Next Article in Journal
α-Diimine Cisplatin Derivatives: Synthesis, Structure, Cyclic Voltammetry and Cytotoxicity
Next Article in Special Issue
A 7-Hydroxy 4-Methylcoumarin Enhances Melanogenesis in B16-F10 Melanoma Cells
Previous Article in Journal
Geniposidic Acid from Eucommia ulmoides Oliver Staminate Flower Tea Mitigates Cellular Oxidative Stress via Activating AKT/NRF2 Signaling
Previous Article in Special Issue
Efficient Synthesis of Fluorescent Coumarins and Phosphorous-Containing Coumarin-Type Heterocycles via Palladium Catalyzed Cross-Coupling Reactions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

New Azido Coumarins as Potential Agents for Fluorescent Labeling and Their “Click” Chemistry Reactions for the Conjugation with closo-Dodecaborate Anion

1
A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Str., 119334 Moscow, Russia
2
M.V. Lomonosov Institute of Fine Chemical Technology, MIREA—Technological University, 86 Vernadsky Avenue, 119571 Moscow, Russia
3
N.S. Kurnakov Institute of General and Inorganic Chemistry, Russian Academy of Sciences, 31 Leninsky Avenue, 119991 Moscow, Russia
4
Basic Department of Chemistry of Innovative Materials and Technologies, G.V. Plekhanov Russian University of Economics, 36 Stremyannyi Line, 117997 Moscow, Russia
5
Department of Chemistry, M.V. Lomonosov Moscow State University, Leninskie Gory 1/3, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2022, 27(23), 8575; https://doi.org/10.3390/molecules27238575
Submission received: 22 November 2022 / Accepted: 30 November 2022 / Published: 5 December 2022
(This article belongs to the Special Issue Coumarin and Its Derivatives II)

Abstract

:
Novel fluorescent 7-methoxy- and 7-(diethylamino)-coumarins modified with azido-group on the side chain have been synthesized. Their photophysical properties and single crystals structure characteristics have been studied. In order to demonstrate the possibilities of fluorescent labeling, obtained coumarins have been tested with closo-dodecaborate derivative bearing terminal alkynyl group. CuI catalyzed Huisgen 1,3-dipolar cycloaddition reaction has led to fluorescent conjugates formation. The absorption–emission spectra of the formed conjugates have been presented. The antiproliferative activity and uptake of compounds against several human cell lines were evaluated.

Graphical Abstract

1. Introduction

Coumarins are an important class of natural products that exhibits various types of biological activity, such as anti-inflammatory [1,2,3], anticonvulsant [4], antibacterial [5,6,7], antiviral [8,9,10], anticancer [11,12,13,14,15] and others [16,17,18,19,20,21,22,23,24,25,26,27,28,29]. Along with biological activity, coumarins have been known to possess unique fluorescent properties, especially when substituted with an electron donating group at the C-7 and an electron withdrawing group at the C-3 position of coumarin moiety. Such coumarins have been namely characterized with large Stokes shifts, high quantum yields of fluorescence, and minimal fluorophore degradation, which resulted to their wide use in medicinal chemistry and various biological study as fluorescent probes and labels [30,31,32,33].
Some commercially available and commonly used for labeling fluorescent coumarins are shown in Figure 1. Functionalized with carboxylic groups or with activated esters, coumarins (Figure 1) have been aimed for detection of amino acids, small sized proteins, and antibodies through the peptide bond formation [34,35,36,37,38,39,40,41,42,43]. 7-Diethylamino-coumarin-3-carbohydrazide is used for fluorescence labeling of carbonylated proteins and lipids [44,45,46,47,48].
Since the discovery by Sharpless the term of “click chemistry” for the [3+2] cycloaddition reactions of alkynes with organic azides, which are high-yielding, practical, operationally safe, avoid by-product formation, and proceed in environment friendly solvents at room temperature and generally under mild conditions [49,50], this approach is widely used for synthesis of various bioactive coumarin derivatives [51,52], as well as fluorescent labeling of nanoparticles and biomolecules using azido derivatives of coumarin, such as coumarin 343 azide (Figure 1) [53,54,55].
However, in some cases labelling requires the use of coumarin azides with a shorter spacer between the fluorescent part and the azide function. Herein, we report the synthesis, fluorescence properties, and crystal structures of two new azido derivatives of coumarin and their use for labelling of the closo-dodecaborate anion, including the evaluation of antiproliferative activity and uptake of the synthesized compounds against several human cell lines.

2. Results and Discussion

2.1. Strategy

The sodium salt of the mercapto derivative of the closo-dodecaborate anion Na2[B12H11SH] (BSH) is one of two clinically approved agents for boron neutron capture therapy (BNCT) for cancer [56,57,58], which is an anti-cancer treatment based on the irradiation of 10B-rich tumours with low energy neutrons. The neutron capture reaction produces high linear energy transfer particles, 4He (α-particle) and 7Li, which cause lethal damage to tumour cells through ionization process [59,60,61,62]. This causes increased interest in the synthesis of conjugates of the closo-dodecaborate anion with various biologically active molecules [57,58,63,64,65,66].
An important step in the evaluation of new agents for BNCT is to investigate their uptake in cells in vitro or in larger organisms in vivo. Molecular imaging using optical or radionuclide tags is ideal for small animal and human imaging. However, the use of radionuclide labels [67,68,69,70] is generally limited by the availability of appropriate radionuclides and the necessary special equipment for working with radioactive substances. Therefore, various fluorescent labels, such as meso-(4-hydroxyphenyl)-4,4-difluoro- 4-bora-3a,4a-diaza-s-indacene (BODIPY) [71], coumarin [72,73,74], and (3,6-diamino-9-(2,5- dicarboxyphenyl)-4,5-disulfo-Xanthylium) (Alexa Flour 488®) [75], are commonly used to label the closo-dodecaborate anion and its derivatives.
An effective approach to the functionalization of the closo-dodecaborate anion is the opening of its cyclic oxonium derivatives by nucleophiles [76]. This approach is widely used for the synthesis of its various functional derivatives and attachment to biomolecules [64,77,78], and has also been used to modify other polyhedral boron hydrides such as the closo-decaborate anion [64], nido-carborane [79], and cobalt and iron bis(dicarbollides) [80]. In particular, the synthesis of the earlier described coumarin derivatives of the closo-dodecaborate anion is based on the opening of oxonium rings with hydroxy groups at various positions of the coumarin heterocycle [72,73,74]. This leads to a significant difference in the photophysical characteristics of the obtained compounds from the commonly used fluorescent labels (see above). Therefore, our goal was to preserve the structure of coumarins traditionally used for labeling. We considered it expedient to use the [3+2]-cycloaddition reaction for the synthesis of coumarin derivatives of the closo-dodecaborate anion. The “click” chemistry reaction has found a wide synthetic application for the preparation of a wide range of conjugates based on the boron cluster [81,82,83,84,85,86,87,88,89,90]. However, the opening of the 1,4-dioxane derivative results in moieties where the boron cluster connected with functional fragment via a biologically compatible flexible di(ethylene glycol) spacer. Thus, in particular, derivatives of the closo-dodecaborate anion with a terminal alkynyl group were obtained [91,92]. Therefore, the use of commercially available coumarin 343 azide will result in a product with an unnecessarily long spacer between the boron cluster and the fluorescent label.
Taking into account these limitations, we decided to modify the known 7-diethylamino- and 7-methoxy-coumarin-3-carboxylic acids [40,41] by forming amides with 2-azidoethylamine.

2.2. Synthesis

The target coumarins functionalized with azido group 2a,b were synthesized as shown in Scheme 1.
The starting 7-methoxycoumarin-3-carboxylic acid 1a and 7-diethylaminocoumarin- 3-carboxylic acid 1b were prepared according to the literature procedures by the Knoevenagel condensation reaction of 4-substituted salicylaldehyde with diethyl malonate [93], or with ethyl cyanoacetate [94], respectively, followed by hydrolysis reactions. Coumarins 1a,b undergo amidation reaction with 2-azidoethylamine hydrochloride in refluxing acetonitrile in the presence of Et3N as a base. The reaction progress was monitored by thin layer chromatography. It should be noted that, after the cooling of the reaction mixture, the products precipitate from the reaction mixture, and in the case of derivative 2a, in the form of crystals suitable for single crystal X-ray diffraction study. The structures of the obtained compounds 2a,b were established using IR, 1H NMR, and 13C NMR spectroscopy (see Supplementary Materials). The 1H NMR spectra of each compound in CDCl3 contain, in addition to characteristic signals of the coumarin moiety, broad triplets of the NH-amide protons at 9.0 ppm and signals from the two side chain methylene groups at ~3.6 and 3.5 ppm. In the IR spectra, the characteristic absorption bands of the N3 group appear at 2125 and 2098 cm−1 for 2a and 2b, respectively.
To demonstrate the ability of novel azido coumarins 2a,b to form conjugates with fluorescent properties, we have tested them in reactions with a closo-dodecaborate derivative containing a terminal alkynyl group in the side chain 3[NBu4], under typical “click” reaction conditions. As a result, new 1,2,3-triazoles 4a[Cs] and 4b[Cs] were obtained in high yields (Scheme 2).
The formation of conjugates 4a,b was monitored by the TLC method, and the full conversion was gained in 16 h. The conjugation products were isolated by precipitation in the form of the Cs+ salts and characterized by 1H, 11B, 13C NMR, and IR spectroscopy as well as high-resolution mass spectrometry (see Supplementary Materials). The formation of the 1,2,3-triazole heterocycle results in the appearance of the corresponding singlets at ~ 8.45 ppm in the 1H NMR spectra. In the IR spectra of 4a,b, the absorption bands of the BH stretching (2487 cm−1) demonstrate the presence of the boron hydride cluster, and absorption bands corresponding to the 1,2,3-triazole heterocycle (1705 and 1710 cm−1, respectively) were detected. In contrast to azido-coumarins 2a,b, fluorescent conjugates 4a,b possess a good solubility in water.

