Next Article in Journal
A New Editor in Chief for Molecules (It’s Been a Fun Ride)
Next Article in Special Issue
Polymer-Supported Synthesis of Various Pteridinones and Pyrimidodiazepinones
Previous Article in Journal
A Comprehensive Study of the Impacts of Oat β-Glucan and Bacterial Curdlan on the Activity of Commercial Starter Culture in Yogurt
Previous Article in Special Issue
Practical Synthesis of Quinoline-Protected Morpholino Oligomers for Light-Triggered Regulation of Gene Function
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Substituted 2-Phenacylbenzoxazole Difluoroboranes: Synthesis, Structure and Properties

by
Agnieszka Skotnicka
1,* and
Przemysław Czeleń
2
1
Faculty of Chemical Technology and Engineering, UTP University of Science and Technology, Seminaryjna 3, 85-326 Bydgoszcz, Poland
2
Department of Physical Chemistry, Collegium Medicum, N. Copernicus University, Kurpińskiego 5, 85-950 Bydgoszcz, Poland
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(22), 5420; https://doi.org/10.3390/molecules25225420
Submission received: 13 October 2020 / Revised: 17 November 2020 / Accepted: 18 November 2020 / Published: 19 November 2020
(This article belongs to the Special Issue Synthetic Heterocyclic Chemistry)

Abstract

:
Novel fluorescent dyes such as benzoxazole-boron complexes, bearing β-ketoiminate ligands, have been synthesized and characterized with a focus on the influence of a substituent on the basic photophysical properties. 1H, 11B, 13C, 15N, and 19F nuclear magnetic resonance (NMR) spectra of substituted 2-phenacylbenzoxazole difluoroboranes have been recorded and discussed. It is worth mentioning that a high correlation coefficient was found between 15N-NMR parameters and substituent constants. The photophysical properties of these new dyes have been investigated by fluorescence and ultraviolet-visible (UV-Vis) absorption spectroscopy. The geometry optimization, vibrational spectra, and the HOMO and LUMO energies were calculated based on density functional theory with the use of the B3LYP functional and 6-311++G(d,p) basis set.

Graphical Abstract

1. Introduction

Boron dipyrromethene (BODIPY) dyes display excellent photophysical and optoelectronic properties, such as high photostability, a large absorption coefficient in the visible and near-infrared (IR) ranges, excellent chemical stability, high fluorescence quantum yields, relatively long excited-state lifetime, good solubility in organic solvents, relative insensitivity to environmental conditions, and good biocompatibility [1,2,3]. Moreover, the BODIPY chromophore is a versatile platform for the construction of fluorescent probes for bioimaging and biosensing applications [4,5,6]. In addition, novel photoinitiators based on the BODIPY are widely used for efficient cationic polymerization of epoxy, epoxy-silicone, and vinyl ether monomers [7,8].
There are mainly three types of these fluorine-boron complexes, classified as N,N bidentate, O,O bidentate, and N,O bidentate compounds. The typical BODIPY core contains two pyrrole units and its structure is usually symmetric. The compounds studied now contain NBF2O moiety. They carry an unsymmetrically chelated BF2 moiety. It was found that some difluoroboron β-ketoiminate boron complexes bearing benzoxazole gave strong emissions in solid states on account of aggregation-induced emission (AIE). The reversible mechanofluorochromism was due to the transformation between crystalline and amorphous states [9,10].
In our previous published work, we have investigated the influence of heteroatom (X = O, S, and NMe) substitution in a five-membered heterocyclic ring on the photophysical properties of novel BF2 complexes containing or not a dimethylamino group (R = NMe2) [11]. It was shown that the fluorescence yields of the dimethylamino derivatives were much larger (up to one order of magnitude) than those of the corresponding unsubstituted compounds. Therefore, using the electron-donating or electron-accepting substituents, the electronic properties of the compounds may be tuned [12,13]. This study provides a report on the basic photophysical properties of a recently synthesized series of substituted 2-phenacylbenzoxazole difluoroboranes (Scheme 1) with a view towards the effect of substitution on their properties.

2. Results and Discussion

2.1. Synthesis and NMR Data

Treatment of 2-methylbenzoxazole with benzoyl chloride and triethylamine causes that the 2-methyl group undergoes double benzoylation, i.e., it is transformed to –CH=C(Ph)–COOPh. Non purified intermediate products 1a8a were always used in the subsequent synthetic step. By refluxing their morpholine solutions, 1a8a can be easily transformed into 2-phenacylbenzoxazoles 1b8b [14]. In solution, compounds 1b8b exist in two tautomeric forms, ketimine form (K) and enolimine form (O), through keto-enol tautomerization. The ketimine tautomer structure was confirmed by the methylene signals observed at the range of 4.54–4.67 ppm (s, 2H) in their 1H-NMR spectra. Enolimine tautomers were also confirmed by the 1H-NMR signals at 6.05–6.43 ppm (s, 1H, C=CH) and 12.4–12.8 ppm (s, 1H, OH) [15]. The tautomeric mixture 1b8b was then allowed to react with the boron trifluoride ethyl etherate in the presence of N-ethyldiisopropylamine in dichloromethane to give the corresponding BF2 complex 18 (Scheme 2).
The 1H, 11B, 13C, 15N, and 19F-NMR spectra that confirm the structures of 18 are maintained in solution. NMR chemical shifts of characteristic nuclei are collected in Table 1.
The 1H-NMR signal of H10 of 2-phenacylbenzoxazole difluoroboranes 18 can be seen at δ = 6.37–6.81 ppm in CDCl3, which is a singlet. The presence of a boron atom is confirmed by signals at δ = 1.79–1.85 ppm (triplet). The resonance of F13 at δ = −134.14 to −135.59 ppm shows as a quartet (two fluorine atoms). Analogously Kubota et al. [16] observed signals at −134.6 ppm (q, 2F) for boron complex of 2-methylbenzothiazole. N3 in the 15N-NMR spectra of 2-phenacylbenzoxazole difluoroboranes 18 resonates in the ranges of δ = −212.43 to −215.72 ppm. These signals were correlated to the Hammett substituent constant σ [17]. It was observed that with increasing electron-withdrawing ability of substituent, the 15N-NMR signal experienced a continuous upfield shift, which was indicated by a linear dependence of the chemical shift against σ (δ(15N) = 4.79 σ–214.12, R2 = 0.909), Figure 1. 1H NMR and 13C NMR spectra of 18 are available in Supplementary Materials.

