Next Article in Journal
Temperature-Sensitive Modified Bentonite Based on NIPAM for Drilling Fluid: Experimental and Molecular Simulation Studies
Next Article in Special Issue
Metallacarborane Synthons for Molecular Construction—Oligofunctionalization of Cobalt Bis(1,2-dicarbollide) on Boron and Carbon Atoms with Extendable Ligands
Previous Article in Journal
Multifunctional Applications of Ionic Liquids in Polymer Materials: A Brief Review
Previous Article in Special Issue
DFT Surface Infers Ten-Vertex Cationic Carboranes from the Corresponding Neutral closo Ten-Vertex Family: The Computed Background Confirming Their Experimental Availability
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Synthesis and Structural Characterization of p-Carboranylamidine Derivatives †

1
Chemisches Institut, Otto-von-Guericke-Universität Magdeburg, Universitätsplatz 2, 39106 Magdeburg, Germany
2
Institut für Anorganische und Analytische Chemie, Friedrich-Schiller-Universität Jena, Humboldtstr. 8, 07743 Jena, Germany
3
Institut für Physik, Otto-von-Guericke-Universität Magdeburg, Universitätsplatz 2, 39106 Magdeburg, Germany
*
Author to whom correspondence should be addressed.
Dedicated to Professor John D. Kennedy on the occasion of his 80th birthday.
Molecules 2023, 28(9), 3837; https://doi.org/10.3390/molecules28093837
Submission received: 6 April 2023 / Revised: 27 April 2023 / Accepted: 27 April 2023 / Published: 30 April 2023

Abstract

:
In this contribution, the first amidinate and amidine derivatives of p-carborane are described. Double lithiation of p-carborane (1) with n-butyllithium followed by treatment with 1,3-diorganocarbodiimides, R–N=C=N–R (R = iPr, Cy (= cyclohexyl)), in DME or THF afforded the new p-carboranylamidinate salts p-C2H10B10[C(NiPr)2Li(DME)]2 (2) and p-C2H10B10[C(NCy)2Li(THF)2]2 (3). Subsequent treatment of 2 and 3 with 2 equiv. of chlorotrimethylsilane (Me3SiCl) provided the silylated neutral bis(amidine) derivatives p-C2H10B10[C{iPrN(SiMe3)}(=NiPr)]2 (4) and p-C2H10B10[C{CyN(SiMe3)}(=NCy)]2 (5). The new compounds 3 and 4 have been structurally characterized by single-crystal X-ray diffraction. The lithium carboranylamidinate 3 comprises a rare trigonal planar coordination geometry around the lithium ions.

1. Introduction

Icosahedral closo-carborane cage compounds of the general composition C2B10H12 can be viewed as 3D molecular analogs of benzene [1]. They are of high scientific and technological interest due to a variety of practical applications, e.g., in materials science [2,3,4,5,6,7,8,9,10,11,12,13], homogeneous catalysis [14,15,16,17,18,19,20,21], and medicinal chemistry [22,23,24,25,26,27,28,29]. Moreover, carborane derivatives are widely employed in coordination chemistry and as building blocks in supramolecular, bio-inorganic, and organometallic chemistry [30,31,32,33,34]. Depending on the position of the two carbon atoms in the carborane cage, three isomers can be distinguished. As shown in Figure 1, these are ortho-carborane (o-carborane, 1,2-C2B10H12), meta-carborane (m-carborane, 1,7-C2B10H12), and para-carborane (p-carborane, 1,12-C2B10H12) [35]. Most readily available among these is o-carborane, while the other two are made by thermal rearrangement of the ortho-isomer. This is why p-carborane is the most expensive precursor and its derivative chemistry is the least studied [35].
In 2010, we reported a novel type of o-carborane-based N-chelating ligands which were named carboranylamidinates. These were obtained in the form of their lithium derivatives by in situ lithiation of o-carborane using n-butyllithium followed by treatment with one equivalent of 1,3-diorganocarbodiimides, R–N=C=N–R (R = iPr, Cy (= cyclohexyl)). As illustrated in Scheme 1, the resulting carboranylamidinate anions were quite unique as they combined the highly versatile amidinate ligand system, [RC(NR′)2] [36,37,38,39,40], with a σ-bond to the carborane cage. Subsequently, the lithium salts served as precursors for a variety of main-group and transition metal complexes comprising carboranylamidinate ligands [41,42,43,44,45]. In all these complexes, the carboranylamidinate ligands adopt the characteristic κC,κN-chelating coordination mode instead of the regular κN,κN′-chelating mode of metal-coordinated amidinate anions.
Thus far, the formation of κC,κN-chelating carboranylamidinate anions has been limited to compounds derived from o-carborane. In 2014, we reported that similar reactions starting from m-carborane take a completely different course. As illustrated in Scheme 2, successive treatment of m-carborane with n-butyllithium and 1,3-di-iso-propylcarbodiimide did not lead to the formation of a related carboranylamidinate anion. Instead, an unprecedented deboronation reaction of the m-carborane took place, in which a BH group was detached from the carborane cage and incorporated into a nido-carborane-anellated diazadiborepine ring system. 1,3-dicyclohexylcarbodiimide reacted in a very similar manner but afforded a slightly modified seven-membered diazadiborepine ring system [46].
Until now, the question remained of how the third isomer, p-carborane (1), would behave in the same reaction sequence of lithiation and carbodiimide addition. Here, we present the answer to this question.

