Next Article in Journal
The Manufacturing Process of Kiwifruit Fruit Powder with High Dietary Fiber and Its Laxative Effect
Next Article in Special Issue
Catalyst-Free Synthesis of Polysubstituted 5-Acylamino-1,3-Thiazoles via Hantzsch Cyclization of α-Chloroglycinates
Previous Article in Journal
Synthesis of Trichodermin Derivatives and Their Antimicrobial and Cytotoxic Activities
Previous Article in Special Issue
Sequential MCR via Staudinger/Aza-Wittig versus Cycloaddition Reaction to Access Diversely Functionalized 1-Amino-1H-Imidazole-2(3H)-Thiones
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Nucleophilic Arylation of Halopurines Facilitated by Brønsted Acid in Fluoroalcohol

1
Faculty of Pharmacy, Meijo University, 150 Yagotoyama, Tempaku-ku, Nagoya 468-8503, Japan
2
College of Pharmaceutical Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525–8577, Japan
3
Department of Applied Chemistry, College of Life Sciences, Ritsumeikan University, 1-1-1 Nojihigashi, Kusatsu, Shiga 525–8577, Japan
*
Authors to whom correspondence should be addressed.
Molecules 2019, 24(21), 3812; https://doi.org/10.3390/molecules24213812
Submission received: 8 October 2019 / Revised: 18 October 2019 / Accepted: 21 October 2019 / Published: 23 October 2019
(This article belongs to the Special Issue Modern Strategies for Heterocycle Synthesis)

Abstract

:
Various aryl-substituted purine derivatives were synthesized through the direct arylation of halopurines with aromatic compounds, facilitated by the combination of triflic acid and fluoroalcohol. This metal-free method is complementary to conventional coupling reactions using metal catalysts and reagents for the syntheses of aryl-substituted purine analogues.

1. Introduction

Biogenic purine bases are heteroaromatic compounds that constitute the basic subunits of DNA and RNA and play a crucial role in biological processes. In addition to these natural nucleosides, various chemically modified purine nucleosides have recently been discovered, and detailed analyses of their bioactivities have attracted much attention. Purine derivatives bearing an aryl substituent are of particular interest among these extensively studied classes of compounds, and their preparation has gained much attention owing to the promising biological properties of the derivatives, such as cytotoxicity and antitumor activity [1,2,3]. In addition, their applications as biological probes have also been consistent with the synthetic advances in artificial purine compounds [4].
The classical methods for the preparation of purines bearing aryl substituents are based on heterocyclization; however, the cyclization methodology usually requires multistep procedures. Consequently, the synthesis of the target aryl purines afforded only moderate to low yields [5]. The recent methods for the synthesis of aryl-substituted purines involve the transition metal-catalyzed cross-coupling reactions of aryl organometallics (Ar-M) with halopurines (Scheme 1A) [6,7]. For example, Suzuki–Miyaura [8,9,10,11,12,13,14,15,16], Stille [17,18], Negishi [19], and Kumada [20] coupling reactions have been frequently used for the preparation of aryl-substituted purines. Indeed, these approaches represent versatile and reliable synthetic methods; however, these coupling reactions require stoichiometric amounts of metallic reagents and the protection of the nucleophilic functional groups—such as the hydroxyl and amino groups—in the substrates. Hence, direct arylation of 6-chloropurines by electron-rich arenes using a three-fold excess of aluminum chloride (AlCl3) was reported by Guo’s group as an alternative method for preparing aryl purines in a short synthetic step (Scheme 1B) [21].
Despite these synthetic advances brought about by the alternative method, there unfortunately remain some limitations regarding the structural diversity of the obtainable aryl purines. To the best of our knowledge, the preparation of N-7-substituted 6-arylpurines has seldom been reported in the scientific literature [22,23,24]. To expand the synthetic scope for obtaining highly functionalized aryl purines with greater structural and steric diversities, developing a new practical method for preparing a variety of aryl-substituted purines is still necessary. In our continuous study on the development of a new method for the synthesis of functionalized nucleobases [25,26,27], we would like to report herein the metal-free arylation of purine derivatives facilitated by the combination of triflic acid and fluoroalcohol (Scheme 2).

