Next Article in Journal
Thiomonosaccharide Derivatives from D-Mannose
Previous Article in Journal
The Reactions of N,N′-Diphenyldithiomalonamide with Michael Acceptors
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Proceeding Paper

Synthesis and Photophysical Characterization of 2′-Aminochalcones †

1
LAQV-REQUIMTE, Campus de Santiago, Universidade de Aveiro, 3810-193 Aveiro, Portugal
2
CICECO-Aveiro Institute of Materials, Universidade de Aveiro, Campus de Santiago, 3810-193 Aveiro, Portugal
*
Author to whom correspondence should be addressed.
Presented at the 24th International Electronic Conference on Synthetic Organic Chemistry, 15 November–15 December 2020; Available online: https://ecsoc-24.sciforum.net/.
Chem. Proc. 2021, 3(1), 29; https://doi.org/10.3390/ecsoc-24-08301
Published: 14 November 2020

Abstract

:
Chalcones present biological activity and are, therefore, currently studied for their therapeutic potential. They have shown antitumor, anti-inflammatory, antifungal and antibacterial properties. These compounds occur in nature as secondary metabolites of plants and are precursors of flavonoid biosynthesis. Generally, they are not luminescent, but derivatives with some particular patterns of substitution can present fair quantum yields. Excited State Intramolecular Proton Transfer (ESIPT) is a particular case of tautomerization, and some molecules can exhibit ESIPT fluorescence if their structures incorporate an intramolecular hydrogen bonding interaction between a hydrogen donor (-OH or -NH) and a hydrogen acceptor. If the dye is fluorescent, emission from both tautomers may be observed, leading to a dual emission. The chalcones with a strong push–pull character, compounds 2c, 3c and 4c, presented dual emission of keto–enol tautomerism as a consequence of ESIPT.

Published: 14 November 2020

1. Introduction

Chalcone (1,3-diphenylprop-2-en-1-one) is an aromatic ketone, where the two aromatic nuclei are joined by an α,β-unsaturated carbonyl bridge, which forms the central nucleus of a variety of important biological compounds [1]. Chalcone exists as trans (E)- or cis (Z)-isomers, with the most stable and predominant stereoisomer being the (E)-isomer [2]. There are several methods for the synthesis of chalcones, with Claisen–Schmidt condensation being the classic method [3]. Chalcones and their derivatives have been extensively studied because of their wide range of applicability [4]. Chalcones have several applications such as anti-inflammatory, Reference [5] anti-HIV, Reference [6] antibacterial, Reference [7] anticancer, Reference [8] antituberculosis, Reference [9] antileishmanial, Reference [10] therapeutical activity, and Reference [11], among others.
Structural changes in chalcone derivatives can modify different optical properties such as refractive index, absorption magnitude and spectral position, as well as the optical energy gap [4]. These examples indicate that structurally modified chalcones can exhibit variable physical properties.
Within this class of compounds, 2’-aminochalcone derivatives still represent a recent field of investigation [12]. They have antitumor activity [13] and have been shown to be photoprotectors [14].
Generally, they are not luminescent, but derivatives with some particular patterns of substitution can present fair quantum yields [15].
Excited State Intramolecular Proton Transfer (ESIPT) is a particular case of tautomerization, and molecules can exhibit ESIPT fluorescence if their structures incorporate an intramolecular hydrogen bonding interaction between a hydrogen donor (-OH or -NH) and a hydrogen acceptor [16]. The intramolecular hydrogen bond plays a crucial role in the proton transfer process [17]. If the dye is fluorescent, emission from both tautomers may be observed, leading to a dual emission [18]. ESIPT fluorophores have some beneficial photophysical properties such as intense luminescence, photostability and an unusually large Stokes Shift [3]. The proton-transfer tautomer emits fluorescence at longer wavelengths and will result in larger Stokes shifts.
ESIPT is a system with four energy levels (L→L*→E*→E). The two forms of the chalcone (L and ESIPT form) are in equilibrium (Scheme 1). In the ground state, the chalcone is in the L form. After absorption of a photon, the L* suffers an ESIPT tautomerization equilibrium to the E*, and both tautomers can emit light. After relaxation to the E form, the chalcone goes back to the keto form through reverse proton transfer (RPT). This equilibrium may be influenced by the environment of the chalcone, especially the polarity and proticity of the solvent, leading to different proportion of both tautomers, and therefore to different emission intensities in the two bands when dual emission is observed. As such, these fluorophores may be found in such applications as molecular probes, giving an insight into the properties of their environment, including the polarity of an organelle.
Here, we report the synthesis and photophysical characterizations of 2′-aminochalcones, substituted with electron donating or withdrawing groups, which could be used as markers to study the environment of cellular organelles.