2.2. Spectroscopic Properties

The measured absorption spectra of all eight compounds obtained in chloroform are given in Figure 2, and the associated data are presented in Table 2. Chloroform is a known solvent to prevent boron-ligand dissociation, exciplex formation, or the photochemical reactions possible in solvents containing Lewis bases, aromatic rings, or double bonds [18].
As shown in Figure 2, 2-phenacylbenzoxazole difluoroboranes 18 have an intense, wide absorption band in the 300–410 nm region, which is attributed to the π → π* transition but also associated with n → π* photochemically forbidden transition. The position of the absorption band depends on the dye structure. Absorption at λmax was found to progressively shift to longer wavelength upon substituting unsubstituted complex 5 by weaker electron-withdrawing (Cl) and then electron-releasing (4-Me, 4-OMe, and 4-NMe2) substituents. It must be highlighted that the electron pulling effect from the boron atom created a strong electronic current in the six-membered ring such as B-O-C-C-C-N, which is responsible for the higher % of bathochromic in compound 2 in chloroform. A considerable red-shift of the absorption band was observed for compound 1 (4-NMe2). The 4-NMe2 substituent causes a 58 nm red shift in absorption relative to the parent compound 5. The molar absorption coefficient (ε) of 1 (42,500) was significantly higher than for other complexes (Table 2). These red-shifts of the absorption maxima position along with the high molar absorption coefficients indicate that the transitions of the R = NMe2 derivative have a π → π* nature associated with a significant charge transfer (CT) [11].
The changes in Stokes shifts (Figure 3) and the molar absorption coefficients related to the nature of the substituent in 2-phenacyl moiety are also reflected in changes in the fluorescence quantum yields. As can be seen in Table 2, the measurements of the fluorescence quantum yields of the investigated dyes remain very small. Only derivatives with electron-releasing groups exhibit fluorescence quantum yields significantly higher (up to one order of magnitude for 1).
From the data presented in Table 3 and Figure 4, it is seen that the polar protic solvents such as methanol, dimethylformamide, acetonitryle, and dimethyl sulfoxide cause a bathochromic shift of the absorption band. The largest shift of the absorption bands towards longer wavelengths is observed for the most protic solvent, which is dimethyl sulfoxide. It is clear from Table 3 that the Stokes shift increases with varying solvent polarity. With an increase in solvent polarity, the magnitude of the Stokes shift varies from 2449 cm−1 to 3043 cm−1 for compound 1 (4-NMe2), and from 4530 cm−1 to 6464 cm−1 for other derivatives.

2.3. Solvatochromism

To verify the effect of solvent polarity, Stokes shifts (Δν) of 18 in a variety of solvents were plotted against the solvent polarity parameter Δf (ε, n) with the general form of the Lippert-Mataga equation given below [19,20]
Δ ν = 2 Δ f 4 π ε 0 ћ c a 3   μ e   μ g 2 +   b
Δ f   = ε 1 2 ε + 1 n 2 1 2 n 2   + 1
in which Δν = νabsνem stands for Stokes shift, νabs and νem are absorption and emission frequency (cm−1), ħ is the Planck’s constant, c is the velocity of light in vacuum, a is the Onsager radius, and b is a constant. Δf is the orientation polarizability, ε is the refractive index, n is the dielectric constant, μe and μg are the dipole moments of the emissive and ground states, respectively, and ε0 is the permittivity of the vacuum. (μeμg)2 is proportional to the slope of the Lippert-Mataga plot.
Lippert-Mataga plots showed higher Stokes shifts for all compounds in acetonitrile as a solvent, nonlinear nature of the plot was observed in acetonitrile for 16 shown in Figure 5, while such behavior for 7 and 8 was observed in chloroform. Compound 7 exhibited similar nature to that of 8.

2.4. Computational Details

An important factor in assessing the practical use of the compounds under consideration is their reactivity. The use of computational chemistry methods allows us to describe this property of the tested molecules in the context of a set of descriptors based on the energy values of the frontal orbitals [21]. In Table 4, there are presented values characterizing all considered substituted 2-phenacylbenzoxazole difluoroboranes, including energy values of HOMO and LUMO orbitals, energy gaps, and hardness (η). Obtained data show that among all considered molecules, the highest reactivity is exhibited by the derivative containing 4-dimethylamino substituent. Such a molecule is characterized by the lowest values of hardness and energy gap. The distribution of HOMO and LUMO orbitals obtained for this molecule is presented in Figure 6. The rest of the considered derivatives exhibit quite similar values of both reactivity descriptors, however, there is observed a trend indicating that substitution in the fourth position to aromatic ring is more favorable than in the third. Taking into account all considered substituents, a decrease in reactivity is observed in the following order 4-NMe2 > 4-OMe > 4-Cl > 4-Me > 3-Cl > H > 3-Me > 3-OMe.