2. Results and Discussion

2.1. Synthesis and Characterization

In the first set of experiments, THF solutions of p-carborane were metalated with 1 equiv. of n-butyllithium and then treated in situ with two different carbodiimides R–N=C=N–R (R = iPr, Cy). Under these conditions, only small amounts (ca. 20% yield) of crystalline products could be isolated, which were difficult to separate from unreacted p-carborane (NMR control). This finding implied that the envisaged mono-amidinate derivatives shown in the upper equation in Scheme 3 were not formed as pure reaction products and that disubstitution was instead the preferred reaction pathway. This assumption was soon confirmed by adjusting the stoichiometric ratio of the reactants to 1:2:2 according to the second equation in Scheme 3. Under these conditions, the new compounds 2 (R = iPr) and 3 (R = Cy) could be isolated as pure crystalline solids in significantly improved yields of 51% (2) and 46% (3), respectively. Both lithium amidinate salts are readily soluble in THF, DME, and diethyl ether. Crystallization from DME (2) and THF (3) afforded the nicely crystalline solvates depicted in Scheme 4.
Both bis(anionic) title compounds 2 and 3 were fully characterized through the usual set of elemental analyses and spectroscopic methods. In the IR spectra, strong bands at 1523 cm−1 (2) and 1543 cm−1 (3) are typical for the stretching vibrations of the delocalized amidinate NCN units [36,37,38,39,40]. Medium strong bands in the range of 2590–2620 cm−1 could be assigned to the B–H stretching vibrations, while the νas(C-O-C) bands of the coordinated solvents appear around 1050 cm−1 as medium or strong bands. The 1H and 13C NMR spectra show the typical signals of the iso-propyl and cyclohexyl substituents, which do not need to be discussed here in detail (cf., Experimental Section and Supplementary Materials). In the 1H NMR spectra, the B–H hydrogens give rise to broad multiplets extending over a range of ca. 1.5 ppm. The 13C NMR chemical shifts of the carbon atoms of the NCN groups are 155.2 ppm (2) and 154.0 ppm (3), respectively. A 13C resonance of the quaternary carbon atoms within the carborane cage could be detected only in the spectrum of 2 ( δ 93.3 ppm). All cage boron atoms give rise to a single resonance around −14 ppm in the 11B NMR spectra of both amidinate salts. Apparently, the centrosymmetric structure leads to very similar chemical shifts of the boron atoms so that the signals could not be further resolved. Finally, 7Li NMR spectra displayed only one signal around 0.1 ppm. As expected for salt-like compounds, the mass spectra of 2 and 3 did not show the respective molecular ions but only fragment peaks of the unsolvated carboranylamidinate anions (cf., Experimental Section).
Remarkably, the formation of the lithium carboranylamidiate represents the first incidence of a “normal” reactivity of a lithiated carborane with carbodiimides. This means that 1,12-dilithiocarborane behaves toward carbodiimides like any other organolithium reagents and adds to the central carbon atom of the N=C=N moiety under the formation of regular amidinate anion of the type [RC(NR′)2] [36,37,38,39,40]. This finding reveals that all three C2B10H12 isomers behave differently in their reactivity toward 1,3-diorganocarbodiimides.
As an initial reactivity study involving the lithium carboranylamidinate salts 2 and 3, we investigated silylation reactions with chlorotrimethylsilane, Me3SiCl, which should lead to the formation of neutral bis-silylated amidine derivatives, as illustrated in Scheme 4.
Both reactions were carried out in THF solutions at r.t. Work-up via extraction with toluene afforded the bis-silylated products 4 and 5 as colorless crystals in moderate yields (4: 54%, 5: 43%). Both compounds dissolve freely in diethyl ether and toluene, and are moderately moisture-sensitive due to the presence of Si–N bonds. Besides an X-ray structural analysis of 4 (see next paragraph), all analytical and spectroscopic data were in excellent agreement with the formation of bis(silylated) p-carboranyl-bis(amidines). Highly characteristic in the IR spectra are the ν C=N bands at 1622 cm−1 (4) and 1627 cm−1 (5), respectively. These bands clearly indicate the transition from the delocalized amidinate NCN units in the salt-like amidinate precursors 2 and 3 (ν NCN 1523 cm−1 (2) and 1543 cm−1 (3)) to N–C=N moieties with localized carbon–nitrogen double and single bonds. Bands at 2590 cm−1 (4) and 2606 cm−1 (5) can be assigned to the B–H stretching vibrations, while typical Si–C stretch bands of the SiMe3 groups appear at νas 726 cm−1 and 750 cm−1 as well as νs 651 cm−1 and 658 cm−1, respectively. The 1H, 13C, and 29Si NMR spectra of 4 and 5 all showed only one singlet resonance for the SiMe3 groups. This is in agreement with the centrosymmetric molecular structure found in the X-ray structural analysis of 4 (see following paragraph). As was observed for the anionic precursors 2 and 3, the 11B NMR spectra of the bis-silylated derivatives also displayed only single resonances around −13.4 ppm.

2.2. Crystal and Molecular Structures

The title compounds 3 and 4 could be structurally characterized through X-ray diffraction studies. The molecular structures are depicted in Figure 2 and Figure 3. Colorless, prism-shaped single-crystals of 3 were grown from concentrated solutions in THF at r.t., while compound 4 was obtained in the form of well-formed, colorless, block-like single-crystals upon slow crystallization from toluene at 4 °C. As illustrated in Figure 2, structure determination of compound 3 confirmed the presence of an anionic p-carboranylamidinate species formed by the addition of dilithiated p-carborane to the central C atom of the carbodiimide reagent. The overall molecular structure is centrosymmetric. With 1.953(3) Å, the Li-N2 bond length is typical for a coordinative bond. As in other typical lithium amidinates such as Li[MeC6H4C(NSiMe2)2](THF)2 [36,47], the C–N distances in the amidinate NCN unit are quite similar (N(1)-C(2) 1.305(2), N(2)-C(2) 1.340(2)), indicating complete delocalization of the negative charge. However, the coordination of the lithium ion to the anionic amidinate moieties differs from the vast majority of other lithium amidinates in that it is not κN,κN′-chelating. Instead, the lithium ions are coordinated to only one nitrogen atom of the NCN moiety, resulting in a nearly trigonal planar coordination geometry around Li. There are only very few examples of similar monodentate amidinate coordination to lithium [48,49], and all of them result from steric crowding around the NCN unit, e.g., through very bulky terphenyl or triptycenyl substituents. Thus it can be assumed that steric congestion is also the reason for the rather unusual trigonal planar coordination of the lithium ions in compound 3.
The neutral p-carboranyl-bis(amidine) derivative 4 crystallizes in the monoclinic space P21/n, and, like 3, the molecule also shows crystallographically imposed centrosymmetry. The transition from the delocalized anionic NCN moieties in the amidinate salts 2 and 3 to a neutral amidine is clearly evidenced by the change in the C–N bond lengths. With distances of C(2)-N(1) 1.265(2) Å and C(2)-N(2) 1.423(2) Å, compound 4 clearly contains N–C=N units with isolated single and double bonds. In this respect, the molecular structure of 4 is closely related to the oxygen analogue p-carborane-1,12-dicarboxylic acid [50]. The N1-C2-N2 angle is 128.9(1)°, and the Si-N2 distance is 1.751(1) Å. Both values are in good agreement with those of the silylated o-carboranylamidine derivative o-C2B10H10[κC,N-C(iPrNSiMe3)(=NiPr)]SiMe3 [51].

