Next Article in Journal
Synthesis by Hydrothermal Treatment of ZnO-Based Varistors Doped with Rare Earth Oxides and Their Characterization by Impedance Spectroscopy
Next Article in Special Issue
Exploration of Bis(nickelation) of 1,1′-Bis(o-carborane)
Previous Article in Journal
A Wave-Based Vibration Analysis of a Finite Timoshenko Locally Resonant Beam Suspended with Periodic Uncoupled Force-Moment Type Resonators
Previous Article in Special Issue
Reactions of Experimentally Known Closo-C2B8H10 with Bases. A Computational Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

“Free of Base” Sulfa-Michael Addition for Novel o-Carboranyl-DL-Cysteine Synthesis †

1
A.N.Nesmeyanov Institute of Organoelement Compounds, Russian Academy of Sciences, Vavilova str. 28, 119991 Moscow, Russia
2
M.V. Lomonosov Institute of Fine Chemical Technology, MIREA—Russian Technological University, 86 Vernadsky Av., 119571 Moscow, Russia
*
Author to whom correspondence should be addressed.
Dedicated to Professor Alan J. Welch in recognition of his outstanding contribution in the chemistry of carboranes.
Crystals 2020, 10(12), 1133; https://doi.org/10.3390/cryst10121133
Submission received: 2 November 2020 / Revised: 1 December 2020 / Accepted: 7 December 2020 / Published: 11 December 2020
(This article belongs to the Special Issue Carborane: Dedicated to the Work of Professor Alan Welch)

Abstract

:
The sulfa-Michael addition reaction was applied for the two-step synthesis of o-carboranyl cysteine 1-HOOCCH(NH2)CH2S-1,2-C2B10H11 from the trimethylammonium salt of 1-mercapto-o-carborane and methyl 2-acetamidoacrylate. To avoid the decapitation of o-carborane into its nido-form, the “free of base” method under mild conditions in a system of two immiscible solvents toluene-H2O was developed. The replacement of H2O by 2H2O resulted in carboranyl-cysteine containing a deuterium label at the α-position of the amino acid 1-HOOCCD(NH2)CH2S-1,2-C2B10H11. The structure of the protected o-carboranyl cysteine was determined by single-crystal X-ray diffraction. The obtained compounds can be considered as potential agents for the Boron Neutron Capture Therapy of cancer.

1. Introduction

Cancer remains one of the major health issues and the second leading cause of death worldwide [1]. Radiation therapy has a central role in the management of malignant brain tumors, especially for the most aggressive ones [2,3,4]. First introduced in 1936 by Locher [5], the Boron Neutron Capture Therapy model (BNCT) is a promising type of radiation therapy for cancer that has the potential to be an important treatment for numerous types of tumors, including for those lying in areas that are difficult to access for surgery intervention, such as high-grade gliomas and metastatic brain tumors [6,7,8] Currently, only two low molecular weight boron-containing compounds, sodium mercapto-undecahydro-closo-dodecaborate (BSH, developed in about 1965) and a water-soluble fructose complex of borylphenylalanine (BPA, discovered in 1958), have been approved and found clinical use in BNCT [9,10,11].
Recent advances in the development of potential agents for BNCT treatment include various boron-containing bioconjugates [12]; there are many with natural and unnatural amino acids among them [13]. It is believed that amino acid conjugates are preferentially taken up by rapidly growing tumor cells [14]. Polyhedral carboranes as a boron-rich compounds are considered to have potential for usage in BNCT [15]. In this regard, the chemistry of dicarbaboranes [C2B10H12] and their derivatives has been well investigated [16]. The synthesis of o-carboranes containing unnatural ω-mercapto-amino acids with side-chain lengths ranging from 4 to 6 methylene units 1(a–c) [17], L-o-carboranyl-alanine 2 [18] as well as o-carboranyl derivatives of (S)-Asparagine 3a and (S)-Glutamine 3b [19], phenylalanine 4 [20] and (S)-m-carboranyl-homocysteine sulfoxide 5 [21], have been published. Our current interest was inspired by the recent synthesis [22] of m-carboranyl-cysteine 6 and its biological evaluation as an agent for Boron Neutron Capture Therapy [23,24] (Figure 1). Therefore, we decided to synthesize its analogue based on readily available 1-mercapto-o-carborane [25,26].
Here, we report the synthesis of o-carboranyl-cysteine via the sulfa-Michael addition reaction of an easily available salt of 1-mercapto-o-carborane and methyl 2-acetamidoacrylate (MAA) using a two-phase system.

2. Materials and Methods

2.1. General

[1-S-1,2-C2B10H11](Me3NH) (7) was prepared according to the literature procedure [27]. Methyl 2-acetamidoacrylate (Sigma Aldrich Chemie GmbH, NJ, USA) was used without purification. 2H2O was received from Carl Roth GmbH. The reaction proceedings were monitored via thin-layer chromatograms (Merck F254 silica gel on aluminums plates). Boron compounds were visualized with PdCl2 stain solution, which, upon heating, gave dark brown spots. Purifications were carried out using column chromatography with Silica gel 60 0.060–0.200 mm (Acros Organics, NJ, USA). 1H, 13C, and 11B NMR spectra were recorded at 400.13, 100.61, and 128.38 MHz, respectively, on a BRUKER-Avance-400 spectrometer. Tetramethylsilane and BF3 × Et2O were used as standards for 1H, 13C NMR, and 11B NMR, respectively. All the chemical shifts are reported in ppm (δ) relative to external standards. IR spectra were recorded on an IR Prestige-21 (SHIMADZU, Kyoto, Japan) instrument. High-resolution mass spectra (HRMS) were measured on a Bruker micrOTOF II instrument using electrospray ionization (ESI), with a mass range from m/z 50 to m/z 3000; external or internal calibration was carried out with ESI Tuning Mix, Agilent. Spectra can be found in the Supplementary Materials.

