• Aucun résultat trouvé

Naphthalenediimides with Cyclic Oligochalcogenides in Their Core

N/A
N/A
Protected

Academic year: 2022

Partager "Naphthalenediimides with Cyclic Oligochalcogenides in Their Core"

Copied!
17
0
0

Texte intégral

(1)

Article

Reference

Naphthalenediimides with Cyclic Oligochalcogenides in Their Core

SHYBEKA, Inga, et al.

Abstract

Naphthalenediimides (NDIs) are privileged scaffolds par excellence, of use in functional systems from catalysts to ion channels, photosystems, sensors, ordered matter in all forms, tubes, knots, stacks, sheets, vesicles, and colored over the full visible range. Despite this extensively explored chemical space, there is still room to discover core‐substituted NDIs with fundamentally new properties: NDIs with cyclic trisulfides (i.e., trisulfanes) in their core absorb at 668 nm, emit at 801 nm, and contract into disulfides (i.e., dithietes) upon irradiation at

SHYBEKA, Inga, et al. Naphthalenediimides with Cyclic Oligochalcogenides in Their Core.

Chemistry - A European Journal, 2020, vol. 26, no. 62, p. 14059-14063

DOI : 10.1002/chem.202003550

Available at:

http://archive-ouverte.unige.ch/unige:144761

Disclaimer: layout of this document may differ from the published version.

1 / 1

(2)

1

Naphthalenediimides with Cyclic Oligochalcogenides in Their Core

Inga Shybeka,[a]# Alexander Aster,[a]# Yangyang Cheng,[a] Naomi Sakai,[a] Antonio Frontera,[b]

Eric Vauthey[a]* and Stefan Matile[a]*

[a] I. Shybeka, A. Aster, Dr. Y. Cheng, Dr. N. Sakai, Prof. E. Vauthey, Prof. S. Matile School of Chemistry and Biochemistry

University of Geneva, Geneva, Switzerland

E-mail: stefan.matile@unige.ch, eric.vauthey@unige.ch

[b] Prof. A. Frontera, Department de Química, Universitat de les Illes Balears, Palma de Mallorca, Spain

# These authors contributed equally

Supporting information for this article is given via a link at the end of the document.

Abstract: Naphthalenediimides (NDIs) are privileged scaffolds par excellence, of use in functional systems from catalysts to ion channels, photosystems, sensors, ordered matter in all forms, tubes, knots, stacks, sheets, vesicles, and colored over the full visible range. Despite this extensively explored chemical space, there is still room to discover core-substituted NDIs with fundamentally new properties: NDIs with cyclic trisulfides (i.e., trisulfanes) in their core absorb at 668 nm, emit at 800 nm, and contract into disulfides (i.e., dithietes) upon irradiation at

< 475 nm. Intramolecular 1,5-chalcogen bonds account for record red shifts with trisulfides, ring-tension mediated chalcogen bond cleavage for blue shifts to 492 nm upon ring contraction.

(3)

2

Cyclic oligochalcogenides (COCs) in the NDI core open faster than strained dithiolanes as in asparagusic acid and are much better retained on thiol exchange affinity columns. This makes COC-NDIs attractive not only within the existing multifunctionality, particularly artificial photosystems, but also for thiol-mediated cellular uptake.

The chemistry of core-substituted NDIs is most colorful.[1–10] Single-atom substitutions already suffice to cover the primary colors.[1,2] In this study, cyclic oligochalcogenides (COCs)[11,12] 1–3 are added to this rich collection (e.g., 4–17, Figure 1). NDI 1 has cyclic trisulfides in the core.

The respective disubstituted benzobis(trisulfane)s have been reported, including absorption around 300 nm and reversible one-electron oxidation waves observed by the cyclic voltammetry (CV).[13] In NDI 2, one trisulfide is replaced by a strained cyclic disulfide, in 3 both. The analogous 1,2-benzodithietes need bulky and electron withdrawing substituents for stabilization, tend to dimerize and might equilibrate into retrocyclized 1,2-dithiones.[14]

COC-NDIs 1-3 were readily accessible from the known[9] NDIs 4 with four sulfides in the core (Figure 1, Schemes S1, S2). With tert-butyl sulfides (and solubilizing triethyleneglycol (TEG) tails as imide substituents), reaction with S2Cl2 at room temperature gave NDI-bis(trisulfane) 1 in 76% yield. NDI 1 is stable in open air and ambient light, ongoing variations confirm that the presence of two TEG tails is not essential. NDI-bis(dithiete) 3 and intermediate 2 could be isolated in low yields from a mixture of mostly polymers obtained by a brief exposure of 1 to tris(2-carboxyethyl)phosphine (TCEP) under basic conditions.