2. Results and Discussions

Fluoroalcohols, such as 1,1,1,3,3,3-hexafluoro-2-propanol (HFIP) and 2,2,2-trifluoroethanol (TFE), possess specific properties that differentiate them from other non-fluorinated alcohols; they are highly polar [28] and weakly nucleophilic [29] and exhibit remarkable hydrogen-bond donor abilities [30]. Owing to their unique physical properties, these fluoroalcohols can dramatically direct the course of reactions; thus, as a means of developing new reactions, the authors utilized HFIP and TFE as attractive and distinctive alternatives to ordinary solvents in hypervalent iodine chemistry [31,32,33]. In these studies, we unexpectedly discovered the metal-free SNAr-type arylation of heteroaromatic diaryliodonium salts by nucleophilic aromatic compounds facilitated by Lewis and Brønsted acids, i.e., boron trifluoride, trimethylsilyl triflate, and triflic acid (TfOH), in fluoroalcohols [34,35,36]. Golding’s group also reported that the combination of trifluoroacetic acid (TFA) and TFE allowed the amination of halopurines by various substituted anilines under metal-free conditions; this method involves C–N bond formation [37,38,39,40]. Meanwhile, the metal-free nucleophilic arylation of halopurines involving C–C bond formation has not been reported.
In a pilot experiment, we first examined the SNAr-type coupling reaction of 6-chloropurine 1a initiated by a Brønsted acid [34,35], using methyl indole 2a as an aromatic nucleophile in HFIP as the model case (Table 1). In order to optimize the coupling reaction, varying equivalents (Entries 1–4) of TfOH were used, and the desired arylation product of purine 3aa was obtained in excellent yield when we used 0.5 to 1.0 equiv. of TfOH for the reactions (Entries 3 and 4). The usage of alternative Brønsted acids as additives, such as H3PO4, p-TsOH, and even TFA [37,38,39,40], was not as effective and provided inferior results in comparison with the use of TfOH. The fluoroalcohol HFIP plays an essential role in the reaction, and a solvent mixture of HFIP and 1,2-dichloroethane (DCE) did not smoothly produce the coupling product 3aa. (Entry 5). Also, the replacement of HFIP with TFE and the use of methanol and acetonitrile as the solvent instead of HFIP yielded low or null amounts of product 3aa.
To evaluate the generality of the reaction system, the substrate scope under the optimized reaction conditions was examined (Table 2). The reaction of N-protected 9-benzyl-6-chloro-9H-purine 1b cleanly favored the corresponding product 3ba with good yield. When non-N-protected indoles 2bd were subjected to analogous reaction conditions, the desired products 3bbbd were also obtained in excellent yields. However, it was revealed that the reaction of indoles bearing electron-withdrawing groups, such as 5-nitroindole 2e, did not proceed under these reaction conditions due to the deactivation of the aromatic nucleophile by hydrogen bonding with HFIP [41,42,43]. Furthermore, other electron-rich arenes were as compatible as the aromatic nucleophiles; similarly, good results were obtained from the coupling reactions with 1-naphthol 2f, 1-methoxynaphthalene 2g, several alkoxybenzenes 2h,i, and resorcinol 2j.
One of the significant advantages of the present reaction system is the production of structural and sterically diverse N-7-substituted 6-arylpurines; these arylpurines are not easily accessible by other synthetic methods [22,23,24]. As a result, the proposed reaction conditions were also utilized for the coupling of 7-benzyl-6-chloro-7H-purine 1c with indole 2k and naphthalene nucleophile 2f to afford the corresponding 6-arylated N-7-substituted purines 3ck and 3cf in good yields (Scheme 3).
When using p-anisidine 4 as a substrate, our reaction system with TfOH became valuable for the chemoselective N-arylation of halopurines at the 6 position under mild temperature (Scheme 4) [37]. We subjected 9H-chloropurine 1d and aniline 4 to our optimized conditions at 60 °C, obtaining selective N-arylation that smoothly provided N-(4-methoxyphenyl)-9H-purine-6-amine 5 in 79% yield, without the formation of the C-arylated purine coupling product 5′. On the other hand, Guo’s group previously reported the reaction of purines and anilines or naphthylamines in the presence of a three-fold excess of AlCl3 in DCE, which alternatively gave the C-arylated coupling products and likewise the biaryl 5′ [21]. Therefore, our reaction system is complementary to the AlCl3-mediated coupling reaction [21] for the syntheses of C6-aryl-substituted purine derivatives in view of product selectivity.
The success of the metal-free coupling reaction relies on the use of HFIP as the solvent. Although the precise role of HFIP [44,45,46] remains unclear, we presume that HFIP can increase the acidity of TfOH (Brønsted acid activation by H-bond donor) to enhance the reactivity of halopurine electrophiles through the purine nitrogen atoms [47,48,49,50]. Importantly, HFIP offers a means of improving the leaving group ability of the chloride atom in the purine substrates through hydrogen bonding as well as solvation [51]. Recently, such unique role of fluoroalcohol as the H-bond donor has been discussed in several Brønsted acid catalyzed reactions in regard to its ability to accelerate substitution processes [37,47,48,49,50,51]. Interestingly, these cases would involve intermediates activated by hydrogen bonding with fluoroalcohol, and, with our present system, the formation of a similar intermediate would also be expected to facilitate the aromatic substitution reactions.

3. Conclusions

In conclusion, we have developed a new metal-free coupling method of halopurines for the syntheses for diverse C6-aryl-substituted purine derivatives based on Brønsted acid activation. The combination of TfOH and HFIP is an efficient and practical methodology for the direct nucleophilic arylation of halopurines under mild conditions. We have elucidated that the unique properties of HFIP (hydrogen-bonding formation and weak nucleophilicity) could facilitate the direct arylation of halopurines by various nucleophilic arene molecules. Further investigations on the utilization of the obtained purine biaryls are currently underway in our research group.

4. Experimental Section

The melting points (mp) are uncorrected. The 1H-NMR (and 13C-NMR) spectra of the coupling products 3 and 5 were recorded by a JEOL JMN-400 spectrometer (JEOL Ltd., Tokyo, Japan) operating at 400 MHz (100 MHz for 13C-NMR) in DMSO-d6 at 25 °C with tetramethylsilane as the internal standard. The data are reported as follows: chemical shift in part per million (δ), multiplicity (s = singlet, d = doublet, t = triplet, q = quartet, br = broad singlet, m = multiplet), integration, and coupling constant (Hz). The infrared spectra (IR) were obtained using a Hitachi 270–50 spectrometer (Hitachi Ltd., Tokyo, Japan); absorptions are reported in reciprocal centimeters (cm−1) for representative peaks. High-resolution mass spectra were measured with a Thermo Scientific Exactive Plus Orbitrap (Thermo Fisher Scientific., Inc., Waltham, MA, USA). All chemicals used in this study are commercially available and were used without further purification. Regarding fluoroalcohol, we used commercial water-containing hexafluoroisopropanol (HFIP) as supplied for the reactions.