2. Experimental

2.1. General Procedure for the Synthesis of 2′-Aminochalcones 2a–c

The 2′-aminochalcone derivatives were obtained via aldol condensation reaction of 2′-aminoacetophenone (2 mmoL) with the appropriate substituted benzaldehyde 1a–c (2 mmoL) catalyzed by potassium hydroxide (4 mmoL) in methanol (5 mL). The reaction mixture was stirred at 60 °C overnight followed by addition of H2O (100 mL), the pH was adjusted to 3–4 with dilute HCl and the solution extracted with CH2Cl2 (3 × 30 mL). The organic layer was collected, dried over anhydrous sodium sulfate, and the solution was concentrated to dryness. The resulting oil was purified by silica gel column chromatography, using dichloromethane as eluent.

2.1.1. (E)-1-(2-aminophenyl)-3-(4-methylphenyl)prop-2-en-1-one (2a)

The product was isolated as a yellow solid (1.139 g, 4.81 mmoL, 80%). M.p. = 91–92 °C. 1H NMR (300 MHz, CDCl3) δ = 2.39 (s, 3H, CH3), 6.44 (br s, 2H, NH2), 6.71 (d, J = 8.1 Hz, 1H, H-3), 6.72 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.22 (d, J = 8.1 Hz, 2H, H-3′,5′), 7.30 (dt, J = 8.1 Hz, 1.2 Hz, 1H, H-4), 7.54 (d, J = 8.1 Hz, 2H, H-2′,6′), 7.61 (d, J = 15.6 Hz, 1H, α-H), 7.78 (d, J = 15.3 Hz, 1H, β-H), 7.89 (dd, J = 8.1 and 1.2 Hz, 1H, H-6) ppm. 13C NMR (75 MHz, CDCl3) δ = 21.6 (CH3), 115.8 (C-5), 117.4 (C-3), 119.2 (C-1), 122.1 (α-C), 128.4 (C-2′,6′), 129.7 (C-3′,5′), 131.1 (C-6), 132.6 (C-1′), 134.3 (C-4), 140.6 (C-4′), 143.0 (β-C), 151.2 (C-2), 191.8 (CO) ppm. MS m/z (ESI %): 238.1 ([M+H]+); HMRS: m/z (ESI) calc. for C16H16NO 238.1226, found 238.1230. UV/Vis (CHCl3, nm): λmax (log ε) = 317 (4.59).

2.1.2. (E)-1-(2-aminophenyl)-3-(4-methoxyphenyl)prop-2-en-1-one (2b)

The product was isolated as a dark yellow solid (0.214 g, 0.81 mmoL, 43%). M.p. = 68–69 °C. 1H NMR (300 MHz, CDCl3) δ = 3.81 (s, 3H, OCH3), 6.25 (br s, 2H, NH2), 6.65–6.70 (m, 2H, H-3 and H-5), 6.90 (d, J = 8.8 Hz, 2H, H-3′,5′), 7.26 (dt, J = 8.4 and 1.5 Hz, 1H, H-4), 7.48 (d, J = 15.6 Hz, 1H, α-H), 7.56 (d, J = 8.8 Hz, 2H, H-2′,6′), 7.71 (d, J = 15.6 Hz, 1H, β-H), 7.84 (dd, J = 8.4 and 1.5 Hz, 1H, H-6) ppm. 13C NMR (75 MHz, CDCl3) δ = 55.4 (OCH3), 114.4 (C-3′,5′), 115.8 (C-5), 117.3 (C-3), 119.3 (C-1), 120.8 (α-C), 128.0 (C-1′), 130.0 (C-2′,6′), 130.9 (C-6), 134.1 (C-4), 142.8 (β-C), 151.0 (C-2), 161.3 (C-4′), 191.8 (CO) ppm. MS m/z (ESI %): 254.1 ([M+H]+); HMRS: m/z (ESI) calc. for C16H16NO2 254.1176, found 254.1172. UV/Vis (CHCl3, nm): λmax (log ε) = 333 (4.51).

2.1.3. (E)-1-(2-aminophenyl)-3-[4-(dimethylamino)phenyl]prop-2-en-1-one (2c)

The product was isolated as an orange solid (1.210 g, 4.51 mmoL, 75%). M.p. = 120–121 °C. 1H NMR (300 MHz, CDCl3) δ = 3.00 (s, 6H, N(CH3)2), 6.26 (br s, 2H, NH2), 6.67 (d, J = 8.7 Hz, 2H, H-3′,5′), 6.66–6.71(m, 2H, H-3 and H-5), 7.25 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.41 (d, J = 15.4 Hz, 1H, α -H), 7.52 (d, J = 8.7 Hz, 2H, H-2′,6′), 7.73 (d, J = 15.4 Hz, 1H, β -H), 7.86 (dd, J = 8.1 and 1.2 Hz, 1H, H-6) ppm. 13C NMR (75 MHz, CDCl3) δ = 40.2 (N(CH3)2), 111.9 (C-3′,5′), 115.8 (C-5), 117.3 (C-3), 117.9 (α -C), 119.9 (C-1), 123.0 (C-1′), 130.2 (C-2′,6′), 130.8 (C-6), 133.7 (C-4), 144.0 (β -C), 150.7 (C-2), 151.8 (C-4′), 192.0 (CO) ppm. MS m/z (ESI %): 267.1 ([M+H]+); HMRS: m/z (ESI) calc. for C17H19N2O 267.1492, found 267.1499. UV/Vis (CHCl3, nm): λmax (log ε) = 418 (4.59).