NDI-bis(trisulfane) 1 was green, 2 blue, and 3 yellow (Figure 1, 2a). The absorption maximum at 668 nm (e = 12.3 mM–1 cm–1) was beyond the record 642 nm of NDI 6 with four benzylamines

(4)

3

Figure 1. COC-NDIs 1–3 (N-- = N(CH2CH2O)3CH3, TEG) compared to known NDIs 4–18, including, as available, absorption maxima (in nm), emission maxima and fluorescence quantum yields (in parenthesis), LUMO energies (from CV, relative to –5.1 eV for Fc+/Fc), and quadrupole moment Qzz (in Buckinghams, B). a) S2Cl2, THF, rt, 12 h, 76% (4: S-- = StBu). b) hn, lex = 340 nm, CH2Cl2, 4 h, quant . c) 1. TCEP, Et3N, H2O/CH2Cl2 1:1, 2. O2, rt, 5 min, 12% 3, 3% 2. d) 5 mM DTT, 10 mM HEPES/MeCN 7:3, pH 7.4, quant.

in the core (Figures 1, 2a).[1,2,7] This bathochromism was intriguing because all previous attempts to crack this record have failed (e.g., 7, 9),[10] excluding, of course, larger extensions of conjugation, including polymers, and any reduction or

λabs (nm)

N N

O O

O O

S S S S

N N

O

O O

O O

O

N N

O

O O

O S

NH

N N

O

O O

O HN

NH

N N

O O

O O

HN NH HN NH

N N

O O

O O

O O O O

N N

O

O O

O S

S

N N O

O O

O

O O

N N

O

O O O

O O O O

B B HO

OH

N N

O

O O

O

1 668 (801, 17%)

–4.16 +8.5 B

N N

O O

O SHO HS

HS SH 2 602 (750, 3%)

+8.1 B 3

492 (540, <1%)

–4.11 +7.8 B

1 6 7 6 642 (687, 13%)

–3.41 5

598 4

577 –3.96

13 544 16

419 –3.83

18 400 –4.00 +18.6 B 15 469 –3.89 +8.0 B

14 528 –3.93

8 620 –3.70 +2.3 B 10

580 8 5 2 9

10 4 12 11

a) d)

668

–4.16 –3.41 640 642

N N

O

O O

O Se

Se 11

553

N N

O O

O O

O O O OH

7 640 620

–3.70 598 602 615

580 553 577

552

–4.13 –3.96

N N

O

O O

O O

NH 12 552

9 615 ELUMO

(eV)

N N

O O

O S O SS

SSS N

N O

O O

O S SS

S S

N N

O O

O O

S S S S

b, c)

N N

O

O O

O P+

P 17

400 +

(5)

4

Figure 2. a) Absorption (solid), emission (dashed) and excitation (dotted) spectra of 1 (green), 2 (blue) and 3 (yellow) in CH2Cl2. b) Change of absorption upon photolysis in CH2Cl2 with irradiation at 340 nm (blue filled spectrum) at time steps illustrated in the inset. The concentration changes over time, shown in the inset, are obtained from the fit with a 1à2à3 sequential reaction model (grey dashed lines). c) Transient absorption spectra upon excitation at 650 and 490 nm of 1 (top) and 3 (bottom) in toluene, respectively.

oxidation.[15-17] For instance, two tertiary arylamine donors in the core, particularly larger dendrimers, produce maxima up to 830 nm because the extra aromatic rings are conjugated.[15]

(6)

5

Already radical anions of unsubstituted NDIs 18 are red and show several maxima,[3] mono- and di-reduced phosphonium NDIs 17 afford shifts up to 720 nm.[17]

Figure 3. a) PBE1PBE-D3/def2-TZVP minimized structures of 3’ (left), 2’ (middle) and 1’ (right, Me imide substituents) with 1,5 O-S distances and chalcogen-bond angles. The electron density at the OS bond critical point (BCP): 1’, ρ = 0.0300 a.u.; 2’, ρ = = 0.0301 a.u.; 3’, no BCP (hydrogen bonds: ρ = 0.002–0.04 a.u.[17]). b) Molecular electrostatic potential (MEP) surfaces (0.001 a.u. isosurface, positive maxima in kcal mol–1), and c) computed HOMO (bottom) and LUMO (top).