4.1. General Procedure for Brønsted Acid Catalyzed Arylation of Halopurines in Fluoroalcohol (Table 2 and Scheme 3)

To a stirred solution of chloropurine 1 (0.50 mmol) in hexafluoroisopropanol (5 mL), aromatic nucleophile 2 (0.55 mmol, 1.1 equiv) and trifluoromethanesulfonic acid (TfOH, 44 μL, 0.5 mmol, 1 equiv) were successively added. The resulting mixture was stirred at 60 °C for 24 h. After completion of the reaction checked by TLC, the reaction mixture was poured into sat. NaHCO3 aqueous. The resultant solution was extracted with ethyl acetate, dried with solid sodium sulfate, and then concentrated. The residue was purified by short-column chromatography on silica gel using hexane-ethyl acetate as the eluent to give the purine aromatic-linked compound 3 in the indicated yield in Table 2 or Scheme 3.
Compound 6-(1-methyl-1H-indol-3-yl)-7H-purine (3aa). A yellow powder, mp 346–350 °C. IR: 3647, 1732 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.95 (s, 3H), 7.21–7.31 (m, 2H), 7.55 (d, J = 7.9 Hz, 1H), 8.49 (s, 1H), 8.81–8.84 (m, 2H), 8.97 (s, 1H), 13.4 (bs, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 33.1, 110.4 (x 2), 121.0, 122.4, 122.9, 126.2, 128.0, 136.2, 137.2, 142.7, 151.3, 152.1, 152.3 ppm; HRMS (DART): Calcd. for C14H12N5 [M + H]+: 250.1087, found: 250.1087.
Compound 6-(1-methyl-1H-indol-3-yl)-9-phenylmethyl)-9H-purine (3ba) [21]. A yellow powder, mp 163–166 °C. IR: 3047, 2932, 1581, 1536, 1498, 1475 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.96 (s, 3H), 5.51(s, 2H), 7.23–7.59 (m, 7H), 7.57 (d, 1H, J = 7.8 Hz), 8.68 (s, 1H), 8.80 (d, 1H, J = 7.8 Hz), 8.86 (s, 1H), 8.96 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 33.0, 54.9, 110.2, 110.3, 121.1, 122.5, 122.8, 126.2, 127.6, 127.8, 128.1, 128.7, 136.3, 136.8, 137.2, 144.4, 150.3, 152.3, 152.7 ppm.
Compound 6-(1H-indol-3-yl)-9-phenylmethyl-9H-purine (3bb) [21]. A yellow powder, mp 178–180 °C. IR: 3631, 1688 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 5.51 (s, 2H), 7.17–7.41 (m, 7H), 7.52 (d, 1H, J = 6.8 Hz), 8.66 (s, 1H), 8.76–8.89 (m, 2H), 8.99 (s, 1H), 12.0 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.3, 111.2, 112.0, 120.9, 122.4, 122.7, 125.7, 127.9, 128.4, 128.8, 132.8, 136.6, 136.9, 144.5, 150.3, 152.3, 153.1 ppm.
Compound 6-(5-methyl-1H-indol-3-yl)-9-phenylmethyl-9H-purine (3bc). A brown liquid, IR: 3649, 1690, 1559, 1540 cm–1. 1H-NMR (400 MHz, DMSO-d6) δ 2.36 (s, 3H), 5.45 (s, 2H), 7.08 (d, J = 8.3 Hz, 1H), 7.26−7.36 (m, 6H), 8.53 (s, 1H), 8.59 (s, 1H), 8.82 (s, 1H), 8.90 (s, 1H), 11.8 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 21.5, 46.0, 110.8, 111.7, 122.3, 124.0, 126.0, 127.6, 127.9, 128.3, 128.8, 129.5, 132.9, 135.0, 136.9, 144.3, 150.2, 152.3, 153.2 ppm; HRMS (DART): Calcd. for C21H18N5+ [M + H]+: 340.1557, found: 340.1557.
Compound 6-(5-methoxy-1H-indol-3-yl)-9-phenylmethyl-9H-purine (3bd). A brown solid, mp 229–231 °C. IR: 3595, 1704, 1559, 1508, 1437 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.89 (s, 3H), 5.57 (s, 2H), 6.92 (d, 1H, J = 8.6 Hz), 7.31–7.49 (m, 6H), 8.38 (s, 1H), 8.71 (s, 1H), 8.93 (s, 1H), 9.01 (s, 1H), 11.9 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.3, 55.4, 104.6, 111.0, 112.2, 112.7, 126.4, 127.6, 127.9, 128.2, 128.8, 131.6, 133.2, 136.9, 144.3, 150.2, 152.3, 153.2, 154.7 ppm; HRMS (DART): Calcd. for C21H18N5O+ [M + H]+: 356.1506, found: 356.1507.
Compound 4-(7H-purin-6-yl)-naphthalene-1-ol (3af). A yellow powder, mp 204–208 °C. IR: 3650, 1541 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 7.09 (d, 1H, J = 8.6 Hz), 7.51–7.57 (m, 2H), 8.06–8.63 (m, 4H), 9.06 (s, 1H), 10.8 (bs, 1H), 13.3 (bs, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 107.8, 122.6, 123.5, 125.0, 125.3, 125.8, 127.3, 131.4, 132.2, 152.0, 155.5 ppm; HRMS (DART): Calcd. for C15H11N4O+ [M + H]+: 263.0927, found: 263.0928.