2.5. Vibrational Analysis

Vibrational spectroscopy is used extensively for the study of molecular conformations, identification of functional groups and reaction kinetics, etc. Spectroscopy FT-IR and quantum chemical calculations using the B3LYP functional and 6-311++G(d,p) basis set have been employed to study the structure of the title compounds. Substituted 2-phenacylbenzoxazoles 1b8b (starting compounds that were converted to BF2 complexes) exist in two tautomeric forms, the ketimine form (K) and enolimine form (O), through keto-enol tautomerization. A strong band observed in the FT-IR spectrum at around 1660 cm1 was assigned to the C=O stretching vibration of ketimine tautomer. The free hydroxyl group absorbs strongly in the region 3700–3584 cm1, whereas the existence of intramolecular hydrogen bond formation can lower the O–H stretching frequency in the range 3500–3200 cm1 with an increase in intensity and breadth [22]. In the present study, a strong band observed at about 3380 cm1 in FT-IR spectrum was assigned to O–H stretching vibration. The observed FT-IR spectrum of 2-(4-dimethylamino)benzoylbenzoxazole (1b), where the intensity was plotted against the vibrational wavenumber, is shown in Figure 7.
For the substituted 2-phenacylbenzoxazole difluoroboranes 18, no O–H and C=O vibrations could be observed in the FT-IR spectra. The theoretical spectrogram for the FT-IR spectrum of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1) was also constructed and compared with the experimental spectrum. The computed vibrational frequencies of this molecule were found in good agreement with experimental results. To improve the numerical agreement, the linear scaling method as given by Yoshida et al. [23] has been used. In this, the calculated harmonic vibrational wavenumbers have been scaled by the formula (νobscal = (1.0087–0.000016 × νcal) cm−1) to facilitate a better agreement with the observed values. The value of the correlation coefficient was found to be R2 = 0.9997, which shows good agreement of the simulated wavenumbers with the observed one for the molecule of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1). The comparison of the experimental and simulated IR spectrum is presented in Figure 8.

3. Experimental

3.1. Materials

All reagents and solvents were purchased from Sigma-Aldrich (Poznań, Poland) and used without further purification. The highest (≥99%) purity of all used chemicals was required for spectroscopic studies.

3.2. Synthesis

The compounds were obtained from 2-phenacylbenzoxazoles 1b8b (Scheme 2) as described earlier [15]. The typical procedure was as follows: BF3 etherate (three equivalents) was added to the magnetically stirred solution (nitrogen atmosphere) of substituted 2-phenacylbenzoxazole (1 g) in dry chloroform (15–20 mL) and N-ethyldiisopropylamine (three equivalents). Then the solution was stirred overnight at room temperature, and concentrated Na2CO3 water solution (20 mL) was added slowly to the mixture. The organic fraction was separated, the water layer extracted with chloroform (two times using ca. 20–30 mL), dried (Na2SO4) and evaporated under reduced pressure. Residual solids were purified by flash chromatography (SiO2) using DCM as an eluent.

