Next Article in Journal
Characterization of Phenolic Compounds, Vitamin E and Fatty Acids from Monovarietal Virgin Olive Oils of “Picholine marocaine” Cultivar
Next Article in Special Issue
Syntheses and Study of a Pyrroline Nitroxide Condensed Phospholene Oxide and a Pyrroline Nitroxide Attached Diphenylphosphine
Previous Article in Journal
Targeting Beta-Blocker Drug–Drug Interactions with Fibrinogen Blood Plasma Protein: A Computational and Experimental Study
Previous Article in Special Issue
Synthesis of Nitroxide Diradical Using a New Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluorinated Organic Paramagnetic Building Blocks for Cross-Coupling Reactions

by
Larisa V. Politanskaya
1,
Pavel A. Fedyushin
1,
Tatyana V. Rybalova
1,
Artem S. Bogomyakov
2,
Nargiz B. Asanbaeva
1 and
Evgeny V. Tretyakov
1,3,*
1
N.N. Vorozhtsov Novosibirsk Institute of Organic Chemistry, 9 Ac. Lavrentiev Avenue, 630090 Novosibirsk, Russia
2
International Tomography Center, Siberian Branch of Russian Academy of Sciences, 3a Institutskaya Str., 630090 Novosibirsk, Russia
3
N.D. Zelinsky Institute of Organic Chemistry, Leninsky Prospect, 47, 119991 Moscow, Russia
*
Author to whom correspondence should be addressed.
Molecules 2020, 25(22), 5427; https://doi.org/10.3390/molecules25225427
Submission received: 21 October 2020 / Revised: 15 November 2020 / Accepted: 18 November 2020 / Published: 19 November 2020
(This article belongs to the Special Issue Stable Organic Radicals)

Abstract

:
New stable polyfluorinated nitroxide radicals for use in cross-coupling reactions, namely, N-tert-butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-iodobenzene and N-tert-butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-ethynylbenzene, were prepared from perfluoroiodobenzene. The reaction of the polyfluoro derivative with tert-butylamine under autoclaving conditions leading to the formation of N-tert-butyl-2,3,5,6-tetrafluoro-4-iodoaniline proved to be the key stage of the whole process. The fluorinated tert-butyl iodophenyl nitroxide was found to form in a solid state via N–O···I halogen bonds, a one-dimensional assembly of the radicals. The acceptor role of the nitroxide group in the halogen bonding changes to a donor role when the nitroxide reacts with Cu(hfac)2. In the last case, zero-dimensional assembly prevails, giving a three-spin complex with axial coordinated nitroxide groups and, as a consequence, causing ferromagnetic intramolecular exchange interactions between Cu(II) and radical spins.

Graphical Abstract

1. Introduction

tert-Butylarylnitroxides and corresponding polynitroxides are an important class of paramagnets that have been actively used in different fields of research. A typical method for the preparation of tert-butylarylnitroxides involves an interaction of an appropriate aromatic organometallic compound with tert-nitrosobutane, thus giving aryl tert-butylhydroxylamine, followed by its oxidation to the target radical product. Less commonly, the oxidation of N-tert-butylanilines is performed to obtain nitroxide radicals [1,2].
Recently, we proposed a new synthetic approach to obtaining functionalized tert-butylphenylnitroxides by sequential substitution of a fluorine atom in polyfluorinated arenes with tert-butylamine or lithium tert-butylamide and oxidation of resultant N-tert-butylanilines with meta-chloroperoxybenzoic acid (m-CPBA). This approach has been found to generate nitroxide mono- and diradicals in high yields (Figure 1) [3,4]. Moreover, it turned out that polyfluorinated tert-butylarylnitroxides are much stable than nonfluorinated analogs. Therefore, the nitroxides of this type can serve as paramagnetic building blocks for diverse applications.
In particular, one of the tasks of our current project aimed at obtaining spin-labeled graphene nanostructures is to search for stable paramagnetic blocks suitable for introduction into polycondensed aromatic structures and capable of efficiently injecting spin density into such structures [5,6,7,8]. Because the above-mentioned polyfluorinated tert-butylarylnitroxides satisfy all these requirements, within the framework of this work, we studied the possibility of obtaining their iodo and ethynyl derivatives that may be next subjected to the Sonogashira cross-coupling reaction.