Compound 6-(4-methoxynaphthalen-1-yl)-7H-purine (3ag). A yellow powder, mp 170–172 °C. IR: 3629, 1704, 1542, 1508 cm–1. 1H-NMR (400 MHz, DMSO-d6) δ 4.06 (s, 3H), 7.16 (d, 1H, J = 8.6 Hz), 7.49–7.59 (m, 2H), 8.22–8.30 (m, 3H), 8.58 (bs, 1H), 9.02 (s, 1H), 13.6 (bs, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 56.0, 103.9, 121.7, 124.8, 125.0, 125.6, 127.1, 131.5, 151.6, 156.3 ppm; HRMS (DART): Calcd. for C16H13N4O+ [M + H]+: 277.1084, found: 277.1082.
Compound 6-(4-methoxynaphthalen-1-yl)-9-phenylmethyl-9H-purine (3bg) [21]. A white solid, mp 197–198 °C. IR: 3672, 2968, 1507 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 4.02 (s, 3H), 5.55 (s, 2H), 7.12 (d, 1H, J = 8.0 Hz), 7.24–7.36 (m, 3H), 7.42 (d, 2H, J = 7.4 Hz), 7.48–7.55 (m, 2H), 8.13 (d, 1H, J = 8.6 Hz), 8.24–8.28 (m, 1H), 8.44–8.49 (m, 1H), 8.75 (s, 1H), 9.09 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.5, 55.8, 103.8, 121.7, 124.5, 125.0, 125.4, 125.9, 127.0, 127.8, 127.9, 128.7, 131.6, 131.8, 136.5, 146.1, 151.7, 151.8, 156.4, 156.5 ppm.
Compound 6-(1,3,5-trimethoxyphen-4-yl)-9-phenylmethyl-9H-purine (3bh) [21]. A white solid, mp 251–253 °C. IR: 3650, 1698 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.58 (s, 6H), 3.84 (s, 3H), 5.49 (s, 2H), 6.35 (s, 2H), 7.27–7.47 (m, 5H), 8.61 (s, 1H), 8.90 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.6, 55.5, 55.7, 91.0, 106.6, 128.1, 128.8, 133.6, 136.6, 145.7, 150.8, 151.8, 154.2, 158.7, 162.0 ppm.
Compound 6-(1,3-dimethoxyphen-4-yl)-9-phenylmethyl-9H-purine (3bi) [52]. A white solid, mp 125–127 °C. IR: 3671, 1707 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.74 (s, 3H), 3.83 (s, 3H), 5.50 (s, 2H), 6.66 (d, 1H, J = 8.8 Hz), 6.73 (s, 1H), 7.24–7.42 (m, 5H), 7.53 (d, 1H, J = 8.3 Hz), 8.66 (s, 1H), 8.93 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.5, 55.4, 55.7, 98.9, 105.2, 117.8, 127.8, 128.0, 128.8, 131.9, 132.5, 136.7, 145.6, 151.1, 151.8, 155.4, 158.8, 162.0 ppm.
Compound 4-(9-phenylmethyl-9H-purin-6-yl)-benzene-1,3-diol (3bj) [21]. A yellow solid, mp 250–253 °C. IR: 3691, 2983, 1686, 1507, 1318 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 5.47 (s, 2H), 6.36 (s, 1H), 6.48 (d, 1H, J = 8.8 Hz), 7.20–7.34 (m, 5H), 8.72 (s, 1H), 8.79 (s, 1H), 9.21 (d, 1H, J = 8.8 Hz), 14.6 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.6, 103.3, 106.4, 108.2, 109.1, 127.8, 128.8, 133.8, 136.5, 145.6, 149.5, 151.1, 154.4, 158.6, 162.5, 163.5 ppm.
Compound 6-(1-phenyl-1H-indol-3-yl)-7-phenylmethyl-7H-purine (3ck). A yellow powder, mp 191–194 °C. IR: 1693, 1521 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 5.54 (s, 2H), 7.26–7.39 (m, 7H), 7.49–7.74 (m, 6H), 8.72 (s, 1H), 8.91–8.95 (m, 2H), 9.14 (s, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 46.4, 110.9, 112.8, 122.1, 123.2, 123.6, 124.5, 126.8, 127.6, 127.7, 127.9, 128.8, 130.1, 134.5, 136.0, 136.7, 138.2, 143.3, 145.0, 150.5, 152.0, 152.3 ppm. HRMS (DART): Calcd. for C26H20N5 [M + H]+: 402.1713, found: 402.1713.
Compound 4-(7-phenylmethyl-7H-purin-6-yl)-naphthalene-1-ol (3cf). A yellow powder, mp 259–263 °C. IR: 3613, 2980, 1697 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 4.96-5.04 (m, 2H), 6.18 (d, 2H, J = 7.3 Hz), 6.77 (t, 2H, J = 7.3 Hz), 6.86 (t, 1H, J = 7.3 Hz), 6.93 (d, 1H, J = 7.8 Hz), 7.18 (d, 1H, J = 8.3 Hz), 7.26 (t, 2H, J = 5.4 Hz), 7.42 (t, 1H, J = 7.3 Hz), 8.22 (d, 1H, J = 8.3 Hz), 8.91 (s, 1H), 9.04 (s, 1H), 10.7 (bs, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 49.8, 107.0, 122.2, 123.4, 123.6, 124.3, 124.7, 124.9, 125.7, 126.9, 127.3, 127.9, 128.8, 132.3, 135.6, 150.8, 151.7, 152.0, 154.9, 161.6 ppm; HRMS (DART): Calcd. for C22H17N4O+ [M + H]+: 353.1397, found: 353.1395.