2.2. Synthesis of 1-CH3O(O)C(CH3(O)CHN)CHCH2S-1,2-C2B10H11 (8)

To a solution of 7 (0.5 g, 2.1 mmol) in toluene (30 mL) methyl 2-acetamidoacrylate (0.3 g, 2.1 mmol) and H2O (30 mL) were added; the resulting system was vigorously stirred under reflux for 24 h. Then, the mixture was cooled to r.t. and the toluene layer was separated, washed with H2O (2 × 20), dried over Na2SO4, and evaporated. The product was purified by silica gel column chromatography using Et2O as an eluent and vacuum-dried to give a light yellow solid. Yield: 0.46 g (68%). 1H NMR (Chloroform-d) δ = 6.29 (d, J = 7.7 Hz, 1H, NH), 4.89 (td, J = 7.2, 4.5 Hz, 1H, α-CH), 3.99 (broad s, 1H, carb-CH), 3.82 (s, 3H, COOCH3), 3.51 (dd, J = 13.4, 4.6 Hz, 1H, CH2CH), 3.18 (dd, J = 13.4, 6.9 Hz, 1H, CH2CH), 2.06 (s, 3H, NHCOCH3), 3.0–1.5 (broad, 10H, BH); 11B NMR (Chloroform-d) δ = −1.5 (d, J = 150 Hz, 1B), −4.7 (d, J = 146 Hz, 1B), −8.8 (d, J = 154 Hz, 4B), −12.2 (d, J = 162 Hz, 4B); 13C NMR (Chloroform-d) δ = 170.21, 170.17 (CO), 73.9, 67.9 (C-carb), 53.2 (OCH3), 51.2 (α-CH), 39.4 (CH2CH), 23.1 (COCH3). ESI-MS, m/z, C8H21B10NO3S calcd. 320.2322, found 320.2322 ([M+H]+). IR-FT (ν, cm−1): 3371(NH), 3060 (CH-carb); 2948, 2926, 2852 (broad CH); 2607, 2584, 2557 (BH); 1736, 16177 (CO).

2.3. Synthesis of 1-HOOC(NH2)CHCH2S-1,2-C2B10H11 × HCl (9)

Water (2 mL) and concentrated HCl (10 mL) were added to a solution of 8 (0.1 g, 0.3 mmol) in glacial acetic acid (10 mL). The resulting mixture was heated at 70 °C for 40 h., cooled to r.t., and evaporated. The residue was suspended in 5 mL of water, formed solid was filtered, washed with water (2 × 5 mL) and vacuum dried to give 9. Light yellow solid (0.093 mg, 70%). 1H NMR (Methanol-d4) δ = 4.89 (s, 1H, carb-CH), 4.09 (m, 1H, α-CH), 3.64 (m, 1H, CH2CH), 3.43 (m, 1H, CH2CH), 3.0–1.5 (broad, 10H, BH); 11B NMR (Methanol-d4) δ = −1.8 (d, J = 151 Hz, 1B), −5.1 (d, J = 148 Hz, 1B), −9.2 (d, J = 148 Hz, 4B), −12.3 (d, J = 172 Hz, 4B). 13C NMR (Methanol-d4) δ 172.0 (COOH), 74.8, 68.4 (C-carb), 51.60 (α-CH), 38.3 (CH2CH). ESI-MS, m/z, C5H17B10NO2S calcd. 264.2059, found 264.2061. IR-FT (ν, cm−1): 3399 (broad NH+), 3058 (CH- carb); 2923, 2854 (CH); 2580 (broad BH), 1728 (CO).

2.4. Synthesis of 1-CH3O(O)C(CH3(O)CHN)CDCH2S-1,2-C2B10H11 (10)

Under an argon atmosphere, methyl 2-acetamidoacrylate (0.06 g, 0.4 mmol) and 2H2O (5 mL) were added to a solution of 7 (0.1 g, 0.4 mmol) in toluene (15 mL); the resulting two-phase system was vigorously stirred under reflux for 30 h. Purification was held in the same manner as for compound 8. Light yellow solid (90 mg, 66%). 1H NMR (Chloroform-d) δ = 6.31 (s, 1H, NH), 4.00 (broad s, 1H, CH-carb), 3.82 (s, 3H, COOCH3), 3.50 (d, J = 13.4 Hz, 1H, CH2CD ), 3.17 (d, J = 13.4 Hz, 1H, CH2CD), 2.06 (s, 3H, NHCOCH3), 3.0–1.5 (broad, 10H, BH); 11B NMR (Chloroform-d) δ = −1.5 (d, J = 151 Hz, 1B), −4.8 (d, J = 150 Hz, 1B), −8.9 (d, J = 151 Hz, 4B), −12.4 (d, J = 176 Hz, 4B); 13C NMR (Chloroform-d) δ = 170.23, 170.17 (CO), 73.9, 67.9 (C-carb), 53.2 (OCH3), 51.0 (t, α-CD), 39.3 (CH2), 23.1 (COCH3). ESI-MS, m/z, C8H20DB10NO3S calcd. 321.2385, found 321.2399. IR-FT (ν, cm−1): 3370 (NH), 3061 (CH- carb); 2953 (CH); 2608, 2586, 2557 (BH), 1733, 1674 (CO).