2. Results and Discussion

Following the newly developed approach, we carried out the reaction of perfluoroiodobenzene with tert-butylamine in chloroform at room temperature (r.t.). As opposed to perfluorotoluene and perfluorobenzonitrile, the iodo derivative did not possess enough reactivity, and the reaction with the amine did not proceed. The use of tert-butylamine under autoclaving conditions at 120 °C led to the substitution of the fluorine atom in perfluoroiodobenzene and the formation of N-tert-butyl-2,3,5,6-tetrafluoro-4-iodoaniline (1) in a high yield. Formation of other isomeric amines was not detectable (Supplementary Materials). The oxidation of diamine 1 with m-CPBA was performed in CHCl3 at r.t. and gave one of the target nitroxide radicals, 4, in a yield of ~60% (Scheme 1).
4-Iodoaniline 1 was cross-coupled with trimethylsilyl acetylene in dry Et3N in the presence of Pd(PPh3)2Cl2 (5 mol%) and CuI (10 mol%) as catalysts. The reaction was carried out at r.t. in a tightly closed Schlenk flask in an argon atmosphere. The process was terminated after the disappearance of 19F NMR signals belonging to the starting compound to obtain alkynyl derivative 2 with a yield of ~92%. The resulting compound 2 was cleaved by the action of K2CO3 in MeOH at r.t., thus producing fluorinated ethynyl-derivative 3 with an almost quantitative yield (Supplementary Materials). At the final stage, amine 3 was oxidized with m-CPBA in CHCl3 at r.t. and gave ethynyl-substituted nitroxide 5 as a red oily compound in a yield of ~58%. Both nitroxide radicals 4 and 5 were found to be stable and were purified by thin-layer chromatography (TLC).
Electron spin resonance (ESR) spectra (Figure 2 and Figure 3) for diluted (~10−4 M) and oxygen-free toluene solutions of radicals 4 and 5 showed triplet patterns at giso = 2.0065 and giso = 2.0064, respectively. The spectra were simulated with the following sets of parameters: For 4, AN = 1.283 mT and A2F = 0.078 mT; for 5, AN = 1.278 mT. In the case of nitroxide 5, a high-resolution ESR spectrum was recorded to achieve more complex splitting of the central line of its triplet. The spectrum was well simulated, taking into account nine hyperfine structure (hfs) constants for the protons of the tert-butyl group (AH = 0.022 mT), two pairs of hfs constants for the distant fluorine atoms (AFortho = 0.089 mT; AFmeta = 0.044 mT) and one hfs constant for the acetylene proton (AH = 0.075 mT).
Even though both freshly prepared radicals 4 and 5 were red oils, repeated efforts were made to obtain them in a crystalline form. Eventually, by crystallization from a cold n-heptane solution, nitroxide radical 4 was isolated as high-quality crystals that allowed to solve its molecular and crystal structure by X-ray diffraction (XRD) analysis (Figure 4). It was revealed that radical 4 crystallizes in the orthorhombic Pbca space group (Table 1). Bond lengths of the tert-butylnitroxide moiety (Table 2) are fully consistent with those of previously described radicals of this family [3]. The nitroxide groups in diradical 4 are twisted at a large angle (~64°) relative to the aromatic ring. The observed twisting is due to, first, steric repulsion between the tert-butyl group and ortho-fluorine atoms and, second, electrostatic repulsion of dipoles C‒F and N‒O. In this regard, the analogous dihedral angle in nonfluorinated tert-butylphenylnitroxides is twofold smaller and has experimental and calculated values of 23–34° [9].
The X-ray analysis revealed an assembly of radicals 4 into zigzag chains via halogen bonds between the nitroxide oxygen and iodine atom (Figure 4b,c). If we consider so-called conjugated nitroxides [2], there are very few cases when halogen bonding induces one-dimensional self- or supra-organization of paramagnetic species. Moreover, the halogen bonding with the participation of an iodine atom has been described only for the family of nitronyl nitroxides [10,11,12,13]. For tert-butylphenylnitroxides, this type of halogen bonding was shown for the first time. In addition, the halogen bonds observed in nitroxide 4 are very strong, as indicated by the short I1⋯O1 distance (2.967 Å, >15% shorter than the sum of van der Waals radii of O and I [14]) and almost linear C–I⋯O angles (~164°). The N–O (1.278(4) Å) bond distance is typical of nitronyl nitroxide moieties, indicating that the presence of the halogen bond has little impact on electronic properties of the free radicals. This is a consequence of the essentially electrostatic nature of the halogen bond interaction.
An interaction of half an equivalent of bis(hexafluoroacetylacetonato)copper(II) (abbreviated as Cu(hfac)2) with radical 4 in CHCl3 gave rise to the [Cu(hfac)2(4)2] complex (Supplementary Materials). Its recrystallization from n-hexane generated well-shaped crystals suitable for X-ray analysis. If an equivalent amount of Cu(hfac)2 was used in the reaction, complex [Cu(hfac)2(4)2] also formed along with the crystallization of excess complex Cu(hfac)2. Attempts to prepare a similar complex with nitroxide 5 every time led to the formation of dense intergrowths of crystals of complex [Cu(hfac)2(5)]. Due to the low quality of the crystals, the structure of the complex was solved with R-factor 8.76% and is considered a tentative result here.
According to XRD analysis, coordination compound [Cu(hfac)2(4)2] crystallizes in the triclinic crystal system (space group P-1) with two independent molecules per asymmetric unit, as shown in Figure 5a. In the distorted octahedron around the copper ion, two ONO atoms occupy axial positions with dCu–O = 2.380 and 2.400 Å in the independent molecules. Equatorial Cu–Ohfac distances in both molecules are in the range of 1.933 to 1.955 Å (Table 2).
In the crystal structure of [Cu(hfac)2(4)2], the independent molecules are bound via two van der Waals interactions of the C–F⋅⋅⋅CPh type (contacts i and ii). One of the fluorine atoms attached to the aromatic cycle engages in an intermolecular contact (iii) with the H atoms of the tert-butyl group, thereby binding the dimers into a layer (Figure 5b).
Temperature dependences of the effective magnetic moment (µeff) and inverse magnetic susceptibility (1/χ) for complex [Cu(hfac)2(4)2] are shown in Figure 6. The µeff value is 3.17 µB at 300 K and increases with decreasing temperature to 3.55 µB at 4 K. The 1/χ(T) dependence is linear in the temperature range 100–300 K and obeys the Curie–Weiss law with best-fit values C = 1.218 ± 0.002 K·cm3/mol and θ = 9.1 ± 0.3 K. The µeff value at 300 K and Curie constant C are in good agreement with theoretical spin-only values of 3.00 µB and 1.125 K·cm3/mol for three paramagnetic centers: Two nitroxides and one Cu(II) ion. The increase in µeff with decreasing temperature and positive Weiss constant θ indicate the domination of ferromagnetic exchange interactions between spins of paramagnetic centers. An analysis of experimental µeff(T) dependences was performed via a model of a three-spin exchange cluster (spin Hamiltonian H = −2JCuR × (SR1SCu + SCuSR2) − 2JRR × SR1SR2), taking into account intermolecular exchange interactions zJ′ in the mean field approximation. Best-fit values of gCu and exchange interaction parameters JCu–R, JR–R and zJ′ are 2.21, 15.9 cm−1, −8.7 cm−1 and 0.15 cm−1, respectively. g-Factors for the nitroxides were fixed at gR = 2 to avoid overparametrization. The observed ferromagnetic exchange interactions between the unpaired electrons in [Cu(hfac)2(4)2] are consistent with the XRD data on the axial coordination of the nitroxide groups to the Cu2+ ion [15,16,17,18].