2.3. Single Crystal X-ray Diffraction Study of Azido Derivatives of Coumarins

The molecular structure of the obtained azido derivatives of coumarins 2a and 2b were determined by a single crystals X-ray diffraction study (Figure 2).
The amide fragments in both compounds are expectedly planar: the mean deviation of non-hydrogen atoms from the corresponding mean-square planes are only 0.02 and 0.04 Å for 2a and 2b, respectively. The most significant deviations from the coumarin-acetylamide planes are observed for the atoms of amide moiety (0.03 Å for N1 in 2a and 0.10 Å for O3 in 2b) that can be rationalized by the formation of intramolecular N1(H)…O1 hydrogen bonds between the amide hydrogen and the coumarin carbonyl (∠(NHO) 138.1° and 131.6°, N…O 2.751 and 2.749 Å in 2a and 2b, respectively), which is typical for primary 3-arylamido- and 3-hydrazidocoumarins [95,96,97,98]. The conformations of the azidoethyl fragments are pronouncedly different in 2a and 2b: while the anti-conformation is observed for the substituents at C11-C12 bond in 2a (the N1-C11-C12-N2 torsion angle is 170.5(1)°), the gosh-conformation is observed in 2b (the corresponding torsion angle is 61.2(1)°). In its turn, the N3 group is antiperiplanar to the C11-C12 bond in 2a (175.0(1)°) and to the C12-H12A bond in 2b (173.8(1)°). This difference can be readily explained by the peculiarities of crystal packing (Figure 3): in 2b, the C12(H) group participates in CH…O contacts with O2 oxygen atoms of two neighboring molecules, whereas the only meaningful interaction for the azidoethyl fragment in 2a is the stacking interaction between N3 moieties (N…N 3.07 Å), which were also observed in some other organic compounds with the azido group [99,100]. In general, the morphology of crystal packings of 2a and 2b are still similar (see Supplementary Materials, Figure S1): the layer-type arrangement of molecules is stabilized by stacking interactions between coumarin acylamide fragments and CH…O hydrogen bonds.

2.4. Photophysical Properties

The UV–vis spectra of compounds 2a and 2b were measured in acetonitrile. In contrast, the derivatives 4a and 4a are soluble in water, and their absorption spectra were measured in water correspondingly (Figure 4). All compounds demonstrate intense bands in the short wavelength region of the π → π* character within the coumarin aromatic system. Differences at the absorption maxima suggest that the donor–acceptor properties of the 7-substituted group affect the excitation energy. Compounds 2a and 4a with -OMe substituent demonstrate bands at 350 and 343 nm respectively. Compounds 2b and 4b containing -NEt2 substituent demonstrate bands in the lower energy region (413 and 430 nm, respectively). Compounds 2a, 2b in CH3CN, and 4a, 4b in water demonstrate unstructured broad bands centered at ca. 400 and 470 nm for OMe and NEt2 substituted analogues, respectively (Figure 5, Table 1). The emission observed is a fluorescence, being typical for substituted coumarins [101,102]. Differences at the emission maxima demonstrate the effects of the donor properties of substituent on the emission. The 7–10 nm of the wavelength difference between 2a and 4a as well as between 2b and 4b may be explained by the solvent effect.

2.5. Antiproliferative Activity and Cell Uptake of Boronated Coumarins

The antiproliferative activity of compounds 4a and 4b against the human HCT116 colorectal carcinoma, MCF7 breast adenocarcinoma, A549 non-small cell lung carcinoma and WI38 nonmalignant lung fibroblast cell lines was evaluated by means of a standard MTT colorimetric assay as the IC50 value after 72 h of incubation. Compounds 4a and 4b were found to be not active up to 200 µM (See Supplementary Materials) and can be considered as non-toxic against cancerous and non-malignant cells. Therefore, they could be of potential interest for use in BNCT for cancer. Cellular accumulation plays a fundamental role in the antiproliferative activity of low-molecular-mass compounds [103,104]. One of the main requirements of effective BNCT is the selective accumulation of boron-containing compounds in the tumor with a fairly high concentration (approx. 109 10B atoms/cell or 20–35 µg/gram [60]). Therefore, we measured uptake of compounds 4a and 4b at 250 µM concentration in A549 and WI38 cell lines for 1, 4, 6 and 24 h; the amount of boron was determined by ICP-MS (Figure 6).
The accumulation of compound 4b in both cancerous and normal cells occurs noticeably faster than compound 4a and reaches a maximum after 6 h for the WI38 cell line and after 24 h for the A549 cell line. At the same time, both compounds accumulated slightly better in the WI38 cell line. The cells of many tumours are known to be much larger than normal cells [105,106]; however, the close size of the A549 cells [107] and red blood cells [108] allows a rough estimate of the accumulation of the compound 4b in the cancer cells, which does not exceed 10 μg/g, which is apparently insufficient for effective BNCT.

3. Materials and Methods

Chemicals were reagent grade and received from commercial vendors. Acetonitrile was distilled from P2O5 and then from CaH2. Compounds 1a [93], 1b [94], 2-azidoethylamine hydrochloride [109] and 3 [92] were prepared according to the literature procedures. All reactions were carried out in air. Analytical TLC was performed on Kieselgel 60 F245 (Merck, Darmstadt, Germany) plates. Boron compounds were visualized with PdCl2 stain solution, which upon heating gave dark brown spots. 1H, 13C and 11B NMR spectra were recorded at 400.13, 100.61 and 128.38 MHz, respectively, on a BRUKER-Avance-400 spectrometer (Bruker, Karlsruhe-Zurich, Switzerland-Germany). Tetramethylsilane and BF3 × Et2O were used as standards for 1H and 13C NMR and 11B NMR, respectively. All chemical shifts are reported in ppm (δ) relative to external standards. IR spectra were recorded on IR Prestige-21 (Shimadzu, Kyoto, Japan) instrument. The UV–vis spectra of solutions were measured on Cary 50 Uv–Vis Spectrophotometer (Varian, Palo Alto, CA, USA). The photoluminescence spectra were recorded on the Spectrofluorometer RF-6000 (Shimadzu). High resolution mass spectra (HRMS) were measured on a Bruker micrOTOF II instrument (Bruker Daltonic, Bremen, Germany) using electrospray ionization (ESI). For anionic boron containing compounds, the measurements were performed in a negative ion mode (3200 V) with a mass range from m/z 50 to m/z 3000; external and internal calibration was performed with ESI Tuning Mix (Agilent, Santa Clara, CA, USA). The boron content in cells was determined by inductively coupled plasma mass spectrometry (ICP-MS) using X-II spectrometer (Thermo Scientific, Waltham, MA, USA) and the following parameters: RF generator power 1400 W, nebulizer PolyCon, spray chamber cooling 3 °C, plasma gas flow rate 12 L/min, auxiliary flow rate 0.9 L/min, nebulizer flow rate 0.9 L/min, sample update 0.8 mL/min, resolution 0.8 M. The main parameters of MS measurements were: detector mode double (pulse counting and analogous) and scanning mode Survey Scan and Peak Jumping. The settings for the Survey Scan were: number of runs 10, dwell time 0.6 ms, channels per mass 10, acquisition 13.2 s. The settings for the Peak Jumping were: sweeps 25, dwell time 10 ms, channels per mass 1, acquisition 34 s.