2.5. Synthesis of 1-HOOC(NH2)CDCH2S-1,2-C2B10H11 × HCl (11)

Water (2 mL) and concentrated HCl (5 mL) were added to a solution of 10 (0.05 g, 0.16 mmol) in glacial acetic acid (5 mL). The resulting mixture was heated at 70 °C for 40 h., cooled to r.t., and evaporated. The residue was suspended in 5 mL of water, formed solid was filtered, washed with water (2 mL) and vacuum dried to yield 11: 38 mg (81%). 1H NMR (Methanol-d4) δ = 4.88 (broad s, 1H, CH-carb), 4.27–4.20 (α-CH, m, 0,1H), 3.64 (d, J = 13.9 Hz, 1H, CH2CD), 3.44 (d, J = 13.9 Hz, 1H, CH2CD), 3.0–1.5 (broad, 10H, BH). 11B NMR (Methanol-d4) δ = -1.9 (d, J = 149 Hz, 1B), −4.9 (d, J = 140 Hz, 1B), −9.0 (d, J = 143 Hz, 4B), −12.3 (d, J = 175 Hz, 4B). 13C NMR (Methanol-d4) δ = 168.0 (CO), 73.6 (C-carb), 68.4 (C-carb), 51.8 (α-C), 36.0 (CH2). ESI-MS, m/z, C5H17B10NO2S calcd. 265.2122, found 265.2127.

2.6. Synthesis of 1-CH3O(O)C(CH3(O)CHN)CHCH2S-2-D-1,2-C2B10H10 (12)

Under an argon atmosphere, methyl 2-acetamidoacrylate (0.06 g, 0.4 mmol), 2H2O (5 mL), and anhydrous K2CO3 (0.055 g, 0.4 mmol) were added to a solution of 7 (0.1 g, 0.4 mmol) in toluene (15 mL); the resulting system was vigorously stirred under reflux for 10 h. The purification was held in the same manner as for compound 8. Light yellow solid (13 mg, 10%). 1H NMR (Chloroform-d) δ = 6.20 (d, J = 7.7 Hz, 1H, NH), 4.90 (td, J = 7.2, 4.7 Hz, 1H, α-CH), 4.00 (undetectable, s, 0H, carb-CD), 3.83 (s, 3H, COOCH3), 3.53 (dd, J = 13.4, 4.7 Hz, 1H, CH2CH), 3.19 (dd, J = 13.4, 6.9 Hz, 1H, CH2CH), 2.07 (s, 3H, NHCOCH3), 3.0–1.5 (broad, 10H, BH); 11B NMR (Chloroform-d) δ = −1.5 (d, J = 152 Hz, 1B), −4.8 (d, J = 147 Hz, 1B), −8.9 (d, J = 152 Hz, 4B), −12.5 (d, J = 173 Hz, 4B); 13C NMR (Chloroform-d) δ = 170.16,170.14 (CO), 73.8 (CH-carb), 68.5–66.8 (t, CD-carb), 53.2 (OCH3), 51.2 (α-CH), 39.4 (CH2), 23.1 (COCH3). ESI-MS, m/z, C8H20DB10NO3S calcd. 321.2385, found 321.2384. IR-FT (ν, cm−1): 3372(NH), 2923, 2851 (CH); 2607, 2584, 2557 (BH); 2289 (CD-carb); 1736, 1677 (CO).

2.7. X-ray Diffraction

Crystals of 8 are triclinic, space group P-1, at 120K: a = 7.6513(5), b = 10.5669(7), c = 11.5392(7), α= 68.378(1)°, β = 85.686(1)°, γ = 72.663(1)°, V = 827.27(9) Å3, Z = 2 (Z’ = 1), dcalc = 1.282 g·cm−3, F(000) = 332. Intensities of 11,058 reflections were measured with a Bruker SMART APEX 2 Duo CCD diffractometer [λ(MoKα) = 0.71072Å, ω-scans, 2θ < 60°] and 4824 independent reflections [Rint = 0.0324] were used in further refinement. 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 refined in the isotropic approximation within the riding model. For 8, the refinement converged to wR2 = 0.1167 and GOF = 0.781 for all independent reflections (R1 = 0.0371 was calculated for 3661 observed reflections with I > 2σ(I)). All the calculations were performed using SHELX2018 [28]. The CCDC 2034871 contains the supplementary crystallographic data for 8. 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]).
Computational details: All the calculations were conducted in the Gaussian09 program (rev. D01) [28]. The geometry optimization procedures were performed using standard criteria on displacements and forces. The DFT optimization were performed using the PBE0 functional [29,30] and the 6-311++G(d,p) basis set (ultrafine integration grids were used). The non-specific solvation was modelled using the self-consistent reaction field approach (PCM model, ε = 72). The influence of specific solvation on the geometry of 8 was accounted for by the optimization of a central molecule in clusters; two models were used: (1) the trimer of molecules from crystal of 8 with fixed coordinates and normalized X-H bond lengths for lateral molecules and (2) the shell of molecules from crystal of 8 with fixed coordinates and normalized X-H bond lengths. The Quantum theory of Atoms in Molecules surface integrals for the trimer were calculated using the AIMAll program [31]. The shell in the second model was generated using the criteria of at least one geometrical contact (within the sum of van der Waals radii plus 5 Å) between the surrounding molecules and the central molecule. The shell was described by the ONIOM approach (PBE0/6-311++G(d,p):PBE0/3-21G), and only the internal layer (the central molecule) was optimized. The Hessian calculations for all the optimized structures revealed their correspondence to energy minimums.

