Next Article in Journal
Experimental and Computational Study of a Liquid Crystalline Dimesogen Exhibiting Nematic, Twist-Bend Nematic, Intercalated Smectic, and Soft Crystalline Mesophases
Next Article in Special Issue
Influence of Electronic Environment on the Radiative Efficiency of 9-Phenyl-9H-carbazole-Based ortho-Carboranyl Luminophores
Previous Article in Journal
Spin Exchanges between Transition Metal Ions Governed by the Ligand p-Orbitals in Their Magnetic Orbitals
Previous Article in Special Issue
The First Nickelacarborane with closo-nido Structure
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Structure of Nido-Carboranyl Azide and Its “Click” Reactions

A.N. Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, 28 Vavilov Str., 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Submission received: 30 December 2020 / Revised: 15 January 2021 / Accepted: 15 January 2021 / Published: 20 January 2021

Abstract

:
Novel zwitter-ionic nido-carboranyl azide 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 was prepared by the reaction of 9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 with NaN3. The solid-state molecular structure of nido-carboranyl azide was determined by single-crystal X-ray diffraction. 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 was used for the copper(I)-catalyzed azide-alkyne cycloaddition with phenylacetylene, alkynyl-3β-cholesterol and cobalt/iron bis(dicarbollide) terminal alkynes to form the target 1,2,3-triazoles. The nido-carborane-cholesterol conjugate 9-3β-Chol-O(CH2)C-CH-N3(CH2)3Me2N-nido-7,8-C2B9H11 with charge-compensated group in a linker can be used as a precursor for preparation of liposomes for Boron Neutron Capture Therapy (BNCT). A series of novel zwitter-ionic boron-enriched cluster compounds bearing a 1,2,3-triazol-metallacarborane-carborane conjugated system was synthesized. Prepared conjugates contain a large amount of boron atom in the biomolecule and potentially can be used for BNCT.

Graphical Abstract

1. Introduction

Nido-Carborane (7,8-dicarba-nido-undecaborate anion) and its derivatives attract attention due to its unique electronic structure, namely, electron delocalization, which is often considered as an unusual three-dimensional aromaticity [1,2], and also due to its electron-withdrawing character of the skeletal carbon atoms. Such structural and physicochemical properties of 7,8-dicarba-nido-undecaborate anion induce the preparation of new functional derivatives of nido-carboranes with a variety of practical applications. These compounds attract the continued interest of researchers working in various fields, such as medical chemistry (boron neutron capture therapy (BNCT) of malignant tumors [3,4,5], HIV protease inhibitors [6], reagents for radioimaging of tumors [7]), and creation of new materials [1] (carborane-containing polymers, ionic liquids, liquid crystals).
It is well known that BNCT is a binary method of cancer treatment, in which it is necessary that boron compounds selectively accumulate in the tumor tissue at the required therapeutic concentration for its subsequent irradiation with thermal neutrons [8,9,10]. The advantage of using derivatives of nido-carboranes in BNCT is their stability and a high content of boron atoms in the molecule. In this regard, the new methods for the synthesis of various functional mono-substituted derivatives of nido-carborane should be developed. Currently, there are several methods for the preparation of mono-substituted nido-carboranes: C-substituted derivatives can be obtained by modification of the carborane cage at carbon atoms of the parent ortho-carborane followed by its deboronation to the corresponding nido-carborane. Last time, the most cited methods of synthesis B-substituted functional derivatives of nido-carboranes are ring opening of their cyclic oxonium derivatives under the action of nucleophilic agents [11,12,13,14], alkylation of their methylsulfide derivatives [15,16,17,18], Cu-promoted synthesis of their ammonium derivatives [19] and nucleophilic addition of alcohols and mercaptans to highly polarized triple bond B-N+≡CR [20].
Synthesis of mono-substituted derivatives of nido-carborane with a functional group makes it possible to attach a carborane fragment to various bio- and macromolecules and thus to obtain compounds with a given set of properties. The substituent introduced can be a biologically active derivative that acts as a tumor targeting vector, or a simple functional group, which can be used for conjugation with high molecular weight biomolecules using standard methods of bioorganic chemistry. Recently the Cu(I)-catalyzed reaction of 1,3-dipolar [3 + 2]-cycloaddition of azides with alkynes (“click” reaction) has found more and more widespread use for the bioconjugation of molecules [21,22,23,24,25]. Such reactions must proceed rapidly under ambient conditions, resulting in a high yield of desired 1,2,3-triazole. The 1,2,3-triazole scaffold is known to be one of the most valuable in the chemistry of biologically active compounds [26], which exhibit anticancer [27,28,29], anti-HIV [27,29,30], antibacterial [27,29,31,32], antioxidant [27,29,33,34] activities. In addition, the 1,2,3-triazole in the molecules between boron-containing and the biologically active fragments (porphyrins, nucleosides) is a linking unit that mimics geometry and electronic properties of the peptide bond is a more stable to hydrolysis reactions [14,35]. However, the preparation of a suitable nido-carborane-containing substrate for a “click” reaction often requires multistage syntheses, for instance, either the incorporation of a distant or cage-bound azide group or a terminal alkyne. Azide is one the most popular biorthogonal functional groups due to its small size coupled with stability to water and inertness towards endogenous biological functionalities [36]. It should be noted that only a few examples of azido-containing nido-carboranes are known to date. Recently we have obtained C-substituted nido-carboranyl azides [7-N3CH2CH2OCH2CH2O-nido-7,8-C2B9H11] and [7-N3CH2CH2OCH2CH2S-nido-7,8-C2B9H11] which were synthesized by alkylation of 1-mercapto-ortho-carborane with bis(2-chloroethyl) ether followed by introduction of azide group and by the conversion of closo-derivative to nido-form [37]. B-substituted [10-N3CH2CH2CH2CH2O-nido-7,8-C2B9H11] and [10-N3CH2CH2OCH2CH2O-nido-7,8-C2B9H11] were prepared by the ring-opening reactions of the corresponding cyclic oxonium derivatives with sodium azide [11,14].
In this contribution, we describe a synthesis of zwitter-ionic B-substituted nido-carborane bearing a functional azido-group 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 and study its behaviour in the copper(I)-catalyzed azide-alkyne cycloaddition reaction with phenylacetylene, alkynyl-3β-cholesterol and terminal alkynes derivatives of cobalt/iron bis(dicarbollide).

2. Results and Discussion

2.1. Synthesis of 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 and Its “Click” Reaction with Phenylacetylene

The reaction of 9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 1 with NaN3 in the presence of NaI as a catalyst in DMF upon prolonged heating under 50 °C for 7 days results in 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 2 with 90% yield (Scheme 1). It was isolated as a white non-hydroscopic solid soluble in common organic solvents like CH2Cl2, CH3CN, alcohols and non-soluble in hydrocarbons and water.
The 11B-NMR spectrum contains eight signals indicating a nonsymmetrical monosubstituted structure. The 1H-NMR spectrum of the signal of the methylene group bonded to the nitrogen atom exhibits a singlet at 3.37 ppm. It should be noted that the splitting reduces as the distance from the nitrogen atom increases resulting in singlet and multiplet at 2.15 and 3.50 ppm for the second and third methylene groups. In addition, the signals of the CHcarb groups and the extra-hydrogen are observed approx. at 2.59 and -3.4 ppm, correspondingly. In the 13C-NMR spectrum the most characteristic is signal of CH2N3 group. As observed earlier [19], the signal of the CH2Cl group in the 13C-NMR spectrum appears at 43.0 ppm for 1. The substitution of chlorine for azide results in the low-field shift to 65.2 ppm. The azide stretching band in the IR spectrum of 2 is located at 2075 cm-1.
Conditions of “click” reactions in the preparation of various boron-containing biomolecules vary particularly wide. Earlier “click” reactions have been successfully used for the synthesis of conjugates of bis(dicarbollide) metallacarboranes and nido-carborane with thymidine [14]. The synthesis was carried out in a mixture of tert-butanol/water (1:1) at ambient temperature using copper(II) sulfate pentahydrate with potassium ascorbate as a catalyst. The same reaction for synthesis of conjugate dodecaborate dianion with thymidine proceed in CH3CN at ambient temperature with copper(II) sulfate pentahydrate with sodium ascorbate [38]. Conjugates of chlorine e6 with a cobalt bis(dicarbollide) anion or closo-dodecaborate dianion were obtained using CuI and Et3N in acetonitrile at ambient temperatures [35]. Series of 1,2,3-triazoles bearing closo-dodecaborate fragment was obtained using CuI as a catalyst and Et3N as a base under reflux in ethanol [39].
In the present work, we studied the behavior of azide 2 in copper(I)-catalyzed azide-alkyne cycloaddition using simple terminal alkyne as a pilot compound. It was showed that it readily reacts with phenylacetylene in ethanol in the presence of diisopropylethylamine (DIPEA) and catalytic amount of CuI to give the corresponding 1,2,3-triazole 3 with 85 % yield (Scheme 2).
The structure of the nido-carborane 3 was confirmed by the data of 1H-, 11B- and 13C-NMR, IR spectroscopy and HRMS. The 1H- and 13C-NMR spectra of compounds 3 along with the signals of the heteroaliphatic chain and the phenyl group contain the characteristic signals of the triazole cycle in the 1H-NMR spectrum signal of the CHtriazole hydrogen appears at 8.62 ppm. In the 13C-NMR spectrum, the signal of the CHtriazole carbon is observed at 122.0 ppm, whereas the signal of the Ctriazole carbon appears at 146.9 ppm. In the 1H-NMR spectrum, the signal of the methylene group next to the triazole cycle is observed at 4.54 ppm and the characteristic signal of the Me2N hydrogens appear at 2.88 ppm. The IR spectrum of 3 demonstrates an absence of the azide band stretching and the appearance of the band of the triazole cycle at 1462 cm−1.