3. Experimental Section

3.1. General Procedures and Instrumentation

All reactions were carried out in oven-dried or flame-dried glassware under an inert atmosphere of dry argon employing standard Schlenk and glovebox (MBraun MBLab) techniques. The solvents n-pentane, toluene, DME, and THF were distilled from sodium/benzophenone under nitrogen atmosphere prior to use. p-carborane was obtained from Katchem spol. s.r.o., 278 01 Kralupy nad Vltavou, Czech Republic (https://katchem.cz/en). Other starting materials were purchased from Sigma-Aldrich and used without further purification. All NMR spectra (1H, 13C, 29Si, 11B, and 7Li) were recorded in THF-d8 solutions on a Bruker DPX 400 spectrometer. IR spectra were measured with a Bruker Vertex 70V spectrometer equipped with a diamond ATR unit between 4000 cm−1 and 50 cm−1. Mass spectra were measured on a MAT 95 apparatus (EI, 70 eV). Microanalyses (C, H, N) were performed using a VARIO EL cube apparatus. Melting/decomposition points were determined using a Büchi Melting Point B-540.

3.2. Synthesis of Compound p-C2H10B10[C(NiPr)2Li(DME)]2 (2)

A total of 0.50 g (3.5 mmol) p-carborane, dissolved in THF (50 mL), was treated at r.t. with 2 equiv. of n-butyllithium (7.0 mmol, 4.40 mL of a 1.6 M solution in n-hexane). After stirring for 1 h, 0.88 g (7.0 mmol) of 1,3-di-iso-propylcarbodiimide was added via syringe, and stirring at r.t. was continued for 12 h. The resulting clear yellow solution was evaporated to dryness and the oily crude product was redissolved in a minimum volume of DME (ca. 10 mL). Product 2 was precipitated by the addition of n-pentane (ca. 50 mL) and isolated after drying under vacuum as a microcrystalline, pale yellow solid in 51% isolated yield (1.06 g). M.p. 215 °C (dec.). Elemental analysis calcd. for C24H58B10N4O4Li2 (M = 588.71 g mol−1): C, 48.96%; H, 9.9%; N, 9.51%; found C, 48.91%; H, 9.77%; N, 9.64%. 1H NMR (400 MHz, THF-d8, 21 °C): δ = 3.53 (sept, 4 H, 3J = 6.40 Hz, CH-iPr), 3.41, 3.25 (DME), 1.55–3.12 (m br, 10 H, BH), 0.78 (d, 24 H, 3J = 6.00 Hz, CH3-iPr) ppm. 13C{1H} NMR (100.6 MHz, THF-d8, 23 °C): δ = 155.2 (NCN), 93.3 (C-NCN), 72.6, 58.9 (DME), 46.3 (CH-iPr), 25.6 (CH3-iPr) ppm. 11B{1H} NMR (128.4 MHz, THF-d8, 23 °C): δ = –14.1 ppm. 7Li{1H} NMR (155.5 MHz, THF-d8, 23 °C): δ = 0.11 ppm. IR (ATR): νmax 2955 m (νs CH3), 2926 m (νas CH2), 2860 w (νas,s CH3/CH2), 2597 m (ν BH), 1523 s (ν NCN), 1461 m (δas,s CH3/CH2), 1367 m (δs CH3), 1357 m (δs CH3), 1312 m, 1285 m, 1264 s, 1191 w, 1157 w, 1112 m, 1077 vs (νas C-O-C), 1021 m, 961 w, 901 w, 870 m, 801 w, 737 m, 679 w, 621 m, 599 m, 566 m, 506 m, 494 m, 460 w, 385 w, 374 w, 242 vs cm−1. MS (EI, 70 eV): m/z (%) 396 (17) [((iPrN)2C)2-C2H10B10]+, 353 (27) [((iPrN)2C)2-C2H10B10-iPr]+, 270 (15) [(iPrN)2C-C2H10B10]+, 227 (49) [(iPrN)NC-C2H10B10]+, 170 (25) [NC-C2H10B10]+, 143 (90) [C2H9B10]+, 58 (100) [iPrN + H]+.

3.3. Synthesis of Compound p-C2H10B10[C(NCy)2Li(THF)2]2 (3)

This compound was prepared in a similar manner as described for 2 but using 1.44 g (7.0 mmol) of N,N’-dicyclohexylcarbodiimide as a precursor. The resulting clear solution was concentrated to a total volume of ca. 30 mL which led to the formation of a white precipitate, which was then redissolved by brief heating. Colorless single crystals suitable for X-ray diffraction were obtained directly by storing the concentrated THF solution at r.t. for a few days. Yield: 1.37 g (46%). M.p. 238 °C (dec.). Elemental analysis calcd. for C44H86B10Li2N4O4 (M = 857.18 g mol−1): C, 61.65%; H, 10.11%; N, 6.54%; found C, 61.59%; H, 10.01%; N, 6.67%. 1H NMR (400 MHz, THF-d8, 24 °C): δ = 3.20–3.29 (m br, 2 H, CH-Cy), 3.60 (THF), 2.99–3.09 (m br, 2 H, CH-Cy), 1.50–2.90 (m br, 10 H, BH), 1.76 (THF), 1.63–1.73 (m, 10 H, CH2-Cy), 1.43–1.61 (m, 10 H, CH2-Cy), 1.12–1.34 (m, 20 H, CH2-Cy) ppm. 13C{1H} NMR (100.6 MHz, THF-d8, 24 °C): δ = 154.2, 154.0 (NCN), 68.1 (THF), 56.4, 56.1, 54.6, 54.4 (CH-Cy), 36.5, 35.8, 35.0, 34.9, 34.8, 34.7, 30.6, 27.8, 27.7, 27.0, 26.9, 26.4 (CH2-Cy), 26.3 (THF), 25.8, 25.7, 25.6, 25.4, 25.0, 24.9 (CH2-Cy) ppm. 11B{1H} NMR (128.4 MHz, THF-d8, 24 °C): δ = −14.0, −14.8, −15.8 ppm. 7Li{1H} NMR (MHz, THF-d8, 24 °C): δ = 0.09 ppm. IR (KBr disk): νmax 3419 w, 3222 w, 3091 w, 2979 m, 2927 vs (νas CH2), 2852 s (νs CH2), 2611 m (ν BH), 2119 w, 1959 w, 1657 m, 1543 s (ν NCN), 1510 m, 1463 m, 1449 m (δs CH2), 1372 w, 1332 m, 1289 m, 1270 m, 1230 m, 1181 w, 1147 w, 1132 m, 1043 m (νas C-O-C), 970 w, 930 w, 888 m, 841 w, 833 w, 810 w, 794 w, 742 w, 699 w, 674 w, 589 w, 561 w, 511 w, 489 w, 477 w, 449 w, 437 w cm−1. MS (EI, 70 eV): m/z (%) 557 (45) [((CyN)2C)2C2H10B10 + H]+, 474 (27) [((CyN)2C)2C2H10B10 − C6H11 + H]+, 392 (100) [((CyN)2C)2C2H10B10 − 2C6H11 + 2H]+, 268 (33) [(CyN)NCC2H10B10 + H]+, 143 (16) [C2H9B10]+.