3.1. Synthetic Procedures

3.1.1. N-(2-azidoethyl)-7-methoxy-2-oxo-2H-chromene-3-carboxamide 2a

To a suspension of 3-carboxy-7-methoxycoumarin 1a (0.5 g, 2.3 mmol) and 2-azidoethanamine hydrochloride (0.31 g, 2.5 mmol) in 50 mL of acetonitrile, NEt3 was added in excess (approximately 0.5 mL). After the amine addition, the solution became clear. The reaction mixture was stirred under reflux for 16 h. Pale yellow crystals were formed after a slow cooling to room temperature; they were filtered washed with water (5 mL), cold CH3CN (5 mL) and air dried to give 0.32 g of 2. Another portion of 2a was purchased from the mother liquor by recrystallisation from EtOH to give while combining 0.40 g of 2a. Colorless solid. Yield: 61%.
1H NMR (Chloroform-d, δ, ppm): δ 9.03 (s, 1H, NH), 8.85 (s, 1H, H-4), 7.61 (d, J = 8.7 Hz, 1H, H-5), 6.67 (dd, J = 8.7, 2.4 Hz, 1H, H-6), 6.88 (d, J = 2.5 Hz, 1H, H-8), 3.93 (s, 3H, OMe), 3.66 (q, J = 5.8 Hz, 2H, NHCH2CH2), 3.56 (t, J = 5.8 Hz, 2H, NHCH2CH2). 13C NMR (Chloroform-d, δ, ppm): 165.0, 162.5 (CO), 161.8 (C*), 156.7 (C-7), 148.5 (C-4), 131.0 (C-5), 114.3 (C-6), 114.1 (C*), 112.3 (C-3), 100.3 (C-8), 56.1 (OCH3), 50.6 (NCH2CH2), 39.1 (NCH2CH2). IR (KBr, ν, cm−1): δ 3353 (NH), 3086, 3050, 2986, 2946 (CH), 2125 (N3), 1713, 1658 (CO), 1609 (C=C).

3.1.2. N-(2-azidoethyl)-7-(diethylamino)-2-oxo-2H-chromene-3-carboxamide 2b

To a suspension of 3-carboxy-7-diethylaminocoumarin 1b (0.5 g, 1.9 mmol) and 2-azidoethanamine hydrochloride (0.26 g, 2.1 mmol) in 50 mL of acetonitrile, NEt3 was added in excess (approximately 0.5 mL). After the amine addition, the solution became clear. The reaction mixture was stirred under reflux for 16 h. Then it was allowed to cool to room temperature. Yellow precipitate was formed after a cooling of the reaction mixture to 0 °C; it was filtered washed with water (5 mL), cold CH3CN (5 mL) and air dried. Mother liquor was evaporated to dryness and the oily residue was recrystallized from the EtOAc-EtOH mixture to give an additional portion of the title product. Orange solid. Yield: 63%. 0.4 g Crystals were obtained from the EtOAc solution while in slow evaporation.
1H NMR (Chloroform-d, δ, ppm): δ 9.06 (s, 1H, NH), 8.72 (s, 1H, H-4), 7.45 (d, J = 9.0 Hz, 1H, H-5), 6.67 (dd, J = 8.9, 2.5 Hz, 1H, H-6), 6.52 (d, J = 2.5 Hz, 1H, H-8), 3.65 (q, J = 5.8 Hz, 2H, NHCH2CH2), 3.55 (t, J = 5.9 Hz, 2H, NHCH2CH2), 3.48 (q, J = 7.1 Hz, 4H, NCH2CH3), 1.26 (t, J = 7.1 Hz, 6H, NCH2CH3). 13C NMR (Chloroform-d, δ, ppm): 163.6, 162.7 (CO), 157.7 (C*), 152.7 (C-7), 148.3 (C-4), 131.2 (C-5), 110.0 (C-6), 109.8 (C*), 108.3 (C-3), 96.6 (C-8), 50.7 (NCH2CH2), 45.1 (CH2CH3), 39.0 (NCH2CH3), 12.4 (CH2CH3). IR (KBr, ν, cm−1): δ 3357 (NH), 2986, 2942, 2883(CH), 2097 (N3), 1697, 1649 (CO), 1626 (C=C), 1590, 1525.

3.1.3. Synthesis of Conjugate 4a

A mixture of alkyne 3 (NBu4) (0.1 g, 0.2 mmol), 2a (0.052 g, 0.2 mmol), Et3N (0.5 mL) and CuI (0.01 g, 0.04 mmol) in 15 mL of CH3CN was stirred under reflux for 16 h. The reaction mixture was allowed to cool to room temperature and passed through a layer of silica gel (2–3 cm) on a Schott filter. The system was washed with CH3CN until the product ceased to be detected by thin layer chromatography. The solvent was removed on a rotary evaporator. The residue was dissolved in MeOH (10 mL) and an excess of CsF solution in MeOH was added. Formed precipitate was filtered, washed with MeOH (20 mL) and air dried to yield title 4a as a pale-yellow solid. 0.12 g. Yield: 91%.
1H NMR (DMSO-d6, δ, ppm): 8.83 (s, 1H, NH), 8.73 (s, 1H, H-4), 8.46 (s, 1H, CH-triazole), 7.93 (d, J = 8.7 Hz, 1H, H-5), 7.13 (d, J = 2.5 Hz, 1H, H-8), 7.05 (dd, J = 8.9 Hz, 1H, H-6), 4.71 (s, 2H, N+CH2), 4.63 (m, 2H, N+CH2-spacer), 3.91 (m, 5H, BOCH2 + OCH3), 3.81 (m, 2H, CONCH2CH2), 3.46 (s, 6H, 2×CH2-spacer + CONCH2CH2), 3.10 (s, 6H, N+CH3). 11B NMR (DMSO-d6, δ, ppm): 6.3 (s,1B, B(1)); −16.8 (d, J = 149 Hz, 5B, B(2-6)); −18.1 (d, J = 158 Hz, 5B, B(7-11)); −22.7 (d, J = not resolved, 1B, B(12)). 13C NMR (DMSO-d6, δ, ppm): 165.0, 162.3 (CO), 161.0 (C*), 156.7 (C-7), 148.4 (C-4), 136.0 (C-triazole), 132.2 (CH-triazole), 129.5 (C-5), 114.9 (C-6), 114.2 (C*), 112.5 (C-3), 100.8 (C-8), 72.3, 68.3, 64.7 (OCH2), 62.8, 58.8 (N+CH2), 56.8 (NCH2CH2), 51.1 (N+CH3), 49.9 (NCH2CH2). IR (KBr, ν, cm−1): 3437 (NH), 2942, 2867 (CH), 2487 (BH), 1705 (triazole), 1617 (CO). ESI-MS, m/z, C22H40B12N5O6 calcd. 600.4185 [M], found 600.4185 [M].

3.1.4. Synthesis of Conjugates 4b

A mixture of alkyne 3 (NBu4) (0.1 g, 0.2 mmol), 2b (0.06 g, 0.2 mmol), Et3N (0.5 mL) and CuI (0.01 g, 0.04 mmol) in 15 mL of CH3CN was stirred under reflux for 16 h. Then it was allowed to cool to room temperature and passed through a layer of silica gel (2–3 cm) on a Schott filter. The system was washed with CH3CN until the product ceased to be detected by thin layer chromatography. The solvent was removed on a rotary evaporator. The residue was dissolved in MeOH (10 mL) and an excess of CsF solution in MeOH was added. Formed precipitate was filtered (attention! very fine powder), washed with MeOH (20 mL) and air dried to yield title 4b as yellow solid. 0.09 g. Yield: 63%.
1H NMR (DMSO-d6, δ, ppm): δ 8.78 (t, J = 5.7 Hz, 1H, NH), 8.54 (s, 1H, H-4), 8.44 (s, 1H, CH-triazole), 7.69 (d, J = 9.0 Hz, 1H, H-5), 6.80 (m, 1H, H-6), 6.62 (d, J = 2.4 Hz, 1H, H-8), 4.71 (s, 2H, N+CH2), 4.62 (t, J = 5.7 Hz, 2H, N+CH2-spacer), 3.92 (m, BOCH2), 3.80 (m, 2H, CONCH2CH2), 3.47 (m, 10H, 2×CH2-spacer, CONCH2CH2, NCH2CH3), 3.10 (s, 6H, N+CH3), 1.15 (t, J = 7.0 Hz, 6H, NCH2CH3), 1.7–0.5 (broad m. 11H, BH). 11B NMR (DMSO-d6, δ, ppm): 6.4 (s,1B, B(1)); −16.8 (d, J = 151 Hz, 5B, B(2-6)); −18.1 (d, J = 152 Hz, 5B, B(7-11)); −22.7 (d, J = not resolved, 1B, B(12)). 13C NMR (DMSO-d6, δ, ppm): 163.1, 161.9 (CO), 157.7 (C*), 153.0 (C-7), 148.2 (C-4), 136.0 (C-triazole), 132.2 (CH-triazole), 129.4 (C-5), 110.7 (C-6), 109.3 (C*), 108.0 (C-3), 96.3 (C-8), 72.3, 68.3, 64.7 (OCH2), 62.8, 58.86 (N+CH2), 51.1 (N+CH3), 50.1 (NCH2CH2), 44.8 (NCH2CH3), 12.8 (NCH2CH3). IR (KBr, ν, cm−1): 3337 (NH), 2978, 2939, 2874 (CH), 2487 (BH), 1710 (triazole). ESI-MS, m/z, C25H47B12N6O5 calcd. 641.4815 [M], found 641.4812 [M].