4.2. General Procedure for Brønsted Acid Catalyzed N-Coupling of Aniline Derivatives to Halopurines in Fluoroalcohol (Scheme 4)

To a stirred solution of 9H-chloropurine 1d (77.3 mg, 0.50 mmol) in hexafluoroisopropanol (5 mL) p-methoxyaniline 4 (67.8 mg, 0.55 mmol, 1.1 equiv) and trifluoromethanesulfonic acid (TfOH, 44 μL, 0.5 mmol, 1 equiv) were successively added. The resulting mixture was stirred at 60 °C for 24 h. After completion of the reaction checked by TLC, the reaction mixture was poured into sat. NaHCO3 aqueous. The resultant solution was extracted with ethyl acetate, dried with solid sodium sulfate, and then concentrated. The residue was purified by short-column chromatography on silica gel using hexane-ethyl acetate as the eluent to give N-(4-methoxyphenyl)-9H-purine-6-amine 5 in 79% yield (95.3 mg, 0.395 mmol) as a white powder.
CompoundN-(4-methoxyphenyl)-9H-purine-6-amine (5) [53]. A white solid, mp 266–267 °C. IR: 3673, 3630 cm−1. 1H-NMR (400 MHz, DMSO-d6) δ 3.33 (s, 3H), 6.89 (d, 2H, J = 9.3 Hz), 7.79 (d, 2H, J = 8.8 Hz), 8.22 (s, 1H), 8.29 (s, 1H), 9.61 (s, 1H), 13.1 (bs, 1H) ppm; 13C-NMR (100 MHz, DMSO-d6) δ 55.2, 113.6, 119.2, 122.4, 132.8, 139.5, 150.2, 151.9, 154.9, 159.7 ppm.

Author Contributions

N.T. and T.D. conceived and designed the experiments and directed the project; N.T., T.S., T.M., A.H., and S.U. performed the experiments; T.D. and K.K. analyzed the data and checked the experimental details; T.H. contributed to critical discussion and presentation of the results; N.T. and T.D. wrote the paper.

Funding

This work was supported by JSPS KAKENHI Grant Number 16K18854. T.D. acknowledges the support from JSPS KAKENHI (C) Grant Number 19K05466 and the Ritsumeikan Global Innovation Research Organization (R-GIRO) project.