This new red shift record was achieved by focusing on chalcogen bonding rather than on “super donors.” CV placed the LUMO of 1 (–4.16 eV) below amines 6 (–3.41 eV), ethers 16 (–3.83 eV), sulfides 4 (–3.96 eV) and even unsubstituted NDI 18 (–4.00 eV, Figures 1, S2). The LUMO of green 1, blue 2 and yellow 3 was almost the same in experiment and theory (Figures 1, 3c, S2). Same for the computed quadrupole moments, identifying the green 1 (Qzz = +8.5 B) as π

b)

c) a)

3.0 Å 2.6 Å

critical point

+16.8 +29.7

+23.2 144º 164º

+24.9 +23.6

+29.3 +17.5

–3.56 eV –3.56 eV –3.59 eV

–6.69 eV –6.11 eV –5.86 eV

(7)

6

acid in the range of hexafluorobenzene or the yellow 15 but clearly stronger than the almost π- neutral blue diamino NDI 8 (Qzz = +2.3 B, Figure 1).[2,8]

This constantly low LUMO confirmed that the record red shift of 1 occurred without injecting electron density into the aromatic core. This seemingly impossible was achieved by shifting the HOMO from the aromatic core in 3 to the trisulfides in 1 (Figure 3c). Namely, four intramolecular 1,5 O-S chalcogen bonds[18] inject electron density from the imide carbonyl into the trisulfide (Figure 3a). As a result, the MEP surface does not show deep s holes on the electron-enriched trisulfide (+16.8 kcal mol–1) but slightly deepened π holes on the NDI surface instead (+24.9 kcal mol–1, Figures 3b, c). From this trisulfide centered HOMO, reversible single-electron oxidation became detectable at –5.8 eV in the CV (Figure S2).

In 3, ring tension[11,14] moves the sulfur atoms away from the imide oxygens, breaking the 1,5 O-S chalcogen bonds (Figure 3a). Interestingly, the O-S distances (3.2 Å) remained below the sum of the vdW radii (3.3 Å), but bond angles (144º) were far from ideal (180º), and the absence of bond path and critical point excluded the presence of a chalcogen bond.[19] These broken chalcogen bonds made the HOMO move back down to the NDI core (ca. –6.6 eV, undetectable by CV, Figure 3c), and deep s holes opened up on the disulfides (+29.7 kcal mol–1), while NDI π holes slightly decreased (+23.2 kcal mol–1, Figure 3b). The result is a hypsochromism of – 176 nm upon removal of two sulfur atoms, one from each ring in 1 (Figure 1). The absorption of 3 (492 nm, e = 11.1 mM–1cm–1) blue-shifted beyond 4 and 14 with 2-4 sulfides in their core.[9]

This supported that chalcogen bonds already account for the relative red shifts of sulfide and selenide NDIs 4, 10, 11[4] and 14 compared to their oxygen analogs 12, 15 and 16.

Intramolecular chalcogen bonds have been used previously to solve spectroscopic challenges, most recently in cascade switched mechanophores.[20]

(8)

7

The emission spectra of 1 and 2 were significantly narrower than the absorption spectra and exhibited an ~1300 cm–1 vibrational progression (Figure 2a). Quantum-chemical calculations predicted a shallow ground-state energy minimum along the dihedral angle between the S–C–

C–S and S–S–S plane (Figure S21). The negligible <0.5 kcal mol–1 barrier to the planar transition state enables the population of a wide range of conformers with differing S1←S0 gaps, leading to a broadening of the absorption band. If this structural coordinate is more restricted in the S1 state, a smaller conformational space is accessible and the emission spectrum is narrower.