3. Results

3.1. o-Carboranyl-Cysteine Synthesis

The reaction conditions for the sulfa-Michael addition of thiophenols to methyl 2-acetamidoacrylate are well-studied [32]. It has been shown that reactions proceed in toluene or THF with catalytic quantity of the basic salt K2CO3 in the presence of solid-liquid phase-transfer catalyst. On the one hand, it was reasonable to apply conditions studied for thiophenols to the mercapto derivative of o-carborane; however, prolonged reaction with a base may cause of a side reaction due to the well-known property of o-carborane to undergo deboronation in the presence of bases, even mild ones, to yield nido-[7,9-C2B9H12] [33,34,35,36,37,38,39]. The deboronation of o-C2B10H12 proceeds fairly easily, whereas m-C2B10H12, which was used for the synthesis of 6, is much more stable under the same conditions. A detailed study of such differences has been presented [40]. On the other hand, more common work up of o-carboranyl thiols includes the conversion of the formed SH-derivative to the more stable triethylammonium [41,42] or trimethylammonium salt [27] 7 (Scheme 1). The usage of the trimethylammonium salt 7 let us to avoid the deprotonation stage during the reaction process to keep the o-carborane structure from degradation under the action of a base. The synthetic route developed for the preparation of the cysteine derivative of o-carborane conjugate with cysteine is outlined in Scheme 1.
Compared to mercapto-o-carborane, o-carboranyl thiolate 7 lacks sufficient acidic hydrogen, which is necessary for the reaction to proceed. We found that reaction of 7 with methyl 2-acetamidoacrylate in a biphasic toluene-H2O system under reflux for 24 h without base or any additional catalyst results in the formation of sulfa-Michael addition reaction product 8.
The structure of 8 was confirmed by 1H, 11B, 13C NMR, and IR spectroscopy, high-resolution mass spectrometry, and single-crystal X-ray diffraction study. The 1H NMR spectrum of 8 in CDCl3 shows two singlets at 2.06 and 3.83 due to the acetamide and the ester groups, a doublet at 6.17 ppm from the NH-group, a multiplet at 4.90 ppm attributed to the proton bonded to the α-carbon, two doublet of doublets at 3.51 and 3.18 ppm assigned to the diastereotopic methylene protons, and a broad singlet of carborane C-H group at 3.99 ppm. The 13C-NMR spectrum contains signals of two carbonyl groups at 170.21 and 170.17 ppm., two cluster carbons at 73.9 and 67.9 ppm, two methyls of ester at 53.2 and acetamide at 23.1 ppm., as well as the CH and CH2 groups at 51.2 and 39.4 ppm. The 11B and 11B{1H} NMR spectra are in agreement with the C-mono-substituted o-carborane structure.
The solid-state structure of 8 was determined by single-crystal X-ray diffraction. The molecule of 8 is crystallized as the racemate in the centrosymmetric P-1 space group and contains a stereocenter at the C4 atom (Figure 2).
While the main structural features of 8 are expected for this class of compounds (the C1-C2, S1-C1, and S1-C3 bond lengths equal 1.669(2), 1.791(1), and 1.826(1) Å, respectively; the sum of valence angles at the amide-type N1 atom equals 358.9°), the rotation of the substituent at the C1 atom with respect to the carborane cage can, however, hardly be rationalized by common intramolecular structural effects. For instance, the lone electron pairs of the S1 atom are nearly periplanar to the C3-H and C3-C4 bonds that contradicts the preferred geometry of expected LP(S)→σ* stereoelectronic interactions (Figure 3a). Indeed, the Cambridge Structural Database search reveals the predominance of staggered conformations of the C-Ssp3-Csp3-Csp3 fragments (Figure 3b), whereas the corresponding torsion angle equals 118.1(2)° in 8.
Based on the geometrical analysis of the crystal packing of 8, it has been found that the substituent conformation could be stabilized by the environment effects. Indeed, there are several shortened intermolecular contacts formed by atoms of the C1 substituent which can be attributed to rather strong interactions – NH…O H-bonds (N1…O1 2.933(2) Å, NHO 161.3° with normalized N-H bond length) and O…π interactions (O1…O3 3.020(2) Å), both bounding molecules into infinite chains (Figure 4). The weak BH…HC and CH…O interactions are the only meaningful interactions between these chains.
In order to reveal the influence of environment effects on the conformation of 8, the DFT calculations were additionally performed. Indeed, both the isolated molecule optimization and the optimization of 8 accounting for non-specific solvation effects (the SCRF-PCM model, relative dielectric permittivity of 72) produced structures being significantly different from those observed in the crystal: the root-mean-square deviations of the best overlap for non-hydrogen atoms are 0.495 and 0.446 Å, respectively. Surprisingly, the explicit DFT treatment of three neighboring molecules from a chain (see Figure 4) forming the mentioned shortened intramolecular contacts did not lead to any pronounced changes: the rmsd value for the central molecule in the calculated trimer equals 0.320 Å. Note that the total energy of mentioned N-H…O H-bonds and O…π interactions exceeds 24 kcal·mol−1 according to the QTAIM electron density analysis [43] carried out for the trimer. Our best result was achieved by the ONIOM calculation, which considers all neighboring molecules at the DFT level (the rmsd value equals to 0.168 Å, Figure 5). Thus, despite the relatively large strength of the H-bonds and O…π interactions, the conformation of 8 in crystal depends heavily on the peculiarities of other, weaker intermolecular interactions such as BH…HC and CH…O.
The treatment of 8 with a mixture of glacial acetic and hydrochloric acids under heating gave the title S-substituted cysteine 9 in the form of hydrochloride with a good yield (Scheme 1). The synthesized amino acid was characterized by 1H, 11B, 13C NMR spectroscopy, IR spectroscopy, and high-resolution mass spectrometry. A singlet at 4.89 ppm corresponding to the C-H from the o-carborane was observed in the 1H NMR; other chemical shifts agreeing with amino acid structure were also observed.

