Aryne Polymerization Enabling Straightforward Synthesis of Elusive

Dec 2, 2014 - Life Sciences, Mochida Memorial Foundation, Tokyo Bio- chemical Research .... Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P...
1 downloads 0 Views 1MB Size
Subscriber access provided by Brought to you by ST ANDREWS UNIVERSITY LIBRARY

Communication

Aryne polymerization enabling straightforward synthesis of elusive poly(ortho-arylene)s Yoshihide Mizukoshi, Koichiro Mikami, and Masanobu Uchiyama J. Am. Chem. Soc., Just Accepted Manuscript • Publication Date (Web): 02 Dec 2014 Downloaded from http://pubs.acs.org on December 2, 2014

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.

Journal of the American Chemical Society 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 5

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

Journal of the American Chemical Society

Aryne  polymerization  enabling  straightforward  synthesis  of  elusive   poly(ortho-­‐arylene)s   Yoshihide Mizukoshi,1 Koichiro Mikami*1 and Masanobu Uchiyama*1,2 1

2

Advanced Elements Chemistry Laboratory, Graduate School of Pharmaceutical Sciences, The University of Tokyo, 7-3-1 Hongo, Bunkyo-ku, Tokyo 113-0033, Japan. Advanced Elements Chemistry Research Team, RIKEN Center for Sustainable Resource Science, and Elements Chemistry Laboratory, RIKEN, 2-1 Hirosawa, Wako-shi, Saitama 351-0198, Japan.

Supporting Information Placeholder ABSTRACT: Herein we present the first straightfor-

ward aryne polymerization that successively connects aromatic rings at their ortho-positions, directly giving poly(ortho-phenylene)s of up to ca. 100-mer size. The polymerization proceeds smoothly in a chain-growth fashion in the presence of monovalent copper reagent. Direct synthesis of poly(orthoarylene)s has been a “missing piece” of aromatic chemistry for a long time (although the discovery of aryne dates back to 1902), in contrast to the wellinvestigated para- and meta-linked polyarylenes. Our achievement reported here is expected to open up new areas of nanocarbon and materials science.

Well-regulated linkage of aromatic rings is a central issue in fundamental synthetic chemistry and materials science.1 Aryne, formed by abstraction of two hydrogen atoms from an aromatic compound, is one of the oldest-known aromatic reactive intermediates,2 and has a distorted triple bond within the ring, resulting in a low-lying LUMO that is associated with extremely high reactivity and a short life time. Aryne chemistry has been extensively studied both experimentally and theoretically since its discovery over 100 years ago, and plays key roles in many fields, including reaction development, total synthesis of natural products and functionalization of nanomaterials.3 However, the polymerization of aryne has remained elusive. In 2005, Ihara et al. reported that all their attempts to polymerize aryne using a wide range of initiators (Grignard reagent, transition metals, radical species, etc.) under various conditions were unsuccessful.4 In 2014, Ito and Nozaki et al. developed a formal aryne polymerization through polymer-

ization of [2.2.1]oxabicyclic alkenes followed by dehydration of the resultant polymers.5 They noted that they adopted this approach because polymerization of aryne is extremely challenging due to its uncontrollable, high reactivity. Now, we report the first polymerization of aryne that enables direct synthesis of poly(ortho-arylene)s, consisting of aromatic units successively connected through the ortho-position. For in situ aryne generation, we selected a well-established protocol, i.e., the combination of otrimethylsilyl triflate precursor and fluorine anion (Scheme 1).6 After extensive experimentation to screen initiators, including various organic/organometallic reagents, we found that the use of a Lipshutz-type cuprate,7 a copper ate complex prepared from one equivalent of CuCN and two equivalents of nBuLi, for the polymerization of 1 in THF gave an insoluble white solid. The attenuated total reflection (ATR) infrared spectrum of the solid showed absorptions attributable to C-H bond and C=C double bond stretching vibrations of the aromatic moiety at around 3000 and 1500 cm-1, respectively (FIGURE S1). This strongly suggested that the insoluble solid was not an inorganic salt, but an organic compound. The insolubility prevented further analysis, but we speculated that the solid might be poly(orthophenylene).

