Next Article in Journal
Dichlorido(η6-p-cymene)[tris(2-cyanoethyl)phosphine]ruthenium(II)
Previous Article in Journal
1-Methyl-3-{4-[(4-(2-oxo-2,3-dihydro-1H-benzimidazol-1-yl)piperidin-1-yl)benzyl]}-2-phenylindole
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Purine-Furan and Purine-Thiophene Conjugates

Faculty of Materials Science and Applied Chemistry, Riga Technical University, P. Valdena Str. 3, LV-1048 Riga, Latvia
*
Author to whom correspondence should be addressed.
Molbank 2018, 2018(4), M1024; https://doi.org/10.3390/M1024
Submission received: 14 September 2018 / Revised: 4 October 2018 / Accepted: 5 October 2018 / Published: 8 October 2018
(This article belongs to the Section Organic Synthesis)

Abstract

:
Furyl and thienyl moieties were introduced into a purine structure to elevate its fluorescence properties, while a trityl group was used to increase the amorphous properties of the purine compounds. The title compounds were prepared by a sequence involving a Mitsunobu, a SNAr and a Suzuki–Miyaura reaction and their photophysical properties were studied. Quantum yields in the solution reached up to 88% but only up to 5% in the thin layer.

Graphical Abstract

1. Introduction

The synthesis and development of novel push–pull purine derivatives with potential as new materials and/or heavy metal sensors are currently in high demand [1]. Purines containing five-membered heterocycles in their structure show fluorescence with good quantum yields in the solution [2,3]. In 2017, functionalized purine chromophores were developed using a Stille cross-coupling between 6-bromopurine and distannyl π-linkers of benzodithiophene, thiophene, or dithienylbenzothiadiazole [4]. These chromophores showed high thermal stability, long excited state lifetimes and high quantum yields, which will permit the use of such purine derivatives in material chemistry in the future.
We have developed novel purine derivatives, containing a group at N(9) to enhance their amorphous properties [5], for study a fluorescent materials in thin films. Our approach for building push–pull structures is based on the introduction of electron donating piperidinyl groups at C(2) and C(6) of the purine cycle and by extending the conjugation through introduction of a furan or thiophene at C(8).
Various approaches have been used for the introduction of thienyl and furyl substituents at the purine C(8) position. Ozola and co-workers used palladium-catalyzed Stille cross-coupling reactions in the presence of CuO to install the 2- and 3-furanyl rings [6]. The Sedlaček group also successfully introduced a 2-thienyl group to the 9-substituted 2,6-diaminopurine using a Stille cross-coupling and a 3-thienyl group using a Suzuki–Miyaura reaction [7]. 2-Iodothiophene was used for direct cross-coupling at the purine C(8) position in the presence of CuI and Pd(OAc)2 giving the 8-substitued product in 15% yield [8].

2. Results and Discussion

The Mitsunobu reaction between 2,6-dichloropurine 1 and 2-hydroxyethyl 3,3,3-triphenylpropanoate 2 led to the C(9)-substituted 2,6-dichloropurine 3 (Scheme 1) which was used as a starting material in subsequent steps.
In the SNAr reaction of 2,6-dichloropurine 3 with piperidine, the most reactive chlorine atom at C(6) was replaced followed by the less active C(2) chlorine. A complete conversion to 2-chloro-6-piperidinylpurine derivative was observed by HPLC over a period of 15 min, followed by the slower substitution at C(2) of the purine giving product 4 in 56% yield. In isopropanol, this reaction runs without any significant formation of side-products.
The 8-bromo derivative 5 was obtained by bromination of the 8-position of the purine ring in 71% yield. Subsequently, the Suzuki–Miyaura reaction with the furyl- and thienylboronic acids resulted in the 2,6,8-tri-substituted purine derivatives 6ac in 50–63% yields (Scheme 2).
Fluorescence quantum yields were measured for compounds 6ac in solution (DCM) and as thin films. The quantum yields were much lower in the films than in solution. Compound 6c exhibited the highest quantum yield (0.88, solution) among the three synthesized compounds. In the case of compounds 6b and 6c, there is a significant drop in fluorescence quantum yields in the thin film in comparison to the solution, from 0.60 and 0.88 to 0.04 and 0.05, respectively (Table 1). Typically, amorphous thin films are characterized by a significant degree of disorder which might result in the concentration induced or aggregation induced fluorescence self-quenching [9,10,11,12,13,14,15,16]. Compounds 6ac exhibited an absorption maximum around 320–350 nm and an emission maximum around 380–450 nm when excited at 320–350 nm (Figure 1).