3.2. The Sulfa-Michel Addition Mechanism Investigation for o-Carboranyl Cysteine and Synthesis of Deuterium Labeled Compounds

The investigation of the reaction conditions for the synthesis of 8 revealed a crucial role of the choice of solvent. Thus, the formation of only an insignificant amount of 8 was observed when the reaction proceeded in homogenous systems EtOH-H2O or THF-H2O, whereas in absolute EtOH no product was detected at all. The best result was achieved when the toluene-H2O system was used without any base. Since the initial salt 7 did not have an acid proton, as 1-mercapto-o-carborane has, this is why we assumed that α-proton comes into intermediate 8 from water. To gain some insight into the process, we carried out the “free of base” reaction with 2H2O, which required strictly anhydrous conditions (Scheme 2).
We found that the two-phase system reaction proceeds in the way we proposed. Compound 10 with deuterium at the α-position of amino acid was isolated and characterized by NMR, IR spectroscopy, and mass spectrometry. The reaction time for the synthesis of 10 in the toluene-2H2O system increased as compared with those in the toluene-H2O system. This difference in reaction time may be attributed to the presence of a deuterium isotope effect. The yield of reaction, which was detected, remains over 65%. The comparison of the 1H NMR and 13C spectra of 8 and 10 is displayed below (Figure 6).
The 1H NMR spectrum of 10 in CDCl3 does not contain the multiplet at 4.90 ppm, which indicates the absence of a proton at the α-position of the amino acid, whereas in the 13C-NMR spectrum the triplet due to the C2H group appears at 51.0 ppm, indicating the substitution of H for 2H. In addition, the doublet of the NH group (6.17 ppm) and two doublets of doublets from the methylene CH2 protons, which were detected in the 1H NMR spectrum of 8, collapsed to a singlet and two doublets in the spectrum of [α-2H]carboranyl-cysteine 10, respectively, due to the absence of an adjacent proton.
The subsequent acid hydrolysis of 10 resulted in the o-carboranyl-cysteine 11, labeled at the α-position of the amino acid (Scheme 3).
The presence of 2H and its position were determined by high-resolution mass spectrometry and 1H NMR spectroscopy. The 1H NMR spectrum showed the presence of an unlabeled compound in an amount of less than 10% (Figure 7).
The method proposed above can be considered as a preparative one for the synthesis of labeled o-carboranyl-DL-[α-2H]-cysteine, which could be useful for study of the metabolism of o-carboranyl-DL-cysteine itself, as well as the metabolism of boron-containing proteins on its base as potential BNCT agents [44,45,46].
It was mentioned that the o-carborane system is sensitive to the presence of a base. To investigate the effect of water-soluble basic salt toward the reaction that proceeds in a two-phase system, we added K2CO3 to the reaction system. We found that the addition of a basic salt accelerates the reaction, which is complete in 8 h, but resulted in an only 17% yield of 8, while the main product is nido-carborane derivatives, as found by 11B NMR. The solvent change from H2O to 2H2O along with the use of anhydrous K2CO3 and a Schlenk technique under an atmosphere of argon increases the reaction time from 8 h to 10 h due to the isotope effect, and the yield of product falls from 17% to 10%. Interestingly, the structure of the resulting product was found to be different from 8 and 10. According to the NMR spectroscopy data, deuterium from 2H2O took up a place of the 1-CH-o-carborane proton, whereas proton appeared at the α-position of the amino acid (Scheme 4, Figure 8). Compound 12 was isolated and characterized, and its structure and composition were confirmed by 1H, 11B, 13C NMR, and IR spectroscopy and high-resolution mass spectrometry.

4. Discussion

To summarize the obtained results, we assumed the mechanism of reaction proceeding (Scheme 5).
The reaction realizes according to a classical Michael addition reaction, where thiolate 7 acts as a nucleophile without additional deprotonation. In the first step, nucleophile reacts with the electrophilic alkene to form A in a conjugate addition reaction. In the next stage, the deuterium abstraction from solvent by the enolate A forms the final conjugate and a 2HO anion from a water molecule. 2HO anion, being a base, may cause a deboronation process, however its cooperation with a cation forms a water-soluble trimethylammonium hydroxide. When the reaction proceeds in the toluene-2H2O system, trimethylammonium hydroxide eliminates from the toluene medium and under the reaction conditions it decomposes in an aqueous medium with the formation of trimethylamine and water. Thus, the final stage process may be described as proton abstraction from cation that results in the deactivation of a base. The proposed mechanism is also supported by the fact that an increase in the reaction time does not have a significant effect on the yield of 8. The reaction between 7 and methyl 2-acetamidoacrylate in EtOH does not give the desired product. According to the proposed mechanism, in this case the proton abstraction from solvent by the enolate A forms an EtO anion, which is not eliminated from the reaction system. The EtO anion being formed may undergo the Michael addition reaction, as well as 7 and/or deboronate o-carborane structure. The explanation above can be applied to a reaction in a system of two miscible liquids, such as THF-H2O. The formed hydroxide anion remains in one system with the reactants to yield mainly a nido-carborane product.

5. Conclusions

Two possibilities of the sulfa-Michael addition reaction in the synthesis of o-carboranyl-DL-cysteine were investigated. The reaction proceeding between the trimethylammonium salt of 1-mercapto-o-carborane and methyl 2-acetamidoacrylate in the two-phase system, toluene-H2O with base, as expected, demonstrates a low yield. Using the “free of base” method under the mild conditions in the two-phase system, the side processes were minimized and after the deprotection 1-HOOCCH(NH2)CH2S-1,2-C2B10H11 was obtained in a good total yield. The developed “free of base” method was applied for the preparation of o-carboranyl-DL-cysteine labeled with deuterium at the α-position of amino acid using cheap and easily available 2H2O as a deuterium source. The obtained compounds can be considered as potential agents for the Boron Neutron Capture Therapy of cancer as well as for protein metabolism studies.

Supplementary Materials

The following are available online at https://www.mdpi.com/2073-4352/10/12/1133/s1.

Author Contributions

Synthesis and writing—original draft preparation, J.L.; synthesis, M.S.; NMR research analysis, I.K.; X-ray diffraction study, I.A.; supervision, V.B. and I.S. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the Russian Science Foundation (№ 19-72-30005).