3. Conclusions

The development of the chemistry of nitroxides opens up new areas of their practical application. Recently, we devised a new synthetic approach to functionalized tert-butylphenylnitroxides via substitution of a fluorine atom in polyfluorinated arenes by tert-butylamine with subsequent oxidation of the obtained N-tert-butylanilines with m-CPBA [3]. With highly activated polyfluorinated compounds, the substitution reaction proceeds smoothly, with the formation of the corresponding aniline. In the case of less activated polyfluorinated aromatics, we proposed the use of lithium tert-butylamide, which allowed, for example, to synthesize a necessary diamine precursor for the N,N′-(perfluorobiphenyl-4,4′-diyl)bis(N-tert-butyl-N-oxyamine) diradical [4]. In the present work, we employed the third version of the reaction, namely, the process was carried out in an autoclave at 120 °C, using tert-butylamine both as a reagent and a solvent. Under these conditions, the reaction of perfluoroiodobenzene with tert-butylamine proceeded regioselectively and gave N-tert-butyl-2,3,5,6-tetrafluoro-4-iodoaniline in high yields. The obtained iodo derivative was converted via a multistep procedure into an ethynyl-substituted aryl nitroxide, which was oxidized into the corresponding tert-butylphenylnitroxide. The latter radical crystalizes with the formation of the structure in which free radicals are bound into one-dimensional chains by relatively strong halogen bonds N–O···I. If upon the formation of the halogen bond the nitroxide group acts as an acceptor, then in the reaction with Cu(hfac)2 it becomes an electron pair donor. The reaction produces a molecular complex of 1:2 composition with axial coordination of the paramagnetic ligands. The result of such coordination gives rise to ferromagnetic intramolecular exchange interactions between Cu(II) and radical spins.
Both synthesized nitroxides, N-tert-butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-iodobenzene and N-tert-butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-ethynylbenzene, represent stable paramagnetic building blocks for the Sonogashira cross-coupling reaction. This reaction is being evaluated for the synthesis of functional paramagnets with potential application in the fields of molecular magnetism and spintronics. In particular, in the course of preliminary experiments, it was shown that the paramagnetic iodo derivative can be successfully used to introduce a spin-bearing group into the planar hexabenzocoronene backbone.