3. Materials and Methods

1H- and 13C-NMR spectra were recorded at 300 and 75.5 MHz, respectively. The proton signals for residual non-deuterated solvents (δ 7.26 for CDCl3 and δ 2.50 for DMSO-d6) and carbon signals (δ 77.1 for CDCl3 and δ 39.5 for DMSO-d6) were used as internal references for 1H- and 13C-NMR spectra, respectively (see Supplementary Materials). Coupling constants are reported in Hz. Infrared spectra were recorded using a Perkin Elmer Spectrum BX spectrometer (PerkinElmer, Inc., Hebron, KY, USA). Analytical thin layer chromatography (TLC) was performed on Merck silica gel 60 F254 aluminum plates precoated with a 0.25 mm layer of silica gel. For HPLC analyses Agilent Technologies 1200 Series system (Agilent Technologies, Foster City, CA, USA) was used (XBridge C18 column, 4.6 × 150 mm, particle size 3.5 µm) with a flow rate of 1 mL/min; eluent system: 0.01% TFA water solution/MeCN (95:5, v/v). The content of MeCN was changed as follows: 20–95–95–20% (0–5–10–12 min). The wavelength of detection was 260 nm. The UV–vis absorption spectra of compounds were acquired using a Perkin-Elmer 35 UV–vis spectrometer. Emission spectra were measured on a QuantaMaster 40 steady state spectrofluorometer (Photon Technology International, Inc., Ford, West Sussex, UK). Absolute photoluminescence quantum yields were determined using a QuantaMaster 40 steady state spectrofluorometer (Photon Technology International, Inc.) equipped with a 6-inch integrating sphere by LabSphere, using a florescence standard of quinine sulfate in 0.1 M H2SO4 as the reference.
Molbank 2018 m1024 i001
2-(2,6-Dichloro-9H-purin-9-yl)ethyl 3,3,3-triphenylpropanoate (3); To a solution of 2,6-dichloropurine 1 (1.00 g, 5.26 mmol) in THF (15 mL), compound 2 (2.00 g, 5.78 mmol), Ph3P (1.66 g, 6.31 mmol) and DIAD (0.41 mL, ρ = 1.03 g/cm3, 6.31 mmol) were added. The resulting solution was stirred at 20 °C for 12 h. The reaction mixture was filtered, washed with cold MeOH (2 × 5 mL) and dried under reduced pressure. Yield: 2.39 g, 87%. Colorless powder, Rf = 0.28 (DCM/MeCN = 20:1). HPLC: tR = 7.31 min, purity 93%. IR (KBr) ν (cm−1): 2928, 1746, 1593, 1551, 1343, 1232, 1142. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.68 (s, 1H, H-C(8)), 7.30–7.12 (m, 15H, 15 × H-C(Ph)), 4.22–4.10 (m, 4H, 2 × (-CH2-)), 3.71 (s, 2H, (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.4, 153.2, 153.1, 152.0, 146.2, 146.1, 130.8, 129.1, 128.1, 126.6, 61.7, 55.9, 46.0, 43.3. HRMS (ESI): calcd for [C28H22Cl2N4O2 + H]+ 517.1193, found 517.1197.
Molbank 2018 m1024 i002
2-[2,6-Di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (4); To a solution of compound 3 (1.50 g, 3.25 mmol) in i-PrOH (8 mL) piperidine (1.52 mL, 15.60 mmol) was added. The resulting reaction mixture was stirred at 100 °C for 48 h, evaporated under reduced pressure and purified by silica gel column chromatography (DCM/MeCN, gradient 0–3%). Yield: 300 mg, 42%. Colorless powder, Rf = 0.56 (DCM/MeCN = 20:1). HPLC: tR = 7.93 min, purity 90%. IR (KBr) ν (cm−1): 2931, 2851, 1742, 1595, 1567, 1482, 1444, 1316, 1245, 1206, 1142, 1022. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.31–7.16 (m, 16H, 15 × H-C(Ph), H-C(8)), 4.27–4.06 (m, 6H, 3 × (-CH2-)), 4.02–3.94 (m, 2H, (-CH2-)), 3.79–3.68 (m, 6H, 3 × (-CH2-)), 1.78–1.54 (m, 12H, 6 × (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.7, 158.9, 154.0, 153.1, 146.4, 135.8, 129.2, 127.8, 126.4, 113.7, 62.3, 55.9, 46.3, 46.2, 45.6, 41.9, 26.2, 25.9, 25.1, 25.0. HRMS (ESI): calcd for [C38H42N6O2 + H]+ 615.3442, found 615.3426.
Molbank 2018 m1024 i003
2-[8-Bromo-2,6-di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (5); To a solution of compound 4 (1.50 g, 3.25 mmol) in DCM (10 mL) bromine was added (2.60 mL, 32.50 mmol). The resulting reaction mixture was stirred at 23 °C for 1 h, evaporated under reduced pressure and purified by silica gel column chromatography (DCM/MeCN, gradient 0–2%). Yield: 200 mg, 71%. Colorless powder, Rf = 0.56 (DCM/MeCN = 20:1). HPLC: tR = 11.79 min, purity 97%. IR (KBr) ν (cm−1): 2931, 2851, 1743, 1596, 1568, 1482, 1444, 1311, 1245, 1213, 1142, 1023. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.33–7.17 (m, 15H, 15 × H-C(Ph)), 4.22–3.98 (m, 8H, 4 × (-CH2-)), 3.77–3.63 (m, 6H, 3 × (-CH2-)), 1.77–1.51 (m, 12H, 6 × (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.7, 158.2, 152.8, 146.5, 129.3, 127.9, 126.4, 120.4, 114.3, 67.2, 61.6, 55.9, 46.3 (2C) (determined by the H-C HSQC spectrum), 45.7, 42.7, 29.8, 26.2, 25.9, 25.1, 25.0. HRMS (ESI): calcd for [C38H41N6O2Br + H]+ 695.2534, found 695.2527.