Elemental Analysis Is as Follows

2-(4-Dimethylamino)benzoylbenzoxazole Difluoroborane (1) Orange solid, yield 41%, m.p. 298–300 °C. 1H-NMR (CDCl3 from TMS) δ (ppm): 7.96 (d, 2H, 3JH,H = 9.12 Hz), 7.85 (d, 1H, 3JH,H = 7.36 Hz), 7.61 (d, 1H, 3JH,H = 7.64 Hz), 7.50 (m, 2H), 6.89 (s, 1H), 6.81 (d, 2H, 3JH,H = 9.16 Hz), 3.07 (s, 6H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.79, (t). 13C-NMR δ (ppm): 172.1, 153.7, 152.4, 148.2, 130.6, 129.6, 127.0, 125.9, 119.1, 114.2, 112.1, 111.8, 77.7, ca. 40.61–39.35 (overlapped with solvent). 15N-NMR (CDCl3 from MeNO2) δ (ppm): −313.97. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −134.64. C17H15BF2N2O2, Calcd. C, 62.23; H, 4.61; N, 8.54. Found C, 62.14; H, 4.64; N, 8.39.
2-(4-Methoxy)benzoylbenzoxazole Difluoroborane (2) Yellow solid, yield 39%, m.p. 242–243 °C (251–253 °C [9]). 1H-NMR (CDCl3 from TMS) δ (ppm): 7.98 (m, 2H), 7.78 (d, 1H, 3JH,H = 7.84 Hz), 7.55 (d, 1H, 3JH,H = 8.16 Hz), 7.47 (m, 1H), 7.40 (m, 1H), 6.98 (m, 2H), 6.37 (s, 1H), 3.89 (s, 3H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.80, (t). 13C-NMR δ (ppm): 172.2 164.9, 163.3, 148.1, 130.6, 129.3, 126.6 125.6, 115.2, 114.1, 111.0, 78.8, 55.5. 15N-NMR (CDCl3 from MeNO2) δ (ppm): −215.72. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −135.59. C16H12BF2NO3, Calcd. C, 60.99; H, 3.84; N, 4.45. Found C, 60.69; H, 3.99; N, 4.60.
2-(4-Methyl)benzoylbenzoxazole Difluoroborane (3) Yellow solid, yield 47%, m.p. 256–258 °C. 1H-NMR (CDCl3 from TMS) δ (ppm): 7.91 (m, 2H), 7.80 (d, 1H, 3JH,H = 8.36 Hz), 7.57 (m, 1H), 7.48 (m, 1H), 7.40 (m, 1H), 7.41 (m, 1H), 7.29 (d, 2H, 3JH,H = 8.00 Hz), 6.44 (s, 1H) 2.44 (s, 3H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.84, (t). 13C-NMR δ (ppm): 172.6, 164.9, 148.2, 143.5, 130.6, 129.5, 127.3, 126.7, 125.8, 115.4, 111.1, 79.7, 21.7. 15N-NMR (CDCl3 from MeNO2) δ (ppm): −215.01. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −135.34. C16H12BF2NO2, Calcd. C, 64.25; H, 4.04; N, 4.68. Found C, 64.35; H, 3.94; N, 4.68.
2-(3-Methyl)benzoylbenzoxazole Difluoroborane (4) Yellow solid, yield 48%, m.p. 230–231 °C. 1H-NMR (CDCl3 from TMS) δ (ppm): 7.85 (s, 1H), 7.76 (m, 2H), 7.57 (m, 1H), 7.49 (m, 1H), 7.49 (m, 1H), 7.42 (m, 1H), 7.37 (m, 2H), 6.47 (s, 1H) 2.44 (s, 3H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.86, (t). 13C-NMR δ (ppm): 172.7, 164.8, 148.2, 138.6, 133.4, 133.1, 130.5, 128.6, 127.9, 126.7, 125.9, 124.4, 115.4, 111.1, 80.3, 21.4. 15N-NMR (DMSO-d6 from MeNO2) δ (ppm): −214.18. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −135.14. C16H12BF2NO2, Calcd. C, 64.25; H, 4.04; N, 4.68. Found C, 63.92; H, 4.1; N, 4.85.
2-Benzoylbenzoxazole Difluoroborane (5) Yellowish-green solid, yield 38%, m.p. 239.2–240.8 °C (244–246 °C [9]). 1H-NMR (CDCl3 from TMS) δ: 8.02 (m, 2H), 7.82 (m, 1H), 7.57 (m, 2H) 7.51 (m, 2H), 7.45 (m, 2H), 6.49 (s, 1H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ: 1.85, (t). 13C-NMR δ: 170.8, 165.1, 148.5, 133.1, 130.1, 129.5, 127.5, 126.9, 126.4, 114.9, 112.5, 82.1. 15N-NMR (CDCl3 from MeNO2) δ: −213.54. 19F-NMR (CDCl3 from CFCl3) δ: −135.07. C15H10BF2NO2, Calcd. C, 63.20; H, 3.54; N, 4.91. Found C, 63.25; H, 3.49; N, 4.83.
2-(3-Methoxy)benzoylbenzoxazole Difluoroborane (6) Yellow solid, yield 42%, m.p. 218–220 °C (235–237 °C [9]). 1H-NMR (CDCl3 from TMS) δ (ppm): 7.82 (d, 1H, 3JH,H = 7.84 Hz), 7.56 (m, 3H), 7.49 (m, 1H), 7.44 (m, 1H), 7.40 (m, 1H), 7.09 (m, 1H), 6.47 (s, 1H) 3.9 (s, 3H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.84, (t). 13C-NMR δ (ppm): 172.3, 164.7, 159.9, 148.2, 134.6, 130.5, 129.7, 126.8, 126.0, 119.6, 118.9, 115.5, 112.0, 111.2, 80.7, 55.6. 15N-NMR (CDCl3 from MeNO2) δ (ppm): −213.97. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −134.98. C16H12BF2NO3, Calcd. C, 60.99; H, 3.84; N, 4.45. Found C, 60.72; H, 4.01; N, 4.95.
2-(4-Chloro)benzoylbenzoxazole Difluoroborane (7) Yellow solid, yield 37%, m.p. 233–235 °C (250 °C [10]). 1H-NMR (CDCl3 from TMS) δ (ppm): 7.94 (m, 2H), 7.82 (d, 1H, 3JH,H = 7.84 Hz), 7.58 (m, 1H), 7.51 (m, 1H), 7.47 (m, 2H), 7.44 (m, 1H), 6.45 (s, 1H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.81, (t). 13C-NMR δ (ppm): 171.0, 164.5, 148.2, 138.8, 131.6, 130.4, 129.1, 128.5, 126.9, 126.2, 115.5, 111.2, 80.6. 15N-NMR (CDCl3 from MeNO2) δ (ppm): −213.00. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −134.98. C15H9BF2NO2, Calcd. C, 56.39; H, 2.84; N, 4.38. Found C, 56.51; H, 2.75; N, 4.41.
2-(3-Chloro)benzoylbenzoxazole Difluoroborane (8) Yellow solid, yield 34%, m.p. 228–230 °C. 1H-NMR (CDCl3 from TMS) δ (ppm): 8.00 (t, 1H), 7.88 (m, 1H), 7.83 (d, 1H, 3JH,H = 7.84 Hz), 7.60 (m, 1H), 7.52 (m, 2H), 7.45 (m, 2H), 6.48 (s, 1H). 11B-NMR (CDCl3 from BF3⋅Et2O) δ (ppm): 1.81, (t). 13C-NMR δ (ppm): 170.6, 164.4, 148.3, 135.1, 135.0, 132.4, 130.4, 130.1, 127.3, 126.9, 126.3, 125.3, 115.6, 111.3, 81.2. 15N-NMR (CDCl3 from MeNO2) δ (ppm): −212.43. 19F-NMR (CDCl3 from CFCl3) δ (ppm): −134.78. C15H9BF2NO2, Calcd. C, 56.39; H, 2.84; N, 4.38. Found C, 56.26; H, 2.81; N, 4.54.