4. Materials and Methods

4.1. Materials and Measurements

Cu(hfac)2 was prepared according to previously described procedures [19]; before use, the Cu(hfac)2 complex was sublimated and stored in a desiccator charged with NaOH. The starting materials were obtained from commercial sources (Merck KGaA, Darmstadt, Germany) and were used without further purification. All solvents were purified according to standard procedures. Preparative TLC was performed on Merck precoated silica gel 60 PF254 containing gypsum. The visualization of developed chromatograms was performed by means of UV light.
NMR spectra were recorded on Bruker Avance-300 (300.13 MHz for 1H and 282.37 MHz for 19F), Avance-400 (400.13 MHz for 1H and 100.62 MHz for 13C) and DRX-500 (500.13 MHz for 1H, 125.76 MHz for 13C) spectrometers (Bruker Corporation, Billerica, MA, USA). CDCl3 served as a solvent, with residual CHCl3H = 7.26, δC = 77.0) acting as an internal standard; C6F6F = 163.0) was used as an external reference for recording 19F NMR spectra. 13C NMR spectra were determined with C–H spin decoupling.
Continuous-wave ESR spectra were acquired on a commercial Bruker X Band (9 GHz) spectrometer, Elexsys E 540 (Bruker Corporation, Billerica, MA, USA), at r.t. in dilute (~10−4 M) toluene solutions degassed by means of repeated freeze–pump–thaw cycles. ESR spectra were recorded with the following settings: Frequency, 9.87 GHz; microwave power, 2.0 mW; modulation amplitude, 0.01 mT for radical 4 and 0.005 mT for radical 5; time constant, 20.5 ms; and conversion time, 20 ms. Simulations of solution ESR lines were carried out in the EasySpin software (5.2.28) [20], which is available at http://www.easypin.org.
Masses of molecular ions were determined by high-resolution mass spectrometry (HRMS) on a DFS Thermo Scientific instrument at an ionization energy of 70 eV (Thermo Fisher Scientific, Waltham, MA, USA). Melting points were recorded on a Mettler-Toledo FP81 Thermosystem apparatus (Mettler-Toledo, Greifensee, Switzerland). IR spectra were acquired by means of a Bruker Vector 22 spectrometer in KBr pellets or thin films (Bruker Optik GmbH in Ettlingen). Elemental analyses were performed on a Carlo Erba 1106 CHN elemental analyzer.
Magnetic susceptibility was measured with a MPMSXL SQUID magnetometer (Quantum Design, San Diego, CA) in the temperature range 2 to 300 K with a magnetic field of up to 5 kOe. Diamagnetic corrections were made via Pascal constants. The effective magnetic moment was calculated as follows: μeff(T) = ((3k/NAμB2)χT)1/2 ≈ (8χT)1/2.

4.2. Synthetic Procedures

4.2.1. N-tert-Butyl-2,3,5,6-tetrafluoro-4-iodoaniline (1)

Pentafluoroiodobenzene (1.50 g, 5.1 mmol) and 2-methylpropan-2-amine (5 mL) were placed into a stainless-steel autoclave (volumetric capacity 50 mL). The reaction mixture was stirred at 120 °C for 90 h. Then, the mixture was allowed to cool down to r.t., placed directly onto a chromatography plate (silica gel) and air-dried. The title product was isolated by TLC with hexane as an eluent (Rf = 0.65) in a 91% yield (1.62 g) as colorless oil. IR (thin), νmax: 3423, 2968, 2121, 2088, 1801, 1732, 1635, 1487, 1460, 1396, 1369, 1288, 1228, 1203, 1130, 1088, 972, 810 cm−1; 1H NMR (300 MHz, CDCl3): δ = 3.46 (br s, 1H, NH), 1.28 (s, 9H, CH3); 13C NMR (126 MHz, CDCl3): δ = 147.1 [dddd, 1J (C3,F3) = 242.3 Hz, 2J (C3,F2) = 15.1 Hz, C3, C5], 140.2 [ddt, 1J (C2,F2) = 243.4 Hz, 2J (C2,F3) = 16.2 Hz, C2, C6], 126.6 [tt, 2J (C1,F2) = 14.1 Hz, 3J (C1,F3) = 2.4 Hz, C1], 60.4 [t, 2J (C4,F3) = 2J (C4,F5) = 28.1 Hz, C4], 54.5 (s, C7), 30.1 [t, J (C8,F2) = J (C8,F6) = 2.7 Hz, C8]; 19F NMR (282 MHz, CDCl3): δ = −124.1 (m, 2F, F3, F5), −149.9 (m, 2F, F2, F6); HRMS (EI): m/z [M]+ calcd. for C10H10F4NI: 346.9789; found 346.9790.

4.2.2. N-tert-Butyl-2,3,5,6-tetrafluoro-4-((trimethylsilyl)ethynyl)aniline (2)

Pd(PPh3)2Cl2 (50 mg, 0.07 mmol) and CuI (27 mg, 0.144 mmol) were added to a stirred solution of iodoaryl 1 (500 mg, 1.44 mmol) and ethynyltrimethylsilane (282 mg, 2.88 mmol) in dry Et3N (20 mL) at r.t. in an argon atmosphere. The mixture was stirred at r.t. for 24 h, placed directly onto a chromatographic plate (silica gel) and air-dried. The title product was isolated by TLC with hexane as an eluent (Rf = 0.80) in a 92% yield (420 mg) as colorless oil. IR (thin), νmax: 3427, 2966, 2160, 1651, 1500, 1471, 1435, 1252, 1230, 1203, 983, 847, 762, 702, 677 cm−1; 1H NMR (300 MHz, CDCl3): δ = 3.34 (br s, 1H, NH), 1.29 [s, 9H, C(CH3)3], 0.24 [s, 9H, Si(CH3)3]; 13C NMR (100 MHz, CDCl3): δ = 147.5 [dddd, 1J (C3,F3) = 249.7 Hz, 2J (C3,F2) = 14.6 Hz, C3, C5], 139.2 [ddt, 1J (C2,F2) = 239.3 Hz, 2J (C2,F3) = 14.6 Hz, C2, C6], 127.1 [tt, 2J (C1,F2) = 13.7 Hz, 3J (C1,F3) = 2.6 Hz, C1], 105.7 [t, J (C7,F3) = J (C7,F5) = 3.4 Hz, C7], 94.4 [tt, 2J (C4,F3) = 2J (C4,F5) = 18.4 Hz, C4], 89.4 [t, J (C8,F5) = J (C8,F3) = 3.9 Hz, C8], 54.3 (s, C10), 30.2 [t, J (C11,F2) = J (C11,F6) = 3.0 Hz, C11], −0.4 (s, C9); 19F NMR (282 MHz, CDCl3): δ = −139.9 (m, 2F, F3, F5), −154.2 (m, 2F, F2, F6); HRMS (EI): m/z [M]+ calcd. for C15H19F4NSi: 317.1217; found 317.1215.