2.2. General Procedure for the Synthesis of 2′-(methanesulfonylamino)chalcones 3a–c

2′-Aminochalcone derivatives 2a–c (1 mmoL), methane sulfonyl chloride (1 mmoL) and triethylamine (1 mmoL) were stirred in dichloromethane (5 mL) overnight at room temperature. Following the addition of H2O (50 mL), the pH was neutralized with dilute HCl and the solution extracted with CH2Cl2 (3 × 20 mL). The organic layer was collected, dried over anhydrous sodium sulfate, and the solution was concentrated to dryness. The resulting solid was purified by silica gel column chromatography, using dichloromethane as eluent.

2.2.1. (E)-N-{2-[3-(4-methylphenyl)acryloyl]phenyl}methanesulfonamide (3a)

The product was isolated as a yellow solid (0.148 g, 0.47 mmoL, 56%) M.p. = 125–126 °C. 1H NMR (300 MHz, CDCl3) δ = 2.40 (s, 3H, CH3), 3.06 (s, 3H, SO2CH3), 7.21 (dt, J = 7.8 and 1.2 Hz, 1H, H-5), 7.24 (d, J = 8.4 Hz, 2H, H-3′,5′), 7.55 (d, J = 15.5 Hz, 1H, α -H), 7.56 (d, J = 8.4 Hz, 2H, H-2′,6′), 7.57 (dt, J = 7.8 and 1.2 Hz, 1H, H-4), 7.79 (dd, J = 8.4 and 1.2 Hz, 1H, H-3), 7.83 (d, J = 15.5 Hz, 1H, β -H), 8.03 (dd, J = 7.8 and 1.2 Hz, 1H, H-6), 11.27 (br s, 1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 21.6 (CH3), 40.1 (SO2CH3), 118.8 (C-3), 120.7 (α -C), 122.8 (C-5), 123.6 (C-1′), 128.8 (C-2′,6′), 129.8 (C-3′,5′), 131.1 (C-6), 131.7 (C-2), 134.8 (C-4), 140.6 (C-1), 141.8 (C-4′), 146.4 (β -C), 192.7 (CO) ppm. MS m/z (ESI %): 316.1 ([M+H]+); HMRS: m/z (ESI) calc. for C17H18NO3S 316.1002, found 316.1008. UV/Vis (CHCl3, nm): λmax (log ε) = 336 (4.56).

2.2.2. (E)-N-{2-[3-(4-methoxyphenyl)acryloyl]phenyl}methanesulfonamide (3b)

The product was isolated as a yellow solid (0.109 g, 0.33 mmoL, 42%). M.p. = 135–136 °C. 1H NMR (300 MHz, CDCl3) δ = 3.06 (s, 3H, SO2CH3), 3.86 (s, 3H, OCH3), 6.94 (d, J = 8.7 Hz, 2H, H-3′,5′), 7.21 (dt, J = 8.1 and 1.2 Hz, 1H, H-5), 7.46 (d, J = 15.3 Hz, 1H, α-H), 7.56 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.60 (d, J = 8.7 Hz, 2H, H-2′,6′), 7.77 (dd, J = 8.1 and 1.2 Hz, 1H, H-3), 7.81 (d, J = 15.3 Hz, 1H, β-H), 8.03 (dd, J = 8.1 and 1.2 Hz, 1H, H-6), 11.32 (br s,1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 40.1 (SO2CH3), 50.5 (OCH3), 114.6 (C-3′,5′), 118.8 (C-3), 119.3 (α-C), 122.8 (C-5), 123.7 (C-1), 127.2 (C-1′), 130.6 (C-2′,6′), 131.0 (C-6), 134.7 (C-4), 140.6 (C-2), 146.2 (β-C), 162.2 (C-4′), 192.6 (CO) ppm. MS m/z (ESI %): 332.1 ([M+H]+); HMRS: m/z (ESI) calc. for C17H18NO4S 332.0951, found 332.0965. UV/Vis (CHCl3, nm): λmax (log ε) = 361 (4.57).