The fluorescence quantum yield (QY) of 1 in CH2Cl2 was 17%, higher than that of the tetraamino NDI 6 (Figure 1). Ring contraction into 2 reduced the brightness to 3% and to even <0.1% for 3. The femtosecond transient absorption spectra of 1 and 3 revealed the reason for this loss (Figures 2c). Whereas the S1 state population of 1 decayed with a lifetime of 4.1 ns (Figure S5), the S1 state of 3 depopulated within the first 50 ps, as reflected by the decrease of the ground state bleach (GSB) between 400 and 490 nm, and of the excited state absorption (ESA) bands at 350, 560 and 1000 nm. After 50 ps, the transient spectra showed the spectral features of the T1 state, which is populated by intersystem crossing (ISC). The spectral shift of the intense ESA of 1 during the first 50 ps was attributed to structural relaxation and arises from different equilibrium geometries of the S1 and S0 states, in line with the above-mentioned narrowing of the emission spectrum (Figure 2c). The excited-state lifetime of 2 was 1.3 ns and therefore intermediate relative to 1 and 3 (Figure S7). Comparison of the QYs, triplet yields and S1

lifetimes of the three COC-NDIs (Table S3) revealed that both the internal conversion and the ISC rate constants increase by more than two orders of magnitude upon ring contraction from 1 to 3. These changes reflect the different nature of the S1 state as predicted by the quantum chemical calculations. The lowest energy transition for 1 and 2 can be classified as n→π* with

(9)

8

considerable charge transfer from the lone pairs of the S atoms to the NDI core (Figure 3c). In 3, the S0→S1 transition has π→π* character, whereas the second band around 458 nm is associated with the n→π* transition. The discrepancy between excitation and absorption spectrum of 1 and 2 (Figure 2a, dotted lines) illustrated that the higher energy π→π* states also feature an ultrafast deactivation channel such as the S1 state of 3.

In contrast to excitation at 650 and 532 nm, irradiation at l < 475 nm in CH2Cl2 with a LED lead to permanent changes of the absorption spectra of 1 and 2 (Figure 2b). These spectral changes could be modelled using the e (l) of 1, 2 and 3 and a consecutive reaction kinetic scheme (1à2à3). Fitting this model to the data yielded the concentration profiles depicted in the inset and quantum yields for the two photolysis steps QY1à2, QY2à3 of <1% (Figure 2b). Even though QY1à2 and QY2à3 were very small, the overall reaction yield was close to unity. If the photolysis was started with 2, isosbestic points could be observed upon irradiation for more than 20 hours, illustrating the photostability of the ensuing 3 (Figure S13).

The wavelength dependence of this photoreaction matched the deviation of the excitation and absorption spectra of 1 (Figure 2a). The ~50% difference between excitation and absorption spectra at 340 nm is much larger than the QY1à2, indicating that only a small fraction of the population undergoing the non-Kasha process[21] leads to ring contraction. This could be due to efficient reformation of the sulfide bond upon photocleavage, which has been invoked to account for the surprising intrinsic photostability of disulfide bridges in proteins.[22]

In neutral buffer, the absorption maximum of 1 further red shifted to 715 nm and broadened beyond 800 nm, possibly due to self-assembly (Figure 4a).[23,24] Ring opening of COC-NDIs with thiolates or TCEP shifted the maximum to 623 nm, corresponding to the first conjugate base of 5 (pKa = 4.2, at pH < 4.2: 598 nm, Figure 1). For trisulfide 1, ring opening by thiol-

(10)

9

trisulfide exchange with glutathione was 6 times faster than for the strained dithiolane of asparagusic acid but 20 times slower than for dithiete 3 (Figures S14-16, Table S5).

Computational simulations confirmed that thiol-di/trisulfide exchange occurs without transition state,[25] and that electrophile and leaving group inactivation by chalcogen bonding and charge repulsion account as much as reduced ring tension for the slower opening of 1 compared to 3 (Figures 4c, S23).

Figure 4. a) Absorption spectra of 50 µM 1 in 10 mM HEPES/MeCN 7:3 pH 7.4 with DTT (0, 50, 100, 150, 200, 250, 300, 350, 400, 450, 500 µM and 5 mM; green→blue). b) Thiol-exchange affinity column chromatogram of 1 in 10 mM Tris, 0.1 M NaCl, 1 mM EDTA, pH 7.5 with a 0-50 mM DTT gradient at t = 60–70 min (solid) and constant 50 mM DTT from t = 0 (dashed). c) PBE1PBE-D3/def2-TZVP minimized structures of reactive intermediates for trisulfide (left, +6 kcal mol–1) and disulfide (right) of 2’ reacting with methylthiolate.