2.2. Synthesis of Nido-Carboranyl Cholesterol Derivative with Charge-Compensated Group

Furthermore, compound 2 was used for synthesis of boronated cholesterol as precursor for the preparation of liposomes. The usage of liposomes is the important approach directed to selective delivery of therapeutics into tumors [40,41,42]. The development of selective, non-toxic boron delivery agents that can preferentially deliver a high concentration of boron to the tumor is probably the greatest need for the future progress of BNCT [43]. Due to the high permeability of the walls of blood vessels inside the tumor, stagnant blood flow occurs and lymphatic outflow is disrupted. These changes lead to the EPR (enhanced permeability and retention) effect, due to which macromolecules and nanoparticles such as liposomes penetrate from the bloodstream of the tumor vessel into the intercellular space and accumulate mainly in the tumor tissue [44,45,46]. Cholesterol is the major component of the cell membrane and most liposomal formulations. Therefore, the development of boronated derivatives of cholesterol is an effective approach for the selective delivery of boron clusters into the cancer cells via liposomes. Recently, using “click” reactions we obtained a series of mono-negative charged conjugates of cobalt bis(dicarbollide) with cholesterol [47], conjugates of cobalt/iron of bis(dicarbollide) and cholesterol with similar length spacer but with zwitter-ionic character of target molecule [48] and conjugates of closo-dodecaborate dianion with cholesterol [49]. It has been only recently shown that the inclusion of lipophilic boron-containing species in the liposome bilayer provides an attractive method to increase the gross boron content of the liposomes in the formulation [50,51]. In addition, it has been found that PEGylated liposome encapsulating nido-carborane by hydrating thin lipid films significantly suppresses tumors in boron neutron capture therapy [52].
In the present work, we use the “click” methodology to obtain new conjugate of nido-carborane with cholesterol suitable for the preparation of boron-containing liposomes as potential drugs for boron neutron capture therapy of cancer. Usage of nido-carborane 2 for synthesis of carborane-cholesterol conjugates leads to zwitter-ionic character of product structure. Its reaction proceeded in a slight excess of alkynyl-3β-cholesterol 4 in the presence of a CuI catalyst and diisopropylethylamine (DIPEA) as a base in ethanol upon prolonged reflux for 8 h to give novel boron conjugate 5 with 85 % yield (Scheme 3).
The 1H-NMR spectrum of complex 5 contains a signal for the proton of the triazole group at 8.01 ppm. The characteristics signal of the alkyne CHst hydrogen of cholesterol in the conjugate 5 is observed at 5.37 ppm. The spectral characteristics of the CH protons of cholesterol are in good agreement with the literature data [53]. The 13C-NMR spectrum of 5 exhibits signals for two carbon atoms of the triazole fragment at 145.8 ppm for CHtriazole atom and at 123.1 ppm for Ctriazole. In the 11H-NMR spectrum, the signal of the extra-hydrogen, as expected, is observed approx. at −3.4 ppm. The IR spectrum of compound 5 exhibits absorption bands characteristic of the BH group 2685 cm−1 and the triazole ring 1392 cm−1.
Based on synthesized compounds the boronated liposomes are planned to prepare in order to deliver boron clusters into a cancer cell for the BNCT experiment.

2.3. Synthesis of Zwitter-Ionic Boron-Enriched Cluster Compounds Bearing a 1,2,3-Triazol-Metallacarborane-Nido-Carborane Conjugated Systems

As was mentioned above, functionalized nido-carboranes can be used as building blocks for design and construction of boron-containing compounds for various medical applications [3,4,54]. For example, one of the important requirements of BNCT is the synthesis of structures with a higher content of boron atoms in the molecule than in the clinically used compounds [55,56]. In this contribution, we propose to combine two boron clusters into one molecule: the bis(dicarbollide) cluster serves as a boron-containing substituent providing low toxicity [57,58] and amphiphilicity [59,60] of the molecule and the nido-carborane cage serves as a boron-containing base for attachment to molecules by the “click” reaction. Penetration of various substances through biological membranes, their accumulation and retention in cells largely depend on their charge. It is known that positively charged particles have better penetration through biological membranes than negatively changed ones [61,62,63]. It motivated us that the synthesis of such compounds is based on the introduction of two ammonium centers in a spacer: the first one compensating the negative charge of the nido-carborane fragment and the second one compensating the charge of the cobalt/iron bis(dicarbollide) moiety. This allows us to double the boron content of the biomolecule as compared to the single cage approach. Moreover, by changing the type and the size of a spacer between these two boron cages, it is possible to control, to some extent, the hydrophilic/hydrophobic balance of the compounds.
At the first step, cobalt and iron bis(dicarbollide) terminal alkynes with charge-compensated group 69 were prepared by the cleavage reactions of oxonium derivatives of cobalt/iron bis(dicarbollide) with N,N-dimethylprop-2-yn-1-amine [48,64]. It was found that alkynes prepared from 1,4-dioxane and tetrahydropyran derivatives of cobalt bis(dicarbollide) 6 and 7 [48] readily undergo “click” reactions with a small excess of azido-derivative of nido-carborane 2 to give novel boron conjugates 10 and 11. High preparative yields (85 %) of the desired products 10 and 11 was achieved using CuI in the presence diisopropylethylamine (DIPEA) as a catalyst and running the reaction for 8 h under reflux. At the same time, usage of the alkyne synthesized from tetrahydropyran derivative of iron bis(dicarbollide) 9 in “click” reaction with 2 leads to dramatically decrease of yield of product 12 (54 %). However, the alkyne synthesized from 1,4-dioxane derivatives of iron bis(dicarbollide) 8 does not react with azido-derivative 2 at all under the same conditions as for the compounds 6, 7, 9 and leads to the recovery of the starting materials. It should be noted that a similar difference of behavior of iron and cobalt complexes was observed earlier in the reaction of cobalt/iron bis(dicarbollide) terminal alkynes 69 with azido-cholesterol [48]. Alkyne 8 had not reacted with azido-cholesterol and had not led to target triazole 12 in contrast to alkynes 6, 7 and 9. These results pushed us to carry out the reaction of alkyne 8 with methyl azidoacetate under the same conditions as for compounds 6, 7 and 9. However, we did not get the desired result. Employing another amine and/or solvent (e.g., Et3N in CH3CN, Et3N in EtOH, DIPEA in CH3CN) did not provide the required material as well. The use of a 3-fold excess of CuI, EtOH as a solvent, DIPEA as a base and running the reaction for 8 h under reflux also did not lead to formation of the target 1,2,3-triazole (Scheme 4).
Structures of the compounds 1012 were established by 1H-, 11B- and 13C-NMR and IR spectra. In IR spectra of them the absorption bands of BH (2524–2531 cm-1) and 1,2,3-triazole (1461–1464 cm−1) were observed. In the 1H-NMR spectra of the obtained compounds, the characteristics signals of the CHtriazole hydrogens appear in the region at 6.52–8.64 ppm. In the 13C-NMR spectra, the signals of the CHtriazole carbons for 1012 are observed in the range of 126.8–129.1 ppm, whereas the signals of the Ctriazole carbons appear in the range 133.8-135.9 ppm. In IR spectra of 1012 the characteristic bands 1,2,3-triazoles (1461–1464 cm−1) were observed.
The compounds obtained are of potential interest for the delivery of boron-enriched drugs in boron neutron capture cancer therapy.