4.2.3. N-tert-Butyl-4-ethynyl-2,3,5,6-tetrafluoroaniline (3)

To a solution of 2 (230 mg, 0.72 mmol) in MeOH (10 mL), K2CO3 (200 mg, 1.45 mmol) was added. The reaction mixture was stirred at r.t. for 3.5 h, then placed directly onto a chromatographic plate (silica gel) and air-dried. The title product was isolated by TLC with hexane as an eluent (Rf = 0.50) in a 98% yield (174 mg) as colorless oil. IR (thin), νmax: 3426, 3307, 2974, 2875, 2117, 1705, 1651, 1500, 1471, 1433, 1471, 1369, 1303, 1232, 1203, 1172, 1041, 985, 974, 893, 806, 673, 605, 432 cm−1; 1H NMR (300 MHz, CDCl3): δ = 3.63 (br s, 1H, NH), 3.45 (s, 1H, H8), 1.30 (t, 9H, C(CH3)3); 13C NMR (126 MHz, CDCl3): δ = 147.9 [dddd, 1J (C3,F3) = 250.2 Hz, 2J (C3,F2) = 14.4 Hz, C3, C5], 139.0 [ddt, 1J (C2,F2) = 239.1 Hz, 2J (C2,F3) = 14.5 Hz, C2, C6], 127.7 [tt, 2J (C1,F2) = 13.5 Hz, 3J (C1,F3) = 2.7 Hz, C1], 92.7 [tt, 2J (C4,F3) = 2J (C4,F5) = 18.2 Hz, C4], 87.0 [t, J (C8,F5) = J (C8,F3) = 3.4 Hz, C8], 69.4 [t, J (C7,F3) = J (C7,F5) = 4.0 Hz, C7], 54.2 (s, C9), 30.3 [t, J (C11,F2) = J (C11,F6) = 3.1 Hz, C10]; 19F NMR (282 MHz, CDCl3): δ = −140.1 (m, 2F, F3, F5), −154.2 (m, 2F, F2, F6); HRMS (EI): m/z [M]+ calcd. for C12H11F4N: 245.0822; found 245.0823.

4.2.4. N-tert-Butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-iodobenzene (4)

A solution of N-tert-butyl-2,3,5,6-tetrafluoro-4-iodoaniline 1 (0.7 mmol) and m-CPBA (1.0 mmol, 172 mg) in CHCl3 (5 mL) was stirred at r.t. for 24 h. Column chromatography (SiO2 or Al2O3 column 3 × 20 cm, CHCl3 as an eluent) afforded a red fraction of radical 4. The solvent was removed under reduced pressure at r.t. to obtain title compound 4. Red crystals; yield 60% (150 mg); mp 89.8 °C (decomp.). IR (KBr), νmax: 2981, 1626, 1483, 1462, 1398, 1350, 1275, 1248, 1196, 970, 800, 773, 721, 579 cm−1; UV-Vis (EtOH) λmax/nm (lg ε): 235 (3.16), 307 (2.60); HRMS (EI): m/z [M]+ calcd. for C10H9F4NOI: 361.9660; found 361.9662; Anal. calcd. for C10H9F4NOI: C, 33.17; H, 2.51; N, 3.87. Found: C, 33.44; H, 2.43; N, 3.87.

4.2.5. N-tert-Butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-ethynylbenzene (5)

The compound was obtained through the oxidation of amine 3 in a manner similar to the preparation of nitroxyl 4. Red oil; yield 58% (105 mg). IR (thin), νmax: 3305, 3257, 2983, 2943, 2123, 1722, 1680, 1643, 1616, 1491, 1396, 1367, 1304, 1252, 1201, 989, 835, 787, 769, 735, 677, 642, 584, 501 cm−1; UV-Vis (EtOH) λmax/nm (lg ε): 239 (3.11), 249 (3.08), 317 (2.41); HRMS (EI): m/z [M]+ calcd. for C12H10F4NO: 260.0693; found 260.0691.