Thiol-trisulfide exchange was further explored by thiol affinity chromatography. First 1 was passed through the column, then DTT was added to the mobile phase to release all retained material. The retention of 1 clearly exceeded that of asparagusic acid and was comparable to benzopentasulfanes (BPS, Figures 4b, solid, S17).[9] Comparable to BPS also was significant

c)

δ

δ δ

δ

δ δ

DTT DTT

0 50 100 150

t / min

400 600 800

λ / nm

b) a)

(11)

10

retention in the permanent presence of DTT in the mobile phase, which is not observed with asparagusic acid (Figure 4b, dashed). This exceptional retention was consistent with ultrafast ring opening followed by slow exchange of the resulting acylic di/trisulfides because charge repulsion from the proximal imide carbonyl inactivates the thiolate leaving group (Figure 4c).

NDIs have functioned as anion-π[5,26–28] catalysts,[25] ion channels,[28,29] sensors,[29] artificial photosystems,[6,8,30] semi-conductors,[6,31] G-quartet stabilizers,[32] and self-assembled matter in all variations, such as vesicles,[33] nanotubes,[34] mono- and multilayers on surfaces,[8,30,33]

dynamic covalent libraries, knotted molecular topologies or donor-acceptor stacks.[33–36] COC- NDIs will be of interest to explore within this existing functional space, particularly with regard to artificial photosystems.[8] Exceptional retention on thiol affinity columns also promises potential with regard to chromogenic cellular uptake,[11] that is to further expand the multifunctionality of this wonderful scaffold.[37,38]

Acknowledgements

We thank the NMR and MS platforms for services, and the University of Geneva, the Swiss National Centre of Competence in Research (NCCR) Chemical Biology, the NCCR Molecular Systems Engineering and the Swiss NSF for financial support. AF thanks the MINECO/AEI (project CTQ2017-85821-R, FEDER funds) for financial support.

Keywords: Naphthalenediimides • cyclic oligochalcogenides • chalcogen bonds • trisulfides • ring tension • ring contraction • photochemistry

[1] F. Würthner, S. Ahmed, C. Thalacker, T. Debaerdemaeker, Chem. Eur. J. 2002, 8, 4742–

4750.

[2] N. Sakai, J. Mareda, E. Vauthey, S. Matile, Chem. Commun. 2010, 46, 4225–4237.

(12)

11

[3] a) S. Maniam, H. F. Higginbotham, T. D. M. Bell, S. J. Langford, Chem. Eur. J. 2019, 25, 7044–7057; b) S. Kumar, J. Shukla, Y. Kumar, P. Mukhopadhyay, Org. Chem. Front. 2018, 5, 2254–2276; c) S.-L. Suraru, F. Würthner, Angew. Chem. Int. Ed. 2014, 53, 7428–7448; d) L. Rocard, A. Berezin, F. De Leo, D. Bonifazi, Angew. Chem. Int. Ed. 2015, 54, 15739–15743;

e) S. Kuila, K. V. Rao, S. Garain, P. K. Samanta, S. Das, S. K. Pati, M. Eswaramoorthy, S. J.

George, Angew. Chem. Int. Ed. 2018, 57, 17115–17119; f) S. Kuila, A. Ghorai, P. K.

Samanta, R. B. K. Siram, S. K. Pati, K. S. Narayan, S. J. George, Chem. Eur. J. 2019, 25, 16007–16011; g) B. Reiß, H.-A. Wagenknecht, Beilstein J. Org. Chem. 2019, 15, 2043–2051;

h) A. Weißenstein, V. Grande, C. R. Saha-Möller, F. Würthner, Org. Chem. Front. 2018, 5, 2641–2651; i) M. Al Kobaisi, S. V. Bhosale, K. Latham, A. M. Raynor, S. V. Bhosale, Chem.

Rev. 2016, 116, 11685–11796; j) N. Pearce, E. S. Davies, R. Horvath, C. R. Pfeiffer, X.-Z.

Sun, W. Lewis, J. McMaster, M. W. George, N. R. Champness, Phys. Chem. Chem. Phys.

2018, 20, 752–764; k) S. V. Bhosale, C. H. Jani, S. J. Langford, Chem. Soc. Rev. 2008, 37, 331–342.