2.2.3. (E)-N-{2-[3-(4-(dimethylamino)phenyl)acryloyl]phenyl}methanesulfonamide (3c)

The product was isolated as a red solid (0.142 g, 0.41 mmoL, 55%). M.p. = 155–156 °C. 1H NMR (300 MHz, CDCl3) δ = 3.03 (s, 3H, SO2CH3), 3.06 (s, 6H, N(CH3)2), 6.69 (d, J = 8.8 Hz, 2H, H-3′,5′), 7.19 (dt, J = 8.1 and 1.2 Hz, 1H, H-5), 7.36 (d, J = 15.0 Hz, 1H, α-H), 7.53 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.55 (d, J = 8.8 Hz, 2H, H-2′,6′), 7.77 (dd, J = 8.1 and 1.2 Hz, 1H, H-3), 7.83 (d, J = 15.0 Hz, 1H, β-H), 8.02 (dd, J = 8.1 and 1.2 Hz, 1H, H-6), 11.40 (br s, 1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 39.9 (SO2CH3), 40.1 (N(CH3)2), 111.8 (C-3′,5′), 116.0 (α-C), 119.1 (C-3), 122.1 (C-1′), 122. 8 (C-5), 124.5 (C-1), 130.7 (C-6), 131.0 (C-2′,6′), 134.1 (C-4), 140.3 (C-2), 147.5 (β-C), 152.5 (C-4′), 192.4 (CO) ppm. MS m/z (ESI %): 345.1 ([M+H]+); HMRS: m/z (ESI) calc. for C18H21N2O3S 345.1267, found 345.1263. UV/Vis (CHCl3, nm): λmax (log ε) = 439 (4.65).

2.3. General Procedure for the Synthesis of 2′-(Acetylamino)Chalcones 4a–c

2′-Aminochalcone derivatives 2a–c (1 mmol), acetyl chloride (1 mmoL) and triethylamine (1 mmoL) in dichloromethane (5 mL) were stirred overnight at room temperature. Following the addition of H2O (50 mL), the pH was neutralized with dilute HCl and the solution extracted with CH2Cl2 (3 × 20 mL). The organic layer was collected, dried over anhydrous sodium sulfate, and the solution was concentrated to dryness. The resulting solid was purified by silica gel column chromatography, using dichloromethane as eluent.

2.3.1. (E)-N-{2-[3-(4-methylphenyl)acryloyl]phenyl}acetamide (4a)

The product was isolated as a yellow solid (0.118 g, 0.42 mmoL, 84%). M.p. = 121–122 °C. 1H NMR (300 MHz, CDCl3) δ = 2.25 (s, 3H, COCH3), 2.41 (s, 3H, CH3), 7.16 (dt, J = 8.2 and 1.2 Hz, 1H, H-5), 7.25 (d, J = 8.4 Hz, 2H, H-3′,5′), 7.52 (d, J = 15.6 Hz, 1H, α-H), 7.55 (d, J = 8.4 Hz, 2H, H-2′,6′), 7.56 (dt, J = 8.2 and 1.2 Hz, 1H, H-4), 7.79 (d, J = 15.6 Hz, 1H, β-H), 7.97 (dd, J = 8.2 and 1.2 Hz, 1H, H-3), 8.70 (d, J = 8.2 Hz, 1H, H-6), 11.55 (br s, 1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 21.6 (CH3), 25.5 (COCH3), 121.2 (C-6), 121.8 (α-C), 122.4 (C-5), 123.5 (C-1), 128.7 (C-2′,6′), 129.8 (C-3′,5′), 130.5 (C-3), 131.9 (C-1′), 134.6 (C-4), 141.0 (C-2), 141.6 (C-4′), 145.8 (β-C), 169.4 (COCH3), 193.5 (CO) ppm. MS m/z (ESI %): 280.1 ([M+H]+); HMRS: m/z (ESI) calc. for C18H18NO2 280.1332, found 280.1337. UV/Vis (CHCl3, nm): λmax (log ε) = 334 (4.49).

2.3.2. (E)-N-{2-[3-(4-methoxyphenyl)acryloyl]phenyl}acetamide (4b)

The product was isolated as a yellow solid (0.101 g, 0.34 mmoL, 43%). M.p. = 121–122 °C. 1H NMR (300 MHz, CDCl3) δ = 2.23 (s, 3H, COCH3), 3.84 (s, 3H, OCH3), 6.93 (d, J = 8.7 Hz, 2H, H-3′,5′), 7.13 (dt, J = 8.1 and 1.2 Hz, 1H, H-5), 7.42 (d, J = 15.6 Hz, 1H, α -H), 7.53 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.59 (d, J = 8.7 Hz, 2H, H-2′,6′), 7.76 (d, J = 15.6 Hz, 1H, β -H), 7.94 (dd, J = 8.1 and 1.2 Hz, 1H, H-3), 8.69 (dd, J = 8.1 and 1.2 Hz, 1H, H-6) 11.58 (br s, 1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 25.5 (COCH3), 55.4 (OCH3), 114.5 (C-3′,5′), 120.3 (α -C), 121.1 (C-6), 122.4 (C-5), 123.6 (C-1), 127.3 (C-1′), 130.4 (C-3), 130.5 (C-2′,6′), 134.4 (C-4), 140.9 (C-2), 145.5 (β -C), 162.0 (C-4′), 169.3 (CO CH3), 193.3 (CO) ppm. MS m/z (ESI %): 296.1 ([M+H]+); HMRS: m/z (ESI) calc. for C18H18NO3 296.1281, found 296.1272. UV/Vis (CHCl3, nm): λmax (log ε) = 359 (4.54).