4.2.6. trans-Bis(1,1,1,5,5,5-hexafluoropentane-2,4-dionato-κ2O,O′)bis{N-tert-butyl-N-oxyamino-2,3,5,6-tetrafluoro-4-iodobenzene-κO}copper(II) ([Cu(hfac)2(4)2])

Cu(hfac)2·H2O (248 mg, 0.5 mmol) was added to a solution of radical 4 (362 mg, 1.0 mmol) in CHCl3 (10 mL). The reaction mixture was stirred for 30 min and then was incubated at −15 °C for 10 h. The solution was filtered and evaporated. The residue was dissolved in n-hexane (5 mL) and the resultant solution was filtered and incubated at −15 °C for 10 h to prepare crystals that were filtered off and air-dried. Brown crystals; yield 78%; IR (KBr) v ~ max, cm−1: 3435, 3001, 2985, 1641, 1597, 1558, 1527, 1485, 1400, 1354, 1261, 1207, 1149, 1107, 1030, 968, 816, 798, 773, 746, 723, 681, 596, 581, 530; calcd. for C30H20N2CuF20I2O6: C 29.98, H 1.68, F, 31.62; N 2.33; found C 30.30, H 1.57, F 31.76, N 2.07.

4.3. Single-Crystal XRD Analyses

XRD data were collected at r.t. on a Bruker Kappa Apex II CCD diffractometer with Mo Kα radiation (λ = 0.71073 Å) and a graphite monochromator (Table 1). Absorption corrections were applied empirically using SADABS programs [21]. The structures were solved by direct methods by means of the SHELXS-97 software suite [22,23] and were refined using the full-matrix least-squares method against all F2 in anisotropic approximation (beside the H atoms) using the SHELXL2014/7 software suite [24]. The H atoms’ positions were calculated with the riding model. The asymmetric unit of [Cu(hfac)2(4)2] contains halves of two independent moieties.
CCDC 2038555, 2,038,556 and 2,039,515 (preliminary data) contain the crystallographic data for 4, [Cu(hfac)2(4)2] and [Cu(hfac)2(5)]. These data can be obtained free of charge via http://www.ccdc.cam.ac.uk/cgi-bin/catreq.cgi or from the Cambridge Crystallographic Data Centre, 12 Union Road, Cambridge CB2 1EZ, UK; fax: (+44) 1223 336 033; or e-mail: [email protected].

Supplementary Materials

The Supplementary Materials are available online. NMR spectra and the IR spectrum of [Cu(hfac)2(4)2].

Author Contributions

conceptualization, E.V.T.; methodology, L.V.P., P.A.F. and A.S.B.; investigation, L.V.P., P.A.F., T.V.R., A.S.B. and N.B.A.; writing—Original draft preparation, E.V.T.; supervision, E.V.T.; funding acquisition, E.V.T. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by the Russian Science Foundation (grant No. 18-13-00173).

Acknowledgments

The authors would like to acknowledge the Multi-Access Chemical Research Center SB RAS for spectral and analytical measurements.