[4] A. Błaszczyk, M. Fischer, C. von Hänisch, M. Mayor, Helv. Chim. Acta 2006, 89, 1986–2005.

[5] Y. Kumar, S. Kumar, K. Mandal, P. Mukhopadhyay, Angew. Chem. Int. Ed. 2018, 57, 16318–

16322.

[6] J. A. Letizia, M. R. Salata, C. M. Tribout, A. Facchetti, M. A. Ratner, T. J. Marks, J. Am. Chem.

Soc. 2008, 130, 9679–9694.

[7] C. Röger, F. Würthner, J. Org. Chem. 2007, 72, 8070–8075.

[8] R. S. K. Kishore, O. Kel, N. Banerji, D. Emery, G. Bollot, J. Mareda, A. Gomez-Casado, P.

Jonkheijm, J. Huskens, P. Maroni, M. Borkovec, T. Sawada, E. Vauthey, N. Sakai, S. Matile, J. Am. Chem. Soc. 2009, 131, 11106–11116.

(13)

12

[9] J. Míšek, A. Vargas Jentzsch, S. Sakurai, D. Emery, J. Mareda, S. Matile, Angew. Chem. Int.

Ed. 2010, 49, 7680–7683.

[10] A. Fin, I. Petkova, D. A. Doval, N. Sakai, E. Vauthey, S. Matile, Org. Biomol. Chem. 2011, 9, 8246–8252.

[11] Y. Cheng, L. Zong, J. López-Andarias, E. Bartolami, Y. Okamoto, T. R. Ward, N. Sakai, S.

Matile, Angew. Chem. Int. Ed. 2019, 58, 9522–9526.

[12] a) R. Steudel, Chem. Rev. 2002, 102, 3905–3946; b) C.-S. Jiang, W. E. G. Müller, H. C.

Schröder, Y.-W. Guo, Chem. Rev. 2012, 112, 2179–2207; c) A. Mahendran, A. Vuong, D.

Aebisher, Y. Gong, R. Bittman, G. Arthur, A. Kawamura, A. Greer, J. Org. Chem. 2010, 75, 5549–5557.

[13] a) S. Ogawa, T. Ohmiya, T. Kikuchi, A. Kawaguchi, S. Saito, A. Sai, N. Ohyama, Y. Kawai, S. Niizuma, S. Nakajo, T. Kimura, R. Sato, J. Organomet. Chem. 2000, 611, 136–145; b) T.

Kimura, S. Mizusawa, A. Yoneshima, S. Ito, K. Tsujimura, T. Yamashita, Y. Kawai, S. Ogawa, R. Sato, Bull. Chem. Soc. Jpn. 2002, 75, 2647–2653; c) B. L. Chenard, R. L. Harlow, A. L.

Johnson, S. A. Vladuchick, J. Am. Chem. Soc. 1985, 107, 3871–3879; d) T. Kimura, S.

Ogawa, R. Sato, Mini-Rev. Org. Chem. 2007, 4, 15–29.

[14] a) J. Nakayama, A. Ishii, in Adv. Heterocycl. Chem., Elsevier, 2000, pp. 221–284; b) T.

Shimizu, H. Murakami, Y. Kobayashi, K. Iwata, N. Kamigata, J. Org. Chem. 1998, 63, 8192–

8199; c) K. S. Choi, I. Akiyama, M. Hoshino, J. Nakayama, Bull. Chem. Soc. Jpn. 1993, 66, 623–629; d) R. B. Boar, D. W. Hawkins, J. F. McGhie, S. C. Misra, D. H. R. Barton, M. F. C.

Ladd, D. C. Povey, J. Chem. Soc. Chem. Commun. 1975, 756–756; e) V. K. Cherkasov, G.

A. Abakumov, G. K. Fukin, S. V. Klementyeva, V. A. Kuropatov, Chem. Eur. J. 2012, 18, 13821–13827; f) K. C. Nicolaou, S. A. DeFrees, C. K. Hwang, N. Stylianides, P. J. Carroll, J.

P. Snyder, J. Am. Chem. Soc. 1990, 112, 3029–3039; g) C. G. Krespan, B. C. McKusick, T.

(14)

13

L. Cairns, J. Am. Chem. Soc. 1960, 82, 1515–1516; h) N. L. Brock, A. Nikolay, J. S. Dickschat, Chem. Commun. 2014, 50, 5487; i) G. Frenking, Angew. Chem. Int. Ed. 1990, 29, 1410–

1412.