2.4. Single-Crystal X-ray Diffraction Studies

The structure of 9-N3CH2Me2N-nido-7,8-C2B9H11 2 was additionally confirmed by the single crystal X-ray diffraction study (Figure 1). Crystals of 2 suitable of single crystal X-ray analysis were grown from the dichloromethane solution layered with hexane.
Structural features of 2 are expected for this class of compounds (see, for instance [19]) with the only exception of rotation of the substituent at the B9 atom (Figure 1). Namely, the compact conformation of substituent is observed that results in rather short intramolecular contact between the N4 atom of azide function and the H1C extra-hydrogen atom of nido-carborane cage (3.089 Å with non-normalized B-H bond lengths). A similar contact of the extra hydrogen atom with the acetylene fragment previously found in the crystal of 10-Me3SiC≡CCH2(Me)S-7,8-C2B9H11 was attributed to the attractive B-H…π interaction which hindered free rotation of substituent in a solution [16]. For 2, the DFT calculation of the isolated molecule was performed to get more insight into the nature of the H1C…N4 contact. It was found that the conformation of 2 is rather rigid and is influenced by media effects only slightly: the best overlap for non-hydrogen atoms of crystal and isolated molecular structures produces a rather small rmsd value of 0.17 Å (Figure 2). Despite the most deviations are observed for the coordinates of azide fragment, the H1C…N4 contact lengthens insignificantly upon the transition into the isolated state (3.217 Å). According to the analysis of theoretical electron density function ρ(r) in terms of the R. Bader’s “Atoms in Molecules” [65] theory, the compact conformation of 2 can be indeed stabilized by intramolecular non-covalent interactions. In particular, there is a bond path connecting the H1C and N4 atoms which can be considered as the privileged channel of interatomic exchange interaction (Figure 3) [66,67]. This bond path is significantly curved in the area of the azide fragment and less curved in the area of the H1C atom. Implying a non-directional character of the interaction the bond path curvature allows to consider at least the π-electrons of the N3-N4 fragment to be involved into the non-covalent bonding with the H1C atom. This is in concordance with the another real-space descriptor of non-covalent interactions based on the reduced density gradient formalism [68]. The distribution of the sign (λ2)·ρ(r) function (λ2 is the intermediate eigenvalue of the ρ(r) Hessian) mapped onto the 0.4 isosurface of reduced density gradient (see Figure 3) is characterized by a rather wide region of negative values of sign (λ2)·ρ(r) which correspond to the concentration of electronic charge between H1C, N4 and N3 atoms. It is interesting to note that the distribution of Bader’s integral atomic charges does not allow to unambiguously treat this interaction as the H-bond (the charges of H1C, N4 and N3 atoms equal to −0.56e, +0.12e and −0.16e, respectively). Finally, according to the ρ(r) surface integrals scheme [69], this non-covalent interaction of the extra-hydrogen atom is rather strong (2.1 kcal·mol−1) and, thus, can be indeed regarded as the important factor which stabilizes the compact conformation of 2 and, probably, relative compounds.

3. Materials and Methods

3.1. General Methods

9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 [19], alkynyl-3β-cholesterol [53], cobalt bis(dicarbollides) alkynes 6 and 7 [48], iron bis(dicarbollides) alkynes 8 and 9 [64] were prepared according to the literature. Cholesterol (Fisher Scientific, Loughborough, U.K.), diisopropylethylamine (Carl Roth GmbH, Karlsruhe, Germany), CuI (PANREAC QUIMICA SA, Barcelona, Spain), sodium iodide (Sigma-Aldrich Chemie GmbH, Steinheim, Germany) were used without further purification. DMF, ethanol, CH3CN, CH2Cl2, CH3Cl, Et3N, phenylacetylene and NaN3 were commercially analytical grade reagents. The reaction progress was monitored by thin-layer chromatography (Merck F245 silica gel on aluminum plates). Acros Organics silica gel (0.060–0.200 mm) was used for column chromatography. The NMR spectra at 400.1 MHz (1H), 128.4 MHz (11B) and 100.0 MHz (13C) were recorded with a Bruker Avance-400 spectrometer (Bruker, Karlsruhe-Zurich, Switzerland-Germany). The residual signal of the NMR solvent relative to Me4Si was taken as the internal reference for 1H- and 13C-NMR spectra. 11B-NMR spectra were reference using BF3*Et2O as external standard. Infrared spectra were recorded on an IR Prestige-21 (SHIMADZU, Kyoto, Japan) instrument. High resolution mass spectra (HRMS) were measured on a mictOTOF II (Bruker Daltonic, Bremen, Germany) instrument using electrospray ionization (ESI). The measurements were done in a positive ion mode (interface capillary voltage −4500 V); mass range from m/z 50 to m/z 3000; external or internal calibration was done with ESI Tuning Mix, Agilent. A syringe injection was used for solutions in acetonitrile (flow rate 3 µL/min). Nitrogen was applied as a dry gas; interface temperature was set at 180 °C. The electron ionization mass spectra were obtained with a Kratos MS 890 instrument operating in a mass range of m/z 50–800.
Single crystal X-ray diffraction experiment for 2 was performed using with a Bruker SMART APEX 2 Duo CCD diffractometer [λ(MoKα) = 0.71072Å, ω-scans, 2θ < 60°] (Bruker AXC, Maddison, Wisconsin, USA). The structure was solved by the direct method and refined by the full-matrix least-squares technique against F2 in the anisotropic-isotropic approximation. The hydrogen atoms were found in the difference Fourier synthesis and were further refined in the isotropic approximation without restraints. All calculations were performed using SHELX2018 [70]. The CCDC 2048027 contains the Supplementary Crystallographic Data for 2. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/conts/retrieving.html (or from the CCDC, 12 Union Road, Cambridge, CB21EZ, UK; or [email protected]).
The DFT calculations were done in the Gaussian09 program (rev. D01) [71] using the PBE0 functional [72,73] with the Grimme’s D3 dispersion correction [74] and Becke–Jonson damping [75]. The standard def2tzvp basis set and ultrafine integration grids were used. The geometry optimization procedure for the isolated molecule of 2 was performed using standard criteria on displacements and forces. The analysis of the Hessian of potential energy surface for 2 revealed the correspondence of the isolated structure to the energy minimum. The analysis of electron density based on the “Atoms in Molecules” theory together with the reduced density gradient and sign(λ2)·ρ(r) computations were performed in the AIMAll program [76].

3.2. Synthesis of 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 2

Compound 1 (1.00 g, 0.0034 mol) was dissolved in 50 mL of DMF and NaN3 (0.44 g, 0.0068 mol) and anhydrous NaI (0.05 g, 0.0003 mol) were added. The reaction mixture was heated at 50 °C for 7 days. Then the reaction mixture was cooled to room temperature, and 50 mL of H2O was added. The formed precipitate was filtered and dried in the air, then it was purified on silica column by using CH2Cl2-CH3CN (5:1) as eluent to give white solid of 2 (0.80 g, yield 90%). 1H-NMR (400 MHz, methanol-d4): δ 3.50 (2H, m, CH2N3), 3.37 (2H, s, Me2NCH2), 2.99 (3H, s, NMe), 2.93 (3H, s, NMe), 2.59 (1H, s, CHcarb), 2.15 (2H, s, CH2CH2CH2), 1.87 (1H, s, CHcarb), −3.42 (1H, br. s, Hextra) ppm; 11B-NMR (128 MHz, methanol-d4): δ 5.2 (1B, s), −5.4 (1B, d, J = 139), −16.9 (2B, unsolved d,), −17.5 (1B, unsolved d,) −25.0 (1B, d, J = 147), −27.4 (1B, d, J = 143), −32.2 (1B, d, J = 135 Hz), −38.7 (1B, d, J = 145 Hz) ppm; 13C-NMR (101 MHz, methanol-d4) δ 65.2 (CH2N3), 52.7 (NMe), 50.4 (NMe), 48.5 (Me2NCH2), 46.3 (br., Ccarb), 33.7 (br., Ccarb), 23.6 (CH2CH2CH2) ppm. IR (solid): ν˜ = 2537 cm−1 (BH), 2075 cm−1 (N3), HRMS-ESI+ m/z for [C7H23B9N4 + NH4]+ calcd 300.2435, found 300.2434. Crystallograhic data: crystals of 2 (C7H23B9N4, M = 260.58) are monoclinic, space group P21/c, at 120K: a = 6.7746(5), b = 26.5914(18), c = 8.8809(6), β = 111.415(2), V = 1489.41(18) Å3, Z = 4 (Z’ = 1), dcalc = 1.162 g·cm−3. wR2 and GOF converged to 0.1292 and 0.991 for all independent reflections; R1 = 0.0461 was calculated for 3133 observed reflections with I > 2σ(I)).