3.2. In Vitro Antiproliferative Assays and Cellular Uptake

The human HCT116 colorectal carcinoma, MCF7 breast adenocarcinoma, A549 non-small cell lung carcinoma, and WI38 nonmalignant lung fibroblast cell lines were obtained from the European collection of authenticated cell cultures (ECACC; Salisbury, UK). All cells were grown in DMEM medium (Gibco™, Irland) supplemented with 10% fetal bovine serum (Gibco™, Brazil). The cells were cultured in an incubator at 37 °C in a humidified 5% CO2 atmosphere and subcultured 2 times a week. The effect of the investigated compounds on cell proliferation was evaluated using a common MTT assay. The cells were seeded in 96-well tissue culture plates («TPP», Switzerland) at a 1 × 104 cells/well in 100 µL of the medium. After overnight incubation at 37 °C, the cells were treated with the tested compounds in the concentration range of 0 to 200 µM. Cisplatin was used as a standard. After 72 h of treatment, solution was removed, a freshly diluted MTT solution (100 µL, 0.5 mg/mL in cell medium) was added to the wells, and the plates were further incubated for 50 min. Subsequently, the medium was removed, and the formazan product was dissolved in 100 μL of DMSO. The number of living cells in each well was evaluated by measuring the absorbance at 570 nm using the «Zenith 200 rt» microplate reader (Biochrom, Cambridge, UK). The cellular accumulation of the studied compounds was determined as described earlier [110]. Briefly, A549 and WI38 cells were seeded in 25 cm2 cell flask (1 × 106 cells per flask). After 48 pre-incubations, the cells were exposed to compounds for 1, 4, 6, and 24 h. After the treatment, the cells were detached by trypsinization, exhaustively washed with ice-cold PBS (10mM, pH 7.4), counted by using cell counter, and collected by centrifugation. Finally, the cell pellets were digested, and the quantity of boron taken up by the cells was determined by ICP-MS.

3.3. Crystallography

X-ray diffraction data for crystals of 2a and 2b were measured using the Bruker APEX II diffractometer (Bruker AXC, Maddison, WI, USA) equipped with the Photon 2 detector (MoKα-radiation, graphite monochromator, ω-scans). The intensity data were integrated by the SAINT program [111] and were corrected for absorption and decay using SADABS [112]. Both structures were solved by direct methods using SHELXS [113] and refined using the full matrix least squares technique against F2 using SHELXL-2018. Positions of hydrogen atoms were found from the difference Fourier synthesis of electron density. Non-hydrogen and hydrogen atoms were refined in the anisotropic and isotropic approximations, respectively. The main refinement details and parameters are given in the Table S1 in Supporting Information. CCDC 2173129-2173130 contain all additional supplementary data.

4. Conclusions

We have reported synthesis, X-ray structure study and photophysical properties of novel coumarins functionalized with azido group which are suitable for the fluorescent labeling of compounds containing terminal alkynyl group. Their representative application was performed on the “click” reaction with the closo-dodecaborate based alkynyl derivative. The antiproliferative activity and uptake of the synthesized boron compounds against several human cell lines were evaluated. Both compounds were found to be non-toxic against cancerous and non-malignant cells, however their cellar uptake is not sufficient for effective BNCT treatment.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules27238575/s1. 1H, 11B{1H}, 11B, 13C NMR and IR spectra of compounds 2a,b and 4a,b; ESI-HRMS spectra of compounds 4a,b; Table S1 with main crystallographic data for 2a and 2b; Table S2 with IC50 values for compounds 4a,b and cisplatin; Figure S1 with crystal packing of 2a,b.

Author Contributions

Conceptualization, J.L. and A.D.; methodology, J.L, A.S. and I.A.; validation, A.S, A.A.A. and E.T.; formal analysis, I.K.; writing—original draft, J.L.; review and editing, I.S., A.A.N. and V.I.B.; funding acquisition, A.D. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Foundation for Basic Research (RFBR) (synthesis and identification of new compounds—project No 20-03-00251) and the Russian Science Foundation (RSF) (biological study of new compounds—project No 22-63-00016).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The data presented in this study are available in Supplementary Materials.

Acknowledgments

NMR studies, spectral characterization, elemental analysis performed using equipment of the Center for Molecular Structure Studies at A.N. Nesmeyanov Institute of Organoelement Compounds operating with financial support of Ministry of Science and Higher Education of the Russian Federation. I.V. Ananyev is grateful to the support of the structural study by the Ministry of Science and Higher Education of the Russian Federation as part of the State Assignment of the N.S. Kurnakov Institute of General and Inorganic Chemistry of the Russian Academy of Sciences.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are available from the authors.