Acknowledgments

We thank Central Glass Co., Ltd., for the generous gift of fluoroalcohol.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Legraverend, M.; Grierson, D.S. The purines: Potent and versatile small molecule inhibitors and modulators of key biological targets. Bioorg. Med. Chem. 2006, 14, 3987–4006. [Google Scholar] [CrossRef] [PubMed]
  2. Bakkestuen, A.K.; Gundersen, L.L.; Utenova, B.T. Synthesis, biological activity, and SAR of antimycobacterial 9-aryl-, 9-arylsulfonyl-, and 9-benzyl-6-(2-furyl)purines. J. Med. Chem. 2005, 48, 2710–2723. [Google Scholar] [CrossRef] [PubMed]
  3. Hocek, M.; Naus, P.; Pohl, R.; Votruba, I.; Furman, P.A.; Tharnish, P.M.; Otto, M.J. Cytostatic 6-arylpurine nucleosides. 6. SAR in anti-HCV and cytostatic activity of extended series of 6-hetarylpurine ribonucleosides. J. Med. Chem. 2005, 48, 5869–5873. [Google Scholar] [CrossRef] [PubMed]
  4. Storr, T.E.; Strohmeier, J.A.; Baumann, C.G.; Fairlamb, I.J.S. A sequential direct arylation/Suzuki–Miyaura cross-coupling transformation of unprotected 20-deoxyadenosine affords a novel class of fluorescent analogues. Chem. Commun. 2010, 46, 6470–6472. [Google Scholar] [CrossRef] [PubMed]
  5. Shaw, G. Comprehensive Heterocyclic Chemistry; Katritzky, A.R., Rees, C.W., Eds.; Pergamon Press: Oxford, UK, 1984; Volume 5, pp. 501–597. [Google Scholar]
  6. Agrofoglio, L.A.; Gillaizeau, I.; Saito, Y. Palladium-assisted routes to nucleosides. Chem. Rev. 2003, 103, 1875–1916. [Google Scholar] [CrossRef]
  7. Hocek, M. Synthesis of purines bearing carbon substituents in positions 2, 6 or 8 by metal- or organometal-mediated C–C bond-forming reaction. Eur. J. Org. Chem. 2003, 245–254. [Google Scholar] [CrossRef]
  8. Lakshman, M.K.; Hilmer, J.H.; Martin, J.Q.; Keeler, J.C.; Dinh, Y.Q.; Ngassa, F.N.; Russon, L.M. Palladium catalysis for the synthesis of hydrophobic C-6 and C-2 aryl 2′-deoxynucleosides. Comparison of C–C versus C–N bond formation as well as C-6 versus C-2 reactivity. J. Am. Chem. Soc. 2001, 123, 7779–7787. [Google Scholar] [CrossRef]
  9. Hocek, M.; Holy, A.; Votruba, I.; Dvorakova, H. Synthesis and cytostatic activity of substituted 6-phenylpurine bases and nucleosides: Application of the Suzuki-Miyaura cross-coupling reactions of 6-chloropurine derivatives with phenylboronic acids. J. Med. Chem. 2000, 43, 1817–1825. [Google Scholar] [CrossRef]
  10. Cerna, I.; Pohl, R.; Klepetarova, B.; Hocek, M. Synthesis of 6,8,9-tri- and 2,6,8,9-tetrasubstituted purines by a combination of the Suzuki cross-coupling, N-arylation, and direct C–H arylation reactions. J. Org. Chem. 2008, 73, 9048–9054. [Google Scholar] [CrossRef]
  11. Lakshman, M.K.; Thomson, P.F.; Nuqui, M.A.; Hilmer, J.H.; Sevova, N.; Boggess, B. Facile Pd-catalyzed cross-coupling of 2′-deoxyguanosine O6-arylsulfonates with arylboronic acids. Org. Lett. 2002, 4, 1479–1482. [Google Scholar] [CrossRef]
  12. Gunda, P.; Russon, L.M.; Lakshman, M.K. Pd-catalyzed amination of nucleoside arylsulfonates to yield N6-aryl-2,6-diaminopurine nucleosides. Angew. Chem. Int. Ed. 2004, 43, 6372–6377. [Google Scholar] [CrossRef] [PubMed]
  13. Lakshman, M.K.; Gunda, P.; Pradhan, P. Mild and room temperature C–C bond forming reactions of nucleoside C-6 arylsulfonates. J. Org. Chem. 2005, 70, 10329–10335. [Google Scholar] [CrossRef] [PubMed]
  14. Liu, J.; Robins, M.J. Fluoro, alkylsulfanyl, and alkylsulfonyl leaving groups in suzuki cross-coupling reactions of purine 2′-deoxynucleosides and nucleosides. Org. Lett. 2005, 7, 1149–1151. [Google Scholar] [CrossRef] [PubMed]
  15. Liu, J.; Robins, M.J. Azoles as Suzuki cross-coupling leaving groups: Syntheses of 6-arylpurine 2′-deoxynucleosides and nucleosides from 6-(imidazol-1-yl)- and 6-(1,2,4-triazol-4-yl)purine derivatives. Org. Lett. 2004, 6, 3421–3423. [Google Scholar] [CrossRef] [PubMed]
  16. Kang, F.A.; Sui, Z.; Murray, W.V. Pd-catalyzed direct arylation of tautomerizable heterocycles with aryl boronic acids via C-OH bond activation using phosphonium salts. J. Am. Chem. Soc. 2008, 130, 11300–11302. [Google Scholar] [CrossRef]
  17. Langli, G.; Gundersen, L.-L.; Rise, F. Regiochemistry in Stille couplings of 2,6-dihalopurines. Tetrahedron 1996, 52, 5625–5638. [Google Scholar] [CrossRef]
  18. Havelková, M.; Dvořák, D.; Hocek, M. Covalent analogues of DNA base-pairs and triplets. Part 3: Synthesis of 1,4- and 1,3-bis(purin-6-yl)benzenes and 1-(1,3-dimethyluracil-5-yl)-3 or 4-(purin-9-yl)benzenes. Tetrahedron 2002, 58, 7431–7435. [Google Scholar] [CrossRef]
  19. Gundersen, L.L.; Langli, G.; Rise, F. Regioselective Pd-mediated coupling between 2,6-dichloropurines and organometallic reagents. Tetrahedron Lett. 1995, 36, 1945–1948. [Google Scholar] [CrossRef]
  20. Furstner, A.; Leitner, A.; Mendez, M.; Krause, H. Iron-catalyzed cross-coupling reactions. J. Am. Chem. Soc. 2002, 124, 13856–13863. [Google Scholar] [CrossRef]
  21. Guo, H.-M.; Li, P.; Niu, H.-Y.; Wang, D.-C.; Qu, G.-R. Direct synthesis of 6-arylpurines by reaction of 6-chloropurines with activated aromatics. J. Org. Chem. 2010, 75, 6016–6018. [Google Scholar] [CrossRef]
  22. Edenhofer, A. Novel route to imidazoles, and their use for the synthesis of purines and 4,6-dihydro-1,2-dimethyl-8-phenylimidazo[4,5-e]-1,4-diazepin-5(1H)-one. Helv. Chim. Acta 1975, 58, 2192–2209. [Google Scholar] [CrossRef]