ACS Paragon Plus Environment

Journal of the American Chemical Society

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 5

a

TMS Ar OTf

CsF 18-crown-6 Solvent

Ar

CuX quench

H Ar nX

Aryne

ppm

TMS Ar

TMS H 3CO

TMS

OTf , H 3CO

OTf

= 1 nBu

7.0

6.5

b

OTf

CuX =

7.5

6.0

DPn = 19

5.0

4.5

4.0

3.5

3.0

2.5

20

18

2

2CuCNLi 2 , CuCN , Lipshutz-type cuprate

5.5

S Cu ,

21

17

22 23

,

Cu , CuCl, CuBr , CuI 24 16

SCHEME 1. Polymerization of arynes.8 Solvent = THF or acetonitrile.

25 26

27 28

Installing dimethoxy side chains on the benzene ring of the precursor allowed us to access additional information about the structure of the product. When 4,5-dimethoxyl 2-(trimethylsilyl)phenyl triflate 2 was used as a precursor, we obtained a yellow material that was soluble in various common organic solvents. Consumption of 2 was complete within 6 hours, as monitored by gas chromatography (GC) coupled with mass spectrometry. Gel permeation chromatography (GPC) indicated that the material included polymeric compounds. The crude product was purified by means of recycling preparative highperformance liquid chromatography (HPLC), affording a polymer in 32% yield (Mn = 2380, Mw = 3250 and PDI = 1.36). These values indicated that ca. 18 monomer units were present in the polymer molecule. As shown in FIGURE 1a, the proton nuclear magnetic resonance (1H NMR) spectrum in deuterated chloroform (CDCl3) exhibited aromatic proton signals in the range from 4.78 ppm to 7.01 ppm and dimethoxy group proton signals at 2.78-4.13 ppm, which are in good agreement with those of oligo(ortho-phenylene)s synthesized in a stepwise manner by Fukushima and Aida et al.9,10 This strongly supported the view that polymerization of aryne had proceeded to afford a poly(orthophenylene). Furthermore, similar signals of long oligomeric products were observed in acetonitrile-d3 (Figure S2),9 providing further support for the formation of ortho-linkages in the polymerization.

15

29

30 m/z

2000

2500

3000

3500

4000

FIGURE 1. (a) 1H NMR spectrum in CDCl3 and (b) MALDI-TOF mass spectrum of poly(ortho-phenylene) (Mn = 2380, Mw = 3250 and PDI = 1.36) obtained by polymerization of 2 with Lipshutz-type cuprate in THF. DPn is degree of polymerization.

Well-regulated polymer growth was confirmed by matrix-assisted laser desorption ionization time-of-flight (MALDI-TOF) mass spectrometry of the obtained polymer (FIGURE 1b). Detailed analysis of the spectrum revealed that the mass-to-charge ratio (m/z) values could be expressed as 136.1 × n + 1.0 (H) + 26.0 (CN), where n is equal to the number of repeating units, indicating that a single cyano group existed in each polymer chain, and the polymer structure was NC-(ortho-phenylene)n-H. The expected peak for the 19-mer would be at m/z 136.1 × 19 (repeating unit) + 1.0 (H) + 26.0 (CN) = 2612.0, and a peak was observed at m/z 2612.2. We had initially supposed that the nBu ligand on the cuprate would initiate the polymerization, leading to the 19mer with nBu/H ends, which should show a peak at m/z 2643.1, but no such peak was observed. The employment of CuCN as the initiator in place of the cuprate resulted in effective and smooth polymerization at room temperature to give the poly(ortho-phenylene) (FIGURE 2). Furthermore, a broad range of monovalent copper salts, such as thiophenolate, mesitylate, chloride, bromide and iodide, also promoted the polymerization to afford the desired poly(ortho-phenylene)s (Table S1, FIGURE S3 and S6-S11).8 It is noteworthy that the use of mesityl copper as an initiator in acetonitrile gave ca.

ACS Paragon Plus Environment

Page 3 of 5

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 5! 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60

Journal of the American Chemical Society

100-mer polymer (Mn = 13830 g/mol, Mw = 20840 g/mol, PDI = 1.50) (FIGURE S4). Polymerization reaction did not occur or scarcely proceeded with CuBr2, KCN, Zn(CN)2, AuCl, AuClSMe2 or Pd2(dba)3 as the initiator, or without any initiator (FIGURE S5). The polymerization with CuCN proceeded even in the presence of 9,10-dihydroxyanthracene as a radical scavenger and took place in the dark, indicating that the polymerization is not initiated/promoted by a radical species or by light. Thus, monovalent copper salts appear to play a crucial role in the aryne polymerization. As shown in Table 1 and FIGURE 2, the value of Mn of the obtained polymers could be regulated by adjusting the feed ratio of 2 and initiator ([2]0/[I]0). In other words, varying the initiator amount enabled us to modulate the molecular weight of the resultant poly(ortho-arylene)s. This result supports the idea that the polymerization proceeds in a chaingrowth fashion, not through step growth.