2.3.3. (E)-N-{2-[3-(4-(dimethylamino)phenyl)acryloyl]phenyl}acetamide (4c)

The product was isolated as an orange solid (0.162 g, 0.52 mmoL, 66%). M.p. = 129–131 °C. 1H NMR (300 MHz, CDCl3) δ = 2.23 (s, 3H, COCH3), 3.04 (s, 6H, N(CH3)2), 6.67 (d, J = 8.7 Hz, 2H, H-3′,5′), 7.13 (dt, J = 8.1 and 1.2 Hz, 1H, H-5), 7.34 (d, J = 15.3 Hz, 1H, α-H), 7.51 (dt, J = 8.1 and 1.2 Hz, 1H, H-4), 7.54 (d, J = 8.7 Hz, 2H, H-2′,6′), 7.78 (d, J = 15.3 Hz, 1H, β-H), 7.95 (dd, J = 8.1 and 1.2 Hz, H-3), 8.68 (dd, J = 8.1 and 1.2 Hz, 1H, H-6), 11.66 (br s, 1H, NH) ppm. 13C NMR (75 MHz, CDCl3) δ = 25.5 (COCH3), 3.04 (N(CH3)2), 111.8 (C-3′,5′), 117.2 (α-C), 121.0 (C-6), 122.3 (C-1′), 122.3 (C-5), 124.3 (C-1), 130.2 (C-3), 130.8 (C-2′,6′), 133.8 (C-4), 140.7 (C-2), 146.8 (β-C), 152.3 (C-4′), 169.3 (COCH3), 193.2 (CO) ppm. MS m/z (ESI %): 309.2 ([M+H]+); HMRS: m/z (ESI) calc. for C19H21N2O2 309.1598, found 309.1595. UV/Vis (CHCl3, nm): λmax (log ε) = 432 (4.61).

2.4. Photophysical Characterization of 2′-aminochalcones 2, 3 and 4

Solutions of 2′-aminochalcones 2a–c, 3a–c and 4a–c (10−4 to 10−6 M) were prepared in solvents with different character (polar-apolar, protic-aprotic), and their absorption and emission spectra were recorded. Relative fluorescence quantum yields were calculated using fluorescein (ΦF = 0.95 in NaOH 0.1 M) and anthracene (ΦF = 0.27 in ethanol) as standard [19].

3. Results and Discussion

3.1. Synthesis

The synthetic route chosen to obtain the 2′-aminochalcones comprised two steps (Scheme 2). First, the aminochalcones were synthesized by base-catalyzed aldol condensation using potassium hydroxide, in 43–80% yields. Then, the electron-withdrawing groups were introduced to obtain the final compounds 3 (sulfonamide formation) and 4 (acylation), in 42–84% yields. All new compounds were fully characterized by 1H-NMR, 13C-NMR and MS.
The main features of the 1H NMR spectra of compounds 2, 3 and 4 are the signals corresponding to the vinylic protons of their characteristic α,β-unsaturated moieties. The protons H-α and H-β appear at δ 7.34–7.61 and 7.71–7.83 ppm, respectively, as two doublets with a coupling constant of ca. 15 Hz, which indicates an (E)-configuration for the vinylic system. Another important characteristic of the 1H NMR spectra of 3 and 4 is a singlet with high frequency values (δ 11.27–11.66 ppm) attributed to 2′-NH2, due to the intramolecular hydrogen bond with the carbonyl group. This signal is not observed for compounds 2a–c that have a broad singlet with lower frequency values, between 6.25 and 6.44 ppm.

