3-diarylisoindolin-1one by copper-catalyzed cascade annulation of 2-formylbenzonitriles, arenes, and diaryliodonium Salts. J. Org. Chem. 2017, 82, 11084− 11090. (c) Liu, L.; Bai, S.; Li, Y.; Ding, X.; Liu, Q.; Li, J. A threecomponent cascade cyclization to construct 3-(2-Oxopropyl)-2arylisoindolinone derivatives via copper-catalyzed annulation. Adv. Synth. Catal. 2018, 360, 1617−1621. (12) Liu, L.; Qiang, J.; Bai, S.; Li, Y.; Li, J. Iron-catalyzed carbon− sulfur bond formation: Atom-economic construction of thioethers with diaryliodonium salts. Appl. Organomet. Chem. 2017, No. e3810. (13) Liu, L.; Qiang, J.; Bai, S.; Li, Y.; Miao, C.; Li, J. Palladiumcatalyzed cyclocarbonylation of cyclic diaryliodoniums: Synthesis of fluorenones. Appl. Organomet. Chem. 2017, No. e3817. (14) Vedernikov, A. N. C-O reductive elimination from high valent Pt and Pd centers. Top. Organomet. Chem. 2010, 31, 101−121. (15) Cao, C.; Sheng, J.; Chen, C. Cu-Catalyzed Cascade Annulation of diaryliodonium salts and nitriles: synthesis of nitrogen-containing heterocycles. Synthesis 2017, 49, 5081−5092. 12928

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929

ACS Omega

Article

(16) (a) Klauk, H. Organic Electronics: Materials, Manufacturing and Applications; Wiley-VCH: Weinheim, Germany, 2006. (b) Dimitrakopoulos, C. D.; Malenfant, P. R. L.; Dimitrakopoulos, C. D.; Malenfant, P. R. L. Organic thin film transistors for large area electronics. Adv. Mater. 2002, 14, 99−117. (17) Zhu, D.; Wu, Z.; Luo, B.; Du, Y.; Liu, P.; Chen, Y.; Hu, Y.; Huang, P.; Wen, S. Heterocyclic iodoniums for the assembly of oxygen-bridged polycyclic heteroarenes with water as the oxygen source. Org. Lett. 2018, 20, 4815−4818. (18) Kawaguchi, K.; Nakano, K.; Nozaki, K. Synthesis of ladder-type π-conjugated heteroacenes via Palladium-catalyzed double NArylation and Intramolecular O-Arylation. J. Org. Chem. 2007, 72, 5119−5128. (19) Wu, B.; Yoshikai, N. Conversion of 2-iodobiaryls into 2,2′diiodobiaryls via oxidation-iodination sequences: a versatile route to ladder-type heterofluorenes. Angew. Chem. 2015, 127, 8860−8863.

12929

DOI: 10.1021/acsomega.8b02345 ACS Omega 2018, 3, 12923−12929