3.4. Synthesis of Compound p-C2H10B10[C(iPrN(SiMe3)(=NiPr)]2 (4)

A solution of compound 2 (3.5 g in 50 mL) THF was prepared as described above and treated in situ with 0.90 mL (7.0 mmol) of chlorotrimethylsilane. After stirring for 12 h at r.t., the solvent was removed in vacuum, and the orange-yellow residue was extracted with toluene (50 mL) and filtered in order to remove the by-product LiCl. The concentration of the filtrate to a total volume of ca. 20 mL followed by cooling to 4 °C for several days led to the formation of colorless, block-like single crystals which were suitable for X-ray diffraction. Yield: 1.03 g (54%). M.p. 265 °C. Elemental analysis calcd. for C22H54B10N4Si2 (M = 538.98 g mol−1): C, 49.03%; H, 10.10%; N, 10.39%; found C, 48.82%; H, 10.54%; N, 10.50%. 1H NMR (400 MHz, THF-d8, 21 °C): δ = 3.60 (sept, 2 H, 3J = 6.00 Hz, CH-iPr), 3.22 (m, 2 H, CH-iPr), 1.70–3.10 (m br, 10 H, BH), 1.27 (d, 12 H, 3J = 5.60 Hz, CH3-iPr), 0.97 (d, 12 H, 3J = 6.00 Hz, CH3-iPr), 0.16 (s, 18 H, CH3-SiMe3) ppm. 13C{1H} NMR (100.6 MHz, THF-d8, 22 °C): δ = 152.4 (NCN), 89.5 (C-NCN), 51.1, 50.6 (CH-iPr), 24.9, 23.6 (CH3-iPr), 3.82 (CH3-SiMe3) ppm. 11B{1H} NMR (128.4 MHz, THF-d8, 22 °C): δ = –13.4 ppm. 29Si{1H} NMR (79.5 MHz, THF-d8, 21 °C): δ = –1.40 ppm. IR (ATR): νmax 2987 w (νs CH3), 2966 m, 2932 w, 2889 w (νas CH3), 2641 w, 2623 m, 2590 m (ν BH), 1622 m (ν C=N), 1468 w (δas CH3), 1450 w, 1405 w, 1375 m (δs CH3), 1362 w (δs CH3), 1319 m, 1255 m, 1207 s, 1159 m, 1139 m, 1117 m, 1092 m, 1042 w, 1008 m, 979 m, 930 w, 887 m, 858 m, 833 vs (ρ CH3), 750 m, 726 m (νas SiC3), 678 m, 651 m (νs SiC3), 623 w, 607 w, 582 w, 548 w, 516 m, 471 m, 423 w, 379 w, 333 w, 260 m, 176 w, 140 w, 87 w, 68 w, 61 m cm−1. MS (EI, 70 eV): m/z (%) 541 (3) [M]+, 526 (5) [M − CH3]+, 498 (100) [M − 3CH3]+, 468 (29) [M − Si(CH3)3]+, 342 (16) [(iPrN)(iPrNSiMe3)CC2H10B10]+.

3.5. Synthesis of Compound p-C2H10B10[C(CyN(SiMe3)(=NCy)]2 (5)

Compound 5 was prepared in the same manner as described above using 3.00 g (3.5 mmol) of 3 and 0.90 mL (7.0 mmol) chlorotrimethylsilane to afford colorless crystals after crystallization from a small amount of toluene at −32 °C. Yield: 1.05 g (43%). M.p. 159 °C. Elemental analysis calcd. for C34H72B10N4Si2 (M = 701.25 g mol−1): C, 59.23%; H, 10.35%; N, 7.99%; found C, 58.20%; H, 10.28%; N, 7.94%. 1H NMR (400 MHz, THF-d8, 38 °C): δ = 3.32–3.36 (m, 2 H, CH-Cy), 1.90–3.15 (m br, 10 H, BH), 2.73–2.79 (m, 2 H, CH-Cy), 1.04–1.86 (m, 24 H, CH2-Cy), 0.15 (s, 18 H, CH3-SiMe3) ppm. 13C{1H} NMR (100.6 MHz, THF-d8, 23 °C): δ = 152.0 (NCN), 89.6 (C-NCN), 60.1, 58.8 (CH-Cy), 35.8, 35.2, 35.1, 33.6. 27.6, 27.0, 26.8, 26.4, 24.6, 24.5 (CH2-Cy), 4.0 (CH3-SiMe3) ppm. 11B{1H} NMR (128.4 MHz, THF-d8, 23 °C): δ = –13.4 ppm. 29Si{1H} NMR (MHz, THF-d8, 22 °C): δ = 0.20 ppm. IR (ATR): νmax 2930 m (νas CH2), 2852 m (νas,s CH3/CH2), 2644 s, 2606 m (ν BH), 2119 m, 1658 w, 1627 m (ν C=N), 1495 w, 1450 m (δas,s CH3/CH2), 1407 w, 1386 w (δs CH3), 1358 w, 1344 w, 1297 w, 1251 m, 1189 s, 1179 m, 1142 m, 1115 m, 1076 m, 1028 m, 996 m, 981 m, 956 m, 890 m, 857 m, 831 vs (ρ CH3), 820 vs, 777 m, 750 m (νas SiC3), 730 m, 701 m, 673 m, 658 m (νs SiC3), 630 m, 564 w, 540 w, 520 w, 497 w, 474 m, 461 m, 409 m, 384 m, 341 m, 322 m, 291 m, 260 m, 207 w, 181 w, 162 w, 136 w, 117 w, 86 w, 69 w, 57 m cm−1. MS (EI, 70 eV): m/z (%) 702 (3) [M]+, 687 (3) [M − CH3]+, 629 (32) [M − Si(CH3)3]+, 619 (71) [M − C6H11]+, 604 (12) [M − C6H11 − CH3]+, 556 (2) [M − 2 Si(CH3)3]+, 530 (67) [M − Si(CH3)3 − C6H11 − CH3]+, 279 (75) [CyN(Si(CH3)3)CNCy]+, 83 (100) [C6H11]+.