3.2. Photophysical Study of 2′-Aminochalcones 2, 3 and 4

The photophysical properties of the chalcones 2a–c, 3a–c and 4a–c were evaluated. First, the absorption and emission spectra of all compounds were measured in acetonitrile (Figure 1 and Figure 2). The 2′-aminocalcones 2-4 show low to moderate fluorescence quantum yields, and large Stokes’ shifts (5058–9883 cm−1) (Table 1). The influence of the electron-donating group is observed, and compounds 2c, 3c and 4c, bearing a dimethylamino substituent, present a red shift on the absorption and emission maxima. They are also more emissive, with higher quantum yields of 0.03–0.35. For some compounds, the emission profile is not symmetric, and a second emission band appears with a lower intensity. This is consistent with the proposed ESIPT model.
Considering the results obtained in acetonitrile, the photophysical properties of 2′-aminochalcones 2, 3 and 4 were evaluated in other organic solvents presenting different polarity. The solvents tested were dimethylsulfoxide, chloroform and diethyl ether, as examples of solvents with different polarity and proticity, with π* values determined by Kamlet and Taft [20] indicating the ability of a particular solvent to stabilize a charge or a dipole due to its dielectric effect, normalized between cyclohexane (π* = 0) and DMSO (π* = 1). The collected data revealed similar maxima wavelengths of absorption, but some differences in the emission maxima wavelengths (two maxima can be observed in some solvents, for compounds 2c, 3a, 3c and 4c) and the fluorescence quantum yield do not vary significantly (Table 1). The general trend revealed that 2′-aminochalcone 4c (para-position of ring B with dimethylamine and in ring A the acetyl-protected amine group) is the most fluorescent (with ΦF in the range of 0.32–0.36).
The absorption and emission spectra of 2c and 4c in different solvent are presented in Figure 3 and Figure 4. For both compounds, a small shift is observed in the absorption spectra in different solvents, probably due to the solvatochromism expected for push–pull dyes. Two emission bands are clearly visible for 2c, with variable intensities depending on the solvent. They are less obvious for 4c. In diethyl ether (less polar solvent), the band at shorter wavelengths is predominant, corresponding to the L form (Scheme 1). In DMSO and ACN (more polar), the band at longer wavelengths is predominant, corresponding to the E form. In Chloroform, both bands are present, witnessing the ESIPT equilibrium. These bands are probably merged into a single one for 4c.

4. Conclusions

The 2′aminochalcones 2, 3 and 4 are emissive, especially derivatives c (N,N-dimethylamine as a substituent) with moderate fluorescence quantum yields (ΦF = 0.03 for 2c; ΦF = 0.10 for 3c; ΦF = 0.35 for 4c) and large Stokes’ shifts (5405, 5058 and 5379 cm−1, respectively) in acetonitrile. Compound 4c (acetyl group as a substituent on the amine) displayed the highest fluorescence. Fluorophores 2c and 4c present an interesting ESIPT, with a dual emission. The intensity of each emission band is strongly influenced by the solvent and witnesses a dramatic shift in the ESIPT equilibrium. These compounds could, therefore, find applications as molecular probes of organelles polarity, and this study is ongoing.

Acknowledgments

Thanks are due to University of Aveiro, FCT/MEC, Centro 2020 and Portugal2020, the COMPETE program, and the European Union (FEDER program) via the financial support to the LAQV-REQUIMTE (UIDB/50006/2020), to the CICECO-Aveiro Institute of Materials (UID/CTM/50011/2019, UIDB/50011/2020 and UIDP/50011/2020), financed by national funds through the FCT/MCTES, to the Portuguese NMR Network, and to the PAGE project “Protein aggregation across the lifespan” (CENTRO-01-0145-FRDER-000003, C.I.C. Esteves research contract). SG is supported by national funds (OE), through FCT, I.P., in the scope of the framework contract foreseen in the numbers 4, 5, and 6 of the article 23, of the Decree-Law 57/2016, of August 29, changed by Law 57/2017, of July 19. L.F.B. Fontes thanks the FCT for his PhD grant (SFRH/BD/150663/2020).