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 the data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Hicks, R.G. Stable Radicals: Fundamentals and Applied Aspects of Odd-Electron Compounds; John Wiley and Sons: Chichester, UK, 2010. [Google Scholar]
  2. Tretyakov, E.V.; Ovcharenko, V.I. The chemistry of nitroxide radicals in the molecular design of magnets. Russ. Chem. Rev. 2009, 78, 971–1012. [Google Scholar] [CrossRef]
  3. Tretyakov, E.; Fedyushin, P.; Panteleeva, E.; Gurskaya, L.; Rybalova, T.; Bogomyakov, A.; Zaytseva, E.; Kazantsev, M.; Shundrina, I.; Ovcharenko, V. Aromatic SNF-approach to fluorinated phenyl tert-butyl nitroxides. Molecules 2019, 24, 4493. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Fedyushin, P.; Rybalova, T.; Asanbaeva, N.; Bagryanskaya, E.; Dmitriev, A.; Gritsan, N.; Kazantsev, M.; Tretyakov, E. Synthesis of nitroxide diradical using a new approach. Molecules 2020, 25, 2701. [Google Scholar] [CrossRef] [PubMed]
  5. Slota, M.; Keerthi, A.; Myers, W.K.; Tretyakov, E.; Baumgarten, M.; Ardavan, A.; Sadeghi, H.; Lambert, C.J.; Narita, A.; Müllen, K.; et al. Magnetic edge states and coherent manipulation of graphene nanoribbons. Nature 2018, 557, 691–695. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Stass, D.; Tretyakov, E. Estimation of absolute spin counts in nitronyl nitroxide-bearing graphene nanoribbons. Magnetochemistry 2019, 5, 32. [Google Scholar] [CrossRef] [Green Version]
  7. Morozov, V.; Tretyakov, E. Spin polarization in graphene nanoribbons functionalized with nitroxide. J. Mol. Model. 2019, 25, 58. [Google Scholar] [CrossRef] [PubMed]
  8. Ten, Y.A.; Troshkova, N.M.; Tretyakov, E.V. From spin-labelled fused polyaromatic compounds to magnetically active graphene nanostructures. Russ. Chem. Rev. 2020, 89, 693–712. [Google Scholar] [CrossRef]
  9. Ohshita, J.; Iida, T.; Ohta, N.; Komaguchi, K.; Shiotani, M.; Kunai, A. Synthesis of Phenylnitroxides Bridged by an sp3-Linkage. Org. Lett. 2002, 4, 403–406. [Google Scholar] [CrossRef] [PubMed]
  10. Liu, Z.-L.; Song, Y.; Du, C.-F. CCDC 244384, Experimental Crystal Structure Determination. 2014. Available online: https://www.ccdc.cam.ac.uk/structures/search?id=doi:10.5517/cc869c3&sid=DataCite (accessed on 21 October 2020).
  11. Wagner, B.; Gompper, R.; Polborn, K. CSD Communication (Private Communication); CCDC 198532; Department of Chemistry, University of Munich: Munchen, Germany, 2002. [Google Scholar]
  12. Suzuki, S.; Nakamura, F.; Naota, T. A direct synthetic method for (nitronyl nitroxide)-substituted π-electronic compounds via a palladium-catalyzed cross-coupling reaction with a zinc complex. Mater. Chem. Front. 2018, 2, 591–596. [Google Scholar] [CrossRef]
  13. Espallargas, G.M.; Recuenco, A.; Romero, F.M.; Brammer, L.; Libri, S. One-dimensional organization of free radicals via halogen bonding. CrystEngComm 2012, 14, 6381–6383. [Google Scholar] [CrossRef]
  14. Bondi, A. van der Waals Volumes and Radii. J. Phys. Chem. 1964, 68, 441–451. [Google Scholar] [CrossRef]
  15. Caneschi, A.; Gatteschi, D.; Grand, A.; Laugier, J.; Pardi, L.; Rey, P. Moderate Ferromagnetic Exchange between Copper(II) and a Nitronyl Nitroxide in a Square-Pyramidal Adduct. MO Interpretation of the Mechanism of Exchange in Copper(II)–Nitroxide Complexes. Inorg. Chem. 1988, 27, 1031–1035. [Google Scholar]
  16. Iwahori, F.; Markosyan, A.S.; Inoue, K. Structures and Magnetic Properties of the Complexes Made up by Cu(hfac)2 and Bisnitroxide Radical Derivatives. Mol. Cryst. Liq. Cryst. 2002, 376, 449–454. [Google Scholar] [CrossRef]
  17. Ishimaru, Y.; Kitano, M.; Kumada, H.; Koga, N.; Iwamura, H. Regiospecificity in the Exchange Coupling of the Spins of Copper(II) Ion Coordinated with the Ring Nitrogen Atoms and N-tert-Butylaminoxyl Radical Attached as a Substituent on the Pyridine and N-Phenylimidazole Rings. Inorg. Chem. 1998, 37, 2273–2280. [Google Scholar] [CrossRef] [PubMed]
  18. Field, L.M.; Lahti, P.M.; Palacio, F.; Paduan-Filho, A. Manganese(II) and Copper(II) Hexafluoroacetylacetonate 1:1 Complexes with 5-(4-[N-tert-Butyl-N-aminoxyl]phenyl)pyrimidine: Regiochemical Parity Analysis for Exchange Behavior of Complexes between Radicals and Paramagnetic Cations. J. Am. Chem. Soc. 2003, 125, 10110–10118. [Google Scholar] [CrossRef] [PubMed]
  19. Bertrand, J.A.; Kaplan, R.I. A study of bis(hexafluoroacetylacetonato)copper(II). Inorg. Chem. 1966, 5, 489–491. [Google Scholar] [CrossRef]
  20. Stoll, S.; Schweiger, A. EasySpin: Simulating cw ESR spectra. Biol. Magn. Reson. 2007, 27, 299–321. [Google Scholar]
  21. Sheldrick, G.M. SADABS, Program for Area Detector Adsorption Correction; Institute for Inorganic Chemistry, University of Göttingen: Göttingen, Germany, 1996. [Google Scholar]
  22. Sheldrick, G.M. A short history of SHELX. Acta Cryst. A 2008, 64, 112–122. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  23. Sheldrick, G.M. SHELX-97. Programs for Crystal Structure Analysis (Release 97-2); University of Göttingen: Göttingen, Germany, 1998. [Google Scholar]