Acknowledgments

The NMR spectroscopy studies and X-ray diffraction experiment were performed using equipment of the Center for Molecular Structure Studies at A.N. Nesmeyanov Institute of Organoelement Compounds operating with financial support of the Ministry of Science and Higher Education of the Russian Federation.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Siegel, R.L.; Miller, K.D.; Jemal, A. Cancer statistics. CA Cancer J. Clin. 2019, 69, 7–34. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  2. Black, P.M. Brain Tumors. N. Engl. J. Med. 1991, 324, 1471–1476. [Google Scholar] [CrossRef] [PubMed]
  3. Shah, H.K.; Mehta, M.P. Radiation Therapy. In High-Grade Gliomas. Current Clinical Oncology; Barnett, G.H., Ed.; Humana Press: Totowa, NJ, USA, 2007; pp. 231–244. [Google Scholar] [CrossRef]
  4. Afseth, J.; Neubeck, L.; Karatzias, T.; Grant, R. Holistic needs assessment in brain cancer patients: A systematic review of available tools. Eur. J. Cancer Care 2018, 28, e12931. [Google Scholar] [CrossRef] [Green Version]
  5. Locher, G.L. Biological effects and therapeutic possibilities of neutrons. Am. J. Roentgenol. Radium Ther. 1936, 36, 1–13. [Google Scholar]
  6. Tolpin, E.I.; Wellum, G.R.; Dohan, F.C., Jr.; Komblith, P.R.; Zamenhof, R.G. Boron Neutron Capture Therapy of Cerebral Gliomas. Oncology 1975, 32, 223–246. [Google Scholar] [CrossRef] [PubMed]
  7. Barth, R.F.; Zhang, Z.; Liu, T. A realistic appraisal of boron neutron capture therapy as a cancer treatment modality. Cancer Commun. 2018, 38, 36. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  8. Xuan, S.; Vicente, M.G.H. Recent Advances in Boron Delivery Agents for Boron Neutron Capture Therapy (BNCT). In Boron-Based Compounds: Potential and Emerging Applications in Biomedicine; Hey-Hawkins, E., Viñas, C., Eds.; John Wiley & Sons: Hoboken, NJ, USA, 2018; pp. 298–342. [Google Scholar]
  9. Henriksson, R.; Capala, J.; Michanek, A.; Lindahl, S.-Å.; Salford, L.G.; Franzén, L.; Blomquist, E.; Westlin, J.-E.; Bergenheim, T.A. Boron Neutron Capture Therapy (BNCT) for Glioblastoma Multiforme: A Phase II Study Evaluating A Prolonged High-Dose of Boronophenylalanine (BPA). Radiother. Oncol. 2008, 88, 183–191. [Google Scholar] [CrossRef]
  10. Kulvik, M.; Vähätalo, J.; Buchar, E.; Färkkilä, M.; Järviluoma, E.; Jääskeläinen, J.; Křiž, O.; Laakso, J.; Rasilainen, M.; Ruokonen, I.; et al. Clinical implementation of 4-dihydroxyborylphenylalanine synthesised by an asymmetric pathway. Eur. J. Pharm. Sci. 2003, 18, 155–163. [Google Scholar] [CrossRef]
  11. Barth, R.F.; Coderre, J.F.; Vicente, M.G.H.; Blue, T.E. Boron Neutron Capture Therapy of Cancer: Current Status and Future Prospects. Clin. Cancer Res. 2005, 11, 3987–4002. [Google Scholar] [CrossRef] [Green Version]
  12. 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]
  13. Kabalka, G.W.; Yao, M.-L. The synthesis and use of boronated amino acids for boron neutron capture therapy. Anticancer Agents Med. Chem. 2006, 6, 111–125. [Google Scholar] [CrossRef] [PubMed]
  14. Hawthorne, M.F. The Role of Chemistry in the Development of Boron Neutron Capture Therapy of Cancer. Angew. Chem. Int. Ed. Engl. 1993, 32, 950–984. [Google Scholar] [CrossRef]
  15. Sivaev, I.B.; Bregadze, V.I. Polyhedral Boranes for Medical Applications: Current Status and Perspectives. Eur. J. Inorg. Chem. 2009, 1433–1450. [Google Scholar] [CrossRef]
  16. Grimes, R.N. Carboranes, 3rd ed.; Academic Press: New York, NY, USA; Elsevier: Amsterdam, The Netherlands, 2016. [Google Scholar]
  17. Stogniy, M.Y.; Zakharova, M.V.; Sivaev, I.B.; Godovikov, I.A.; Chizov, A.O.; Bregadze, V.I. Synthesis of new carborane-based amino acids. Polyhedron 2013, 55, 117–120. [Google Scholar] [CrossRef]
  18. Leukart, O.; Caviezel, M.; Eberle, A.; Escher, E.; Tun-Kyi, A.; Schwyzer, R. L-o-Carboranylalanine, a Boron Analogue of Phenylalanine. Helv. Chim. Acta 1976, 59, 2184–2187. [Google Scholar] [CrossRef]
  19. Gruzdev, D.A.; Levit, G.L.; Olshevskaya, V.A.; Krasnov, V.P. Synthesis of ortho-Carboranyl Derivatives of (S)-Asparagine and (S)-Glutamine. Russ. J. Org. Chem. 2017, 5, 769–776. [Google Scholar] [CrossRef]
  20. Prashar, J.K.; Moore, D.E. Synthesis of Carboranyl Phenylalanine for Potential Use in Neutron Capture Therapy of Melanoma. J. Chem. Soc. Perkin Trans. 1 1993, 1051–1053. [Google Scholar] [CrossRef]
  21. Gruzdev, D.A.; Ustinov, V.O.; Levit, G.L.; Ol’shevskaya, V.A.; Krasnov, V.P. Synthesis of meta-Carboranyl-(S)-homocysteine Sulfoxide. Russ. J. Org. Chem. 2018, 54, 1579–1582. [Google Scholar] [CrossRef]
  22. He, T.; Misuraca, J.C.; Musah, R.A. “Carboranyl-Cysteine”-Synthesis, Structure and Self-Assembly Behavior of a Novel α-Amino Acid. Sci. Rep. 2017, 7, 16995. [Google Scholar] [CrossRef] [Green Version]