[15] a) J. Shukla, M. R. Ajayakumar, P. Mukhopadhyay, Org. Lett. 2018, 20, 7864–7868.

[16] b) X. Gao, C. Di, Y. Hu, X. Yang, H. Fan, F. Zhang, Y. Liu, H. Li, D. Zhu, J. Am. Chem. Soc.

2010, 132, 3697–3699; c) A. Insuasty, S. Maniam, S. J. Langford, Chem. Eur. J. 2019, 25, 7058–7073.

[17] S. Kumar, J. Shukla, K. Mandal, Y. Kumar, R. Prakash, P. Ram, P. Mukhopadhyay, Chem.

Sci. 2019, 10, 6482–6493.

[18] a) P. Scilabra, G. Terraneo, G. Resnati, Acc. Chem. Res. 2019, 52, 1313–1324; b) N. Biot, D. Bonifazi, Coord. Chem. Rev. 2020, 413, 213243; c) M. S. Taylor, Coord. Chem. Rev. 2020, 413, 213270; d) L. Vogel, P. Wonner, S. M. Huber, Angew. Chem. Int. Ed. 2019, 58, 1880–

1891; e) J. Y. C. Lim, I. Marques, A. L. Thompson, K. E. Christensen, V. Félix, P. D. Beer, J.

Am. Chem. Soc. 2017, 139, 3122–3133; f) A. Bauzá, T. J. Mooibroek, A. Frontera, ChemPhysChem 2015, 16, 2496–2517; g) H. Huang, L. Yang, A. Facchetti, T. J. Marks, Chem. Rev. 2017, 117, 10291–10318; h) L. Chen, J. Xiang, Y. Zhao, Q. Yan, J. Am. Chem.

Soc. 2018, 140, 7079–7082; i) R. Gleiter, G. Haberhauer, D. B. Werz, F. Rominger, C.

Bleiholder, Chem. Rev. 2018, 118, 2010–2041.

[19] C. B. Aakeroy, D. L. Bryce, G. R. Desiraju, A. Frontera, A. C. Legon, F. Nicotra, K. Rissanen, S. Scheiner, G. Terraneo, P. Metrangolo, G. Resnati, Pure Appl. Chem. 2019, 91, 1889–

1892.

[20] M. Macchione, A. Goujon, K. Strakova, H. V. Humeniuk, G. Licari, E. Tajkhorshid, N. Sakai, S. Matile, Angew. Chem. Int. Ed. 2019, 58, 15752–15756.

(15)

14

[21] O. Yushchenko, G. Licari, S. Mosquera-Vazquez, N. Sakai, S. Matile, E. Vauthey, J. Phys.

Chem. Lett. 2015, 6, 2096–2100.

[22] A. B. Stephansen, R. Y. Brogaard, T. S. Kuhlman, L. B. Klein, J. B. Christensen, T. I. Sølling, J. Am. Chem. Soc. 2012, 134, 20279–20281.

[23] Consistent with the literature,[24] additional peaks appearing in CV and DPV recorded at too high concentrations revealed similar self-assembly effects.

[24] a) A.-B. Bornhof, A. Bauzà, A. Aster, E. Vauthey, A. Frontera, N. Sakai, S. Matile, J. Am.

Chem. Soc. 2018, 140, 4884–4892; b) A.-J. Avestro, D. M. Gardner, N. A. Vermeulen, E. A.

Wilson, S. T. Schneebeli, A. C. Whalley, M. E. Belowich, R. Carmieli, M. R. Wasielewski, J.

F. Stoddart, Angew. Chem. Int. Ed. 2014, 53, 4442−4449.

[25] S. M. Bachrach, J. T. Woody, D. C. Mulhearn, J. Org. Chem. 2002, 67, 8983–8990.

[26] Y. Zhao, Y. Cotelle, L. Liu, J. López-Andarias, A.-B. Bornhof, M. Akamatsu, N. Sakai, S.

Matile, Acc. Chem. Res. 2018, 51, 2255–2263.