3.6. X-ray Crystallography

Single-crystal X-ray intensity data of 3 and 4 were collected on a STOE IPDS 2T diffractometer [52] equipped with a 34 cm image plate detector, using graphite-monochromated Mo-Kα radiation. The structure was solved by dual-space methods (SHELXT-2014/5) [53] and refined by full-matrix least-squares methods on F2 using SHELXL-2017/1 [54]. Crystallographic data for the title compounds have been deposited at the CCDC, 12 Union Road, Cambridge CB21EZ, U.K. Copies of the data can be obtained free of charge via the depository numbers 2248726 (3) and 2248725 (4) (e-mail: deposit@ccdc.cam.ac.uk, http://www.ccdc.cam.ac.uk).

4. Conclusions and Future Outlook

In summary, we succeeded in the synthesis and full characterization of the first amidinate and amidine derivatives of para-carborane. Lithium carboranylamidinates based on p-carborane are readily accessible by the addition of in situ-prepared 1,12-dilithio-p-carborane to 1,3-diorganocarbodiimides, R–N=C=N–R (R = iPr, Cy). This result showed that all three isomers of C2B10H12 react in completely different manners with carbodiimides. An initial reactivity study involving treatment of 2 and 3 with 2 equiv. of Me3SiCl revealed that neutral bis-silylated amidine derivatives are also easily prepared. It should be noted here that the oxygen analogue p-carborane-1,12-dicarboxylic acid has been successfully utilized as a linker in the design of carborane-based MOFs (=metal-organic frameworks) [55,56]. One could easily foresee that the amidinate and amidine derivatives of p-carborane reported here will play a similar fruitful role in MOF chemistry in the future.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28093837/s1: IR, NMR, and mass spectra for all title compounds as well as X-ray diffraction data for 3 and 4.

Author Contributions

N.H. performed the experimental work; V.L. supervised the experimental work; P.L. and F.E. carried out the crystal structure determinations; L.H. measured the IR and NMR spectra; S.B. measured the mass spectra and performed the elemental analyses; F.T.E. conceived and supervised the experiments; R.G. provided the research infrastructure and read and edited the paper; F.T.E. wrote the manuscript. All authors have read and agreed to the published version of the manuscript.