3.3. Measurements

The 1H-NMR spectra were recorded using an Ascend III spectrometer operating at 400 MHz, Bruker (Bydgoszcz, Poland). Chloroform was used as a solvent and tetramethylsilane (TMS) as the internal standard. Chemical shifts (δ) were reported in ppm relative to TMS and coupling constants (J) in Hz.
The elemental analysis was made with a Vario MACRO 11.45–0000, Elemental Analyser System GmbH, operating with the VARIOEL software (version 5.14.4.22).
The melting point was measured on the Melting Point M-565 Apparatus (Buchi) with the measuring speed 5 °C/min.
The absorption and emission spectra were measured at room temperature in a quartz cuvette (1 cm) using an Agilent Technology UV-Vis Cary 60 Spectrophotometer and a Hitachi F-7000 Spectrofluorometer, respectively.
The fluorescence quantum yields for the compounds in chloroform were determined as follows, the fluorescence spectrum of diluted (A ≈ 0.1) boranes solution was recorded by excitation at the absorption band maximum of the reference. Diluted 9,10-diphenylanthracene in cyclohexane (ϕ = 0.93) [24] was used as reference. The fluorescence spectrum of 9,10-diphenylanthracene was obtained by excitation at its absorption peak at 355 nm. The quantum yield of the tested compounds (ϕdye) was calculated using the following equation [25]:
ϕ d y e = ϕ r e f   ·   I d y e I r e f A r e f A d y e   ·   n d y e    2 n r e f    2
where ϕref is the fluorescence quantum yield of the reference sample (9,10-diphenylanthracene) in cyclohexane, Adye and Aref are the absorbance of the dye and reference samples at the excitation wavelengths (355 nm), Idye and Iref are the integrated emission intensity for the dyes and references sample, ndye and nref are the refractive indices of the solvents used for the dyes and reference, respectively. Coumarine 153 in cyclohexane (ϕ = 0.90 [26]; λex = 393 nm) was used as reference standard for compound carrying −NMe2 group.
Infrared spectra have been recorded in the region 360–7000 cm1 on a Bruker Alpha FT-IR spectrometer with a spectral resolution of 2 cm1 using the ATR (attenuated total reflectance) method.
The geometry optimization, vibrational spectra, and the HOMO and LUMO energies were calculated based on density functional theory with the use of B3LYP [27,28,29] functional and 6-311++G(d,p) basis set [30,31]. All calculations were carried out using Gaussian 09 software [32]. The analysis of the frontier orbitals and IR spectra including extraction of frequencies and a visual presentation of the vibrational modes and orbitals was realized with the use of Avogadro 1.2.0 application [33].

4. Conclusions

The indicated aims were achieved through a synthesis of a series of 2-phenacylbenzoxazole difluoroboranes substituted by a weaker electron-withdrawing (Cl) and electron-releasing (4-Me, 4-OMe, and 4-NMe2) substituents. Complexes have been identified based on a magnetic atomic nucleus 1H, 11B, 13C, 15N, and 19F isotope resonance spectra. 15N-NMR shifts correlate with the substituent character of the phenyl ring with R2 = 0.91. The spectral data show that the absorption and fluorescence properties of 2-phenacylbenzoxazole difluoroboranes 18 depend on the character of the substituent. The experimental results indicate the positive solvatochromism of studied compounds with increasing solvent polarity. Among all derivatives, only 2-(4-dimethylamino)-benzoylbenzoxazole difluoroborane 1 (the NMe2 group was the strongest electron donor in the series) presented strong and relatively red-shifted fluorescence in the chloroform. Other complexes hardly demonstrated any fluorescence, indicating that probably only benzoxazole-boron complexes with strongly electron-releasing groups like amino groups could be attractive for their use as new fluorescent molecules for future applications in bio-labeling, medicine, or fluorescence microscopy, and they might be alternative or supplementary to popular BODIPY dyes. Furthermore, the study using these compounds as photosensitizers in the photopolymerization process is still in progress.

Supplementary Materials

The following are available online for related NMR spectra.