General Procedure for the Suzuki–Miyaura Reaction

Molbank 2018 m1024 i004
2-[8-(Furan-3-yl)-2,6-di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6a); To a solution of compound 5 (500 mg, 0.72 mmol) in anhydrous toluene (10 mL) 3-furanylboronic acid (161 mg, 1.40 mmol), K2CO3 (200 mg, 1.40 mmol), and Pd(PPh3)4 (41 mg, 0.08 mmol) were added. The resulting reaction mixture was stirred for 4 h at 110 °C, then evaporated to dryness. The residue was dissolved in DCM (20 mL), washed with saturated aqueous NaHCO3 (5 mL) and water (5 mL). The organic layer was dried over anhyd. Na2SO4, evaporated under reduced pressure and purified by silica gel column chromatography (DCM/MeCN, gradient 0–2%). Yield: 300 mg, 52%. Yellow powder, Rf = 0.55 (DCM/MeCN = 9:1). HPLC: tR = 8.83 min, purity 91%. IR (KBr) ν (cm−1): 2929, 2850, 1743, 1596, 1563, 1480, 1443, 1314, 1208, 1142, 1022. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.79 (s, 1H, H-C(furyl)), 7.49 (s, 1H, H-C(furyl)), 7.34–7.11 (m, 15H, H-C(Ph)), 6.84 (s, 1H, H-C(furyl)), 4.29–4.02 (m, 8H, 4 × (-CH2-)), 3.82–3.65 (m, 4H, 2 × (-CH2-)), 3.63 (s, 2H, (-CH2-)), 1.80–1.52 (m, 12H, 6 × (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.7, 158.7, 154.8, 153.6, 146.3, 143.3, 141.1, 139.1, 129.2, 127.9, 126.3, 117.2, 113.7, 110.7, 61.7, 55.9, 46.2, 46.1, 45.6, 41.3, 26.2, 25.9, 25.2, 25.1. HRMS (ESI): calcd for [C42H44N6O3 + H]+ 681.3548, found 681.3545.
Molbank 2018 m1024 i005
2-[2,6-Di(piperidin-1-yl)-8-(3-thienyl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6b); Product 6b was obtained by the general synthetic procedure for the Suzuki–Miyaura reaction: compound 5 (500 mg, 0.96 mmol), 3-thienylboronic acid (113 mg, 1.00 mmol), K2CO3 (200 mg, 1.40 mmol) and Pd(PPh3)4 (55 mg, 0.05 mmol). The reaction mixture was stirred for 4 h at 110 °C and the crude product was purified by silica gel column chromatography (DCM/MeCN, gradient 0–3%). Yield: 320 mg, 63%. Colorless powder, Rf = 0.62 (DCM/MeCN = 20:1). HPLC: tR = 9.14 min, purity 95%. IR (KBr) ν (cm−1): 2930, 2850, 1743, 1595, 1580, 1548, 1480, 1443, 1313, 1209, 1142, 1087. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.59 (d, 1H, 4J = 2.8 Hz, H-C(thienyl)), 7.53 (d, 1H, 3J = 4.6 Hz, H-C(thienyl)), 7.39 (dd, 1H, 4J = 2.8 Hz, 3J = 4.6 Hz H-C(thienyl)), 7.30–7.13 (m, 15H, 15 × H-C(Ph)), 4.30–4.09 (m, 8H, 4 × (-CH2-)), 3.83–3.71 (m, 4H, 2 × (-CH2-)), 3.59 (s, 2H, (-CH2-)), 1.79–1.56 (m, 12H, 6 × (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.7, 158.7, 154.8, 153.8, 146.4, 141.8, 131.7, 129.2, 128.5, 127.9, 126.4, 126.0, 124.5, 113.6, 61.9, 55.9, 46.1, 46.1, 45.7, 41.6, 26.3, 25.9, 25.2, 25.1. HRMS (ESI): calcd for [C42H44N6O2S + H]+ 697.3319, found 697.3332.
Molbank 2018 m1024 i006
2-[2,6-Di(piperidin-1-yl)-8-(2-thienyl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6c); Product 6c was obtained by the general synthetic procedure for the Suzuki–Miyaura reaction: compound 5 (600 mg, 0.86 mmol), 2-thienylboronic acid (220 mg, 1.70 mmol), K2CO3 (238 mg, 1.70 mmol) and Pd(PPh3)4 (50 mg, 0.08 mmol). The reaction mixture was stirred for 4 h at 110 °C and the crude product was purified by silica gel column chromatography (DCM/MeCN, gradient 0–1%). Yield: 300 mg, 50%. Colorless powder, Rf = 0.60 (DCM/MeCN = 9:1). HPLC: tR = 9.63 min, purity 94%. IR (KBr) ν (cm−1): 2928, 2842, 1742, 1594, 1577, 1545, 1481, 1442, 1314, 1243, 1206, 1141, 1023. 1H-NMR (300 MHz, CDCl3) δ (ppm): 7.39–7.32 (m, 2H, H-C(thienyl)), 7.30–7.13 (m, 15H, 15 × H-C(Ph),), 7.41–7.35 (m, 1H, H-C(thienyl)), 4.35–4.10 (m, 8H, 4 × (-CH2-)), 3.82–4.69 (m, 4H, 2 × (-CH2-)), 3.59 (s, 2H, (-CH2-)), 1.79–1.53 (m, 12H, 6 × (-CH2-)). 13C-NMR (75.5 MHz, CDCl3) δ (ppm): 170.6, 158.7, 155.0, 153.6, 146.4, 140.0, 133.3, 129.2, 127.8, 127.7, 127.1, 126.4, 126.3, 113.8, 61.8, 55.8, 46.2, 46.1, 45.6, 41.6, 26.2, 25.9, 25.1, 25.1. HRMS (ESI): calcd for [C42H44N6O2S + H]+ 697.3319, found 697.3319.

4. Conclusions

A method for the synthesis of purine derivatives modified with furan and thiophene heterocycles at C(8) has been developed. The key step was a Suzuki–Miyaura reaction and the products were obtained in 50–63% yields.
Target compounds exhibited strong fluorescence in solution with emission maxima at 380–480 nm. The fluorescence quantum yields in DCM solution reached up to 88% and in the thin layer films up to 5%.