nism. The complexation of CuCN to benzyne occurs first to provide a triangular complex (CP1), which serves as the precursor for the polymerization. The stabilization energy is very large (29.6 kcal/mol) because of an energetically and geometrically favorable interaction between the high-lying d-orbital of copper(I) and the low-lying LUMO of aryne, and this appears to be the reason why a wide range of Cu(I) reagents can promote this unique polymerization. Through TS1, the benzyne triple bond inserts into the Cu-CN bond to form CP2 with an activation energy of 25.3 kcal/mol and a large energy gain (–61.3 kcal/mol). Intermediate CP2 then forms a π-complex CP3 with the second benzyne, and insertion of the benzyne proceeds with an activation barrier of 18.9 kcal/mol to give a biphenyl copper(I) intermediate CP4 with large exothermicity. ∆G / kcal/mol

NC Cu

CuCN +

NC

Table 1. Polymerization of 2 with CuCN as an initiator in THF using CsF/18-crown-6 at various feed ratios. entry [2]0/[I]0 M wa Mn a 1 20 3220 4970 2 45 4050 6890 3 91 5100 11160 a These values were determined by GPC, with polystyrene standards. entry

a

PDI a 1.54 1.70 2.19 calibrated

5! 4.5!

5000!

4!

4000 4000!

3.5!

Mn 3000!

2! 2.0

15! 15

time (min) Retention Time [min]

20! 20

25!

TS1 (–4.3)

Cu Cu

CP1 (–29.6) CP2 (–65.6)

NC Cu

NC Cu

+ CP3 (–88.0)

NC

TS2 (–69.1)

CP4 (–138.0)

Cu

3! 2.5! 2.5

2000! 2000

10!

SM (0.0)

FIGURE 3. A plausible reaction pathway of the polymerization. Energy diagram of the coordination/insertion reaction pathway based on density functional theory calculation at the B3LYP/6-31+G(d)LANL2DZ(Cu) level.

b 6000! 6000

NC

1000!

1.5 1.5!

0!

1! 1.0 100! 100

0! 0

20! 20 30!

40! 40

60! 60 [2]0/[I]0

80! 80

PDI

FIGURE 2. (a) GPC profile of the crude products obtained by polymerization of 2 with CuCN as an initiator in THF using CsF/18-crown-6 ether at various feed ratios; [2]0/[I] = 20 (black line), 45 (red line) and 91 (blue line). (b) Plots of Mn (magenta diamond) and PDI (blue square) versus [2]0/[I].

We next performed density functional theory (DFT) calculations to understand the polymerization mechanism.11 Global reaction route mapping (GRRM) analysis12 indicated that the reaction pathway of polymerization of benzyne initiated by CuCN (FIGURE 3) is the most probable. According to this pathway, the present aryne polymerization proceeds through a coordination/insertion mecha-

Unique properties and potential utility of various oligo-/poly(ortho-arylene)s have recently been partially demonstrated through indirect approaches.9,13,14 Ito et al. developed a synthesis of poly(quinoxaline-2,3-diyl)s through an aromatization approach with a palladium catalyst, and reported that the products showed helical geometry with potential application in dynamic chiral catalysts and in solid polymer films to selectively reflect circularly polarized light.13 Fukushima and Aida et al. reported an iterative stepwise approach to access poly(orthophenylene)s, which showed a redox-responsive spring motion based on an accumulated π-cloud along the helical axis and acted as surface modifiers for homeotropic columnar ordering of discotic liquid crystals.9,14b,c However, the lack of simple and straightforward synthetic methodology for orthosubstituted aromatic materials has hindered the de-

ACS Paragon Plus Environment

Journal of the American Chemical Society

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

velopment of innovative materials, in marked contrast to the ready accessibility of para- and metasubstituted materials,15 and a direct polymerization method for poly(ortho-arylene)s has long been regarded as a “missing piece” of aromatic chemistry.14a,b,15a,16 We believe the direct approach, aryne polymerization reported here, will provide robust access to a range of so-far-unexplored or poorly accessible poly(ortho-arylene) materials. In conclusion, after more than 100 years of aryne chemistry, we present the first aryne polymerization, providing straightforward methodology for direct synthesis of elusive poly(ortho-arylene)s, which have a range of potential applications in nanocarbon chemistry and materials science. ASSOCIATED  CONTENT     Supporting Information. Detailed experimental and theoretical section, GPC, 1H NMR, 13C NMR, IR, MALDI-TOF MS, and discussion for the polymerization mechanism. This material is available free of charge via the Internet at http://pubs.acs.org.