Author Contributions

Conceptualization, Methodology, investigation, writing—review, A.S.; performed DFT calculations, P.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported in part by PLGrid Infrastructure.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Lu, H.; Mack, J.; Yang, Y.; Shen, Z. Structural modification strategies for the rational design of red/NIR region BODIPYs. Chem. Soc. Rev. 2014, 43, 4778–4823. [Google Scholar] [CrossRef] [Green Version]
  2. Jean-Gérard, L.; Vasseur, W.; Scherninski, F.; Andrioletti, B. Recent advances in the synthesis of [a]-benzo-fused BODIPY fluorophores. Chem. Commun. 2018, 54, 12914–12929. [Google Scholar] [CrossRef]
  3. Zhao, J.; Xu, K.; Yang, W.; Wang, Z.; Zhong, F. The triplet excited state of Bodipy: Formation, modulation and application. Chem. Soc. Rev. 2015, 44, 8904–8939. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Sansalone, L.; Tang, S.; Garcia-Amoros, J.; Zhang, Y.; Nonell, S.; Baker, J.D.; Captain, B.; Raymo, F.M. A photoactivatable far-red/near-infrared BODIPY to monitor cellular dynamics in vivo. ACS Sens. 2018, 3, 1347–1353. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  5. Boens, N.; Leen, V.; Dehaen, W. Fluorescent indicators based on BODIPY. Chem. Soc. Rev. 2012, 41, 1130–1172. [Google Scholar] [CrossRef] [PubMed]
  6. Ni, Y.; Wu, J. Far-red and near infrared BODIPY dyes: Synthesis and applications for fluorescent pH probes and bio-imaging. Org. Biomol. Chem. 2014, 12, 3774–3791. [Google Scholar] [CrossRef] [PubMed]
  7. Telitel, S.; Blanchard, N.; Schweizer, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P.; Lalevee, J. BODIPY derivatives and boranil as new photoinitiating systems of cationic polymerization exhibiting a tunable absorption in the 400–600 nm spectral range. Polymer 2013, 54, 2071–2076. [Google Scholar] [CrossRef]
  8. Telitel, S.; Lalevee, J.; Blanchard, N.; Kavalli, T.; Tehfe, M.-A.; Schweizer, S.; Morlet-Savary, F.; Graff, B.; Fouassier, J.-P. Photopolymerization of cationic monomers and acrylate/divinylether blends under visible light using pyrromethene dyes. Macromolecules 2012, 45, 6864–6868. [Google Scholar] [CrossRef]
  9. Zhang, Z.; Wu, Z.; Sun, J.; Xue, P.; Lu, R. Multi-color solid-state emission of β-iminoenolate boron complexes tuned by methoxyl groups: Aggregation-induced emission and mechanofluorochromism. RSC Adv. 2016, 6, 43755–43766. [Google Scholar] [CrossRef]
  10. Zhao, J.; Peng, J.; Chen, P.; Wang, H.; Xue, P.; Lu, R. Mechanofluorochromism of difluoroboron β-ketoiminate boron complexes functionalized with benzoxazole and benzothiazole. Dyes Pigm. 2018, 149, 276–283. [Google Scholar] [CrossRef]
  11. Grabarz, A.M.; Jędrzejewska, B.; Skotnicka, A.; Murugan, N.A.; Patalas, F.; Bartkowiak, W.; Jacquemin, D.; Ośmiałowski, B. The impact of the heteroatom in a five-membered ring on the photophysical properties of difluoroborates. Dyes Pigm. 2019, 170, 1074812. [Google Scholar] [CrossRef]
  12. Zakrzewska, A.; Zaleśny, R.; Kolehmainen, E.; Ośmiałowski, B.; Jędrzejewska, B.; Ågren, H.; Pietrzak, M. Substituent effects on the photophysical properties of fluorescent 2-benzoylmethylenequinoline difluoroboranes: A combined experimental and quantum chemical study. Dyes Pigm. 2013, 99, 957–965. [Google Scholar] [CrossRef]
  13. Ośmiałowski, B.; Zakrzewska, A.; Jędrzejewska, B.; Grabarz, A.M.; Zaleśny, R.; Bartkowiak, W.; Kolehmainen, E. Influence of substituent and benzoannulation on photophysical properties of 1-benzoylmethyleneisoquinoline difluoroborates. J. Org. Chem. 2015, 80, 2072–2080. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Dzvinchuk, I.B.; Lozinskii, M.O.; Vypirailenko, A.V. C-Mono- and dibenzoylation of 2-methylbenzimidazole with use of benzoyl chloride. Zh. Org. Khim. 1994, 30, 909–914. [Google Scholar]
  15. Skotnicka, A.; Kolehmainen, E.; Czeleń, P.; Valkonen, A.; Gawinecki, R. Synthesis and structural characterization of substituted 2-phenacylbenzoxazoles. Int. J. Mol. Sci. 2013, 14, 4444–4460. [Google Scholar] [CrossRef]
  16. Kubota, Y.; Tanaka, S.; Funabiki, K.; Matsui, M. Synthesis and fluorescence properties of thiazole-boron complexes bearing a β-ketoiminate ligand. Org. Lett. 2012, 14, 4682–4685. [Google Scholar] [CrossRef]
  17. Hansch, C.; Leo, A.; Taft, W. A survey of Hammett substituent constants and resonance and field parameters. Chem. Rev. 1991, 91, 165–195. [Google Scholar] [CrossRef]
  18. Xu, S.; Evans, R.E.; Liu, T.; Zhang, G.; Demas, J.N.; Trindle, C.O.; Frser, C.L. Aromatic difluoroboron β-diketonate complexes: Effects of π-conjugation and media on optical properties. Inorg. Chem. 2013, 52, 3597–3610. [Google Scholar] [CrossRef] [Green Version]