  23. He, T.; Musah, R.A. Evaluation of the Potential of 2-Amino-3-(1,7-dicarba-closododecaboranyl-1-thio)propanoic Acid as a Boron Neutron Capture Therapy Agent. ACS Omega 2019, 4, 3820–3826. [Google Scholar] [CrossRef] [Green Version]
  24. He, T.; Chittur, S.V.; Musah, R.A. Impact on Glioblastoma U87 Cell Gene Expression of a Carborane Cluster bearing Amino Acid: Implications for Carborane Toxicity in Mammalian Cells. ACS Chem. Neurosci. 2019, 10, 1524–1534. [Google Scholar] [CrossRef] [PubMed]
  25. Vinas, C.; Benakki, R.; Teixidor, F.; Casabo, J. Dimethoxyethane as a Solvent for the Synthesis of C-Monosubstituted o-Carborane Derivatives. Inorg. Chem. 1995, 34, 3844–3845. [Google Scholar] [CrossRef]
  26. Plešek, J.; Heřmanek, S. Syntheses and properties of substituted icosahedral carborane thiols. Collect. Czech. Chem. Commun. 1981, 46, 687–692. [Google Scholar] [CrossRef]
  27. 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. [Google Scholar] [CrossRef]
  28. 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]
  29. Perdew, J.; Ernzerhof, M.; Burke, K. Rationale for mixing exact exchange with density functional approximations. J. Chem. Phys. 1996, 105, 9982–9985. [Google Scholar] [CrossRef]
  30. Carlo, A.; Barone, V. Toward reliable density functional methods without adjustable parameters: The PBE0 model. J. Chem. Phys. 1999, 110, 6158–6170. [Google Scholar] [CrossRef]
  31. Keith, T.A. TK Gristmill Software; AIMAll (Version 19.10.12): Overland Park, KS, USA, 2019. [Google Scholar]
  32. Cossec, B.; Cosnier, F.; Burgart, M. Methyl Mercapturate Synthesis: An Efficient, Convenient and Simple Method. Molecules 2008, 13, 2394–2407. [Google Scholar] [CrossRef] [Green Version]
  33. Davidson, M.G.; Fox, M.A.; Hibbert, T.G.; Howard, J.A.K.; Mackinnon, A.; Neretin, I.S.; Wade, K. Deboronation of orthocarborane by an iminophosphorane: Crystal structures of the novel carborane adduct nido-C2B10H12·HNP(NMe2)3 and the borenium salt [(Me2N)3PNHBNP(NMe2)3]2O2+ (C2B9H12)2. Chem. Commun. 1999, 1649–1650. [Google Scholar] [CrossRef]
  34. Svantesson, E.; Pettersson, J.; Olin, Å.; Markides, K.E. A kinetic study of the self-degradation of o-carboranylalanine to nido-carboranylalanine in solution. Acta Chem. Scand. 1999, 53, 731–736. [Google Scholar] [CrossRef] [Green Version]
  35. Taoda, Y.; Sawabe, T.; Endo, Y.; Yamaguchi, K.; Fujii, S.; Kagechika, H. Identification of an intermediate in the deboronation of ortho-carborane: An adduct of ortho-carborane with two nucleophiles on one boron atom. Chem. Commun. 2008, 2049–2051. [Google Scholar] [CrossRef]
  36. Kononov, L.O.; Orlova, A.V.; Zinin, A.I.; Kimel, B.G.; Sivaev, I.B.; Bregadze, V.I. Conjugates of polyhedral boron compounds with carbohydrates. 2. Unexpected easy closo- to nido-transformation of a carborane–carbohydrate conjugate in neutral aqueous solution. J. Organomet. Chem. 2005, 690, 2769–2774. [Google Scholar] [CrossRef]
  37. Likhosherstov, L.M.; Novikova, O.S.; Chizhov, A.O.; Sivaev, I.B.; Bregadze, V.I. Conjugates of polyhedral boron compounds with carbohydrates 8. Synthesis and properties of nido-ortho-carborane glyco conjugates containing one to three β-lactosylamine residues. Russ. Chem. Bull. 2011, 60, 2359–2364. [Google Scholar] [CrossRef]
  38. Núñez, R.; Teixidor, F.; Kivekäs, R.; Sillanpää, R.; Viñas, C. Influence of the solvent and R groups on the structure of (carboranyl)R2PI2 compounds in solution. Crystal structure of the first iodophosphonium salt incorporating the anion [7,8-nido-C2B9H10]. Dalton Trans. 2008, 1471–1480. [Google Scholar] [CrossRef]
  39. Teixidor, F.; Viñas, C.; Mar Abad, M.; Nuñez, R.; Sillanpäa, R. Procedure for the degradation of 1,2-(PR2)2-1,2-dicarba-closo-dodecaborane(12) and 1-(PR2)-2-R′-1,2-dicarba-closo-dodecaborane(12). J. Organomet. Chem. 1995, 503, 193–203. [Google Scholar] [CrossRef]
  40. Poater, J.; Viñas, C.; Bennour, I.; Escayola, S.; Solà, M.; Teixidor, F. Too Persistent to Give up: Aromaticity in Boron Clusters Survives Radical Structural Changes. J. Am. Chem. Soc. 2020, 142. [Google Scholar] [CrossRef] [PubMed]
  41. Boyd, L.A.; Clegg, W.; Copley, R.C.B.; Davidson, M.G.; Fox, M.A.; Hibbert, T.G.; Howard, J.A.; Mackinnon, K.A.; Peace, R.J.; Wade, K. Exo-π-bonding to an ortho-carborane hypercarbon atom: Systematic icosahedral cage distortions reflected in the structures of the fluoro-, hydroxy- and amino-carboranes, 1-X-2-Ph-1,2-C2B10H10 (X = F, OH or NH2) and related anions. Dalton Trans. 2004, 2786–2799. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  42. Stogniy, M.Y.; Sivaev, I.B.; Petrovskii, P.V.; Bregadze, V.I. Synthesis of monosubstituted functional derivatives of carboranes from1-mercapto-ortho-carborane: 1-HOOC(CH2)nS-1,2-C2B10H11 and [7-HOOC(CH2)nS-7,8-C2B9H11] (n = 1–4). Dalton Trans. 2010, 39, 1817–1822. [Google Scholar] [CrossRef]