Funding

This work was financially supported by the Otto-von-Guericke-Universität Magdeburg.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Poater, J.; Solà, M.; Viñas, C.; Teixidor, F. π Aromaticity and Three-Dimensional Aromaticity: Two sides of the Same Coin? Angew. Chem. Int. Ed. 2014, 53, 12191–12195. [Google Scholar] [CrossRef]
  2. Brown, A.D.; Colquhoun, H.M.; Daniels, A.J.; MacBride, J.A.H.; Stephenson, I.R.; Wade, K. Polymers and ceramics based on icosahedral carboranes. Model studies of the formation and hydrolytic stability of aryl ether, ketone, amide and borane linkages between carborane units. J. Mater. Chem. 1992, 2, 793–804. [Google Scholar] [CrossRef]
  3. Murphy, D.M.; Mingos, D.M.P.; Haggitt, J.L.; Poell, H.R.; Westcott, S.A.; Marder, T.B.; Taylor, N.J.; Kanis, D.R. Synthesis of icosahedral carboranes for second-harmonic generation. Part 2. J. Mater. Chem. 1993, 3, 139–148. [Google Scholar] [CrossRef]
  4. Koshino, M.; Tanaka, T.; Solin, N.; Suenaga, K.; Isobe, H.; Nakamura, E. Imaging of Single Organic Molecules in Motion. Science 2007, 316, 853. [Google Scholar] [CrossRef]
  5. Villagómez, C.J.; Sasaki, T.; Tour, J.M.; Grill, L. Bottom-up Assembly of Molecular Wagons on a Surface. J. Am. Chem. Soc. 2010, 132, 16848–16954. [Google Scholar] [CrossRef]
  6. Dash, B.P.; Satapathy, R.; Gaillard, E.R.; Maguire, J.A.; Hosmane, N.S. Synthesis and Properties of Carborane-Appended C3-Symmetrical Extended π Systems. J. Am. Chem. Soc. 2010, 132, 6578–6587. [Google Scholar] [CrossRef]
  7. 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, 52998–55300. [Google Scholar]
  8. Cioran, A.M.; Musteti, A.D.; Teixidor, F.; Krpetic, Ž.Č.; Prior, I.A.; He, Q.; Kiely, C.J.; Brust, M.; Viñas, C. Mercaptocarborane-Capped Gold Nanoparticles: Electron Pools and Ion Traps with Switchable Hydrophilicity. J. Am. Chem. Soc. 2012, 134, 212–221. [Google Scholar] [CrossRef]
  9. Schwartz, J.J.; Mendoza, A.M.; Wattanatorn, N.; Zhao, Y.; Nguyen, V.T.; Spokoyny, A.M.; Mirkin, C.A.; Baše, T.; Weisse, P.S. Surface Dipole Control of Liquid Crystal Alignment. J. Am. Chem. Soc. 2016, 138, 5957–5967. [Google Scholar] [CrossRef]
  10. Li, Z.; Núñes, R.; Light, M.E.; Ruiz, E.; Teixidor, F.; Viñas, C.; Ruiz-Molina, D.; Roscini, C.; Planas, J.G. Water-Stable Carborane-Based Eu3+/Tb3+ Metal–Organic Frameworks for Tunable Time-Dependent Emission Color and Their Application in Anticounterfeiting Bar-Coding. Chem. Mater. 2022, 34, 4795–4808. [Google Scholar] [CrossRef]
  11. Zhang, K.; Song, R.; Qi, J.; Zhang, Z.; Yu, C.; Li, K.; Zhang, Z.; Li, B. Colossal Barocaloric Effect in Carboranes as a Performance Tradeoff. Adv. Funct. Mater. 2022, 32, 2112622. [Google Scholar] [CrossRef]
  12. Minnyaylo, E.O.; Kudryavtseva, A.I.; Zubova, V.Y.; Anisimov, A.A.; Zaitsev, A.V.; Ol’shevskaya, V.A.; Dolgushin, F.M.; Peregudov, A.S.; Muzafarov, A.M. Synthesis of mono- and polyfunctional organosilicon derivatives of polyhedral carboranes for the preparation of hybrid polymer materials. New J. Chem. 2022, 46, 11143–11148. [Google Scholar] [CrossRef]
  13. Liu, K.; Zhang, J.; Shi, Q.; Ding, L.; Kiu, T.; Fang, Y. Precise Manipulation of Excited State Intramolecular Proton Transfer via Incorporating Charge Transfer toward High-Performance Film-Based Fluorescence Sensing. J. Am. Chem. Soc. 2023, 145, 7408–7415. [Google Scholar] [CrossRef]
  14. Belmont, J.A.; Soto, J.; King III, R.E.; Donaldson, A.J.; Hewes, J.D.; Hawthorne, M.F. Metallacarboranes in catalysis. 8. I: Catalytic hydrogenolysis of alkenyl acetates. II: Catalytic alkene isomerization and hydrogenation revisited. J. Am. Chem. Soc. 1989, 111, 7475–7486. [Google Scholar] [CrossRef]
  15. Teixidor, F.; Flores, M.A.; Viñas, C.; Kivekäs, R.; Sillanpää, R. [Rh(7-SPh-8-Me-7,8-C2B9H10)(PPh3)2]: A New Rhodacarborane with Enhanced Activity in the Hydrogenation of 1-Alkenes. Angew. Chem. Int. Ed. 1996, 35, 2251–2253. [Google Scholar] [CrossRef]
  16. Ferlekidis, A.; Goblet-Stachow, M.; Liégeois, J.F.; Pirotte, B.; Delarge, J.; Demonceau, A.; Fontaine, M.; Noels, A.F.; Chizhevsky, I.T.; Zinevich, T.V.; et al. Ligand effects in the hydrogenation of methacycline to doxycycline and epi-doxycycline catalysed by rhodium complexes molecular structure of the key catalyst [closo-3,3-(η2,3-C7H7CH2)-3,1,2-RhC2B9H11]. J. Organomet. Chem. 1997, 536–537, 405–412. [Google Scholar] [CrossRef]
  17. Gozzi, M.; Schwarze, B.; Hey-Hawkins, E. Half- and Mixed-sandwich Metallacarboranes in Catalysis. In Boron Chemistry in Organometallics, Catalysis, Materials and Medicine; Hosmane, N.S., Eagling, R., Eds.; World Scientific: Singapore, 2018; pp. 27–80. [Google Scholar]
  18. Fisher, S.P.; Tomich, A.W.; Lovera, S.O.; Kleinsasser, J.F.; Guo, J.; Asay, M.J.; Nelson, H.M.; Lavallo, V. Nonclassical Applications of closo-Carborane Anions: From Main Group Chemistry and Catalysis to Energy Storage. Chem. Rev. 2019, 119, 8262–8290. [Google Scholar] [CrossRef]
  19. Gunther, S.O.; Lai, Q.; Senecal, T.; Huacuja, R.; Bremer, S.; Pearson, D.M.; DeMott, J.C.; Bhuvanesh, N.; Ozerov, O.V.; Klosin, J. Highly Efficient Carborane-Based Activators for Molecular Olefin Polymerization Catalysts. ACS Catal. 2021, 11, 3335–3342. [Google Scholar] [CrossRef]
  20. Grishin, I.V.; Zimina, A.M.; Anufriev, S.A.; Knyazeva, N.A.; Piskunov, A.V.; Dolgushin, F.M.; Sivaev, I.B. Synthesis and Catalytic Properties of Novel Ruthenacarboranes Based on nido-[5-Me-7,8-C2B9H10]2− and nido-[5,6-Me2-7,8-C2B9H9]2−Dicarbollide Ligands. Catalysts 2021, 11, 1409. [Google Scholar] [CrossRef]