  23. Gundersen, L.-L.; Bakkestuen, A.K.; Aasen, J.; Overas, H.; Rise, F. 6-Halopurines in palladium-catalyzed coupling with organotin and organozinc reagents. Tetrahedron 1994, 50, 9743–9756. [Google Scholar] [CrossRef]
  24. Havelkova, M.; Hocek, M.; Cesnek, M.; Dvorak, D. The Suzuki-Miyaura cross-coupling reactions of 6-halopurines with boronic acids leading to 6-aryl- and 6-alkenylpurines. Synlett 1999, 1145–1147. [Google Scholar] [CrossRef]
  25. Takenaga, N.; Ueda, S.; Hayashi, T.; Dohi, T.; Kitagaki, S. Facile synthesis of stable uracil-iodonium(III) salts with various counterions. Heterocycles 2018, 97, 1248–1256. [Google Scholar] [CrossRef]
  26. Takenaga, N.; Ueda, S.; Hayashi, T.; Dohi, T.; Kitagaki, S. Vicinal functionalization of uracil heterocycles with base activation of iodonium(III) salts. Heterocycles 2019, 99, 865–874. [Google Scholar] [CrossRef]
  27. Takenaga, N.; Hayashi, T.; Ueda, S.; Satake, H.; Yamada, Y.; Kodama, T.; Dohi, T. Synthesis of uracil-iodonium(III) salts for practical utilization as nucleobase synthetic modules. Molecules 2019, 24, 3034. [Google Scholar] [CrossRef]
  28. Reichardt, C. Empirical parameters of solvent polarity as linear free-energy relationships. Angew. Chem. Int. Ed. Engl. 1979, 18, 98–110. [Google Scholar] [CrossRef]
  29. Schadt, F.L.; Bentley, T.W.; Schleyer, P.V.R. The SN2-SN1 spectrum. 2. Quantitative treatments of nucleophilic solvent assistance. A scale of solvent nucleophilicities. J. Am. Chem. Soc. 1976, 98, 7667–7675. [Google Scholar] [CrossRef]
  30. Kamlet, M.J.; Abbound, J.L.M.; Abraham, M.H.; Taft, R.W. Linear solvation energy relationships. 23. A comprehensive collection of the solvatochromic parameters, .pi*., alpha., and .beta., and some methods for simplifying the generalized solvatochromic equation. J. Org. Chem. 1983, 48, 2877–2887. [Google Scholar] [CrossRef]
  31. Dohi, T.; Yamaoka, N.; Kita, Y. Fluoroalcohols: Versatile solvents in hypervalent iodine chemistry and syntheses of diaryliodonium(III) salts. Tetrahedron 2010, 66, 5775–5785. [Google Scholar] [CrossRef]
  32. Kamitanaka, T.; Morimoto, K.; Tsuboshima, K.; Koseki, D.; Takamuro, H.; Dohi, T.; Kita, Y. Efficient coupling reaction of quinone monoacetal with phenols leading to phenol biaryls. Angew. Chem. Int. Ed. 2016, 55, 15535–15538. [Google Scholar] [CrossRef] [PubMed]
  33. Dohi, T.; Ito, M.; Morimoto, K.; Minamitsuji, Y.; Takenaga, N.; Kita, Y. Versatile direct dehydrative approach for diaryliodonium(III) salts in fluoroalcohol media. Chem. Commun. 2007, 4152–4154. [Google Scholar] [CrossRef] [PubMed]
  34. Dohi, T.; Ito, M.; Yamaoka, N.; Morimoto, K.; Fujioka, H.; Kita, Y. Unusual ipso substitution of diaryliodonium bromides initiated by a single-electron-transfer oxidizing process. Angew. Chem. Int. Ed. 2010, 49, 3334–3337. [Google Scholar] [CrossRef] [PubMed]
  35. Yamaoka, N.; Sumida, K.; Itani, I.; Kubo, H.; Ohnishi, Y.; Sekiguchi, S.; Dohi, T.; Kita, Y. Single-electron-transfer (SET)-induced oxidative biaryl coupling by polyalkoxybenzene-derived diaryliodonium(III) salts. Chem. Eur. J. 2013, 19, 15004–15011. [Google Scholar] [CrossRef]
  36. Dohi, T.; Ueda, S.; Hirai, A.; Kojima, Y.; Morimoto, K.; Kita, Y. Selective aryl radical transfers into N-heteroaromatics from diaryliodonoium salts with trimethoxybenzene auxiliary. Heterocycles 2017, 95, 1272–1284. [Google Scholar] [CrossRef]
  37. Carbin, B.; Coxon, C.R.; Lebraud, H.; Elliott, K.J.; Matheson, C.J.; Meschini, E.; Roberts, A.R.; Turner, D.M.; Wong, C.; Cano, C.; et al. Trifluoroacetic acid in 2,2,2-trifluoroethanol facilitates SNAr reactions of heterocycles with arylamines. Chem. Eur. J. 2014, 20, 2311–2317. [Google Scholar] [CrossRef]
  38. Whitfield, H.J.; Griffin, R.J.; Hardcastle, I.R.; Henderson, A.; Meneyrol, J.; Mesguiche, V.; Sayle, K.L.; Golding, B.T. Facilitation of addition–elimination reactions in pyrimidines and purines using trifluoroacetic acid in trifluoroethanol. Chem. Commun. 2003, 2802–2803. [Google Scholar] [CrossRef]
  39. Marchetti, F.; Cano, C.; Curtin, N.J.; Golding, B.T.; Griffin, R.J.; Haggerty, K.; Newell, D.R.; Parsons, R.J.; Payne, S.L.; Wang, L.Z.; et al. Synthesis and biological evaluation of 5-substituted O4-alkylpyrimidines as CDK2 inhibitors. Org. Biomol. Chem. 2010, 8, 2397–2407. [Google Scholar] [CrossRef]
  40. Wong, C.; Griffin, R.J.; Hardcastle, I.R.; Northen, J.S.; Wang, L.Z.; Golding, B.T. Synthesis of sulfonamide-based kinase inhibitors from sulfonates by exploiting the abrogated SN2 reactivity of 2,2,2-trifluoroethoxysulfonates. Org. Biomol. Chem. 2010, 8, 2457–2464. [Google Scholar] [CrossRef]
  41. Bégué, J.; Bonnet-delpon, D.; Crousse, B. Fluorinated alcohols: A new medium for selective and clean reaction. Synlett 2004, 18–29. [Google Scholar] [CrossRef]
  42. Shuklov, I.A.; Dubrovina, N.V.; Boerner, A. Fluorinated alcohols as solvents, cosolvents and additives in homogeneous catalysis. Synthesis 2007, 2925–2943. [Google Scholar] [CrossRef]
  43. Baeza, A.; Najera, C. Recent advances in the direct nucleophilic substitution of allylic alcohols through SN1-type reactions. Synthesis 2014, 46, 25–34. [Google Scholar] [CrossRef]