References

  1. Zhuang, C.; Zhang, W.; Sheng, C.; Zhang, W.; Xing, C.; Miao, V. Chalcone: A privileged structure in medicinal chemistry. Chem. Rev. 2017, 117, 7762–7810. [Google Scholar] [CrossRef] [PubMed]
  2. Gomes, M.N.; Braga, R.C.; Grzelak, E.M.; Neves, B.J.; Muratov, E.; Ma, R.; Klein, L.L.; Cho, S.; Oliveira, G.R.; Franzblau, S.G.; et al. QSAR-driven design, synthesis and discovery of potent chalcone derivatives with antitubercular activity. Eur. J. Med. Chem. 2017, 137, 126–138. [Google Scholar] [CrossRef] [PubMed]
  3. Patil, C.B.; Mahajan, S.K.; Katti, S.A. Chalcone: A Versatile Molecule. J. Pharm. Sci. Res. 2009, 1, 11–22. [Google Scholar]
  4. Custodio, J.M.F.; Gotardo, F.; Vaz, W.F.; D’Oliveira, G.D.C.; de Almeida, L.R.; Fonseca, R.D.; Cocca, L.H.Z.; Perez, C.N.; Oliver, A.G.; de Boni, L.; et al. Benzenesulfonyl incorporated chalcones: Synthesis, structural and optical properties. J. Mol. Struct. 2020, 1208, 127845. [Google Scholar] [CrossRef]
  5. Li, Y.-Y.; Huang, S.-S.; Lee, M.-M.; Deng, J.-S.; Huang, G.-J. Anti-inflammatory activities of cardamonin from Alpinia katsumadai through heme oxygenase-1 induction and inhibition of NF-κB and MAPK signaling pathway in the carrageenan-induced paw edema. Int. Immunopharmacol. 2015, 25, 332–339. [Google Scholar] [CrossRef] [PubMed]
  6. Cole, A.L.; Hossain, S.; Cole, A.M.; Phanstiel, O. Synthesis and bioevaluation of substituted chalcones, coumaranones and other flavonoids as anti-HIV agents. Bioorg. Med. Chem. 2016, 24, 2768–2776. [Google Scholar] [CrossRef] [PubMed]
  7. Rudrapal, M.; Satyanandam, R.S.; Swaroopini, T.S.; Lakshmi, T.N.; Jaha, S.K.; Zaheera, S. Naga Lakshmi. Med. Chem. Res. 2013, 22, 2840–2846. [Google Scholar] [CrossRef]
  8. Custodio, J.M.F.; Michelini, L.J.; de Castro, M.R.C.; Vaz, W.F.; Neves, B.J.; Cravo, P.V.L.; Barreto, F.S.; Filho, M.O.M.; Pereza, C.N.; Napolitano, H.B. Structural insights into a novel anticancer sulfonamide chalcone. New J. Chem. 2018, 42, 3426–3434. [Google Scholar] [CrossRef]
  9. Lin, Y.-M.; Zhou, Y.; Flavin, M.T.; Zhou, L.-M.; Nie, W.; Chen, F.-C. Chalcones and flavonoids as anti-tuberculosis agents. Bioorg. Med. Chem. 2002, 10, 2795–2802. [Google Scholar] [CrossRef]
  10. Boeck, P.; Falcao, C.A.B.; Leal, P.C.; Yunes, R.A.; Filho, V.C.; Torres-Santos, E.C.; Rossi-Bergmann, B. Synthesis of chalcone analogues with increased antileishmanial activity. Bioorg. Med. Chem. 2006, 14, 1538–1545. [Google Scholar] [CrossRef] [PubMed]
  11. Mahapatra, D.K.; Bharti, S.K. Therapeutic potential of chalcones as cardiovascular agents. Life Sci. 2016, 148, 154–172. [Google Scholar] [CrossRef] [PubMed]
  12. Sulpizio, C.; Roller, A.; Giester, G.; Rompel, A. Synthesis, structure, and antioxidant activity of methoxy- and hydroxyl-substituted 2′-aminochalcones. Mon. Chem. 2016, 147, 1747–1757. [Google Scholar] [CrossRef] [PubMed]
  13. Pati, H.N.; Holt, H.L., Jr.; LeBlanc, R.; Dickson, J.; Stewart, M.; Brown, T.; Lee, M. Synthesis and Cytotoxic Properties of Nitro- and Aminochalcones. Med. Chem. Res. 2005, 14, 19–25. [Google Scholar] [CrossRef]
  14. Lazópulos, S.Q.; Svarc, F.; Sagrera, G.; Dicelio, L. Absorption and photo-stability of substituted dibenzoylmethanes and chalcones as UVA filters. Cosmetics 2018, 5, 33. [Google Scholar] [CrossRef]
  15. Vaz, P.A.A.M.; Rocha, J.; Silva, A.M.S.; Guieu, S. Aggregation-induced emission enhancement in halochalcones. New J. Chem. 2016, 40, 8198–8201. [Google Scholar] [CrossRef]
  16. Sedgeick, A.C.; Wu, L.; Han, H.; Bull, S.D.; He, X.; James, T.D.; Sessler, J.L.; Tang, V.; Tian, H.; Yoon, J. Excited-state intramolecular proton-transfer (ESIPT) based fluorescence sensors and imaging agents. Chem. Soc. Rev. 2018, 47, 8842–8880. [Google Scholar] [CrossRef] [PubMed]
  17. Yin, H.; Li, H.; Xia, G.; Ruan, C.; Shi, Y.; Wang, H.; Jin, M.; Ding, D. A novel non-fluorescent excited state intramolecular proton transfer phenomenon induced by intramolecular hydrogen bonds: an experimental and theoretical investigation. Sci. Rep. 2016, 6, 19774. [Google Scholar] [CrossRef] [PubMed]
  18. Fontes, L.F.B.; Nunes da Silva, R.; Silva, A.M.S.; Guieu, S. Unsymmetrical 2,4,6-Triarylpyridines as Versatile Scaffolds for Deep-Blue and Dual-Emission Fluorophores. ChemPhotoChem 2020, accepted. [Google Scholar] [CrossRef]