  24. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed]
Figure 1. Synthesized polyfluorinated tert-butylarylnitroxides.
Figure 1. Synthesized polyfluorinated tert-butylarylnitroxides.
Molecules 25 05427 g001
Scheme 1. Synthesis of polyfluorinated tert-butylarylnitroxides 4 and 5.
Scheme 1. Synthesis of polyfluorinated tert-butylarylnitroxides 4 and 5.
Molecules 25 05427 sch001
Figure 2. Experimental (black curve) and simulated (red curve) electron spin resonance (ESR) spectra of 4.
Figure 2. Experimental (black curve) and simulated (red curve) electron spin resonance (ESR) spectra of 4.
Molecules 25 05427 g002
Figure 3. Experimental (black curves) and simulated (red curves) ESR spectra of 5; (a) the whole spectrum; (b) the central component.
Figure 3. Experimental (black curves) and simulated (red curves) ESR spectra of 5; (a) the whole spectrum; (b) the central component.
Molecules 25 05427 g003
Figure 4. Molecular structure (a) and fragments of crystal structure (b,c) of nitroxide radical 4 (the thermal ellipsoids are drawn at the 30% probability level).
Figure 4. Molecular structure (a) and fragments of crystal structure (b,c) of nitroxide radical 4 (the thermal ellipsoids are drawn at the 30% probability level).
Molecules 25 05427 g004
Figure 5. Molecular structure ((a) one of the independent molecules) and fragments of crystal structure (b) of the [Cu(hfac)2(4)2] complex (the thermal ellipsoids are drawn at the 30% probability level).
Figure 5. Molecular structure ((a) one of the independent molecules) and fragments of crystal structure (b) of the [Cu(hfac)2(4)2] complex (the thermal ellipsoids are drawn at the 30% probability level).
Molecules 25 05427 g005
Figure 6. The µeff(T) (●) and 1/χ(T) (□) dependences for the [Cu(hfac)2(4)2] complex (solid curves are theoretical ones).
Figure 6. The µeff(T) (●) and 1/χ(T) (□) dependences for the [Cu(hfac)2(4)2] complex (solid curves are theoretical ones).
Molecules 25 05427 g006
Table 1. XRD data for compounds 4, [Cu(hfac)2(4)2] and [Cu(hfac)2(5)].
Table 1. XRD data for compounds 4, [Cu(hfac)2(4)2] and [Cu(hfac)2(5)].
Compound4[Cu(hfac)2(4)2][Cu(hfac)2(5)]
Empirical formulaC10H9F4INOC30H20CuF20I2N2O6C22H12CuF16NO5
Formula weight362.081201.82737.87
Crystal systemOrthorhombicTriclinicTriclinic
Space groupPbcaP¯1P¯1
Unit cell dimensions a, Å11.5462(4)10.0763(9)9.132(1)
b, Å12.1680(5)13.784(1)10.472(1)
c, Å17.8601(7)16.009(2)17.273 (2)
α, °9096.469(3)75.370(6)
β, °90102.396(4)83.253(6)
γ, °90104.941(3)65.505(5)
Volume, Å32509.2(2)2064.7(3)1454.3(3)
Z822
F(000)13841154728
Density (calcd.), Mg·m−31.9171.9331.685
Abs. coefficient, mm−12.592.160.89
Crystal size, mm0.90 × 0.30 × 0.100.30 × 0.15 × 0.010.25 × 0.24 × 0.14
Tmin, Tmax0.544, 0.8620.722, 0.8620.825, 0.928
Reflections collected282153413428365
Independent reflections [Rint]351673225688
Reflections observed (I > 2σ(I))274138502856
Rint0.0480.0890.074
θ range for data collection (°)2.3–30.41.6–25.22.194–26.088
Range of h,−13→16−12→11−11→11
k,−16→16−16→16−12→12
l−23→24−19→18−21→21
Completeness to θ 50° (%)10099.199.7
Data/restraints/parameters3516/0/1577322/0/5535688/6/503
Final R [F2 > 2 σ (F2)]0.0400.0620.087
Final wR(F2) (all data)0.1150.1730.324
Goodness-of-fit on F21.031.011.05
Largest diff. peak/hole, e·Å−31.44, −1.120.71, −0.640.64, −0.56
Table 2. Selected geometrical parameters for 4 and [Cu(hfac)2(4)2] (two independent molecules).
Table 2. Selected geometrical parameters for 4 and [Cu(hfac)2(4)2] (two independent molecules).
Compound4[Cu(hfac)2(4)2][Cu(hfac)2(4)2]′
Bond lengths, Å
Cu1–O1-2.401(8)2.380(7)
Cu1–O2-1.933(6)1.954(6)
Cu1–O3-1.955(6)1.954(5)
C4–I12.079(4)2.062(9)2.060(8)
N1–O11.278(4)1.29(1)1.283(9)
N1–C11.428(4)1.43(1)1.41(1)
N1–C71.493(5)1.47(1)1.50(1)
Bond angles, °
O2–Cu1–O1-95.7(2)98.6(2)
O3–Cu1–O1-91.5(3)91.2(2)
O2–Cu1–O3-92.3(2)92.3(2)
O1–N1–C1115.7(3)115.8(7)115.9(6)
O1–N1–C7118.0(3)117.2(7)116.8(6)
C1–N1–C7125.3(3)125.7(7)126.4(6)
Sample Availability: Samples of all described compounds are available from the authors.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Politanskaya, L.V.; Fedyushin, P.A.; Rybalova, T.V.; Bogomyakov, A.S.; Asanbaeva, N.B.; Tretyakov, E.V. Fluorinated Organic Paramagnetic Building Blocks for Cross-Coupling Reactions. Molecules 2020, 25, 5427. https://doi.org/10.3390/molecules25225427

AMA Style

Politanskaya LV, Fedyushin PA, Rybalova TV, Bogomyakov AS, Asanbaeva NB, Tretyakov EV. Fluorinated Organic Paramagnetic Building Blocks for Cross-Coupling Reactions. Molecules. 2020; 25(22):5427. https://doi.org/10.3390/molecules25225427

Chicago/Turabian Style

Politanskaya, Larisa V., Pavel A. Fedyushin, Tatyana V. Rybalova, Artem S. Bogomyakov, Nargiz B. Asanbaeva, and Evgeny V. Tretyakov. 2020. "Fluorinated Organic Paramagnetic Building Blocks for Cross-Coupling Reactions" Molecules 25, no. 22: 5427. https://doi.org/10.3390/molecules25225427

Article Metrics

Back to TopTop