References

  1. Bansal, Y.; Sethi, P.; Bansal, G. CoumarIn A potential nucleus for anti-inflammatory molecules. Med. Chem. Res. 2013, 22, 3049–3060. [Google Scholar] [CrossRef]
  2. Grovera, J.; Jachak, S.M. Coumarins as privileged scaffold for anti-inflammatory drug development. RSC Adv. 2015, 5, 38892–38905. [Google Scholar] [CrossRef]
  3. Rostom, B.; Karaky, R.; Kassab, I.; Sylla-Iyarreta Veitía, M. Coumarins derivatives and inflammation: Review of their effects on the inflammatory signaling pathways. Eur. J. Pharm. 2022, 922, 174867. [Google Scholar] [CrossRef] [PubMed]
  4. Keri, R.S.; Budagumpi, S.; Somappa, S.B. Synthetic and natural coumarins as potent anticonvulsant agents: A review with structure–activity relationship. J. Clin. Pharm. Therap. 2022, 47, 915–931. [Google Scholar] [CrossRef]
  5. Qin, H.-L.; Zhang, Z.-W.; Ravindar, L.; Rakesh, K.P. Antibacterial activities with the structure-activity relationship of coumarin derivatives. Eur. J. Med. Chem. 2020, 207, 112832. [Google Scholar] [CrossRef]
  6. Feng, D.; Zhang, A.; Yang, Y.; Yang, P. Coumarin-containing hybrids and their antibacterial activities. Arch. Pharm. 2020, 353, e1900380. [Google Scholar] [CrossRef]
  7. Sahoo, C.R.; Sahoo, J.; Mahapatra, M.; Lenka, D.; Sahu, P.K.; Dehury, B.; Padhy, R.N.; Paidesetty, S.K. Coumarin derivatives as promising antibacterial agent(s). Arab. J. Chem. 2021, 14, 102922. [Google Scholar] [CrossRef]
  8. Hassan, M.Z.; Osman, H.; Ali, M.A.; Ahsan, M.J. Therapeutic potential of coumarins as antiviral agents. Eur. J. Med. Chem. 2016, 123, 236–255. [Google Scholar] [CrossRef]
  9. Mishra, S.; Pandey, A.; Manvati, S. CoumarIn An emerging antiviral agent. Heliyon 2020, 6, e03217. [Google Scholar] [CrossRef] [Green Version]
  10. Li, Z.; Kong, D.; Liu, Y.; Li, M. Pharmacological perspectives and molecular mechanisms of coumarin derivatives against virus disease. Genes Deseases 2022, 9, 80–94. [Google Scholar] [CrossRef]
  11. Lacy, A.; O’Kennedy, R. Studies on coumarins and coumarin-related compounds to determine their therapeutic role in the treatment of cancer. Curr. Pharm. Des. 2004, 10, 3797–3811. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Musa, M.A.; Cooperwood, J.S.; Khan, M.O.F. A review of coumarin derivatives in pharmacotherapy of breast cancer. Curr. Med. Chem. 2008, 15, 2664–2679. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Thakur, A.; Singla, R.; Jaitak, V. Coumarins as anticancer agents: A review on synthetic strategies, mechanism of action and SAR studies. Eur. J. Med. Chem. 2015, 101, 476–495. [Google Scholar] [CrossRef] [PubMed]
  14. Menezes, J.C.J.M.D.S.; Diederich, M. Translational role of natural coumarins and their derivatives as anticancer agents. Futur. Med. Chem. 2019, 11, 1057–1082. [Google Scholar] [CrossRef] [PubMed]
  15. Dorababu, A. Coumarin-heterocycle framework: A privileged approach in promising anticancer drug design. Eur. J. Med. Chem. Rep. 2021, 2, 100006. [Google Scholar] [CrossRef]
  16. Kostova, I.; Bhatia, S.; Grigorov, P.; Balkansky, S.; Parmar, V.S.; Prasad, A.K.; Saso, L. Coumarins as antioxidants. Curr. Med. Chem. 2011, 18, 3929–3951. [Google Scholar] [CrossRef]
  17. Peng, X.-M.; Damu, G.L.V.; Zhou, C.-H. Current developments of coumarin compounds in medicinal chemistry. Curr. Pharm. Design 2013, 19, 3884–3930. [Google Scholar] [CrossRef]
  18. Sandhu, S.; Bansal, Y.; Silakari, O.; Bansal, G. Coumarin hybrids as novel therapeutic agents. Bioorg. Med. Chem. 2014, 22, 3806–3814. [Google Scholar] [CrossRef]
  19. Revankar, H.M.; Bukhari, S.N.A.; Kumar, G.B.; Qin, H.-L. Coumarins scaffolds as COX inhibitors. Bioorg. Chem. 2017, 71, 146–159. [Google Scholar] [CrossRef]
  20. Fotopoulos, I.; Hadjipavlou-Litina, D. Hybrids of coumarin derivatives as potent and multifunctional bioactive agents: A review. Med. Chem. 2020, 16, 272–306. [Google Scholar] [CrossRef]
  21. Bouhaoui, A.; Eddahmi, M.; Dib, M.; Khouili, M.; Aires, A.; Catto, M.; Bouissane, L. Synthesis and biological properties of coumarin derivatives. A review. Chem. Select 2021, 6, 5848–5870. [Google Scholar] [CrossRef]
  22. Katopodi, A.; Tsotsou, E.; Iliou, T.; Deligiannidou, G.-E.; Pontiki, E.; Kontogiorgis, C.; Tsopelas, F.; Detsi, A. Synthesis, bioactivity, pharmacokinetic and biomimetic properties of multi-substituted coumarin derivatives. Molecules 2021, 26, 5999. [Google Scholar] [CrossRef] [PubMed]
  23. Wang, X.; Zhou, H.; Wang, X.; Lei, K.; Wang, S. Design, synthesis, and in vivo and in silico evaluation of coumarin derivatives with potential antidepressant effects. Molecules 2021, 26, 5556. [Google Scholar] [CrossRef]
  24. Koyiparambath, V.P.; Rajappan, K.P.; Rangarajan, T.M.; Al-Sehemi, A.G.; Pannipara, M.; Bhaskar, V.; Nair, A.S.; Sudevan, S.T.; Kumar, S.; Mathew, B. Deciphering the detailed structure–activity relationship of coumarins as monoamine oxidase enzyme inhibitors—An updated review. Chem. Biol. Drug Design 2021, 98, 655–673. [Google Scholar] [CrossRef]
  25. Agić, D.; Karnaš, M.; Šubarić, D.; Lončarić, M.; Tomić, S.; Karačić, Z.; Bešlo, D.; Rastija, V.; Molnar, M.; Popović, B.M.; et al. Coumarin derivatives act as novel inhibitors of human dipeptidyl peptidase III: Combined in vitro and in silico study. Pharmaceuticals 2021, 14, 540. [Google Scholar] [CrossRef]
  26. Dar’in, D.; Kantin, G.; Kalinin, S.; Sharonova, T.; Bunev, A.; Ostapenko, G.I.; Nocentini, A.; Sharoyko, V.; Supuran, C.T.; Krasavin, M. Investigation of 3-sulfamoyl coumarins against cancer-related IX and XII isoforms of human carbonic anhydrase as well as cancer cells leads to the discovery of 2-oxo-2H-benzo[h]chromene-3-sulfonamide—A new caspase-activating proapoptotic agent. Eur. J. Med. Chem. 2021, 222, 113589. [Google Scholar] [CrossRef] [PubMed]
  27. Balewski, L.; Szulta, S.; Jalińska, A.; Kornicka, A. A mini-review: Recent advances in coumarin-metal complexes with biological properties. Front. Chem. 2021, 9, 781779. [Google Scholar] [CrossRef]
  28. Patil, S.B. Medicinal significance of novel coumarin analogs: Recent studies. Results Chem. 2022, 4, 100313. [Google Scholar] [CrossRef]
  29. Pisani, L.; Catto, M.; Muncipinto, G.; Nicolotti, O.; Carrieri, A.; Rullo, M.; Stefanachi, A.; Leonetti, F.; Altomare, C. A twenty-year journey exploring coumarin-based derivatives as bioactive molecules. Front. Chem. 2022, 10, 1002547. [Google Scholar] [CrossRef]
  30. Liu, X.; Xu, Z.; Cole, J.M. Molecular design of UV–vis absorption and emission properties in organic fluorophores: Toward larger bathochromic shifts, enhanced molar extinction coefficients, and greater Stokes shifts. J. Phys. Chem. C 2013, 117, 16584–16595. [Google Scholar] [CrossRef]
  31. Fu, Y.; Finney, N.S. Small-molecule fluorescent probes and their design. RSC Adv. 2018, 8, 29051–29061. [Google Scholar] [CrossRef] [Green Version]
  32. Cao, D.; Liu, Z.; Verwilst, P.; Koo, S.; Jangjili, P.; Kim, J.S.; Lin, W. Coumarin-based small-molecule fluorescent chemosensors. Chem. Rev. 2019, 119, 10403–10519. [Google Scholar] [CrossRef] [PubMed]
  33. Pereira, A.; Martins, S.; Caldeira, A.T. Coumarins as fluorescent labels of biomolecules. In Phytochemicals in Human Health; Rao, V., Mans, D., Rao., L., Eds.; IntechOpen: London, UK, 2020. [Google Scholar] [CrossRef] [Green Version]
  34. Berthelot, T.; Talbot, J.-C.; Laïn, G.; Déleris, G.; Latxague, L. Synthesis of Nε-(7-diethylaminocoumarin-3-carboxyl)- and Nε-(7-methoxycoumarin-3-carboxyl)-L-fmoc lysine as tools for protease cleavage detection by fluorescence. J. Peptide Sci. 2005, 11, 153–160. [Google Scholar] [CrossRef] [PubMed]
  35. Song, H.Y.; Ngai, M.H.; Song, Z.Y.; MacAry, P.A.; Hobley, J.; Lear, M.J. Practical synthesis of maleimides and coumarin-linked probes for protein and antibody labelling via reduction of native disulfides. Org. Biomol. Chem. 2009, 7, 3400–3406. [Google Scholar] [CrossRef]