General Procedure for the Synthesis of the Compounds 3, 5, 1012

A mixture of compound 2 (1,2 eq.), alkene (1 eq.), diisopropylethylamine (0,5–1 mL) and CuI (0,1 eq.) in 10–20 mL ethanol was heated under reflux for 8 h. Then the reaction mixture was cooled to room temperature and was passed through ca. 2–3 cm of silica. Then solvent was removed in vacuo. The crude product was purified on a silica column using CH2Cl2-CH3CN as an eluent to give the desired products 3, 5, 1012.

3.3. Synthesis of 9-(Ph)C-CH-N3(CH2)3Me2N-nido-7,8-C2B9H11 3

Prepared from compound 2 (0.10 g, 0.38 mmol), phenylacetylene (0.035 mL, 0.033 g, 0.32 mmol), diisopropylethylamine (1 mL, 0.74 g, 5.73 mmol) and CuI (0.006 g, 0.03 mmol) in 20 mL of ethanol. The product was obtained as a a white solid of 3 (0.10 g, yield 85%). 1H-NMR (400 MHz, DMSO-d6) δ 8.62, (1H, s, CCHN3), 7.84 (2H, d, Ph, J = 7.6 Hz), 7.46 (2H, t, Ph, J = 7.5 Hz), 7.34 (1H, t, Ph, J = 7.4 Hz), 4.54 (2H, t, -CH2-CCHN3, J = 6.9 Hz,), 3.34 (2H, t, CH2NMe2, J = 8.5 Hz), 2.88 (6H, s, NMe2), 2.77 (1H, s, CHcarb), 2.41 (2H, m, CH2-CH2-CH2), 1.91 (1H, s, CHcarb), −3.45 (1H, br. s, Hextra) ppm; 11B-NMR (128 MHz, DMSO-d6) δ 5.7 (1B, s), −6.0 (1B, unsolved d), −18.0 (3B, unsolved d), −25.1 (1B, unsolved d), −27.5 (1B, d, J = 117 Hz), −32.7 (1B, unsolved d), −38.7 (1B, d, J = 129 Hz) ppm; 13C-NMR (101 MHz, DMSO-d6): δ 146.9 (CCHN3), 131.1 (Ph), 129.4 (Ph), 128.4 (Ph), 125.6 (Ph), 122.0 (CCHN3), 64.4 (CH2-NMe2), 53.2 (NMe2), 51.4 (NMe2), 47.5 (CH2-CCHN3), 34.1 (Ccarb), 25.0 (CH2-CH2-CH2) ppm. IR (solid): ν˜ = 2546 cm−1 (BH), 1462 cm−1 (triazole), HRMS-ESI+ m/z for [C15H29B9N4 + H]+ calcd 363.3377, found 363.3386.

3.4. Synthesis of 9-3β-Chol-O(CH2)C-CH-N3(CH2)3Me2N-nido-7,8-C2B9H11 5

Prepared from compound 2 (0.072 g, 0.28 mmol), alkynyl-cholesterol 4 (0.10 g, 0.23 mmol), diisopropylethylamine (1 mL, 0.74 g, 5.73 mmol) and CuI (0.002 g, 0.03 mmol) in 20 mL of ethanol. The product was obtained as a white solid of 5 (0.10 g, yield 85 %). 1H-NMR (400 MHz, acetone-d6) δ 8.01 (1H, s, CCHN3), 5.37 (1H, m, Cst(6)H), 4.64 (2H, s, -CH2-CCHN3), 4.63 (2H, m, CCHN3-CH2), 3.54 (2H, m, CH2NMe2), 3.31 (1H, m, Cst(3)H), 3.08 (3H, s, NMe2), 3.07 (3H, s, NMe2), 2.64 (3H, m, CH2, CHcarb), 2.42 (1H, m, Cst(H)), 2.19 (1H, m, Cst(H)), 1.88 (6H, m, Cst(H), CHcarb), 1.33 (20H, m, Cst(H)), 1.04 (3H, s, Cst(19)H3), 0.97 (3H, d, J = 6.4, Cst(21)H3,), 0.90 (3H, d, J = 1.4, Cst(26)H3), 0.88 (3H, d, J = 1.4, Cst(27)H3), 0.74 (3H, s, Cst(18)H3), −3.39 (1H, br. s., Hextra) ppm; 11B (128 MHz, acetone-d6) δ 5.4 (1B, s), −5.5 (1B, d, J = 141), −17.2 (2B, d, J = 180), −18.9 (1B, d, J = 120), −24.9 (1B, d, J = 149), −27.3 (1B, d, J = 142), −32.3 (1B, d, J = 166), −38.6 (1B, d, J = 144) ppm; 13C-NMR (101 MHz, acetone-d6) δ 145.8 (CCHN3), 140.8 (Cst(5)), 123.1 (CCHN3), 121.3 (Cst(6)), 78.2 (Cst(3)), 64.7 (CCHN3-CH2), 61.1 (CH2-CCHN3), 56.7 (Cst(14)),56.2 (Cst(17)), 52.8 (NMe2), 51.1 (NMe2), 50.3 (Cst(9)), 47.1 (CH2NMe2), 42.2 (Cst(4)), 39.8 (Ccarb), 39.4 (Cst(12)), 39.0 (Cst(13)), 37.1 (Cst(24)), 36.7 (Cst(1)), 36.1 (Cst(10)), 35.7 (Cst(22)), 34.0 (Ccarb), 31.9 (Cst(20)), 31.8 (Cst(8)), 28.2 (Cst(2)), 28.0 (Cst(7)), 27.8 (Cst(16)), 25.0 (Cst(25)), 24.1 (Cst(15)), 23.6 (CH2), 22.2 (Cst(23)), 22.0 (Cst(26), Cst(27)), 20.9 (Cst(11)), 18.9 (Cst(19)), 18.3 (Cst(21)), 11.4 (Cst(18)) ppm. IR (solid): ν˜ = 2685 cm−1 (BH), 1392 cm−1 (triazole), HRMS-ESI+ m/z for [C37H71B9N4O + H]+ calcd 685.6614, found: 685.6627.

3.5. Synthesis of 9-[(8′-Me2N(CH2)3N3CCH(CH2)NMe2(CH2CH2O)2-1′,2′-C2B9H10)-3′,3″-Co(1″,2″-C2B9H11)]-nido-7,8-C2B9H11 10

Prepared from compound 2 (0.06 g, 0.24 mmol), compound 6 (0.10 g, 0.20 mmol), diisopropylethylamine (1 mL, 0.74 g, 5.73 mmol) and CuI (0.004 g, 0.02 mmol) in 20 mL of ethanol. The product was obtained as an orange solid of 10 (0.13 g, yield 85%). 1H-NMR (400 MHz, acetone-d6) δ 8.57 (1H, s, CCHN3), 4.97 (2H, s, Me2N-CH2-CCHN3), 4.74 (2H, q, J = 6.3 Hz, CCHN3-CH2), 4.13 (4H, m, B-OCH2, CHcarb), 4.00 (2H, br. s, CHcarb), 3.74 (2H, m, B-OCH2), 3.66 (4H, m, 2[CH2NMe2]), 3.55 (2H, m, B-OCH2), 3.40 (6H, s, NMe2), 3.08 (3H, s, NMe2), 3.06 (3H, s, NMe2), 2.68 (3H, s, CH2, CHcarb), 1.89 (1H, s, CHcarb), −3.45 (1H, br. s., Hextra) ppm; 11B-NMR (128 MHz, acetone-d6) δ 24.5 (1B, s, B-O), 5.3 (1B, s, B-N), 0.2 (1B, d, J = 138), −2.8 (1B, d, J = 143), −6.9 (2B, d, J = 178), −6.9 (4B, d, J = 137), −9.1 (2B, unsolved d), −17.2 (3B, d, J = 148), −18.9 (2B, unsolved d), −20.2 (2B, d, J = 163), −22.4 (1B, unsolved d), −24.9 (1B, d, J = 150), −27.4 (1B, J = 140), −28.9 (1B, unsolved d), −32.4 (1B, d, J = 135), −38.6 (1B, d, J = 144) ppm; 13C-NMR (101 MHz, acetone-d6) δ 135.9 (CCHN3), 129.1 (CCHN3), 72.2 (OCH2), 68.8 (OCH2), 64.8 (CH2NMe2), 64.5 (CH2-N3CCH), 62.7 (Ccarb), 59.5 (Ccarb), 52.9 (NMe2), 52.2 (NMe2), 51.5 (NMe2), 47.5 (B-Me2NCH2), 46.5 (Ccarb), 34.0 (Ccarb), 25.1 (CH2) ppm. IR (solid): ν˜ = 2530 cm−1 (BH), 1463 cm−1 (triazole), HRMS-ESI+ m/z for [C20H61B27CoN5O2 + H]+ calcd 755.6936, found: 755.6929.