  44. Berkessel, A.; Adrio, J.A. Dramatic acceleration of olefin epoxidation in fluorinated alcohols: Activation of hydrogen peroxide by multiple H-bond networks. J. Am. Chem. Soc. 2006, 128, 13412–13420. [Google Scholar] [CrossRef] [PubMed]
  45. Colomer, I.; Chamberlain, A.E.R.; Haughey, M.B.; Donohoe, T.J. Hexafluoroisopropanol as a highly versatile solvent. Nat. Rev. Chem. 2017, 1, 0088. [Google Scholar] [CrossRef]
  46. Zhou, Z.; Cheng, Q.-Q.; Kürti, L. Aza-Rubottom oxidation: Synthetic access to primary α-aminoketones. J. Am. Chem. Soc. 2019, 141, 2242–2246. [Google Scholar] [CrossRef] [PubMed]
  47. Liu, J.; Robins, M.J. SNAr displacements with 6-(fluoro, chloro, bromo, iodo, and alkylsulfonyl)purine nucleosides: Synthesis, kinetics, and mechanism. J. Am. Chem. Soc. 2007, 129, 5962–5968. [Google Scholar] [CrossRef] [PubMed]
  48. Vukovic, V.D.; Richmond, E.; Wolf, E.; Moran, J. Catalytic Friedel-Crafts reactions of highly electronically deactivated benzylic alcohols. Angew. Chem. Int. Ed. 2017, 56, 3085–3089. [Google Scholar] [CrossRef]
  49. Liu, W.; Wang, H.; Li, C.-J. Metal-free markovnikov-type alkyne hydration under mild conditions. Org. Lett. 2016, 18, 2184–2187. [Google Scholar] [CrossRef]
  50. Dohi, T.; Hu, Y.; Kamitanaka, T.; Washimi, N.; Kita, Y. [3 + 2] Coupling of quinone monoacetals by combined acid-hydrogen bond donor. Org. Lett. 2011, 13, 4814–4817. [Google Scholar] [CrossRef]
  51. Lee, J.W.; Oliveira, M.T.; Jang, H.B.; Lee, S.; Chi, D.Y.; Kim, D.W.; Song, C.E. Hydrogen-bond promoted nucleophilic fluorination: Concept, mechanism and applications in positron emission tomography. Chem. Soc. Rev. 2016, 45, 4638–4650. [Google Scholar] [CrossRef]
  52. Braendvang, M.; Gundersen, L.-L. Selective anti-tubercular purines: Synthesis and chemotherapeutic properties of 6-aryl- and 6-heteroaryl-9-benzylpurines. Bioorg. Med. Chem. 2005, 13, 6360–6373. [Google Scholar] [CrossRef]
  53. Wang, X.; Han, C.; Wu, K.; Luo, L.; Wang, Y.; Du, X.; He, Q.; Ye, F. Design, synthesis and ability of non-gold complexed substituted purine derivatives to inhibit LPS-induced inflammatory response. Eur. J. Med. Chem. 2018, 149, 10–21. [Google Scholar] [CrossRef]
Sample Availability: Samples of the products are available from the authors.
Scheme 1. General synthetic routes to obtain aryl-substituted purines from halopurines. (A) Stepwise synthesis. (B) Direct synthesis.
Scheme 1. General synthetic routes to obtain aryl-substituted purines from halopurines. (A) Stepwise synthesis. (B) Direct synthesis.
Molecules 24 03812 sch001
Scheme 2. Direct arylation of halopurines facilitated by Brønsted acid in fluoroalcohol. ArH: aryl nucleophile, TfOH: triflic acid.
Scheme 2. Direct arylation of halopurines facilitated by Brønsted acid in fluoroalcohol. ArH: aryl nucleophile, TfOH: triflic acid.
Molecules 24 03812 sch002
Scheme 3. Synthesis of 7-substituted 6-aryl purines.
Scheme 3. Synthesis of 7-substituted 6-aryl purines.
Molecules 24 03812 sch003
Scheme 4. Formation of different products using Brønsted acid and aluminum chloride as activators.
Scheme 4. Formation of different products using Brønsted acid and aluminum chloride as activators.
Molecules 24 03812 sch004
Table 1. Optimization of the reaction conditions a.
Table 1. Optimization of the reaction conditions a.
Molecules 24 03812 i001
EntrySolventTfOHYield (%) b
1HFIP0.1 equiv.8
2HFIP0.2 equiv.54
3HFIP0.5 equiv.89
4HFIP1.0 equiv.quant.
5HFIP/DCE = 9:11.0 equiv.60
a Reaction conditions: The reactions were performed using chloropurine 1a (0.50 mmol), 1-methylindole 2a (0.55 mmol), and TfOH in solvent (0.1 M). b Determined by 1H-NMR using nitromethane as an internal standard. HFIP: hexafluoroisopropanol, DCE: 1,2-dichloroethane.
Table 2. Scope of substrates a.
Table 2. Scope of substrates a.
Molecules 24 03812 i002
Molecules 24 03812 i003 Molecules 24 03812 i004 Molecules 24 03812 i005 Molecules 24 03812 i006
3aa: quant.3ba: 82%3bb: quant.3bc: 96%
Molecules 24 03812 i007 Molecules 24 03812 i008 Molecules 24 03812 i009 Molecules 24 03812 i010
3bd: 90%3be: n.r3af: 97%3ag: 85%
Molecules 24 03812 i011 Molecules 24 03812 i012 Molecules 24 03812 i013 Molecules 24 03812 i014
3bg: 92%3bh: quant.3bi: quant.3bj: quant.
a All the reactions were performed using chloropurine 1 (0.50 mmol), indole or aromatic nucleophile 2 (0.55 mmol), and TfOH (0.5 mmol, 1.0 equiv.) in HFIP (0.1 M) at 60 °C. The yields after isolation are indicated.

Share and Cite

MDPI and ACS Style

Takenaga, N.; Shoji, T.; Menjo, T.; Hirai, A.; Ueda, S.; Kikushima, K.; Hanasaki, T.; Dohi, T. Nucleophilic Arylation of Halopurines Facilitated by Brønsted Acid in Fluoroalcohol. Molecules 2019, 24, 3812. https://doi.org/10.3390/molecules24213812

AMA Style

Takenaga N, Shoji T, Menjo T, Hirai A, Ueda S, Kikushima K, Hanasaki T, Dohi T. Nucleophilic Arylation of Halopurines Facilitated by Brønsted Acid in Fluoroalcohol. Molecules. 2019; 24(21):3812. https://doi.org/10.3390/molecules24213812

Chicago/Turabian Style

Takenaga, Naoko, Toshitaka Shoji, Takayuki Menjo, Akiko Hirai, Shohei Ueda, Kotaro Kikushima, Tomonori Hanasaki, and Toshifumi Dohi. 2019. "Nucleophilic Arylation of Halopurines Facilitated by Brønsted Acid in Fluoroalcohol" Molecules 24, no. 21: 3812. https://doi.org/10.3390/molecules24213812

Article Metrics

Back to TopTop