  36. Zhou, Y.; Yoon, J. Recent progress in fluorescent and colorimetric chemosensors for detection of amino acids. Chem. Soc. Rev. 2012, 41, 52–67. [Google Scholar] [CrossRef] [PubMed]
  37. Kohl, F.; Schmitz, J.; Furtmann, N.; Schulz-Fincke, A.-C.; Mertens, M.D.; Küppers, J.; Benkhoff, M.; Tobiasch, E.; Bartz, U.; Bajorath, J.; et al. Design, characterization and cellular uptake studies of fluorescence-labeled prototypic cathepsin inhibitors. Org. Biomol. Chem. 2015, 13, 10310–10323. [Google Scholar] [CrossRef] [PubMed]
  38. Häußler, D.; Gütschow, M. Fluorescently labeled amino acids as building blocks for bioactive molecules. Synthesis 2016, 48, 245–255. [Google Scholar] [CrossRef]
  39. Keuler, T.; Gatterdam, K.; Akbal, A.; Lovotti, M.; Marleaux, M.; Geyer, M.; Latz, E.; Gütschow, M. Development of fluorescent and biotin probes targeting NLRP3. Front. Chem. 2021, 9, 642273. [Google Scholar] [CrossRef]
  40. Sabnis, R.W. 7-Methoxycoumarin-3-carboxylic acid. In Handbook of Fluorescent Dyes and Probes; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 281–284. [Google Scholar] [CrossRef]
  41. Sabnis, R.W. 7-Diethylaminocoumarin-3-carboxylic acid (DAC) (DEAC). In Handbook of Fluorescent Dyes and Probes; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 149–152. [Google Scholar] [CrossRef]
  42. Sabnis, R.W. 7-Methoxycoumarin-3-carboxylic acid succinimidyl ester. In Handbook of Fluorescent Dyes and Probes; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 285–286. [Google Scholar] [CrossRef]
  43. Sabnis, R.W. 7-Diethylaminocoumarin-3-carboxylic acid succinimidyl ester (DEAC SE). In Handbook of Fluorescent Dyes and Probes; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 155–156. [Google Scholar] [CrossRef]
  44. Grayeski, M.L.; DeVasto, J.K. Coumarin derivatizing agents for carboxylic acid detection using peroxyoxalate chemiluminescence with liquid chromatography. Anal. Chem. 1987, 57, 1203. [Google Scholar] [CrossRef]
  45. Vemula, V.; Ni, Z.; Fedorova, M. Fluorescence labeling of carbonylated lipids and proteins in cells using coumarin-hydrazide. Redox Biol. 2015, 5, 195–204. [Google Scholar] [CrossRef] [Green Version]
  46. Sabnis, R.W. 7-Diethylaminocoumarin-3-carboxylic acid hydrazide (DCCH). In Handbook of Fluorescent Dyes and Probes; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2015; pp. 153–154. [Google Scholar] [CrossRef]
  47. Milic, I.; Fedorova, M. Derivatization and detection of small aliphatic and lipid-bound carbonylated lipid peroxidation products by ESI-MS. In Methods in Molecular Biology; Advanced Protocols in Oxidative Stress, III.; Armstrong, D., Ed.; Humana Press: New York, NY, USA, 2015; Volume 1208, pp. 3–20. [Google Scholar] [CrossRef]
  48. Mukherjee, K.; Chio, T.I.; Bane, S.L. Visualization of oxidative stress-induced carbonylation in live mammalian cells. In Methods in Enzymology; Chemical Tools for Imaging, Manipulating, and Tracking Biological Systems: Diverse Chemical, Optical and Bioorthogonal Methods; Chenoweth, D.M., Ed.; Academic Press: Cambridge, UK, 2020; Volume 641, pp. 165–181. [Google Scholar] [CrossRef]
  49. Kolb, H.C.; Finn, M.G.; Sharpless, B.K. Click chemistry: Diverse chemical function from a few good reactions. Angew. Chem. Int. Ed. 2001, 40, 2004–2021. [Google Scholar] [CrossRef]
  50. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide–alkynecycloaddition (CuAAC) and beyond: New reactivity of copper(i) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef] [PubMed]
  51. Zeydi, M.M.; Kalantarian, S.J.; Kazeminejad, Z. Overview on developed synthesis procedures of coumarin heterocycles. J. Iran. Chem. Soc. 2020, 17, 3031–3094. [Google Scholar] [CrossRef]
  52. Upadhyay, H.C. Coumarin-1,2,3-triazole hybrid molecules: An emerging scaffold for combating drug resistance. Curr. Top. Med. Chem. 2021, 21, 737–752. [Google Scholar] [CrossRef] [PubMed]
  53. Musiol-Kroll, E.M.; Zubeil, F.; Schafhauser, T.; Härtner, T.; Kulik, A.; McArthur, J.; Koryakina, I.; Wohlleben, W.; Grond, S.; Williams, G.J.; et al. Polyketide bioderivatization using the promiscuous acyltransferase KirCII. ACS Synth. Biol. 2017, 6, 421–427. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  54. Zhu, L.; Pelaz, B.; Chakraborty, I.; Parak, W.J. Investigating possible enzymatic degradation on polymer shells around inorganic nanoparticles. Int. J. Mol. Sci. 2019, 20, 935. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  55. Rangel, P.X.M.; Moroni, E.; Merlier, F.; Gheber, L.A.; Vago, R.; Tse Sum Bui, B.; Haupt, K. Chemical antibody mimics inhibit cadherin-mediated cell-cell adhesion: A promising strategy for cancer therapy. Angew. Chem. Int. Ed. Eng. 2020, 59, 2816–2822. [Google Scholar] [CrossRef]
  56. Soloway, A.H.; Hatanaka, H.; Davis, M.A. Penetration of brain and brain tumor. VII. Tumor-binding sulfhydryl boron compounds. J. Med. Chem. 1967, 10, 714–717. [Google Scholar] [CrossRef]
  57. Sivaev, I.B.; Bregadze, V.I.; Kuznetsov, N.T. Derivatives of the closo-dodecaborate anion and their application in medicine. Russ. Chem. Bull. 2002, 51, 1362–1374. [Google Scholar] [CrossRef]
  58. Lamba, M.; Goswami, A.; Bandyopadhyay, A. A periodic development of BPA and BSH based derivatives in boron neutron capture therapy (BNCT). Chem. Commun. 2021, 57, 827–839. [Google Scholar] [CrossRef]
  59. Hawthorne, M.F. The role of chemistry in the development of boron neutron capture therapy of cancer. Angew. Chem. Int. Ed. 1993, 32, 950–984. [Google Scholar] [CrossRef]
  60. Soloway, A.H.; Tjarks, W.; Barnum, B.A.; Rong, F.-G.; Barth, R.F.; Codogni, I.M.; Wilson, J.G. The chemistry of neutron capture therapy. Chem. Rev. 1998, 98, 1515–1562. [Google Scholar] [CrossRef] [PubMed]
  61. Bregadze, V.I.; Sivaev, I.B. Polyhedral boron compounds for BNCT. In Boron Science: New Technologies and Applications; Hosmane, N.S., Ed.; CRC Press: Boca Raton, FL, USA, 2012; pp. 181–207. [Google Scholar]
  62. Hu, K.; Yang, Z.; Zhang, L.; Xie, L.; Wang, L.; Xu, H.; Josephson, L.; Liang, S.H.; Zhang, M.-R. Boron agents for neutron capture therapy. Coord. Chem. Rev. 2020, 405, 213139. [Google Scholar] [CrossRef]
  63. Sivaev, I.B.; Bregadze, V.I.; Sjöberg, S. Chemistry of closo-dodecaborate anion [B12H12]2−: A review. Collect. Czech. Chem. Commun. 2002, 67, 679–727. [Google Scholar] [CrossRef]
  64. Sivaev, I.B. Polyhedral boranes and carboranes. In Comprehensive Organometallic Chemistry IV; Elsevier: Kidlington, UK, 2022; Volume 9, pp. 196–262. [Google Scholar] [CrossRef]
  65. Hattori, Y.; Ishimura, M.; Ohta, Y.; Takenaka, H.; Kawabata, S.; Kirihata, M. Dodecaborate conjugates targeting tumor cell overexpressing translocator protein for boron neutron capture therapy. ACS Med. Chem. Lett. 2022, 13, 50–54. [Google Scholar] [CrossRef]
  66. Hirase, S.; Aoki, A.; Hattori, Y.; Morimoto, K.; Noguchi, K.; Fujii, I.; Takatani-Nakase, T.; Futaki, S.; Kirihata, M.; Nakase, I. Dodecaborate-encapsulated extracellular vesicles with modification of cell-penetrating peptides for enhancing macropinocytotic cellular uptake and biological activity in boron neutron capture therapy. Mol. Pharm. 2022, 19, 1135–1145. [Google Scholar] [CrossRef] [PubMed]
  67. Tolmachev, V.; Koziorowski, J.; Sivaev, I.; Lundqvist, H.; Carlsson, J.; Orlova, A.; Gedda, L.; Olsson, P.; Sjöberg, S.; Sundin, A. closo-dodecaborate(2-) as a linker for iodination of macromolecules. Aspects on conjugation chemistry and biodistribution. Bioconjug. Chem. 1999, 10, 338–345. [Google Scholar] [CrossRef]
  68. Tolmachev, V.; Bruskin, A.; Sivaev, I.; Lundqvist, H.; Sjöberg, S. Radiobromination of closo-dodecaborate anion. Aspects of labelling chemistry in aqueous solution using Chloramine-T. Radiochim. Acta 2002, 90, 229–235. [Google Scholar] [CrossRef]