Supplementary Materials

The following are available online at https://www.mdpi.com/1422-8599/2018/4/M1024/s1. 1H- and 13C-NMR spectra of 2-(2,6-dichloro-9H-purin-9-yl)ethyl 3,3,3-triphenylpropanoate (3), 2-[2,6-di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (4), 2-[8-bromo-2,6-di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (5), 2-[8-(furan-3-yl)-2,6-di(piperidin-1-yl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6a), 2-[2,6-di(piperidin-1-yl)-8-(3-thienyl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6b) and 2-[2,6-di(piperidin-1-yl)-8-(2-thienyl)-9H-purin-9-yl]ethyl 3,3,3-triphenylpropanoate (6c); IR spectra of compounds 3, 4, 5, 6a, 6b, and 6c; results tables of absorption and emission spectra of compounds 6a, 6b, and 6c in DCM solution and in thin film.

Author Contributions

M.T. designed the experiments; Z.K. performed the experiments; Z.K., I.N., and M.T. analyzed the IR, HRMS, and NMR spectral data and wrote the manuscript. All authors read and approved the final manuscript.

Funding

This research was funded by ERDF activity project no 1.1.1.1/16/A/131 “Design and Investigation of Light Emitting and Solution Processable Organic Molecular Glasses”.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Yang, Y.; Cohn, P.; Eom, S.H.; Abboud, K.A.; Castellano, R.K.; Xue, J. Ultraviolet-Violet Electroluminescence from Highly Fluorescent Purines. J. Mater. Chem. C 2013, 1, 2867–2874. [Google Scholar] [CrossRef]
  2. Greco, N.J.; Tor, Y. Furan Decorated Nucleoside Analogues as Fluorescent Probes: Synthesis, Photophysical Evaluation and Site-Specific Incorporation. Tetrahedron 2007, 63, 3515–3527. [Google Scholar] [CrossRef] [PubMed]
  3. Dumas, A.; Luedtke, N.W. Site-Specific Control of N7-Metal Coordination in DNA by a Fluorescent Purine Derivative. Chem. Eur. J. 2012, 18, 245–254. [Google Scholar] [CrossRef] [PubMed]
  4. Collier, G.S.; Brown, L.A.; Boone, E.S.; Kaushal, M.; Ericson, M.N.; Walter, M.G.; Long, B.K.; Kilbey, S.M. Linking Design and Properties of Purine-Based Donor-Acceptor Chromophores as Optoelectronic Materials. J. Mater. Chem. C 2017, 5, 6891–6898. [Google Scholar] [CrossRef]
  5. Traskovskis, K.; Mihailovs, I.; Tokmakovs, A.; Jurgis, A.; Kokars, V.; Rutkis, M. Triphenyl Moieties as Building Blocks for Obtaining Molecular Glasses with Nonlinear Optical Activity. J. Mater. Chem. 2012, 22, 11268–11276. [Google Scholar] [CrossRef]
  6. Ozola, V.; Persson, T.; Gronowitz, S.; Hӧrnfeldt, A.-B. On the Syntheses of 8-Heteroaryl-Substituted 9-(β-d-Ribofuranosyl)-2,6-Diaminopurines through Pd-Catalyzed Coupling in the Presence of Cupric Oxide. J. Heterocycl. Chem. 1995, 32, 863–866. [Google Scholar] [CrossRef]
  7. Sedláček, O.; Břehová, P.; Pohl, R.; Holý, A.; Janeba, Z. The Synthesis of the 8- C-Substituted 2,6-Diamino-9-[2-(phosphonomethoxy)ethyl]purine (PMEDAP) Derivatives by Diverse Cross-Coupling Reactions. Can. J. Chem. 2011, 89, 488–498. [Google Scholar] [CrossRef]
  8. Vaňková, B.; Krchňák, V.; Soural, M.; Hlavác, J. Direct C-H Arylation of Purine on Solid Phase and Its Use for Chemical Libraries Synthesis. ACS Comb. Sci. 2011, 13, 496–500. [Google Scholar] [CrossRef] [PubMed]