3.6. Synthesis of 9-[(8′-Me2N(CH2)3N3CCH(CH2)NMe2(CH2)5-1′,2′-C2B9H10)-3′,3″-Co(1″,2″-C2B9H11)]-nido-7,8-C2B9H11 11

Prepared from compound 2 (0.06 g, 0.24 mmol), compound 7 (0.10 g, 0.20 mmol), diisopropylethylamine (1 mL, 0.79 g, 6.10 mmol) and CuI (0.004 g, 0.02 mmol) in 20 mL of ethanol. The product was obtained as an orange solid of 11 (0.13 g, yield 85 %). 1H-NMR (400 MHz, acetone-d6) δ 8.64 (1H, s, CCHN3), 4.91 (2H, s, Me2N-CH2-CCHN3), 4.76 (2H, q, J = 6.4 Hz, CCHN3-CH2), 4.20 (2H, br. s, CHcarb), 4.10 (2H, br. s, CHcarb), 3.56 (2H, m, B-OCH2), 3.49 (4H, m, 2[CH2NMe2]), 3.33 (6H, s, NMe2), 3.09 (3H, s, NMe2), 3.08 (3H, s, NMe2), 2.71 (3H, s, CH2, CHcarb), 1.88 (1H, s, CHcarb), 1.57 (2H, m, CH2CH2CH2), 1.48 (4H, m, CH2CH2CH2), −3.45 (1H, br. s., Hextra) ppm; 11B-NMR (128 MHz, acetone-d6) δ 23.8 (1B, s, B-O), 5.1 (1B, s, B-N), 0.0 (1B, d, J = 138), −2.7 (1B, d, J = 150), 4.6 (1B, unsolved d), −5.1 (1B, unsolved d), −7.7 (8B, m), −17.3 (3B, d, J = 134), −18.8 (2B, unsolved d), −20.3 (2B, d, J = 153), −22.4 (1B, unsolved d), −25.0 (1B, d, 171), −27.4 (1B, d, J = 151), −28.8 (1B, d, J = 143), −32.4 (1B, d, J = 118), −38.6 (1B, d, J = 171) ppm; 13C-NMR (101 MHz, acetone-d6) δ 135.5 (CCHN3), 128.3 (CCHN3), 68.2 (OCH2), 64.4 (CH2NMe2), 64.4 (CH2-N3CCH), 58.1 (Ccarb), 53.2 (Ccarb), 53.0 (NMe2), 51.6 (NMe2), 50.3 (NMe2), 47.6 (B-Me2NCH2), 46.4 (Ccarb), 34.0 (Ccarb), 25.1 (CH2), 23.1 (CH2), 22.1 (CH2) ppm. IR (solid): ν˜ = 2531 cm−1 (BH), 1461 cm−1 (triazole), HRMS-ESI+ m/z for [C21H63B27CoN5O + H]+ calcd 753.7143, found 753.7138.

3.7. Synthesis of 9-[(8′-Me2N(CH2)3N3CCH(CH2)NMe2(CH2)5-1′,2′-C2B9H10)-3′,3″-Fe(1″,2″-C2B9H11)]-nido-7,8-C2B9H11 12

Prepared from compound 2 (0.03 g, 0.12 mmol), compound 9 (0.05 g, 0.10 mmol), diisopropylethylamine (0.5 mL, 0.37 g, 2.87 mmol) and CuI (0.002 g, 0.01 mmol) in 20 mL of ethanol. The product was obtained as an orange solid of 12 (0.04 g, yield 54%). 1H-NMR (400 MHz, acetone-d6) δ 72.2 (br. s, BH), 67.38 (br. s, BH), 45.15 (br. s, BH), 43.16 (br. s, BH), 37.74 (br. s, CHcarb), 6.52 (1H, s, CCHN3), 4.27 (2H, br s, Me2N-CH2-CCHN3), 3.15 (2H, t, J = 7.7, CH2NMe2), 2.96 (3H, s, NMe2), 2.89 (3H, s, NMe2), 2.79 (6H, s, NMe2), 2.53 (1H, s, CHcarb), 2.28 (2H, m, CH2), 1.64 (1H, br s, CHcarb), −3.88 (1H, br s, Hextra), −5.25 (2H, m, OCH2), −10.22 (2H, br s, CH2), −11.26 (2H, br s, CH2), −11.85 (2H, br s, CH2) ppm; 11B-NMR (128 MHz, acetone-d6) δ 118.8, 100.8, 36.9, 27.1, 5.0, −1.8, −6.1, −17.5, −19.1, −25.2, −27.8, −32.8, −38.9, −370.6, −382.7, −448.1, −486.7 ppm; 13C-NMR (101 MHz, acetone-d6) δ 133.8 (CCHN3), 126.8 (CCHN3), 64.0 (CH2-N3CCH), 59.5 (CH2NMe2), 56.0 (CH2NMe2), 52.7 (NMe2), 51.4 (NMe2), 48.1 (B-Me2NCH2), 47.1 (Ccarb), 35.9 (Ccarb), 24.5 (CH2), 14.9 (CH2), 14.5 (CH2), 10.8 (B-OCH2) ppm. IR (solid): ν˜ = 2524 cm−1 (BH), 1464 cm−1 (triazole), HRMS-ESI+ m/z for [C21H63B27FeN5O + H]+ calcd 750.7161, found 750.7154.

4. Conclusions

In this work we prepared and characterized novel nido-carboranyl azide 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 derived from the reaction of 9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 with NaN3 in the presence of NaI as a catalyst in strong conditions. The possibility of using “click” approach in regard to the obtained compound was demonstrated on the reaction of 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 with phenylacetylene. We also studied the behaviour of nido-carboranyl azide in the copper(I)-catalyzed azide-alkyne cycloaddition reaction with alkynyl-cholesterol and obtained new the nido-carborane-cholesterol conjugate with charge-compensated group in a linker in a good yield. Based on the synthesized compound, the boronated liposomes are planned for preparation in order to deliver boron clusters into a cancer cell for a BNCT experiment. It should be concluded that ”click“ reactions between 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 and cobalt/iron bis(dicarbollide) terminal alkynes lead to novel zwitter-ionic boron-enriched cluster compounds bearing a 1,2,3-triazol-metallacarborane-carborane conjugated system. By changing the type and the size of a spacer between these two boron cages, it is possible to control the hydrophilic/hydrophobic balance of the compounds. Prepared conjugates contain a large amount of boron atom in the biomolecule potentially can be used for boron neutron capture therapy of cancer (BNCT).
The solid-state molecular structure of the novel nido-carboranyl azide 9-N3(CH2)3Me2N-nido-7,8-C2B9H11 was determined by single-crystal X-ray diffraction studies. Structural feature of nido-carboranyl azide is rather rigid and compact conformation of substituent is observed that results in rather short intramolecular contact between the N4 atom of azide function and the H1C extra-hydrogen atom of nido-carborane cage (3.089 Å with non-normalized B-H bond lengths).

Supplementary Materials

The following are available online, 1H-, 11B- and 13C-NMR, IR and high-resolution mass spectra of compounds 2, 3, 5, 1012.

Author Contributions

Supervision and manuscript concept, A.A.D.; synthesis, O.B.Z.; synthesis, N.V.D.; NMR spectroscopy studies, I.D.K.; single crystal X-ray diffraction, I.V.A.; synthesis, S.V.T.; supervision and manuscript concept, V.I.B. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (Grant No. 20-73-00228).