AUTHOR  INFORMATION   Corresponding  Author  

To whom correspondence should be addressed. Email: K.M. ([email protected]) and M.U. ([email protected]) Funding  Sources  

The authors declare no competing financial interests.

ACKNOWLEDGMENT     This work was partly supported by Grant-in-Aid for JSPS Fellows (No. 13J10930) (to K.M.), and grants from JSPS KAKENHI (S) (No. 24229011), Takeda Science Foundation, The Asahi Glass Foundation, DaiichiSankyo Foundation of Life Sciences, Mochida Memorial Foundation, Tokyo Biochemical Research Foundation, Foundation NAGASE Science Technology Development, Yamada Science Foundation and Sumitomo Foundation (to M.U.). The authors thank RIKEN Integrated Cluster of Clusters (RICC) for providing computer resources for the DFT calculations.

REFERENCES   1. (a) Granda, M.; Blanco, C.; Alvarez, P.; Patrick, J. W. and Menéndez, R. Chem. Rev. 2014, 114, 1608-1636. (b) Yokozawa, T. and Yokoyama, A. Chem. Rev. 2009, 109, 5595-5619. (c) Thomas III, S. W.; joly, G. D. and Swager T. M. Chem. Rev. 2007, 107, 1339-1386. (d) Hawker C. J. and Wooley, K. L. Science 2005, 309, 1200-1205. 2. (a) Stoermer, R. and Kahlert, B. Ber. Dtsch. Chem. Ges. 1902, 35, 1633-1640. (b) Wenk H. H.; Winkler, M. and Sander, W. Angew. Chem. Int. Ed. 2003, 5, 502-528. 3. (a) Hoffmann, R. W. Dehydrobenzene and Cycloalkynes; Academic Press: New York, 1967. (b) Nakamura, Y.;