[27] a) L. Li, Y.-J. Hong, D.-Y. Chen, W.-C. Xiao, M.-J. Lin, Chem. Commun. 2019, 55, 2364–

2367; b) T. L. D. Tam, C. K. Ng, X. Lu, Z. L. Lim, J. Wu, Chem. Commun. 2018, 54, 7374–

7377.

[28] V. Gorteau, G. Bollot, J. Mareda, A. Perez-Velasco, S. Matile, J. Am. Chem. Soc. 2006, 128, 14788–14789.

[29] S. Hagihara, H. Tanaka, S. Matile, J. Am. Chem. Soc. 2008, 130, 5656–5657.

[30] N. Sakai, A. L. Sisson, T. Bürgi, S. Matile, J. Am. Chem. Soc. 2007, 129, 15758–15759.

[31] B. A. Jones, A. Facchetti, M. R. Wasielewski, T. J. Marks, J. Am. Chem. Soc. 2007, 129, 15259–15278.

[32] F. Doria, E. Salvati, L. Pompili, V. Pirota, C. D’Angelo, F. Manoli, M. Nadai, S. N. Richter, A. Biroccio, I. Manet, M. Freccero, Chem. Eur. J. 2019, 25, 11085–11097.

(16)

15

[33] M. R. Molla, D. Gehrig, L. Roy, V. Kamm, A. Paul, F. Laquai, S. Ghosh, Angew. Chem. Int.

Ed. 2014, 20, 760–771.

[34] a) M. Ji, M. L. Mason, D. A. Modarelli, J. R. Parquette, Chem. Sci. 2019, 10, 7868–7877; b) G. D. Pantoş, J.-L. Wietor, J. K. M. Sanders, Angew. Chem. Int. Ed. 2007, 46, 2238–2240.

[35] M. Hellstern, M. Gantenbein, L. L. Pleux, G. Puebla-Hellmann, E. Lörtscher, M. Mayor, Adv.

Mater. Interfaces 2019, 6, 1801917.

[36] a) F. B. L. Cougnon, J. K. M. Sanders, Acc. Chem. Res. 2012, 45, 2211–2221; b) B. A.

Ikkanda, S. A. Samuel, B. L. Iverson, J. Org. Chem. 2014, 79, 2029–2037; c) M. B. Avinash, T. Govindaraju, Adv. Funct. Mater. 2011, 21, 3875–3882; d) G. D. Pantos, Naphthalenediimide and Its Congeners: From Molecules to Materials, Royal Society Of Chemistry, London, 2017.

[37] Red-shifting cyclic trisulfides in the core was unique for NDIs: Reaction of the perylenediimides (PDIs) with four thiols in the core gave only the already known[38] bis- dithiano analog to 3 (Scheme S3). Disulfides are preferred with PDIs because the six- membered rings in the PDI core are nearly tension free, and stabilizing chalcogen-bond are missing in PDIs.

[38] a) R. A. Cormier, B. A. Gregg, RSC Adv. 2014, 4, 2368–2373; b) L. Chen, P. Xia, T. Du, Y.

Deng, Y. Xiao, Org. Lett. 2019, 21, 5529–5532; c) G. Li, D. Li, X. Liu, H. Xu, J. Zhang, S.

Wang, Z. Liu, B. Tang, Chem. Commun. 2019, 55, 9661–9664.

(17)

Références

Documents relatifs

It appears at first in a patent filled by Tonello [4] for a scheme named Cyclic Block Filtered MultiTone (CB-FMT) which corresponds to a cyclic transmultiplexer (TMUX) using a

We propose a branch and bound method that finds the schedule which minimizes the mean cycle time with respect to variations.. We show that the evaluation of a schedule can be

Section 3 reviews basic notions from Mumford’s theta theory such as theta groups, theta structures, symmetric and totally symmetric line bundles as well as the main tool

In particular, it has an undecidable satisfiability problem and the expressive power of the logic depends on whether simultaneous fixed point definitions are allowed (the

However, both the 13 C NMR and ATR-FTIR spectra of P2-2 indicate CvC double bond formation (Fig. S7†) which could be explained by the transfer reaction to the monomer at

Géométrie: reproduction de figure sur quadrillage (identique, rétrécissement, agrandissement, symétrie) Liste des épreuves : catégorie français. Épreuve rouge Épreuve verte

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

While it is difficult to reason about exactly how much more difficult analysis becomes as additional cores are included at the HI criticality level, we may make an assumption in