Data Availability Statement

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

Acknowledgments

The NMR spectroscopy and X-ray diffraction data were obtained by using equipment from the Center for Molecular Structure Studies at A.N. Nesmeyanov Institute of Organoelement Compounds, operating with support from the Ministry of Science and Higher Education of the Russian Federation.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: London, UK, 2016; 905p, ISBN 978-0-12-801894-1. [Google Scholar]
  2. Hosmane, N.S.; Maguire, J.A. Comprehensive Organometallic Chemistry III.; Michael, D., Mingos, P., Crabtree, R.H., Eds.; Elsevier: Amsterdam, The Netherlands, 2007; 264p, ISBN 978-0-0804-4601-1. [Google Scholar]
  3. Hawthorne, M.F. The role of chemistry in the development of boron neutron capture therapy of cancer. Angew. Chem. Int. Ed. Engl. 1993, 32, 950–998. [Google Scholar] [CrossRef]
  4. Valliant, J.F.; Guenther, K.J.; King, A.S.; Morel, P.; Schaffer, P.; Sogbein, O.O.; Stephenson, K.A. The medicinal chemistry of carboranes. Coord. Chem. Rev. 2002, 232, 173–230. [Google Scholar] [CrossRef]
  5. Armstrong, A.F.; Valliant, J.F. The bioinorganic and medicinal chemistry of carboranes: From new drug discovery to molecular imaging and therapy. Dalton Trans. 2007, 4240–4251. [Google Scholar] [CrossRef] [PubMed]
  6. Řezačova, P.; Cigler, P.; Matejiček, P.; Lipšik, M.; Pokorna, J.; Grüner, B.; Konvalinka, J. Medicinal application of carboranes: Inhibition of HIV protease. Boron Science: New Technologies and Applications; Hosmane, N.S., Ed.; CRC Press: Boca Raton, FL, USA, 2012; 70p, ISBN 978-1-43-982662-1. [Google Scholar]
  7. Tolmachev, V.V.; Sjöberg, S. Polyhedral boron compounds as potential linkers for attachment of radiohalogens to targeting proteins and peptides. A review. Collect. Czech. Chem. Commun. 2002, 67, 913–935. [Google Scholar] [CrossRef]
  8. Lesnikowski, Z.J. New Opportunities in Boron Chemistry for Medical Applications. Boron Science: New Technologies and Applications; Hosmane, N.S., Ed.; CRC Press: Boca Raton, FL, USA, 2012; 3p, ISBN 978-1-43-982662-1. [Google Scholar]
  9. Sivaev, I.B.; Bregadze, V.I. Polyhedral boranes for medical applications: Current status and perspectives. Eur. J. Inorg. Chem. 2009, 11, 1433–1450. [Google Scholar] [CrossRef]
  10. Tsygankova, A.R.; Kanygin, V.V.; Kasatova, A.I.; Zavyalov, E.L.; Guselnikova, T.Y.; Kichigin, A.I.; Mukhamadiyarov, R.A. Determination of boron by inductively coupled plasma atomic emission spectroscopy. Biodistribution of 10B in tumor-bearing mice. Russ. Chem. Bull. 2020, 69, 601–607. [Google Scholar] [CrossRef]
  11. Stogniy, M.; Abramova, E.N.; Lobanova, I.A.; Sivaev, I.B.; Bragin, V.I.; Petrovskii, P.V.; Tsupreva, V.N.; Sorokina, O.V.; Bregadze, V.I. Synthesis of Functional Derivatives of 7,8-Dicarba-nido-undecaborate Anion by Ring-Opening of Its Cyclic Oxonium Derivatives. Collect. Czechoslov. Chem. Commun. 2007, 72, 1676–1688. [Google Scholar] [CrossRef]
  12. Stogniy, M.Y.; Kazakov, G.S.; Sivaev, I.B.; Bregadze, V.I. Synthesis of podands with nido-carboranyl groups as a basis for construction of crown ethers with an incorporated metallacarborane moiety. Russ. Chem. Bull. 2013, 62, 699–704. [Google Scholar] [CrossRef]
  13. Kazakov, G.S.; Stogniy, M.Y.; Sivaev, I.B.; Suponitsky, K.Y.; Godovikov, I.A.; Kirilin, A.D.; Bregadze, V.I. Synthesis of crown ethers with the incorporated cobalt bis(dicarbollide) fragment. J. Organomet. Chem. 2015, 798, 196–203. [Google Scholar] [CrossRef]
  14. Wojtczak, B.A.; Andrysiak, A.; Grüner, B.; Lesnikowski, Z.J. “Chemical ligation”: A versatile method for nucleoside modification with boron clusters. Chem. Eur. J. 2008, 14, 10675–10682. [Google Scholar] [CrossRef]
  15. Zakharova, M.V.; Sivaev, I.B.; Anufriev, S.A.; Timofeev, S.V.; Suponitsky, K.Y.; Godovikov, I.A.; Bregadze, V.I. A new approach to the synthesis of functional derivatives of nido-carborane: Alkylation of [9-MeS-nido-7,8-C2B9H11]". Dalton Trans. 2014, 43, 5044–5053. [Google Scholar] [CrossRef]
  16. Anufriev, S.A.; Sivaev, I.B.; Suponitsky, K.Y.; Godovikov, I.A.; Bregadze, V.I. Synthesis of 10-methylsulfide and 10-alkylmethylsulfonium nido-carborane derivatives: B–H···π Interactions between the B–H–B hydrogen atom and alkyne group in 10-RC≡CCH2S(Me)-7,8-C2B9H11. Eur. J. Inorg. Chem. 2017, 38–39, 4436–4443. [Google Scholar] [CrossRef] [Green Version]
  17. Anufriev, S.A.; Zakharova, M.V.; Sivaev, I.B.; Bregadze, V.I. New carborane-containing acids and amines. Russ. Chem. Bull. 2017, 66, 1643–1649. [Google Scholar] [CrossRef]
  18. Erokhina, S.A.; Stogniy, M.Y.; Suponitsky, K.Y.; Kosenko, I.D.; Sivaev, I.B.; Bregadze, V.I. Experimental crystal structure determination. Polyhedron. 2018, 153, 145–151. [Google Scholar] [CrossRef]
  19. Timofeev, S.V.; Zhidkova, O.B.; Prikaznova, E.A.; Sivaev, I.B.; Semioshkin, A.; Godovikov, I.A.; Starikova, Z.A.; Bregadze, V.I. Direct synthesis of nido-carborane derivatives with pendant functional groups by copper-promoted reactions with dimethylalkylamines. J. Organomet. Chem. 2014, 757, 21–27. [Google Scholar] [CrossRef]
  20. Stogniy, M.; Erokhina, S.A.; Suponitsky, K.Y.; Anisimov, A.A.; Sivaev, I.B.; Bregadze, V.I. Nucleophilic addition reactions to the ethylnitrilium derivative of nido-carborane 10-EtCN-7,8-C2B9H11. New J. Chem. 2018, 42, 17958–17967. [Google Scholar] [CrossRef]
  21. Stanislav, I.; Presolski, V.; Hong, P.; Finn, M.G. Copper-catalyzed azide-alkyne click chemistry for bioconjugation. Curr. Protoc. Chem. Biol. 2011, 3, 153–162. [Google Scholar] [CrossRef] [Green Version]
  22. Liang, L.; Astruc, D. The copper(I)-catalyzed alkyne-azide cycloaddition (CuAAC) “click” reaction and its applications. An overview. Coord. Chem. Rev. 2011, 255, 2933–2945. [Google Scholar] [CrossRef]
  23. Jiang, Z.; He, H.; Liu, H.; Thayumanavan, S. Azide-Terminated RAFT Polymers for Biological Applications. Curr. Protoc. Chem. Biol. 2020, 12, e85. [Google Scholar] [CrossRef]
  24. Meldal, M.; Diness, F. Recent Fascinating Aspects of the CuAAC Click Reaction. Trends Chemistry. 2020. [Google Scholar] [CrossRef]
  25. Tiwari, V.K.; Mishra, B.B.; Mishra, K.B.; Mishra, N.; Singh, A.S.; Chen, X. Cu-Catalyzed Click Reaction in Carbohydrate Chemistry. Chem. Rev. 2016, 116, 3086–3240. [Google Scholar] [CrossRef] [PubMed]
  26. Yet, L. Five-Membered Ring Systems: With More than One N Atom. Progress in Heterocyclic Chemistry; Gribble, G.W., Joule, J.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; 229p, ISBN 978-0-08-044005-7. [Google Scholar]
  27. Rani, A.; Singh, G.; Singh, A.; Maqbool, U.; Kaur, G.; Singh, J. CuAAC-ensembled 1,2,3-triazole-linked isosteres as pharmacophores in drug discovery: Review. RSC Adv. 2020, 10, 5610–5635. [Google Scholar] [CrossRef]
  28. Xu, Z.; Zhao, S.-J.; Liu, Y. 1,2,3-Triazole-containing hybrids as potential anticancer agents: Current developments, action mechanisms and structure-activity relationships. Eur. J. Med. Chem. 2019, 183, 111700–111737. [Google Scholar] [CrossRef] [PubMed]