  43. 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]
  44. Mutlib, A.E. Application of Stable Isotope-Labeled Compounds in Metabolism and in Metabolism-Mediated Toxicity Studies. Chem. Res. Toxicol. 2008, 21, 1672–1689. [Google Scholar] [CrossRef]
  45. Wilkinson, D.J. Historical and contemporary stable isotope tracer approaches to studying mammalian protein metabolism. Mass Spectrom. Rev. 2018, 37, 57–80. [Google Scholar] [CrossRef]
  46. Pająk, M.; Pałka, K.; Winnicka, E.; Kańska, M. The chemo- enzymatic synthesis of labeled l-amino acids and some of their derivatives. J. Radioanal. Nucl. Chem. 2018, 317, 643–666. [Google Scholar] [CrossRef] [PubMed]
Figure 1. o- and m- carborane–amino acid conjugates.
Figure 1. o- and m- carborane–amino acid conjugates.
Crystals 10 01133 g001
Scheme 1. Synthesis of o-carboranyl cysteine.
Scheme 1. Synthesis of o-carboranyl cysteine.
Crystals 10 01133 sch001
Figure 2. General view of the molecule 8 in crystal. Non-hydrogen atoms are given as probability ellipsoids of atomic displacements (p = 0.5).
Figure 2. General view of the molecule 8 in crystal. Non-hydrogen atoms are given as probability ellipsoids of atomic displacements (p = 0.5).
Crystals 10 01133 g002
Figure 3. The conformation of the C1-S1-C3-C4 fragment in 8 (LP–positions of lone electronic pairs, (a)) and the distribution of corresponding torsion angles in the C-Ssp3-Csp3-Csp3 fragments retrieved from the Cambridge Structural Database (b).
Figure 3. The conformation of the C1-S1-C3-C4 fragment in 8 (LP–positions of lone electronic pairs, (a)) and the distribution of corresponding torsion angles in the C-Ssp3-Csp3-Csp3 fragments retrieved from the Cambridge Structural Database (b).
Crystals 10 01133 g003
Figure 4. A fragment of the infinite chain formed in crystal of 8 along the a axis. Non-hydrogen atoms are given as probability ellipsoids of atomic displacements (p = 0.5).
Figure 4. A fragment of the infinite chain formed in crystal of 8 along the a axis. Non-hydrogen atoms are given as probability ellipsoids of atomic displacements (p = 0.5).
Crystals 10 01133 g004
Figure 5. The best root-mean-square overlap for non-hydrogen atoms between crystal (solid lines) and ONIOM-modelled (dashed lines) conformations of 8.
Figure 5. The best root-mean-square overlap for non-hydrogen atoms between crystal (solid lines) and ONIOM-modelled (dashed lines) conformations of 8.
Crystals 10 01133 g005
Scheme 2. Synthesis of the labeled with deuterium o-carboranyl conjugate.
Scheme 2. Synthesis of the labeled with deuterium o-carboranyl conjugate.
Crystals 10 01133 sch002
Figure 6. Comparison of 1H NMR and fragments of the 13C NMR spectra of 8 and 10 in CDCl3.
Figure 6. Comparison of 1H NMR and fragments of the 13C NMR spectra of 8 and 10 in CDCl3.
Crystals 10 01133 g006
Scheme 3. O-carboranyl-DL-[α-2H]-cysteine synthesis.
Scheme 3. O-carboranyl-DL-[α-2H]-cysteine synthesis.
Crystals 10 01133 sch003
Figure 7. 1H NMR spectrum of 11 in methanol-d4.
Figure 7. 1H NMR spectrum of 11 in methanol-d4.
Crystals 10 01133 g007
Scheme 4. The sulfa-Michael addition reaction in the presence of the basic salt.
Scheme 4. The sulfa-Michael addition reaction in the presence of the basic salt.
Crystals 10 01133 sch004
Figure 8. 1H and a fragment of 13C NMR spectra of 12 in CDCl3.
Figure 8. 1H and a fragment of 13C NMR spectra of 12 in CDCl3.
Crystals 10 01133 g008
Scheme 5. Proposed mechanism of the reaction of trimethylammonium salt of 1-mercapto-o-carborane with methyl 2-acetamidoacrylate.
Scheme 5. Proposed mechanism of the reaction of trimethylammonium salt of 1-mercapto-o-carborane with methyl 2-acetamidoacrylate.
Crystals 10 01133 sch005
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.; Kosenko, I.; Ananyev, I.; Stogniy, M.; Sivaev, I.; Bregadze, V. “Free of Base” Sulfa-Michael Addition for Novel o-Carboranyl-DL-Cysteine Synthesis. Crystals 2020, 10, 1133. https://doi.org/10.3390/cryst10121133

AMA Style

Laskova J, Kosenko I, Ananyev I, Stogniy M, Sivaev I, Bregadze V. “Free of Base” Sulfa-Michael Addition for Novel o-Carboranyl-DL-Cysteine Synthesis. Crystals. 2020; 10(12):1133. https://doi.org/10.3390/cryst10121133

Chicago/Turabian Style

Laskova, Julia, Irina Kosenko, Ivan Ananyev, Marina Stogniy, Igor Sivaev, and Vladimir Bregadze. 2020. "“Free of Base” Sulfa-Michael Addition for Novel o-Carboranyl-DL-Cysteine Synthesis" Crystals 10, no. 12: 1133. https://doi.org/10.3390/cryst10121133

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

Article Metrics

Back to TopTop