  69. Bruskin, A.; Sivaev, I.; Persson, M.; Lundqvist, H.; Carlsson, J.; Sjöberg, S.; Tolmachev, V. Radiobromination of monoclonal antibody using potassium [76Br](4 isothiocyanatobenzyl-ammonio)-bromo-decahydro-closo-dodecaborate (Bromo-DABI). Nucl. Med. Biol. 2004, 31, 205–211. [Google Scholar] [CrossRef]
  70. Wilbur, D.S.; Chyan, M.-K.; Hamlin, D.K.; Perry, M.A. Reagents for Astatination of Biomolecules. 3. Comparison of closo-decaborate(2-) and closo-dodecaborate(2-) moieties as reactive groups for labeling with astatine-211. Bioconjug. Chem. 2009, 20, 591–602. [Google Scholar] [CrossRef] [Green Version]
  71. Chaari, M.; Gaztelumendi, N.; Cabrera-González, J.; Peixoto-Moledo, P.; Viñas, C.; Xochitiotzi-Flores, E.; Farfán, N.; Ben Salah, A.; Nogués, C.; Núñez, R. Fluorescent BODIPY-anionic boron cluster conjugates as potential agents for cell tracking. Bioconjug. Chem. 2018, 29, 1763–1773. [Google Scholar] [CrossRef] [PubMed]
  72. Justus, E.; Izteleuova, D.T.; Kasantsev, A.V.; Axartov, M.M.; Lork, E.; Gabel, D. Preparation of carboranyl and dodecaborate derivatives of coumarin. Collect. Czech. Chem. Commun. 2007, 72, 1740–1754. [Google Scholar] [CrossRef]
  73. Kosenko, I.; Laskova, J.; Kozlova, A.; Semioshkin, A.; Bregadze, V.I. Synthesis of coumarins modified with cobalt bis(1,2-dicarbolide) and closo-dodecaborate boron clusters. J. Organomet. Chem. 2020, 921, 121370. [Google Scholar] [CrossRef]
  74. Serdyukov, A.; Kosenko, I.; Druzina, A.; Grin, M.; Mironov, A.F.; Bregadze, V.I.; Laskova, J. Anionic polyhedral boron clusters conjugates with 7-diethylamino-4-hydroxycoumarin. Synthesis and lipophilicity determination. J. Organomet. Chem. 2021, 946–947, 121905. [Google Scholar] [CrossRef]
  75. Genady, A.R.; Ioppolo, J.A.; Azaam, M.M.; El-Zaria, M.E. New functionalized mercaptoundecahydrododecaborate derivatives for potential application in boron neutron capture therapy: Synthesis, characterization and dynamic visualization in cells. Eur. J. Med. Chem. 2015, 93, 574–583. [Google Scholar] [CrossRef]
  76. Sivaev, I.B.; Semioshkin, A.A.; Brellochs, B.; Sjöberg, S.; Bregadze, V.I. Synthesis of oxonium derivatives of the dodecahydro- closo-dodecaborate anion [B12H12]2−. Tetramethylene oxonium derivative of [B12H12]2− as a convenient precursor for the synthesis of functional compounds for boron neutron capture therapy. Polyhedron 2000, 19, 627–632. [Google Scholar] [CrossRef]
  77. Semioshkin, A.A.; Sivaev, I.B.; Bregadze, V.I. Cyclic oxonium derivatives of polyhedral boron hydrides and their synthetic applications. Dalton Trans. 2008, 977–992. [Google Scholar] [CrossRef]
  78. Sivaev, I.B.; Bregadze, V.I. Cyclic oxonium derivatives as an efficient synthetic tool for the modification of polyhedral boron hydrides. In Boron Science: New Technologies and Applications; Hosmane, N.S., Ed.; CRC Press: Boca Raton, FL, USA, 2012; pp. 623–637. [Google Scholar]
  79. Stogniy, M.Y.; Sivaev, I.B. Synthesis and reactivity of cyclic oxonium derivatives of nido-carborane: A review. Reactions 2022, 3, 172–191. [Google Scholar] [CrossRef]
  80. Druzina, A.A.; Shmalko, A.V.; Sivaev, I.B.; Bregadze, V.I. Cyclic oxonium derivatives of cobalt and iron bis(dicarbollides) and their use in organic synthesis. Russ. Chem. Rev. 2021, 90, 785–830. [Google Scholar] [CrossRef]
  81. Bregadze, V.I.; Semioshkin, A.A.; Las’kova, J.N.; Berzina, M.Y.; Lobanova, I.A.; Sivaev, I.B.; Grin, M.A.; Titeev, R.A.; Brittal, D.I.; Ulybina, O.V.; et al. Novel types of boronated chlorine6 conjugates via ‘click chemistry’. Appl. Organomet. Chem. 2009, 23, 370–374. [Google Scholar] [CrossRef]
  82. Semioshkin, A.; Laskova, J.; Wojtczak, B.; Andrysiak, A.; Godovikov, I.; Bregadze, V.; Lesnikowski, Z.J. Synthesis of closo-dodecaborate based nucleoside conjugates. J. Organomet. Chem. 2009, 694, 1375–1379. [Google Scholar] [CrossRef]
  83. El-Zaria, M.E.; Nakamura, H. New strategy for synthesis of mercaptoundecahydrododecaborate derivatives via click chemistry: Possible boron carriers and visualization in cells for neutron capture therapy. Inorg. Chem. 2009, 48, 11896–11902. [Google Scholar] [CrossRef]
  84. Koganei, H.; Tachikawa, S.; El-Zaria, M.E.; Nakamura, H. Synthesis of oligo-closo-dodecaborates by Hüisgen click reaction as encapsulated agents for the preparation of high-boron-content liposomes for neutron capture therapy. New J. Chem. 2015, 39, 6388–6394. [Google Scholar] [CrossRef]
  85. Goswami, L.N.; Chakravarty, S.; Lee, M.W.; Jalisatgi, S.S.; Hawthorne, M.F. Extensions of the icosahedral closomer structure by using azide–alkyne click reactions. Angew. Chem. Int. Ed. 2011, 50, 4689–4691. [Google Scholar] [CrossRef]
  86. Goswami, L.N.; Ma, L.; Kueffer, P.J.; Jalisatgi, S.S.; Hawthorne, M.F. Synthesis and relaxivity studies of a DOTA-based nanomolecular chelator assembly supported by an icosahedral closo-B122− -core for MRI: A click chemistry approach. Molecules 2013, 18, 9034–9048. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  87. Tsurubuchi, T.; Shirakawa, M.; Kurosawa, W.; Matsumoto, K.; Ubagai, R.; Umishio, H.; Suga, Y.; Yamazaki, J.; Arakawa, A.; Maruyama, Y.; et al. Evaluation of a novel boron-containing α-d-mannopyranoside for BNCT. Cells 2020, 9, 1277. [Google Scholar] [CrossRef]
  88. Druzina, A.A.; Zhidkova, O.B.; Kosenko, I.D. Synthesis of conjugates of closo-dodecaborate dianion with cholesterol using a “click” reaction. Russ. Chem. Bull. 2020, 69, 1080–1084. [Google Scholar] [CrossRef]
  89. Novopashina, D.S.; Vorobyeva, M.A.; Lomzov, A.A.; Silnikov, V.N.; Venyaminova, A.G. Terminal mono- and bis-conjugates of oligonucleotides with closo-dodecaborate: Synthesis and physico-chemical properties. Int. J. Mol. Sci. 2021, 22, 182. [Google Scholar] [CrossRef] [PubMed]
  90. Druzina, A.A.; Grammatikova, N.E.; Zhidkova, O.B.; Nekrasova, N.A.; Dudarova, N.V.; Kosenko, I.D.; Grin, M.A.; Bregadze, V.I. Synthesis and antibacterial activity studies of the conjugates of curcumin with closo-dodecaborate and cobalt bis(dicarbollide) boron clusters. Molecules 2022, 27, 2920. [Google Scholar] [CrossRef]
  91. Sivaev, I.B.; Kulikova, N.Y.; Nizhnik, E.A.; Vichuzhanin, M.V.; Starikova, Z.A.; Semioshkin, A.A.; Bregadze, V.I. Practical synthesis of 1,4-dioxane derivative of the closo-dodecaborate anion and its ring opening with acetylenic alkoxides. J. Organomet. Chem. 2008, 693, 519–525. [Google Scholar] [CrossRef]
  92. Semioshkin, A.A.; Osipov, S.N.; Grebenyuk, J.N.; Nizhnik, E.A.; Godovikov, I.A.; Shchetnikov, G.T.; Bregadze, V.I. An effective approach to 1,2,3-triazole containing 12-vertex closo-dodecaborates. Collect. Czech. Chem. Commun. 2007, 72, 1717–1724. [Google Scholar] [CrossRef]
  93. Prashanth, T.; Avin, B.R.V.; Thirusangu, P.; Ranganatha, V.L.; Prabhakar, B.T.; Sharath Chandra, J.N.N.; Khanum, S.A. Synthesis of coumarin analogs appended with quinoline and thiazole moiety and their apoptogenic role against murine ascitic carcinoma. Biomed. Pharm. 2019, 112, 108707. [Google Scholar] [CrossRef] [PubMed]
  94. He, X.; Yan, Z.; Hu, X.; Zuo, Y.; Jiang, C.; Jin, L.; Shang, Y. FeCl3-Catalyzed cascade reaction: An efficient approach to functionalized coumarin derivatives. Synth. Commun. 2014, 44, 1507–1514. [Google Scholar] [CrossRef]
  95. Jotani, M.M.; Baldaniya, B.B.; Tiekink, E.R.T. N-(2-Oxo-2H-chromen-3-yl)benzamide. Acta Cryst. E 2010, 66, o778. [Google Scholar] [CrossRef] [Green Version]
  96. Matos, M.J.; Uriarte, E.; Santana, L.; Vilar, S. Synthesis, NMR characterization, X-ray structural analysis and theoretical calculations of amide and ester derivatives of the coumarin scaffold. J. Mol. Struct. 2013, 1041, 144–150. [Google Scholar] [CrossRef]