  29. Bozorov, K.; Zhao, J.; Aisa, H.A. 1,2,3-Triazole-containing hybrids as leads in medicinal chemistry: A recent overview. Bioorg. Med. Chem. 2019, 27, 3511–3531. [Google Scholar] [CrossRef] [PubMed]
  30. Karypidou, K.; Ribone, S.R.; Quevedo, M.A.; Persoons, L.; Pannecouque, C.; Helsen, C.; Claessens, F.; Dehaen, W. Synthesis, biological evaluation and molecular modeling of a novel series of fused 1,2,3-triazoles as potential anti-coronavirus agents. Bioorg. Med. Chem. Lett. 2018, 28, 3472–3476. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, B. Comprehensive review on the anti-bacterial activity of 1,2,3-triazole hybrids. Eur. J. Med. Chem. 2019, 168, 357–372. [Google Scholar] [CrossRef] [PubMed]
  32. El Malah, T.; Nour, H.F.; Satti, A.A.E.; Hemdan, B.A.; ElSayed, W.A. Design, synthesis and antimicrobial activities of 1,2,3- triazole glycoside clickamers. Molecules 2020, 25, 790. [Google Scholar] [CrossRef] [Green Version]
  33. Srinivas, B.; Kumar, P.V.; Reddy, P.N.; Venu, S.; Shyam, P.; Krupadanam, G.L.D. Design, synthesis, antioxidant and antibacterial activities of novel 2-((1-benzyl-1h-1,2,3-triazol-4- yl)methyl)-5-(2hchromen- 3-yl)-2h-tetrazoles. Russ. J. Bioorg. Chem. 2018, 44, 244–251. [Google Scholar] [CrossRef]
  34. Kaushik, C.P.; Luxmi, R. Synthesis, antibacterial and antioxidant activities of naphthyl-linked disubstituted 1,2,3-triazoles. J. Heterocycl. Chem. 2020, 57, 2400–2409. [Google Scholar] [CrossRef]
  35. 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 chlorine6conjugates via ‘click chemistry’. Appl. Organomet. Chem. 2009, 23, 370–374. [Google Scholar] [CrossRef]
  36. Best, M.D. “Click” chemistry and bioorthogonal reactions: Unprecedented selectivity in the labeling of biological molecules. Biochemistry 2009, 48, 6571–6584. [Google Scholar] [CrossRef] [PubMed]
  37. Stogniy, M.Y.; Erokhina, S.A.; Druzina, A.A.; Sivaev, I.B.; Bregadze, V.I. Synthesis of novel carboranyl azides and “click” reactions thereof. J. Organomet. Chem. 2019, 904, 121007–121014. [Google Scholar] [CrossRef]
  38. 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]
  39. 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]
  40. Kahl, S.B.; Pate, D.; Laster, B.H.; Popenoe, E.A.; Fairchild, R.G. In Vitro Biological Efficacy of Boronated Low Density Lipoproteins for NCT. Progress in Neutron Capture Therapy for Cancer; Allen, B.J., Moore, D.E., Harrington, B.V., Eds.; Plenum Press: New York, NY, USA, 1992; 365p, ISBN 978-1-4615-3384-9. [Google Scholar]
  41. Feakes, D.A.; Spinler, J.K.; Harris, F.R. Synthesis of boron-containing cholesterol derivatives for incorporation into unilamellar liposomes and evaluation as potential agents for BNCT. Tetrahedron 1999, 55, 11177–11186. [Google Scholar] [CrossRef]
  42. Ji, B.; Peacock, G.; Lu, D.R. Synthesis of cholesterol-carborane conjugate for targeted drug delivery. Bioorgan. Med. Chem. Lett. 2002, 12, 2455–2458. [Google Scholar] [CrossRef]
  43. Barth, R.F. Boron neutron capture therapy at the crossroads: Challenges and opportunities. Appl. Radiat. Isot. 2009, 67, S3–S6. [Google Scholar] [CrossRef]
  44. Maeda, H.; Nakamura, H.; Fang, J. The EPR effect for macromolecular drug delivery to solid tumors: Improvement of tumor uptake, lowering of systemic toxicity, and distinct tumor imaging in vivo. Adv. Drug. Deliv. Rev. 2013, 65, 71–79. [Google Scholar] [CrossRef]
  45. Eloy, J.O.; Claro de Souza, M.; Petrilli, R.; Abriata Barcellos, J.P.; Lee, R.J.; Marchetti, J.M. Liposomes as carriers of hydrophilic small molecule drugs: Strategies to enhance encapsulation and delivery. Colloids. Surf. B Biointerfaces 2014, 123, 345–363. [Google Scholar] [CrossRef]
  46. Jain, A.; Jain, S.K. Advances in tumor targeted liposomes. Curr. Mol. Med. 2018, 18, 44–57. [Google Scholar] [CrossRef]
  47. Druzina, A.A.; Shmalko, A.V.; Andreichuk, E.P.; Zhidkova, O.B.; Kosenko, I.D.; Semioshkin, A.; Sivaev, I.B.; Mandal, S.; Shen, Z.; Bregadze, V. ‘Click’ synthesis of cobalt bis(dicarbollide)–cholesterol conjugates. Mendeleev Commun. 2019, 29, 628–630. [Google Scholar] [CrossRef]
  48. Druzina, A.A.; Kosenko, I.D.; Zhidkova, O.B.; Ananyev, I.V.; Timofeev, S.V.; Bregadze, V.I. Novel cobalt bis(dicarbollide) based on terminal alkynes and their “click” reactions. Eur. J. Inorg. Chem. 2020, 27, 2658–2665. [Google Scholar] [CrossRef]
  49. 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]
  50. Bregadze, V.I.; Sivaev, I.B.; Semioshkin, A.; Shmal’ko, A.V.; Kosenko, I.D.; Lebedeva, K.V.; Mandal, S.; Sreejyothi, P.; Dubey, R.D.; Sarkar, A.; et al. Boron-containing lipids and liposomes: New conjugates of cholesterol with polyhedral boron hydrides. Chem. Eur. J. 2020, 26, 13832–13841. [Google Scholar] [CrossRef]
  51. Dubey, R.D.; Sarkar, A.; Shen, Z.; Bregadze, V.I.; Sivaev, I.B.; Druzina, A.A.; Zhidkova, O.B.; Shmal’ko, A.V.; Kosenko, I.D.; Sreejyothi, P.; et al. Effects of linkers on the development of liposomal formulation of cholesterol conjugated cobalt bis(dicarbollides). J. Pharm. Sci. 2020. [Google Scholar] [CrossRef]
  52. Lee, W.; Sarkar, S.; Ahn, H.; Kim, J.Y.; Lee, Y.J.; Chang, Y.; Yoo, J. PEGylated liposome encapsulating nido-carborane showed significant tumor suppression in boron neutron capture therapy (BNCT). Biochem. Biophys. Res. Commun. 2020, 522, 669–675. [Google Scholar] [CrossRef]
  53. Aly, M.R.E.; Saad, H.A.; Mohamed, M.A.N. Difluoromethyl ketones: Potent inhibitors of wild type and carbamate-insensitive G119S mutant anopheles gambiae acetylcholinesterase. Bioorg. Med. Chem. Lett. 2015, 25, 4405–4411. [Google Scholar] [CrossRef] [Green Version]
  54. Šícha, V.; Farràs, P.; Štíbr, B.; Teixidor, F.; Grüner, B.; Viñas, C. Syntheses of C-substituted icosahedral dicarbaboranes bearing the 8-dioxane-cobalt bisdicarbollide moiety. J. Organomet. Chem. 2009, 694, 1599–1601. [Google Scholar] [CrossRef]
  55. Mishima, Y.; Ichihashi, M.; Hatta, S.; Honda, C.; Yamamura, K.; Nakagawa, T. New thermal neutron capture therapy for malignant melanoma: Melanogenesis-seeking 10B molecule-melanoma cell interaction from in vitro to first clinical trial. Pigm. Cell Res. 1989, 2, 226–234. [Google Scholar] [CrossRef]
  56. Barth, R.F.; Mi, P.; Yang, W. Boron delivery agents forb neutron capture therapy of cancer. Cancer Commun. 2018, 38, 35–46. [Google Scholar] [CrossRef] [Green Version]
  57. Fuentes, I.; Garcia-Mendiola, T.; Sato, S.; Pita, M.; Nakamura, H.; Lorenzo, E.; Teixidor, F.; Marques, F.; Vinas, C. Metallacarboranes on the road to anticancer therapies: Cellular uptake, DNA interaction, and biological evaluation of cobaltabisdicarbollide [COSAN]. Chem. Eur. J. 2018, 24, 17239–17254. [Google Scholar] [CrossRef] [PubMed]
  58. Tarrés, M.; Canetta, E.; Paul, E.; Forbes, J.; Azzouni, K.; Viñas, C.; Teixidor, F.; Harwood, A. Biological interaction of living cells with COSAN-based synthetic vesicles. J. Sci. Rep. 2015, 5, 7804–7811. [Google Scholar] [CrossRef] [PubMed]