  19. Patil, R.S.; Patil, A.S.; Patil, V.S.; Jirimali, H.D.; Mahulikar, P.P. Synthesis, photophysical, solvatochromic and DFT studies of (Z)-2-(2-Phenyl-4H-benzo[4,5]thiazolo[3,2-a]pyrimidin-4-ylidene)acetonitrile derivatives. J. Lumin. 2019, 210, 303–310. [Google Scholar] [CrossRef]
  20. Chitnis, D.; Thejokalyani, N.; Dhoble, S.J. Exploration of spectroscopic properties of solvated tris(thenoyltrifluoroacetonate)(2,2′-bipyridine)europium(III)red hybrid organic complex for solution processed OLEDs and displays. J. Lumin. 2017, 185, 61–71. [Google Scholar] [CrossRef]
  21. Vennila, P.; Govindaraju, M.; Venkatesh, G.; Kamal, C. Molecular structure, vibrational spectral assignments (FT-IR and FT-RAMAN), NMR, NBO, HOMO-LUMO and NLO properties of O-methoxybenzaldehyde based on DFT calculations. J. Mol. Struct. 2016, 1111, 151–156. [Google Scholar] [CrossRef]
  22. Silverstein, R.M.; Webster, F.X. Spectroscopic Identification of Organic Compound, 6th ed.; John Willey & Sons: New York, NY, USA, 1998. [Google Scholar]
  23. Yoshida, H.; Takeda, K.; Okamura, J.; Ehara, A.; Matsurra, H. A new approach to vibrational analysis of large molecules by density functional theory: Wavenumber-linear scaling method. J. Phys. Chem. A 2002, 106, 3580–3586. [Google Scholar] [CrossRef]
  24. Meech, S.R.; Phillips, D. Photophysics of some common fluorescence standards. J. Photochem. 1983, 23, 193–217. [Google Scholar] [CrossRef]
  25. Brouwer, A.M. Standards for photoluminescence quantum yield measurements in solution (IUPAC Technical Report). Pure Appl. Chem. 2011, 83, 2213–2228. [Google Scholar] [CrossRef] [Green Version]
  26. Jones, G., II; Jackson, W.R.; Choi, C.Y.; Bergmark, W.R. Solvent effects on emission yield and lifetime for coumarin laser dyes. Requirements for a rotatory decay mechanism. J. Phys. Chem. 1985, 89, 294–300. [Google Scholar] [CrossRef]
  27. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  28. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  29. Ong, B.K.; Woon, K.L.; Ariffin, A. Evaluation of various density functionals for predicting the electrophosphorescent host HOMO, LUMO and triplet energies. Synth. Met. 2014, 195, 54–60. [Google Scholar] [CrossRef]
  30. Petersson, G.A.; Bennett, A.; Tensfeldt, T.G.; Al-Laham, M.A.; Shirley, W.A.; Mantzaris, J. A complete basis set model chemistry. I. The total energies of closed-shell atoms and hydrides of the first-row elements. J. Chem. Phys. 1988, 89, 2193–2218. [Google Scholar] [CrossRef]
  31. Petersson, G.A.; Al-Laham, M.A. A complete basis set model chemistry. II. Open-shell systems and the total energies of the first-row atoms. J. Chem. Phys. 1991, 94, 6081–6090. [Google Scholar] [CrossRef]
  32. Frisch, D.J.; Trucks, M.J.; Schlegel, G.W.; Scuseria, H.B.; Robb, G.E.; Cheeseman, M.A.; Scalmani, J.R.; Barone, G.; Petersson, V.; Nakatsuji, G.A.; et al. Gaussian 09 (Revision A. 02); Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  33. Hanwell, M.D.; Curtis, D.E.; Lonie, D.C.; Vandermeerschd, T.; Zurek, E.; Hutchison, G.R. Avogadro: An advanced semantic chemical editor, visualization, and analysis platform. J. Cheminform. 2012, 4, 17. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Scheme 1. Structure and atom numbering in 18.
Scheme 1. Structure and atom numbering in 18.
Molecules 25 05420 sch001
Scheme 2. The schematic representation of the synthesis of 2-phencylbenzoxazole difluoroboranes where R = 4-NMe2 (1), 4-OMe (2), 4-Me (3), 3-Me (4), H (5), 3-OMe (6), 4-Cl (7), 3-Cl (8).
Scheme 2. The schematic representation of the synthesis of 2-phencylbenzoxazole difluoroboranes where R = 4-NMe2 (1), 4-OMe (2), 4-Me (3), 3-Me (4), H (5), 3-OMe (6), 4-Cl (7), 3-Cl (8).
Molecules 25 05420 sch002
Figure 1. A plot of N3 chemical shift [ppm] vs. Hammett constant for 18.
Figure 1. A plot of N3 chemical shift [ppm] vs. Hammett constant for 18.
Molecules 25 05420 g001
Figure 2. Normalized absorption spectra of 18 in chloroform (values were converted to have the maximum normalized at 1).
Figure 2. Normalized absorption spectra of 18 in chloroform (values were converted to have the maximum normalized at 1).
Molecules 25 05420 g002
Figure 3. The normalized UV-Vis absorption (a) fluorescence and (b) spectra of 1 and 5 in CHCl3.
Figure 3. The normalized UV-Vis absorption (a) fluorescence and (b) spectra of 1 and 5 in CHCl3.
Molecules 25 05420 g003
Figure 4. (a) Normalized absorption (a) and fluorescence (b) spectra of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1) in solvents of different polarity.
Figure 4. (a) Normalized absorption (a) and fluorescence (b) spectra of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1) in solvents of different polarity.
Molecules 25 05420 g004
Figure 5. Lippert–Mataga plot of compounds 18.