  97. Mague, J.T.; Mohamed, S.K.; Akkurt, M.; Younese, S.H.H.; Albayati, M.R. Crystal structure of 2-oxo-N’-phenyl-2H-chromene- 3-carbohydrazide. Acta Cryst. E 2015, 71, o1005–o1006. [Google Scholar] [CrossRef] [Green Version]
  98. Zhang, L.-J.; Yin, B.-Z. 7-Diethylamino-2-oxo-2H-chromene-3-carbohydrazide. Acta Cryst. E 2011, 67, o1107. [Google Scholar] [CrossRef] [Green Version]
  99. Olejniczak, A.; Katrusiak, A.; Podsiadło, M.; Katrusiak, A. Crystal design by CH···N and N···N interactions: High-pressure structures of high-nitrogen-content azido-triazolopyridazines compounds. Acta Cryst. B 2020, 76, 1136–1142. [Google Scholar] [CrossRef]
  100. Raunio, H.; Pentikainen, O.; Juvonen, R.O. Coumarin-Based Profluorescent and Fluorescent Substrates for Determining Xenobiotic-Metabolizing Enzyme Activities In Vitro. Int. J. Mol. Sci. 2020, 21, 4708. [Google Scholar] [CrossRef]
  101. Titov, A.A.; Smol’yakov, A.F.; Godovikov, I.A.; Chernyadyev, A.Y.; Molotkov, A.P.; Loginov, D.A.; Filippov, O.A.; Belkova, N.V.; Shubina, E.S. The role of weak intermolecular interactions in photophysical behavior of isocoumarins on the example of their interaction with cyclic trinuclear silver(I) pyrazolate. Inorg. Chim. Acta 2022, 539, 121004. [Google Scholar] [CrossRef]
  102. Verlinden, B.; Van Hoecke, K.; Aerts, A.; Daems, N.; Dobney, A.; Janssens, K.; Cardinaels, T. Quantification of boron in cells for evaluation of drug agents used in boron neutron capture therapy. J. Anal. At. Spectrom. 2021, 36, 598–606. [Google Scholar] [CrossRef]
  103. Kasparkova, J.; Kostrhunova, H.; Novohradsky, V.; Logvinov, A.A.; Temnov, V.V.; Borisova, N.E.; Podrugina, T.A.; Markova, L.; Starha, P.; Nazarov, A.A.; et al. Novel cis-Pt(II) complexes with alkylpyrazole ligands: Synthesis, characterization, and unusual mode of anticancer action. Bioinorg. Chem. Appl. 2022, 2022, 1717200. [Google Scholar] [CrossRef] [PubMed]
  104. Shutkov, I.A.; Okulova, Y.N.; Tyurin, V.Y.; Sokolova, E.V.; Babkov, D.A.; Spasov, A.A.; Gracheva, Y.A.; Schmidt, C.; Kirsanov, K.I.; Shtil, A.A.; et al. Ru(III) Complexes with lonidamine-modified ligands. Int. J. Mol. Sci. 2021, 22, 13468. [Google Scholar] [CrossRef] [PubMed]
  105. Del Monte, U. Does the cell number 109 still really fit one gram of tumor tissue? Cell Cycle 2009, 8, 505–506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  106. Hao, S.-J.; Wan, Y.; Xia, Y.-Q.; Zou, X.; Zheng, S.-Y. Size-based separation methods of circulating tumor cells. Adv. Drug Deliv. Rev. 2018, 125, 3–20. [Google Scholar] [CrossRef]
  107. Jiang, R.; Shen, H.; Piao, Y. The morphometrical analysis on the ultrastructure of A549 cells. Rom. J. Morphol. Embryol. 2010, 51, 663–667. [Google Scholar] [PubMed]
  108. Phillips, K.G.; Jacques, S.L.; McCarty, O.J.T. Measurement of single cell refractive index, dry mass, volume, and density using a transillumination microscope. Phys. Rev. Lett. 2012, 109, 118105. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Di Antonio, M.; Biffi, G.; Mariani, A.; Raiber, E.A.; Rodriguez, R.; Balasubramanian, S. Selective RNA versus DNA G-quadruplex targeting by in situ click chemistry. Angew. Chem. Int. Ed. 2012, 124, 11235–11240. [Google Scholar] [CrossRef] [Green Version]
  110. Kasparkova, J.; Kostrhunova, H.; Novohradsky, V.; Ma, L.; Zhu, G.; Milaeva, E.R.; Shtill, A.A.; Vinck, R.; Gasser, G.; Brabec, V.; et al. Is antitumor Pt(IV) complex containing two axial lonidamine ligands a true dual- or multi-action prodrug? Metallomics 2022, 14, mfac048. [Google Scholar] [CrossRef]
  111. Bruker. APEX-III; Bruker AXS Inc.: Madison, WI, USA, 2018. [Google Scholar]
  112. Krause, L.; Herbst-Irmer, R.; Sheldrick, G.M.; Stalke, D. Comparison of silver and molybdenum microfocus X-ray sources for single-crystal structure determination. J. Appl. Cryst. 2015, 48, 3–10. [Google Scholar] [CrossRef] [Green Version]
  113. Sheldrick, G.M. SHELXT-Integrated space-group and crystal-structure determination. Acta Cryst. A 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Some commercially available fluorescent coumarins.
Figure 1. Some commercially available fluorescent coumarins.
Molecules 27 08575 g001
Scheme 1. Synthesis of coumarins 2a,b.
Scheme 1. Synthesis of coumarins 2a,b.
Molecules 27 08575 sch001
Scheme 2. “Click” reactions of coumarins 2a and 2b with the alkyne closo-dodecaborate derivative 3[NBu4].
Scheme 2. “Click” reactions of coumarins 2a and 2b with the alkyne closo-dodecaborate derivative 3[NBu4].
Molecules 27 08575 sch002
Figure 2. Molecular crystal structures of 2a (top) and 2b (bottom). Non-hydrogen atoms are presented by probability ellipsoids of atomic displacements (p = 0.5). Dotted lines correspond to the intramolecular hydrogen bonds.
Figure 2. Molecular crystal structures of 2a (top) and 2b (bottom). Non-hydrogen atoms are presented by probability ellipsoids of atomic displacements (p = 0.5). Dotted lines correspond to the intramolecular hydrogen bonds.
Molecules 27 08575 g002
Figure 3. Fragments of crystal packing of 2a (top) and 2b (bottom) demonstrating the N3…N3 stacking interactions in 2a (dashed bold lines) and bifurcate C12(H)…O2 H-bond in 2b (dotted lines).
Figure 3. Fragments of crystal packing of 2a (top) and 2b (bottom) demonstrating the N3…N3 stacking interactions in 2a (dashed bold lines) and bifurcate C12(H)…O2 H-bond in 2b (dotted lines).
Molecules 27 08575 g003
Figure 4. UV–vis (dashed line) and normalized emission (solid line) spectra of coumarin 2a in CH3CN (c = 1.5 × 10−4 M, blue) and conjugate 4a in water (c = 0.5 × 10−4 M, red).
Figure 4. UV–vis (dashed line) and normalized emission (solid line) spectra of coumarin 2a in CH3CN (c = 1.5 × 10−4 M, blue) and conjugate 4a in water (c = 0.5 × 10−4 M, red).
Molecules 27 08575 g004
Figure 5. UV–vis (dashed line) and normalized emission (solid line) spectra of coumarin 2b in CH3CN (c = 1 × 10−4 M, black) and conjugate 4b in water (c = 0.5 × 10−4 M, green).
Figure 5. UV–vis (dashed line) and normalized emission (solid line) spectra of coumarin 2b in CH3CN (c = 1 × 10−4 M, black) and conjugate 4b in water (c = 0.5 × 10−4 M, green).
Molecules 27 08575 g005
Figure 6. Cellular boron accumulation of compounds 4a and 4b in A549 and WI38 cells.
Figure 6. Cellular boron accumulation of compounds 4a and 4b in A549 and WI38 cells.
Molecules 27 08575 g006
Table 1. Photophysical parameters for compounds 2a, 2b, 4a and 4b.
Table 1. Photophysical parameters for compounds 2a, 2b, 4a and 4b.
λmaxabs (ε∙10−3, M−1 cm−1), nmλmaxemex), nm
2a350 (5.0)398 (340)
4a343 (13.4)405 (340)
2b413 (9.1)464 (410)
4b430 (12.6)474 (410)
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Laskova, J.; Serdyukov, A.; Kosenko, I.; Ananyev, I.; Titova, E.; Druzina, A.; Sivaev, I.; Antonets, A.A.; Nazarov, A.A.; Bregadze, V.I. New Azido Coumarins as Potential Agents for Fluorescent Labeling and Their “Click” Chemistry Reactions for the Conjugation with closo-Dodecaborate Anion. Molecules 2022, 27, 8575. https://doi.org/10.3390/molecules27238575

AMA Style

Laskova J, Serdyukov A, Kosenko I, Ananyev I, Titova E, Druzina A, Sivaev I, Antonets AA, Nazarov AA, Bregadze VI. New Azido Coumarins as Potential Agents for Fluorescent Labeling and Their “Click” Chemistry Reactions for the Conjugation with closo-Dodecaborate Anion. Molecules. 2022; 27(23):8575. https://doi.org/10.3390/molecules27238575

Chicago/Turabian Style

Laskova, Julia, Alexander Serdyukov, Irina Kosenko, Ivan Ananyev, Ekaterina Titova, Anna Druzina, Igor Sivaev, Anastasia A. Antonets, Alexey A. Nazarov, and Vladimir I. Bregadze. 2022. "New Azido Coumarins as Potential Agents for Fluorescent Labeling and Their “Click” Chemistry Reactions for the Conjugation with closo-Dodecaborate Anion" Molecules 27, no. 23: 8575. https://doi.org/10.3390/molecules27238575

Article Metrics

Back to TopTop