  59. Bauduin, P.; Prevost, S.; Farràs, P.; Teixidor, F.; Diat, O.; Zemb, T. A theta-shaped amphiphilic cobaltabisdicarbollide anion: Transition from monolayer vesicles to micelles. Angew. Chem. Int. Ed. 2011, 50, 5298–5300. [Google Scholar] [CrossRef] [PubMed]
  60. Malaspina, D.C.; Viñas, C.; Teixidor, F.; Faraudo, J. Atomistic simulations of COSAN: Amphiphiles without a head-and-tail design display «head and tail» surfactant behavior. J. Angew. Chem. Int. Ed. 2020, 59, 3088–3092. [Google Scholar] [CrossRef] [PubMed]
  61. Gratton, S.E.A.; Ropp, P.A.; Pohlhaus, P.D.; Luft, J.C.; Madden, V.J.; Napier, M.E.; DeSimone, J.M. The effect of particle design on cellular internalization pathways. Proc. Natl. Acad. Sci. USA 2008, 105, 11613–11618. [Google Scholar] [CrossRef] [Green Version]
  62. Chen, L.; Mccrate, J.M.; Li, H.; Lee, J.C.-M. The role of surface charge on the uptake and biocompatibility of hydroxyapatite nanoparticles with osteoblast cells. Nanotechnology 2011, 22, 105708–105717. [Google Scholar] [CrossRef] [Green Version]
  63. Tatur, S.; MacCarini, M.; Barker, R.; Nelson, A.; Fragneto, G. Effect of functionalized gold nanoparticles on floating lipid bilayers. Langmuir 2013, 29, 6606–6614. [Google Scholar] [CrossRef]
  64. Lobanova, I.; Kosenko, I.; Laskova, J.; Ananyev, I.; Druzina, A.; Godovikov, I.; Bregadze, V.; Qi, S.; Lesnikowski, Z.J.; Semioshkin, A. Synthesis and the structure of 8-tetrahydrofuronium and 8-tetrahydropyronium derivatives of iron bis(dicarbollide)(-I) and their cleavage reactions. Dalton Trans. 2015, 44, 1571–1584. [Google Scholar] [CrossRef]
  65. Matta, C.F.; Boyd, R.J. The Quantum Theory of Atoms in Molecules: From Solid State to DNA, Drug Design; Wiley-VCH Verlag GmbH & Co. KgaA: Weinheim, Germany, 2007; 567p, ISBN 978-3-527-30748-7. [Google Scholar]
  66. Pendas, A.M.; Francisco, E.; Blanco, M.A.; Gatti, C. Bond paths as privileged exchange channels. Chem. Eur. J. 2007, 13, 9362–9371. [Google Scholar] [CrossRef]
  67. Anisimov, A.A.; Ananyev, I.V. Interatomic exchange-correlation interaction energy from a measure of quantum theory of atoms in molecules topological bonding: A diatomic case. J. Comput. Chem. 2020, 41, 2213–2222. [Google Scholar] [CrossRef]
  68. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing non-covalent iinteractions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  69. Romanova, A.; Lyssenko, K.; Ananyev, I. Estimations of energy of noncovalent bonding from integrals over interatomic zero-flux surfaces: Correlation trends and beyond. J. Comput. Chem. 2018, 39, 1607–1616. [Google Scholar] [CrossRef] [PubMed]
  70. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, C71, 3–8. [Google Scholar] [CrossRef]
  71. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H.; et al. Gaussian 09, Revision D.01; Gaussian, Inc.: Wallingford, CT, USA, 2016. [Google Scholar]
  72. Perdew, J.P.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  73. Adamo, C.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  74. Grimme, S.; Antony, J.; Ehrlich, S.; Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. J. Chem. Phys. 2010, 132, 154104. [Google Scholar] [CrossRef] [Green Version]
  75. Grimme, S.; Ehrlich, S.; Goerigk, L. Effect of the damping function in dispersion corrected density functional theory. J. Comput. Chem. 2011, 32, 1456–1465. [Google Scholar] [CrossRef]
  76. Keith, T.A. AIMAll, version 13.05.06; TK Gristmill Software: Overland Park, KS, USA, 2019. [Google Scholar]
Scheme 1. Reaction of 9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 with NaN3.
Scheme 1. Reaction of 9-Cl(CH2)3Me2N-nido-7,8-C2B9H11 with NaN3.
Molecules 26 00530 sch001
Scheme 2. “Click” reaction of nido-carboranyl azide with phenylacetylene.
Scheme 2. “Click” reaction of nido-carboranyl azide with phenylacetylene.
Molecules 26 00530 sch002
Scheme 3. Synthesis of conjugate of nido-carborane with cholesterol.
Scheme 3. Synthesis of conjugate of nido-carborane with cholesterol.
Molecules 26 00530 sch003
Scheme 4. Synthesis of conjugates of nido-carborane with cobalt/iron bis(dicarbollide)s.
Scheme 4. Synthesis of conjugates of nido-carborane with cobalt/iron bis(dicarbollide)s.
Molecules 26 00530 sch004
Figure 1. General view of 2 in crystal with the representation of non-hydrogen atoms by ellipsoids of atomic displacement (p = 0.5). The intramolecular H1C…N4 contact is shown by the dashed line. Selected bond lengths: C7-C8 1.551(2) Å, B10-B11 1.862(2) Å, B9-N1 1.583(2) Å, H1C…N4 3.089 Å (non-normalized B-H bond lengths).
Figure 1. General view of 2 in crystal with the representation of non-hydrogen atoms by ellipsoids of atomic displacement (p = 0.5). The intramolecular H1C…N4 contact is shown by the dashed line. Selected bond lengths: C7-C8 1.551(2) Å, B10-B11 1.862(2) Å, B9-N1 1.583(2) Å, H1C…N4 3.089 Å (non-normalized B-H bond lengths).
Molecules 26 00530 g001
Figure 2. The best root-mean-square overlap of non hydrogen atoms of crystal (solid lines) and isolated (dashed lines) structures of 2.
Figure 2. The best root-mean-square overlap of non hydrogen atoms of crystal (solid lines) and isolated (dashed lines) structures of 2.
Molecules 26 00530 g002
Figure 3. The atomic connectivity graph of the isolated molecule of 2 determined by means of the “Atoms in Molecules” theory and the sign(λ2)·ρ(r) distribution mapped onto the 0.4 reduced density gradient isosurface between the H1C, N4 and N3 atoms (from −2.6e−3 a.u. in red to +2.6e−3 a.u. in blue). Bond paths for intramolecular non-covalent interactions are shown by dashed curves.
Figure 3. The atomic connectivity graph of the isolated molecule of 2 determined by means of the “Atoms in Molecules” theory and the sign(λ2)·ρ(r) distribution mapped onto the 0.4 reduced density gradient isosurface between the H1C, N4 and N3 atoms (from −2.6e−3 a.u. in red to +2.6e−3 a.u. in blue). Bond paths for intramolecular non-covalent interactions are shown by dashed curves.
Molecules 26 00530 g003
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Druzina, A.A.; Zhidkova, O.B.; Dudarova, N.V.; Kosenko, I.D.; Ananyev, I.V.; Timofeev, S.V.; Bregadze, V.I. Synthesis and Structure of Nido-Carboranyl Azide and Its “Click” Reactions. Molecules 2021, 26, 530. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26030530

AMA Style

Druzina AA, Zhidkova OB, Dudarova NV, Kosenko ID, Ananyev IV, Timofeev SV, Bregadze VI. Synthesis and Structure of Nido-Carboranyl Azide and Its “Click” Reactions. Molecules. 2021; 26(3):530. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26030530

Chicago/Turabian Style

Druzina, Anna A., Olga B. Zhidkova, Nadezhda V. Dudarova, Irina D. Kosenko, Ivan V. Ananyev, Sergey V. Timofeev, and Vladimir I. Bregadze. 2021. "Synthesis and Structure of Nido-Carboranyl Azide and Its “Click” Reactions" Molecules 26, no. 3: 530. https://0-doi-org.brum.beds.ac.uk/10.3390/molecules26030530

Article Metrics

Back to TopTop