Figure 5. Lippert–Mataga plot of compounds 18.
Molecules 25 05420 g005
Figure 6. Distribution of HOMO (a) and LUMO (b) orbitals for 2-(4-dimethylamino)benzoyl-benzoxazole difluoroborane (1).
Figure 6. Distribution of HOMO (a) and LUMO (b) orbitals for 2-(4-dimethylamino)benzoyl-benzoxazole difluoroborane (1).
Molecules 25 05420 g006
Figure 7. The observed IR absorption spectra of 2-(4-dimethylamino)benzoylbenzoxazole (1b) in the region 500–2000 cm−1 and 2800–3200 cm−1.
Figure 7. The observed IR absorption spectra of 2-(4-dimethylamino)benzoylbenzoxazole (1b) in the region 500–2000 cm−1 and 2800–3200 cm−1.
Molecules 25 05420 g007
Figure 8. Comparison of the observed with simulated (scaled) IR absorption spectra of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1) in the region 500–2000 cm−1 and 2800–3200 cm−1.
Figure 8. Comparison of the observed with simulated (scaled) IR absorption spectra of 2-(4-dimethylamino)benzoylbenzoxazole difluoroborane (1) in the region 500–2000 cm−1 and 2800–3200 cm−1.
Molecules 25 05420 g008
Table 1. Selected 1H, 11B, 13C, 15N and 19F-NMR chemical shifts of 2-phenacylbenzoxazole difluoroboranes 18 for 0.1–0.2 M solution in CDCl3 at 30 °C.
Table 1. Selected 1H, 11B, 13C, 15N and 19F-NMR chemical shifts of 2-phenacylbenzoxazole difluoroboranes 18 for 0.1–0.2 M solution in CDCl3 at 30 °C.
No.SubstituentC10C11H10B13F13N13
14-NMe277.74172.136.811.79−134.64-
24-OMe78.85172.236.371.80−135.59−215.72
34-Me79.74172.636.441.84−135.34−215.01
43-Me80.33172.706.471.85−135.14−214.18
5H82.11170.796.491.85−135.07−213.54
63-OMe80.56172.286.471.84−134.98−213.97
74-Cl80.63171.016.451.81−134.98−213.00
83-Cl81.17170.646.481.81−134.78−212.43
Table 2. Photophysical properties of examined compounds measured in chloroform.
Table 2. Photophysical properties of examined compounds measured in chloroform.
No.Substituentλabs
(nm)
λfl
(nm)
ε
(M−1·cm−1)
Stokes Shift (cm−1)ϕfl
(× 10−2)
14-NMe241045842,500244998.18
24-OMe36243332,90045301.82
34-Me35442831,21948840.66
43-Me35242322,96147690.48
5H35042427,00049050.54
63-OMe35442926,72749380.57
74-Cl35542829,61548050.58
83-Cl35343029,64850730.50
Table 3. Spectroscopic properties of 2-phenacylbenzoxazole difluoroboranes 18 in solvents of different polarity.
Table 3. Spectroscopic properties of 2-phenacylbenzoxazole difluoroboranes 18 in solvents of different polarity.
No. TolueneCHCl3THFAcOEtAcetoneMeOHDMFMeCNDMSO
1λabs (nm)407410408407420419426415431
λflu (nm)453458465462427473479475486
Stokes (cm−1)249524493004292526232725259730432626
2λabs (nm)362362361359360358363359375
λflu (nm)450433432435432431435463438
Stokes (cm−1)540245304553486746304731456062573836
3λabs (nm)355354354352352351355351356
λflu (nm)434428429432426471431455431
Stokes (cm−1)512748844938526149357259496765124888
4λabs (nm)354352353351351349353350355
λflu (nm)435423427433430424429448430
Stokes (cm−1)526147694909539552345068501962504913
5λabs (nm)354350351349349349352349354
λflu (nm)433424425447433425427447456
Stokes (cm−1)515449055071628255595124499062824993
6λabs (nm)356354354352352351355351357
λflu (nm)436429428431432426432434455
Stokes (cm−1)515449384884520752615016502154484916
7λabs (nm)357355355352353352355352358
λflu (nm)458428456448456463459455460
Stokes (cm−1)643148056239608856016545638264476462
8λabs (nm)354353352351351350353350354
λflu (nm)456430454454454454456452459
Stokes (cm−1)631950376383646464646545639964476462
Table 4. The values of hardness (η), energy gap, and energies of HOMO and LUMO orbitals were estimated for considered molecules.
Table 4. The values of hardness (η), energy gap, and energies of HOMO and LUMO orbitals were estimated for considered molecules.
No.SubstituentHOMO (eV)LUMO (eV)Energy gap (eV)η (eV)
14-NMe2−5.664−2.0903.5741.787
24-OMe−6.168−2.3233.8461.923
34-Me−6.328−2.4233.9061.953
43-Me−6.389−2.4543.9351.967
5H−9.438−2.5043.9341.967
63-OMe−6.390−2.4253.9641.982
74-Cl−6.523−2.6633.8601.930
83-Cl−6.595−2.6843.9101.955
Sample Availability: Samples of the compounds 18 are available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Skotnicka, A.; Czeleń, P. Substituted 2-Phenacylbenzoxazole Difluoroboranes: Synthesis, Structure and Properties. Molecules 2020, 25, 5420. https://doi.org/10.3390/molecules25225420

AMA Style

Skotnicka A, Czeleń P. Substituted 2-Phenacylbenzoxazole Difluoroboranes: Synthesis, Structure and Properties. Molecules. 2020; 25(22):5420. https://doi.org/10.3390/molecules25225420

Chicago/Turabian Style

Skotnicka, Agnieszka, and Przemysław Czeleń. 2020. "Substituted 2-Phenacylbenzoxazole Difluoroboranes: Synthesis, Structure and Properties" Molecules 25, no. 22: 5420. https://doi.org/10.3390/molecules25225420

Article Metrics

Back to TopTop