  21. Cheng, R.; Zhang, J.; Zhang, H.; Qiu, Z.; Xie, Z. Ir-catalyzed enantioselective B–H alkenylation for asymmetric synthesis of chiral-at-cage o-carboranes. Nat. Commun. 2021, 12, 7146. [Google Scholar] [CrossRef]
  22. Vaillant, J.F.; Guenther, K.J.; King, A.S.; Morel, P.; Schaffer, P.; Sogbein, O.O.; Stephenson, K. The medicinal chemistry of carboranes. Coord. Chem. Rev. 2002, 232, 173–230. [Google Scholar] [CrossRef]
  23. Armstrong, A.F.; Vaillant, J.F. The bioinorganic and medicinal chemistry of carboranes: From new drug discovery to molecular imaging and therapy. Dalton Trans. 2007, 38, 4240–4251. [Google Scholar] [CrossRef]
  24. Scholz, M.; Hey-Hawkins, E. Carbaboranes as Pharmacophores: Properties, Synthesis, and Application Strategies. Chem. Rev. 2011, 111, 7035–7062. [Google Scholar] [CrossRef]
  25. Issa, F.; Kassiou, M.; Rendina, L.M. Boron in Drug Discovery: Carboranes as Unique Pharmacophores in Biologically Active Compounds. Chem. Rev. 2011, 111, 5701–5722. [Google Scholar] [CrossRef]
  26. Stockmann, P.; Gozzi, M.; Kuhnert, R.; Sarosi, M.-B.; Hey-Hawkins, E. New keys for old locks: Carborane-containing drugs as platforms for mechanism-based therapies. Chem. Soc. Rev. 2019, 48, 3497–3512. [Google Scholar] [CrossRef]
  27. Gozzi, M.; Schwarze, B.; Hey-Hawkins, E. Preparing (Metalla)carboranes for Nanomedicine. ChemMedChem 2021, 16, 1533–1565. [Google Scholar] [CrossRef]
  28. Marfavi, A.; Kavianpour, P.; Rendina, L.M. Carboranes in drug discovery, chemical biology and molecular imaging. Nat. Rev. Chem. 2022, 6, 486–504. [Google Scholar] [CrossRef]
  29. Chen, Y.; Du, F.; Tang, L.; Xu, J.; Zhao, Y.; Wu, X.; Li, M.; Shen, J.; Wen, Q.; Cho, C.H.; et al. Carboranes as unique pharmacophores in antitumor medicinal chemistry. Mol. Ther. Oncolytics 2022, 24, 400–416. [Google Scholar] [CrossRef]
  30. Waddington, M.A.; Zheng, X.; Stauber, J.M.; Moully, E.H.; Montgomery, H.R.; Saleh, L.M.A.; Král, P.; Spokoyny, A.M. An Organometallic Strategy for Cysteine borylation. J. Am. Chem. Soc. 2021, 143, 8661–8668. [Google Scholar] [CrossRef]
  31. Gazvoda, M.; Dhanjee, H.H.; Rodriguez, J.; Brown, J.S.; Farquhar, C.E.; Truex, N.L.; Loas, A.; Buchwald, S.L.; Pentelute, B.L. Palladium-Mediated Incorporation of Carboranes into Small Molecules, Peptides, and Proteins. J. Am. Chem. Soc. 2022, 144, 7852–7860. [Google Scholar] [CrossRef]
  32. Ma, Y.-N.; Gao, Y.; Ma, Y.; Wang, Y.; Ren, H.; Chen, X. Palladium-Catalyzed Regioselective B(9)-Amination of o-Carboranes and m-Carboranes in HFIP with Broad Nitrogen sources. J. Am. Chem. Soc. 2022, 144, 8371–8378. [Google Scholar] [CrossRef]
  33. Ma, Y.-N.; Ren, H.; Wu, Y.; Li, N.; Chen, F.; Chen, X. B(9)-OH-o-Carboranes: Synthesis, Mechanism, and Property exploration. J. Am. Chem. Soc. 2023, 145, 7331–7342. [Google Scholar] [CrossRef]
  34. Ren, H.; Zhang, P.; Xu, J.; Ma, W.; Tu, D.; Lu, C.; Yan, H. Direct B–H Functionalization of icosahedral Carboranes via Hydrogen Atom Transfer. J. Am. Chem. Soc. 2023, 145, 7638–7647. [Google Scholar] [CrossRef]
  35. Bregadze, V. Dicarba-closo-dodecaboranes C2B10H12 and their derivatives. Chem. Rev. 1992, 92, 209–223. [Google Scholar] [CrossRef]
  36. Junk, P.C.; Cole, M.L. Alkali-metal bis(aryl)formamidinates: A study of coordinative versatility. Chem. Commun. 2007, 16, 1579–1590. [Google Scholar] [CrossRef]
  37. Edelmann, F.T. Chapter 3—Advances in the Coordination Chemistry of Amidinate and Guanidinate Ligands. Adv. Organomet. Chem. 2008, 57, 183–352. [Google Scholar]
  38. Edelmann, F.T. Lanthanide amidinates and guanidinates in catalysis and materials science: A continuing success story. Chem. Soc. Rev. 2012, 41, 7657–7672. [Google Scholar] [CrossRef]
  39. Deacon, G.B.; Hossain, M.E.; Junk, P.C.; Salehisaki, M. Rare-earth N,N′-diarylformamdinate complexes. Coord. Chem. Rev. 2017, 340, 247–265. [Google Scholar] [CrossRef]
  40. Sengupta, D.; Gómez-Torres, A.; Fortier, S. Guanidinate, Amidinate, and Formamidinate Ligands. In Comprehensive Coordination Chemistry; University of Texas at El Paso: El Paso, TX, USA, 2021; Volume 3, pp. 366–405. [Google Scholar]
  41. Edelmann, F.T. Carboranylamidinates. Z. Anorg. Allg. Chem. 2013, 639, 655–667. [Google Scholar] [CrossRef]
  42. Yao, Z.-J.; Jin, G.-X. Transition metal complexes based on carboranyl ligands containing N, P, and S donors: Synthesis, reactivity and applications. Coord. Chem. Rev. 2013, 257, 2522–2535. [Google Scholar] [CrossRef]
  43. Rädisch, T.; Harmgarth, N.; Liebing, P.; Beltrán-Leiva, M.J.; Páez-Hernández, D.; Arratia-Pérez, R.; Engelhardt, F.; Hilfert, L.; Oehler, F.; Busse, S.; et al. Three new types of transition metal carboranylamidinate complexes. Dalton Trans. 2018, 47, 6666–6671. [Google Scholar] [CrossRef]
  44. Liebing, P.; Harmgarth, N.; Zörner, F.; Engelhardt, F.; Hilfert, L.; Busse, S.; Edelmann, F.T. Synthesis and Structural Characterization of Two New Main Group Element Carboranylamidinates. Inorganics 2019, 7, 41. [Google Scholar] [CrossRef]
  45. Wang, H. Recent advances on carborane-based ligands in low-valent group 13 and group 14 elements chemistry. Chin. Chem. Lett. 2022, 33, 3672–3680. [Google Scholar] [CrossRef]
  46. Harmgarth, N.; Hrib, C.G.; Lorenz, V.; Hilfert, L.; Edelmann, F.T. Unprecedented formation of polycyclic diazadiborepine derivatives through cage deboronation of m-carborane. Chem. Commun. 2014, 50, 13239–13242. [Google Scholar] [CrossRef]