  19. Brouwer, A.M. Standards for photoluminescence quantum yield measurements in solution (IUPAC Technical Report). Pure Appl. Chem. 2011, 83, 2213–2228. [Google Scholar] [CrossRef]
  20. Kamlet, M.J.; Abboud, 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]
Scheme 1. Description of the Excited State Intramolecular Proton Transfer (ESIPT) process [16].
Scheme 1. Description of the Excited State Intramolecular Proton Transfer (ESIPT) process [16].
Chemproc 03 00029 sch001
Scheme 2. Synthesis of the 2′-aminochalcones 2a–c, 3a–c and 4a–c.
Scheme 2. Synthesis of the 2′-aminochalcones 2a–c, 3a–c and 4a–c.
Chemproc 03 00029 sch002
Figure 1. Normalized absorption spectra of 2′aminochalcones 2, 3 and 4 in acetonitrile.
Figure 1. Normalized absorption spectra of 2′aminochalcones 2, 3 and 4 in acetonitrile.
Chemproc 03 00029 g001
Figure 2. Normalized emission spectra of 2′aminochalcones 2, 3 and 4 in acetonitrile.
Figure 2. Normalized emission spectra of 2′aminochalcones 2, 3 and 4 in acetonitrile.
Chemproc 03 00029 g002
Figure 3. Normalized absorption and emission spectra of chalcone 2c in acetonitrile, diethyl ether, chloroform and dimethyl sulfoxide (λexc = 418 nm, 404 nm, 418 nm and 429 nm, respectively; absorption, full line; emission, dotted line).
Figure 3. Normalized absorption and emission spectra of chalcone 2c in acetonitrile, diethyl ether, chloroform and dimethyl sulfoxide (λexc = 418 nm, 404 nm, 418 nm and 429 nm, respectively; absorption, full line; emission, dotted line).
Chemproc 03 00029 g003
Figure 4. Normalized absorption and emission spectra of chalcone 4c in acetonitrile, diethyl ether, chloroform and dimethyl sulfoxide (λexc = 428 nm, 417 nm, 432 nm and 437 nm, respectively; absorption, full line; emission, dotted line).
Figure 4. Normalized absorption and emission spectra of chalcone 4c in acetonitrile, diethyl ether, chloroform and dimethyl sulfoxide (λexc = 428 nm, 417 nm, 432 nm and 437 nm, respectively; absorption, full line; emission, dotted line).
Chemproc 03 00029 g004
Table 1. UV-visible absorption and emission data for 2′-aminochalcones 2, 3 and 4 in organic solvents with different characteristics.
Table 1. UV-visible absorption and emission data for 2′-aminochalcones 2, 3 and 4 in organic solvents with different characteristics.
CpdSolvent (π*)UV/VisFluorescence
λmax (nm)log ελem (nm)Stokes’
Shift (cm−1)
ΦF
2aDiethyl ether (0.27)3084.58---------
Chloroform (0.58)3174.594599759<0.01
ACN (0.75)3114.574499883<0.01
DMSO (1.00)3174.574609807<0.01
2bDiethyl ether (0.27)3244.51---------
Chloroform (0.58)3334.514558052<0.01
ACN (0.75)3274.524488260<0.01
DMSO (1.00)3354.524547824<0.01
2cDiethyl ether (0.27)4044.584934468<0.01
Chloroform (0.58)4184.59497/5323803/51260.03
ACN (0.75)4184.5754054050.03
DMSO (1.00)4294.5854047920.03
3aDiethyl ether (0.27)3294.57417/4416414/7719<0.01
Chloroform (0.58)3364.564407035<0.01
ACN (0.75)3334.594387199<0.01
DMSO (1.00)3324.58413/4385907/7289<0.01
3bDiethyl ether (0.27)3544.53---------
Chloroform (0.58)3614.574545674<0.01
ACN (0.75)3584.504455461<0.01
DMSO (1.00)3634.524555570<0.01
3cDiethyl ether (0.27)4244.62---------
Chloroform (0.58)4394.6653540870.10
ACN (0.75)4374.6056150580.10
DMSO (1.00)4454.6155745190.11
4aDiethyl ether (0.27)3254.51---------
Chloroform (0.58)3344.494407213<0.01
ACN (0.75)3264.494327527<0.01
DMSO (1.00)3274.504588747<0.01
4bDiethyl ether (0.27)3494.53---------
Chloroform (0.58)3594.544374972<0.01
ACN (0.75)3524.514315207<0.01
DMSO (1.00)3524.524375526<0.01
4cDiethyl ether (0.27)4174.594923656<0.01
Chloroform (0.58)4324.6155350650.32
ACN (0.75)4284.6055653790.35
DMSO (1.00)4374.6055849620.36
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Esteves, C.I.C.; Fontes, L.F.B.; Borges, A.F.N.; Rocha, J.; Silva, A.M.S.; Guieu, S. Synthesis and Photophysical Characterization of 2′-Aminochalcones. Chem. Proc. 2021, 3, 29. https://doi.org/10.3390/ecsoc-24-08301

AMA Style

Esteves CIC, Fontes LFB, Borges AFN, Rocha J, Silva AMS, Guieu S. Synthesis and Photophysical Characterization of 2′-Aminochalcones. Chemistry Proceedings. 2021; 3(1):29. https://doi.org/10.3390/ecsoc-24-08301

Chicago/Turabian Style

Esteves, Cátia I. C., Luís F. B. Fontes, A. Filipa N. Borges, João Rocha, Artur M. S. Silva, and Samuel Guieu. 2021. "Synthesis and Photophysical Characterization of 2′-Aminochalcones" Chemistry Proceedings 3, no. 1: 29. https://doi.org/10.3390/ecsoc-24-08301

Article Metrics

Back to TopTop