Page 4 of 5

Takano, N.; Nishimura, T.; Yashima, E.; Sato, M.; Kudo, T. and Nishimura, J. Org. Lett. 2001, 3, 1193-1196. (c) Pellissier, H. and Santelli, M. Tetrahedron 2003, 59, 701-730. (d) Dyke, A. M.; Hester, A. J. and Lloyd-Jones, G. C. Synthesis 2006, 24, 4093-4112. (e) Gampe, C. M. and Carreira, E. M. Angew. Chem. Int. Ed. 2012, 51, 3766-3778. (f) Tadross, P. M. and Stoltz, B. M. Chem. Rev. 2012, 112, 3550-3577. (g) Hoye, T. R.; Baire, B.; Niu, D.; Willoughby, P. H. and Woods, B. P. Nature 2012, 490, 208-212. (h) Niu, D.; Willoughby, P. H.; Woods, B. P.; Baire, B. and Hoye, T. R. Nature 2013, 501, 531-534. (i) Wu C. and Shi, F. Asian J. Org. Chem. 2013, 2, 116-125. (j) Baire, B.; Niu, D.; Willoughby, P. H.; Woods, B. P. and Hoye, T. R. Nat. Protoc. 2013, 8, 501-508. (k) Niu, D. and Hoye, T. R. Nat. Chem. 2014, 6, 34-40. 4. Ihara, E.; Kurokawa, A.; Koda, T.; Muraki, T.; Itoh, T. and Inoue, K. Macromolecules 2005, 38, 2167-2172. 5. Ito, S.; Takahashi, K. and Nozaki, K. J. Am. Chem. Soc. 2014, 136, 7547-7550. 6. Himeshima, Y.; Sonoda, T. and Kobayashi, H. Chem. Lett. 1983, 12, 1211-1214. 7. Yoshikai, N. and Nakamura, E. Chem. Rev. 2012, 112, 2339-2372. 8. The initiating/end groups shown in Scheme 1 were estimated from the results of MALDI-TOF mass analysis. 9. Ando, S.; Ohta, E.; Kosaka, A.; Hashizume, D.; Koshino, H.; Fukushima, T. and Aida, T. J. Am. Chem. Soc. 2012, 134, 11084-11087. 10. These complicated signals are probably attributable to slow conformational exchange on the NMR time scale.9,14a 11. Frisch, M. J.; Trucks, G. W.; Schlegel, H. B.; Scuseria, G. E.; Robb, M. A.; Cheeseman, J. R.; Scalmani, G.; Barone, V.; Mennucci, B.; Petersson, G. A.; Nakatsuji, H.; Caricato, M.; Li, X.; Hratchian, H. P.; Izmaylov, A. F.; Bloino, J.; Zheng, G.; Sonnenberg, J. L.; Hada, M.; Ehara, M.; Toyota, K.; Fukuda, R.; Hasegawa, J.; Ishida, M.; Nakajima, T.; Honda, Y.; Kitao, O.; Nakai, H.; Vreven, T.; Montgomery, J. A., Jr.; Peralta, J. E.; Ogliaro, F.; Bearpark, M.; Heyd, J. J.; Brothers, E.; Kudin, K. N.; Staroverov, V. N.; Kobayashi, R.; Normand, J.; Raghavachari, K.; Rendell, A.; Burant, J. C.; Iyengar, S. S.; Tomasi, J.; Cossi, M.; Rega, N.; Millam, N. J.; Klene, M.; Knox, J. E.; Cross, J. B.; Bakken, V.; Adamo, C.; Jaramillo, J.; Gomperts, R.; Stratmann, R. E.; Yazyev, O.; Austin, A. J.; Cammi, R.; Pomelli, C.; Ochterski, J. W.; Martin, R. L.; Morokuma, K.; Zakrzewski, V. G.; Voth, G. A.; Salvador, P.; Dannenberg, J. J.; Dapprich, S.; Daniels, A. D.; Farkas, Ö.; Foresman, J. B.; Ortiz, J. V.; Cioslowski, J.; Fox, D. J. Gaussian 09, Revision D.01; Gaussian, Inc., Wallingford, CT, 2009. 12. Maeda, S.; Osada, Y.; Morokuma, K. and Ohno, K. GRRM11, Version 11.01, 2011 13. (a) Yamamoto, T. and Suginome, M. Angew. Chem. Int. Ed. 2009, 48, 539-542. (b) Nagata, Y.; Takagi, K. and Suginome, M. J. Am. Chem. Soc. 2014, 136, 9858-9861. 14. (a) He, J.; Crase, J. L.; Wadumethrige, S. H.; Thakur, K.; Dai, L.; Zou, S.; Rathore, R. and Hartley, C. S. J. Am. Chem. Soc. 2010, 132, 13848-13857. (b) Ohta, E.; Sato, H.; Ando, S.; Kosaka, A.; Fukushima, T.; Hashizume, D.; Yamasaki, M.; Hasegawa, K.; Muraoka, A.; Ushiyama, H.; Yamashita, K. and Aida, T. Nat. Chem. 2011, 3, 68-73. (c) Kajitani, T.; Suna, Y.; Kosaka, A.; Osawa, T.; Fujikawa, S.; Takata, M.; Fukushima, T. and Aida, T. J. Am. Chem. Soc. 2013, 135, 14564-14567. 15. (a) Berresheim, A. J.; Müller, M. and Müllen, K. Chem. Rev. 1999, 99, 1747-1785. (b) Miwa, K.; Furusho, Y. and Yashima, E. Nat. Chem. 2010, 2, 444-449. 16. (a) Kovacic, P.; Uchic, J. T. and Hsu, L.-C. J. Polym. Sci., Part A: Polym. Chem. 1967, 5, 945–964. (b) Kovacic, P. and Ramsey, J. S. J. Polym. Sci., Part A: Polym. Chem. 1969, 7, 111–125. (c) Hsing, C.-F.; Jones, M. B. and Kovacic, P. J. Polym. Sci., Part A: Polym. Chem. 1981, 19, 973–984.

ACS Paragon Plus Environment

Page 5 of 5

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

Journal of the American Chemical Society

      Ar “Aryne” Since 1902

“Cu(I)” Ar First polymerization n Since 2014 Poly(ortho-arylene) up to ca. 100 mer size  

Table of Contents artwork

ACS Paragon Plus Environment