  47. Stalke, D.; Wedler, M.; Edelmann, F.T. Dimere Alkalimetallbenzamidinate: Einfluß des Metallions auf die Struktur. J. Organomet. Chem. 1992, 431, C1–C5. [Google Scholar] [CrossRef]
  48. Schmidt, J.A.R.; Arnold, J. Synthesis and characterization of a series of sterically-hindered amidines and their lithium and magnesium complexes. Dalton Trans. 2002, 14, 2890–2899. [Google Scholar] [CrossRef]
  49. Baker, R.J.; Jones, C. Synthesis and characterisation of sterically bulky lithium amidinate and bis-amidinate complexes. J. Organomet. Chem. 2006, 691, 65–71. [Google Scholar] [CrossRef]
  50. Kahl, S.B.; Kasar, R.A. Simple, High-Yield Synthesis of Polyhedral Carborane Amino Acids. J. Am. Chem. Soc. 1996, 118, 1223–1224. [Google Scholar] [CrossRef]
  51. Harmgarth, N.; Gräsing, D.; Dröse, P.; Hrib, C.G.; Jones, P.G.; Lorenz, V.; Hilfert, S.; Busse, S.; Edelmann, F.T. Novel inorganic heterocycles from dimetalated carboranylamidinates. Dalton Trans. 2014, 43, 5001–5013. [Google Scholar] [CrossRef]
  52. Stoe & Cie. X-Area and X-Red; Stoe & Cie: Darmstadt, Germany, 2002. [Google Scholar]
  53. Sheldrick, G.M. SHELXT—Integrated space-group and crystal-structure determination. Acta Cryst. 2015, A71, 3–8. [Google Scholar] [CrossRef]
  54. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. 2015, 71, 3–8. [Google Scholar]
  55. Farha, O.K.; Spokoyny, A.M.; Mulfort, K.L.; Hawthorne, M.F.; Mirkin, C.A.; Hupp, J.T. Synthesis and Hydrogen Sorption Properties of Carborane Based Metal-Organic Framework Materials. J. Am. Chem. Soc. 2007, 129, 12680–12681. [Google Scholar] [CrossRef]
  56. Dash, B.P.; Satapathy, R.; Maguire, J.A.; Hosmane, N.S. Polyhedral boron clusters in materials science. New J. Chem. 2011, 35, 19551972. [Google Scholar] [CrossRef]
Figure 1. Three isomers of the icosahedral closo-carborane cage C2B10H12.
Figure 1. Three isomers of the icosahedral closo-carborane cage C2B10H12.
Molecules 28 03837 g001
Scheme 1. Preparation of lithium carboranylamidinates derived from o-carborane.
Scheme 1. Preparation of lithium carboranylamidinates derived from o-carborane.
Molecules 28 03837 sch001
Scheme 2. Formation of polycylic diazadiborepines from m-carborane.
Scheme 2. Formation of polycylic diazadiborepines from m-carborane.
Molecules 28 03837 sch002
Scheme 3. Synthetic route to the title compounds 2 and 3.
Scheme 3. Synthetic route to the title compounds 2 and 3.
Molecules 28 03837 sch003
Scheme 4. Preparation of the silylated bis(amidine) derivatives 4 and 5.
Scheme 4. Preparation of the silylated bis(amidine) derivatives 4 and 5.
Molecules 28 03837 sch004
Figure 2. Molecular structure of 3 in the crystal. Displacement ellipsoids of the heavier atoms are drawn with 50% probability; H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg.): C(1)-C(2) 1.551(2), N(1)-C(2) 1.305(2), N(2)-C(2) 1.340(2), N(2)-Li 1.953(3), O(1)-Li 1.949(3), O(2)-Li 1.933(3), N(1)-C(2)-N(2) 136.9(1), N(1)-C(2)-C(1) 109.7(1), N(2)-C(2)-C(1) 113.4(1), C(2)-N(2)-Li 134.9(1). Symmetry code to generate equivalent atoms: ‘2–x, 1–y, 1–z.
Figure 2. Molecular structure of 3 in the crystal. Displacement ellipsoids of the heavier atoms are drawn with 50% probability; H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg.): C(1)-C(2) 1.551(2), N(1)-C(2) 1.305(2), N(2)-C(2) 1.340(2), N(2)-Li 1.953(3), O(1)-Li 1.949(3), O(2)-Li 1.933(3), N(1)-C(2)-N(2) 136.9(1), N(1)-C(2)-C(1) 109.7(1), N(2)-C(2)-C(1) 113.4(1), C(2)-N(2)-Li 134.9(1). Symmetry code to generate equivalent atoms: ‘2–x, 1–y, 1–z.
Molecules 28 03837 g002
Figure 3. Molecular structure of 4 in the crystal. Displacement ellipsoids of the heavier atoms with 50% probability; H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg.): C(1)-C(2) 1.539(2), C(2)-N(1) 1.265(2), C(2)-N(2) 1.423(2), Si(1)-N(2) 1.751(1), N(1)-C(2)-N(2) 128.9(1), N(1)-C(2)-C(1) 113.8(1), N(2)-C(2)-C(1) 117.3(1), C(2)-N(2)-Si(1) 121.00(9). Symmetry code to generate equivalent atoms: ‘1–x,1–y, 1–z.
Figure 3. Molecular structure of 4 in the crystal. Displacement ellipsoids of the heavier atoms with 50% probability; H atoms are omitted for clarity. Selected bond lengths (Å) and angles (deg.): C(1)-C(2) 1.539(2), C(2)-N(1) 1.265(2), C(2)-N(2) 1.423(2), Si(1)-N(2) 1.751(1), N(1)-C(2)-N(2) 128.9(1), N(1)-C(2)-C(1) 113.8(1), N(2)-C(2)-C(1) 117.3(1), C(2)-N(2)-Si(1) 121.00(9). Symmetry code to generate equivalent atoms: ‘1–x,1–y, 1–z.
Molecules 28 03837 g003
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Harmgarth, N.; Liebing, P.; Lorenz, V.; Engelhardt, F.; Hilfert, L.; Busse, S.; Goldhahn, R.; Edelmann, F.T. Synthesis and Structural Characterization of p-Carboranylamidine Derivatives. Molecules 2023, 28, 3837. https://doi.org/10.3390/molecules28093837

AMA Style

Harmgarth N, Liebing P, Lorenz V, Engelhardt F, Hilfert L, Busse S, Goldhahn R, Edelmann FT. Synthesis and Structural Characterization of p-Carboranylamidine Derivatives. Molecules. 2023; 28(9):3837. https://doi.org/10.3390/molecules28093837

Chicago/Turabian Style

Harmgarth, Nicole, Phil Liebing, Volker Lorenz, Felix Engelhardt, Liane Hilfert, Sabine Busse, Rüdiger Goldhahn, and Frank T. Edelmann. 2023. "Synthesis and Structural Characterization of p-Carboranylamidine Derivatives" Molecules 28, no. 9: 3837. https://doi.org/10.3390/molecules28093837

Article Metrics

Back to TopTop