  9. Mikhnenko, O.V.; Blom, P.W.M.; Nguyen, T.-Q. Exciton Diffusion in Organic Semiconductors. Energy Environ. Sci. 2015, 8, 1867–1888. [Google Scholar] [CrossRef]
  10. Kobayashi, H.; Sasaki, M.; Ohsawa, N.; Yasuda, K.; Kotani, M. Fluorescence and Its Density-Dependent Quenching of a Sub-Monolayer Film of Methylene Blue Prepared by Dip Coating. J. Phys. Chem. C 2007, 111, 268–271. [Google Scholar] [CrossRef]
  11. Penzkofer, A.; Lu, Y. Fluorescence Quenching of Rhodamine 6G in Methanol at High Concentration. Chem. Phys. 1986, 103, 399–405. [Google Scholar] [CrossRef]
  12. Musser, A.J.; Rajendran, S.K.; Georgiou, K.; Gai, L.; Grant, R.T.; Shen, Z.; Cavazzini, M.; Ruseckas, A.; Turnbull, G.A.; Samuel, I.D.W.; et al. Intermolecular States in Organic Dye Dispersions: Excimers: Vs. aggregates. J. Mater. Chem. C 2017, 5, 8380–8389. [Google Scholar] [CrossRef]
  13. Jenekhe, S.A.; Osaheni, J.A. Excimers and Exciplexes of Conjugated Polymers. Science 1994, 265, 765–768. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Currie, M.J.; Mapel, J.K.; Heidel, T.D.; Goffri, S.; Baldo, M.A. High-Efficiency Organic Solar Concentrators for Photovoltaics. Science 2008, 321, 226–228. [Google Scholar] [CrossRef] [PubMed]
  15. Jiang, Z.C.; Lin, T.N.; Lin, H.T.; Talite, M.J.; Tzeng, T.T.; Hsu, C.L.; Chiu, K.P.; Lin, C.A.J.; Shen, J.L.; Yuan, C.T. A Facile and Low-Cost Method to Enhance the Internal Quantum Yield and External Light-Extraction Efficiency for Flexible Light-Emitting Carbon-Dot Films. Sci. Rep. 2016, 6, 1–6. [Google Scholar] [CrossRef] [PubMed]
  16. Banal, J.L.; White, J.M.; Ghiggino, K.P.; Wong, W.W.H. Concentrating Aggregation-Induced Fluorescence in Planar Waveguides: A Proof-of-Principle. Sci. Rep. 2014, 4, 1–5. [Google Scholar] [CrossRef] [PubMed]
Scheme 1. The Mitsunobu reaction.
Scheme 1. The Mitsunobu reaction.
Molbank 2018 m1024 sch001
Scheme 2. Synthesis of 8-furyl- and 8-thienylpurine derivatives.
Scheme 2. Synthesis of 8-furyl- and 8-thienylpurine derivatives.
Molbank 2018 m1024 sch002
Figure 1. (a) Absorption and (b) emission spectra for 0.5 × 10−5 M 6ac in DCM.
Figure 1. (a) Absorption and (b) emission spectra for 0.5 × 10−5 M 6ac in DCM.
Molbank 2018 m1024 g001
Table 1. Photophysical properties of target compounds 6ac.
Table 1. Photophysical properties of target compounds 6ac.
CompoundSolvent/Thin Layerλabs, nmlog ελem, nmQY
6aDCM3234.43800.18
Thin layer326-382<0.01
6bDCM3334.34080.60
Thin layer336-4150.04
6cDCM3534.34490.88
Thin layer356-4450.05

Share and Cite

MDPI and ACS Style

Kapilinskis, Z.; Novosjolova, I.; Turks, M. Purine-Furan and Purine-Thiophene Conjugates. Molbank 2018, 2018, M1024. https://doi.org/10.3390/M1024

AMA Style

Kapilinskis Z, Novosjolova I, Turks M. Purine-Furan and Purine-Thiophene Conjugates. Molbank. 2018; 2018(4):M1024. https://doi.org/10.3390/M1024

Chicago/Turabian Style

Kapilinskis, Zigfrīds, Irina Novosjolova, and Māris Turks. 2018. "Purine-Furan and Purine-Thiophene Conjugates" Molbank 2018, no. 4: M1024. https://doi.org/10.3390/M1024

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop