Next Article in Journal
Dimethyl 3,7-diamino-4,8-bis((2-methoxy-2-oxoethyl)thio)benzo[1,2-b:4,5-b’]dithiophene-2,6-dicarboxylate
Previous Article in Journal
1-Phenyl-3,3-di(1H-pyrazol-1-yl)propan-1-one
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

One-Pot Synthesis of Dioxime Oxalates

by
Laura Adarve-Cardona
1,
David Ezenarro-Salcedo
2,
Mario A. Macías
3 and
Diego Gamba-Sánchez
1,*
1
Laboratory of Organic Synthesis Bio- and Organocatalysis, Chemistry Department, Universidad de Los Andes, Cra. 1 No 18A-12 Q:305, Bogotá 111711, Colombia
2
Inorganic Chemistry, Catalysis and Bio-Inorganic Group, Chemistry Department, Universidad de Los Andes, Cra. 1 No. 18A-12, Bogotá 111711, Colombia
3
Crystallography and Chemistry of Materials, CrisQuimMat, Chemistry Department, Universidad de Los Andes, Cra. 1 No. 18A-12, Bogotá 111711, Colombia
*
Author to whom correspondence should be addressed.
Molbank 2022, 2022(4), M1473; https://doi.org/10.3390/M1473
Submission received: 29 September 2022 / Revised: 18 October 2022 / Accepted: 21 October 2022 / Published: 26 October 2022
(This article belongs to the Section Organic Synthesis)

Abstract

:
Dioxime oxalates, a type of carbonyl oximes, are well-known as clean sources of iminyl radicals that undergo key organic chemistry transformations. A series of dioxime oxalates is reported in this manuscript, obtained by the reaction of the corresponding oximes with oxalyl chloride and Et3N at room temperature. This one-pot method afforded three novel dioxime oxalates and the crystal structure of cyclopentanone dioxime oxalate analysis is also presented.

1. Introduction

Since the synthesis and isolation of the first organic free radical by Moses Gomberg at the beginning of the nineteenth century, chemists have been interested in studying and forming radical species emphasizing their applications in life and material sciences [1]. Eventually, in the 1970s, preliminary studies on the chemistry of nitrogen-centered free radicals indicated that iminyl radicals were slightly unreactive and no suitable methods to generate them were available [2,3,4]. That changed during the last few decades. Today, iminyl radicals are valuable intermediates involved in many organic reactions, including additions to π-systems to produce reactive radicals (Scheme 1a) [5,6], transformation into carbon-centered radicals by H transfers (Scheme 1b) [7], and the formation of new C-N bonds in the synthesis of nitrogen heterocycles [8,9,10,11,12,13] (Scheme 1c).
Iminyl radicals can also be formed by thermal fragmentation of N-O bonds and under milder photoredox methodologies in which visible light is used for the excitation process [14], or by triplet energy transfer (EnT) [15].
Oximes and their derivates are arguably the most used compounds for homolytic N-O bond fragmentation to obtain N• and O• centered radicals [16]. For instance, oxime esters form iminyl radicals via single electron transfer (SET) or triplet energy transfer (EnT) [17] (Scheme 2a). Consequently, dioxime oxalates have been recognized as atom-efficient and clean sources of iminyl radicals, considering that the main by-product in the processes is CO2 (Scheme 2b). Examples have been described by Forrester et al. to produce radicals during UV photolysis [18], and by Walton and co-workers during the study of photochemical reactions of a range of symmetrical and unsymmetrical oxime oxalates for synthesizing dihydropyrroles and phenanthridines [19]. Generally, the reported syntheses of dioxime oxalates consist of two steps. The first step requires a nucleophilic substitution between an oxime and oxalyl chloride at low temperatures (−60 °C) to afford an O-(chlorooxalyl)oxime, which can be isolated and undergo a Beckmann rearrangement for the synthesis of nitrilium salts. In contrast, if a second molecule of oxime is added, it reacts with the O-(chlorooxalyl)oxime, yielding the dioxime oxalate (Scheme 2c) [20].
We report herein the one-pot synthesis of dioxime oxalates under simple reaction conditions; a method that is helpful with cyclic and acyclic oximes. We also report three new oxalates and a brief structural analysis of the cyclopentanone dioxime oxalate crystal structure.

2. Results

Non-commercially available ketoximes (1) may be obtained from the classic condensation reaction between the corresponding ketones and hydroxylamine hydrochloride under basic conditions. Acetophenone oxime 1a was employed as a model substrate in developing the one-pot method for preparing dioxime oxalate. The first experiment was carried out by adding dropwise 1.8 mmol of oxalyl chloride to a solution of (1a, 1.5 mmol) and Et3N (1.65 mmol) in anhydrous MeCN at room temperature. Whereas the reaction was complete after 1 h, the crude was a brown viscous oil that showed difficulties in the purification processes. A second reaction was made under the same conditions, replacing MeCN with dry DCM. The result of this experiment was an easily handled crude and the desired dioxime oxalate 2a was obtained in 92% yield; the NMR data agreed with previous descriptions [20] (Scheme 3).
With those results in hand, the same conditions were explored on cyclic oximes, affording another four dioxime oxalates with moderate to good yields (Table 1, Entries 1–4). Remarkably, to our knowledge, compounds (2b), (2d), and (2e) were not reported until now.
It is well known that dioxime oxalates can be stored for long periods under ambient conditions, and are stable to moderate heat and hydrolysis [9]. Moreover, they are solids. Considering these physicochemical properties, crystallization was attempted. A successful crystallization process was achieved with cyclopentanone dioxime oxalate via slow diffusion of Et2O through a saturated dissolution of the product in AcOEt. The obtained crystal confirmed the molecular structure by X-ray crystal analysis.

3. Discussion

The crystal structure of cyclopentanone dioxime oxalate was determined from X-ray single-crystal measurements. A search in the CSD database version 5.41 (date of the search: September 2022) using the ConQuest software version 2020.1 for molecules with the same molecular core did not generate any results about its crystallographic analysis. Crystal data, data collection, and structure refinement details are summarized in Table 2.
The molecular structure drawn as ellipsoids described with anisotropic thermal vibrations (ORTEP style) is shown in Figure 1a. Cyclopentane moieties are out of the plane defined by the dioxime oxalate fragment, with dihedral angles between their weighted least-squares mean planes of 23.53° (Figure 1b). In the crystal, the supramolecular structure lacks classic hydrogen bonds. For this reason, no short contacts are detected. However, C5-H5···O1 hydrogen interactions (H···O distance: 2.70 Å; symmetry code: 3/2 − x,1/2 + y,1 − z) connect molecules to form sheets stacked along the [101] direction (Figure 1c), which are further connected by C6-H6···O2 hydrogen interactions along the [101] direction (H···O distance: 2.73 Å; symmetry code: 1 − x,1 − y,1 − z). The molecular orientations in the crystal leave the dioxime oxalate fragments oriented along the [100] direction (Figure 1d), with distances between neighboring fragments of 4.49 Å.
The supramolecular structure is mainly driven by long C-H···O interactions observed in the Hirshfeld surface (HS) mapped over dnorm (Figure 2). No red spots are detected, suggesting no interactions smaller than the sum of two consecutive atoms’ van der Waals radii. These sorts of structures are considered networks built by weak interactions. Nevertheless, from the 2D-fingerprint plots, H···O/O···H interactions represent 26.0% of the total HS, supporting the importance of these contacts in the formation of the crystal. Other interactions, such as H···N/N···H, constitute 11.3% of the total HS and can be observed due to the closeness caused by the shorter C5-H5···O1 interactions in the formation of the molecular sheets (Figure 1c). The high proportion of H···H non-covalent interactions (51.2% of the HS) results from the critical role that dispersion forces play in crystal growth.

4. Materials and Methods

4.1. Materials

All the reactions were performed under a nitrogen atmosphere. Dichloromethane was dried over anhydrous CaH2 and distilled before use. All the reagents were used as received from commercial suppliers (Alfa Aesar, Massachusetts, United States). Reaction progress was monitored by thin-layer chromatography (TLC) performed on aluminum plates coated with silica gel F254, with 0.2 mm thick TLC plates visualized using ultraviolet (UV) light at 254 nm or stained with p-anisaldehyde, vanillin, or KMnO4 solutions. Flash column chromatography was performed using silica gel 60 (230–400 mesh). Acetophenone, cyclopentanone, and cyclohexanone were obtained from commercial suppliers (Merck/MilliporeSigma, Burlington, MA, United States s). 3-phenylcyclobutan-1-one [21] and 2,3-dihydro-1H-inden-1-one [22] were synthesized according to the previously reported procedures. Some compounds were purified using flash chromatography (FC). The solvents used for purification are described as follows: FC: (Solvent).

4.2. Preparation of Ketoximes

Ketone (1.25 mmol, 1 equiv.), hydroxylamine hydrochloride (3 equiv.), and sodium acetate trihydrate (3 equiv.) were mixed in absolute EtOH (12.5 mL) and then, refluxed until complete conversion was observed by TLC (pentane:AcOEt 4:1). The solvent was then evaporated under reduced pressure. Saturated NaCl solution (50 mL) was added and the aqueous phase was extracted with CH2Cl2 (3 × 50 mL). The combined organic phases were dried over anhydrous MgSO4, filtered, and concentrated. The product was purified by flash column chromatography on silica gel (pentane:AcOEt 4:1) to give the desired oxime. Oximes NMR data agree with previous literature reports descriptions.

4.2.1. Acetophenone Oxime (1a)

According to the general procedure, acetophenone (1.20 g, 10 mmol) was used to obtain the acetophenone oxime as a white solid (1.15 g, 85%) [23].
1H NMR (400 MHz, CDCl3): δ (ppm) 9.01 (brs, 1H), 7.63 (m, 2H), 7.39 (m, 3H), 2.31 (s, 3H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 156.2, 136.5, 129.4, 128.7, 126.1, 12.5.

4.2.2. Cyclopentanone Oxime (1b)

According to the general procedure, cyclopentanone (1.00 g, 11.9 mmol) was used to obtain the cyclopentanone oxime as a white solid (1.02 g, 87%) [23].
1H NMR (400 MHz, CDCl3): δ (ppm) 8.44 (brs, 1H), 2.46 (t, J = 7.0 Hz, 2H), 2.36 (t, J = 6.7 Hz, 2H), 1.70–1.81 (m, 4H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 167.7, 31.0, 27.2, 25.3, 24.7.

4.2.3. Cyclohexanone Oxime (1c)

According to the general procedure, cyclohexanone (986 mg, 10.05 mmol) was used to obtain the cyclohexanone oxime as a white solid (1.04 g, 91%)[23].
1H NMR (400 MHz, CDCl3): δ (ppm) 2.50 (t, J = 6.0 Hz, 2H), 2.21 (t, J = 6.2 Hz, 2H), 1.54–1.72 (m, 6H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 161.1, 32.3, 27.0, 25.9, 25.8, 24.5.

4.2.4. 3-Phenylcyclobutan-1-one Oxime (1d)

According to the general procedure, 3-phenylcyclobutan-1-one (300 mg, 2.05 mmol) was used to obtain the 3-phenylcyclobutan-1-one oxime as pale-yellow solid (272 mg, 82%) [24].
1H NMR (400 MHz, CDCl3): δ (ppm) 8.26 (brs, 1H), 7.21–7.39 (m, 5H), 3.66–3.56 (m, 1H), 3.27–3.54 (m, 2H), 2.99– 3.11 (m, 2H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 156.7, 144.1, 128.7, 126.7, 39.5, 38.3, 32.9.

4.2.5. 2,3-Dihydro-1H-Inden-1-one Oxime (1e)

According to the general procedure, 2,3-dihydro-1H-inden-1-one (500 mg, 3.78 mmol) was used to obtain the 2,3-dihydro-1H-inden-1-one oxime as pale-yellow solid (446 mg, 80.1%) [22].
1H NMR (400 MHz, CDCl3): δ (ppm)7.67–7.66 (d, 1H), 7.50 (brs, 1H), 7.30–7.38 (m, 2H), 7.20–7.30 (m, 1H), 3.02–3.12 (m, 2H), 2.92–3.01 (m, 2H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 164.50, 148.64, 136.19, 130.55, 127.15, 125.77, 121.62, 28.66, 25.91.

4.3. Preparation of Dioxime Oxalates

To a solution of 1.5 mmol ketoxime in 5 mL of anhydrous dichloromethane, the Et3N (1.65 mmol) was added. Then, oxalyl chloride (1.8 mmol) was added dropwise at room temperature under stirring and a nitrogen atmosphere. The reaction was stirred until completion was indicated by TLC (pentane:AcOEt 9:1) (the total reaction time is around 1 h). Water (15 mL) was added and the mixture was extracted with DCM (3 × 10 mL). The combined extracts were washed with brine (2 × 25 mL), dried over anhydrous Na2SO4, filtered, and concentrated in vacuo. Purification of the crude by flash chromatography (pentane:AcOEt 9:1) afforded the corresponding oxalate.

4.3.1. O,O′-Oxalylbis(1-phenylacetophenone oxime) (2a)

According to the general procedure, 1-phenylethan-1-one oxime (203 mg, 1.5 mmol) was used to obtain acetophenone dioxime oxalate as a white solid (180 mg, 74%) [20].
1H NMR (400 MHz, CDCl3): δ (ppm) 7.71 (d, J = 7.5 Hz, 2H), 7.39–7.47 (m, 3H), 2.48 (s, 3H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 164.4, 133.7, 131.2, 128.7, 127.1, 14.7.
FT-IR (neat) ν (cm−1 ): 2927, 2858, 1789, 1307, 763, 694.
mp = 50–52 °C

4.3.2. O,O′-Oxalyldicyclopentanone Oxime (2b)

According to the general procedure, cyclopentanone oxime (297 mg, 3.0 mmol) was used to obtain the cyclopentanone dioxime oxalate oxime as a colorless solid (230 mg, 60%).
The product was crystallized through a liquid–liquid slow diffusion technique. In total, 20 mg of 2b was dissolved in 5 mL of ethyl acetate, and 10 mL of diethyl ether was layered and cooled to −5 °C. After 15 days, the single crystals grow colorless. Finally, the crystals were washed with cold diethyl ether.
1H NMR (400 MHz, CDCl3): δ (ppm) 2.60–2.68 (m, 2H), 2.54–2.60 (m, 2H), 1.79–1.88 (m, 4H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 178.4, 31.7, 30.0, 25.3, 24.6.
mp = 98–100 °C
FT-IR (neat) ν (cm−1): 2962, 2885, 1789, 1654, 1261, 1126, 840, 790.
HRMS (ESI): calculated for C12H17N2O4 (M+H) 253.1183; found 253.1183.

4.3.3. O,O′-Oxalyldicyclohexanone Oxime (2c)

According to the general procedure, cyclohexanone oxime (300 mg, 2.65 mmol) was used to obtain the cyclohexanone dioxime oxalate as a pale-yellow solid (222 mg, 60%) [20].
1H NMR (400 MHz, CDCl3): δ (ppm) 2.61 (t, J = 6.4 Hz, 2H), 2.39 (t, J = 6.3 Hz, 2H), 1.60–1.81 (m, 6H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 171.6, 32.1, 27.4, 26.9, 26.0, 25.4.
mp = 68–69 °C
FT-IR (neat) ν (cm−1): 2642, 1666, 1616, 1431, 1234, 713, 559.
HRMS (ESI): calculated for C15H25N2O5 (M + CH3OH + H) 313.1758; found, 313.1758.

4.3.4. O,O′-Oxalylbis(3-phenyl-3-phenylcyclobutan-1-one oxime) (2d)

According to the general procedure, 3-phenylcyclobutan-1-one oxime (161 mg, 1 mmol) was used to obtain the 3-phenylcyclobutanone dioxime oxalate as a colorless solid (168 mg, 89%).
1H NMR (400 MHz, CDCl3): δ (ppm) 7.31–7.41 (m, 2H), 7.23–7.31 (m, 3H), 3.65–3.74 (m, 1H), 3.48–3.65 (m, 2H), 3.09–3.28 (m, 2H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 168.6, 142.6, 128.9, 127.2, 126.4, 39.8, 39.5, 32.4.
FT-IR (neat) ν (cm−1): 3028, 2927, 1755, 1496, 1107, 698.
mp = 118–120 °C
HRMS (ESI): calculated for C22H24N3O4 (M + NH4), 394, 1761; found, 394, 1766.

4.3.5. O,O′-Oxalylbis(2,3-dihydro-1H-2,3-dihydro-1H-inden-1-one oxime) (2e)

According to the general procedure, 2,3-dihydro-1H-inden-1-one (60 mg, 0.408 mmol) was used to obtain the 2,3-dihydro-1H-inden-1-one dioxime oxalate as a colorless solid (67 mg, 94%).
1H NMR (400 MHz, CDCl3): δ (ppm) 7.83–7.93 (m, 1H), 7.41–7.51 (m, 1H), 7.27–7.41 (m, 2H), 3.04–3.22 (m, 4H).
13C NMR {1H} (101 MHz, CDCl3): δ (ppm) 172.65, 150.5, 133.6, 132.9, 127.6, 125.9, 123.6, 28.7, 28.3.
FT-IR (neat) ν (cm−1): 1762, 1631, 1338, 1123, 1045, 883, 760.
mp = 140–145 °C
HRMS (ESI): calculated for C20H17N2O4 (M + H), 349.1183; found, 349.1187.

4.4. Characterization Techniques

All the 1H NMR and 13C NMR spectra were recorded using a BRUKER Avance III HD Ascend 400 spectrometer. Chemical shifts are given in parts per million (ppm, δ), referenced to the TMS (1H and 13C). When necessary, the solvent peak of residual CDCl3 was defined at δ = 7.26 ppm (1H NMR) and δ = 77.16 (13C NMR). Coupling constants are quoted in Hz (J). Splitting patterns of 1H NMR were designated as singlet (s), doublet (d), triplet (t), quartet (q), or multiplet (m). Splitting patterns that could not be interpreted or easily visualized were designated multiplet (m) or broad (br). Neat infrared spectra were recorded using a THERMO NICOLET-NEXUS (FT-IR) with PIKE MIRacle ATR cell. Wave numbers (νmax) are reported in cm−1. High-resolution mass spectrometry was recorded using an Agilent 5973 (70 eV) spectrometer, using electrospray ionization (ESI). GC-MS was recorded in a Thermo Scientific Trace 1300. Copies of NMR spectra are provided as Supplementary Materials. The X-ray intensity data were measured at room temperature, 298 (2) K, using CuKα radiation (λ = 1.54184 Å), and ω scans in an Agilent SuperNova, Dual, Cu at Zero, Atlas four-circle diffractometer equipped with a CCD plate detector (Rigaku Americas Corporation, The Woodlands, Texas, USA). The collected frames were integrated with the CrysAlis PRO software package (CrysAlisPro 1.171.39.46e, Rigaku Oxford Diffraction, 2018). Data were corrected for the absorption effect using the CrysAlis PRO software package by the empirical absorption correction using spherical harmonics, implemented in the SCALE3 ABSPACK scaling algorithm (CrysAlisPro 1.171.39.46e, Rigaku Oxford Diffraction, 2018). The structure of the cyclopentanone dioxime oxalate 2b was solved using an iterative algorithm [25] and then, completed by a difference Fourier map.

5. Conclusions

We developed a one-pot method for synthesizing dioxime oxalates, starting from the corresponding oximes. Compared with the previously described methods, our methodology does not isolate the O-(chlorooxalyl)oxime intermediate and does not need cryogenic baths. The procedure is applicable for acyclic and cyclic ketoximes, allowing three novel dioxime oxalate compounds. Finally, we report the structural analysis of the cyclopentanone dioxime oxalate from X-ray single-crystal data. The crystal growth is controlled by a combination of dispersion forces and weak C-H···O hydrogen interactions, equal to or longer than the sum of the van der Waals radii. Hirshfeld’s surface maps support this observation.

Supplementary Materials

The NMR spectra of all oximes and oxalate oximes are reported herein. Crystallographic information for the structural analysis was deposited in the Cambridge Crystallographic Data Center, CCDC, with a deposition number 2209301. A copy of this information may be obtained free of charge from CCDC, 12 Union Road, Cambridge, CB2 1EZ, UK (Fax: +44-1223-336033; e-mail: https://www.ccdc.cam.ac.uk/structures/ or http://www.ccdc.cam.ac.uk accessed on 24 September 2022).

Author Contributions

Conceptualization, D.G.-S.; methodology, L.A.-C.; formal analysis, D.G.-S. and L.A.-C.; investigation, L.A.-C. and D.E.-S.; data curation, L.A.-C.; writing—original draft preparation, L.A.-C., M.A.M., and D.G.-S.; writing—review and editing, D.G.-S., M.A.M., and L.A.-C.; supervision, D.G.-S.; project administration, D.G.-S.; funding acquisition, D.G.-S.; recrystallization, D.E.-S.; X-ray diffraction analysis, M.A.M. All authors have read and agreed to the published version of the manuscript.

Funding

This research was funded by Universidad de Los Andes, Faculty of Science, grant number INV-2021-126-2269.

Data Availability Statement

The data presented in this study are available upon request to the corresponding author.

Acknowledgments

L.A.-C. and D.E.-S. acknowledge the Universidad de Los Andes and especially, the Chemistry Department for their fellowships. M.A.M. acknowledge the support from Facultad de Ciencias at the Universidad de Los Andes, Colombia (FAPA-P18.160422.043).

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. American Chemical Society National Historic Chemical Landmarks. Moses Gomberg and Organic Free Radicals. Available online: http://www.acs.org/content/acs/en/education/whatischemistry/landmarks/freeradicals.html (accessed on 1 September 2022).
  2. Symons, M.C.R. The importance of imino radicals (R2C=N) as reaction intermediates. Tetrahedron 1973, 29, 615. [Google Scholar] [CrossRef]
  3. Roberts, B.P.; Winter, J.N. Electron spin resonance studies of radicals derived from organic azides. J. Chem. Soc. Perkin Trans. 1979, 2, 1353. [Google Scholar] [CrossRef]
  4. Griller, D.; Mendenhall, G.D.; Van Hoof, W.; Ingold, K.U. Kinetic applications of electron paramagnetic resonance spectroscopy. XV. Iminyl radicals. J. Am. Chem. Soc. 1974, 96, 6068. [Google Scholar] [CrossRef]
  5. Boivin, J.F.E.; Zard, S.Z. A New and Synthetically Useful Source of Iminyl Radicals. Tetrahedron Lett. 1991, 32, 4299–4302. [Google Scholar] [CrossRef]
  6. Boivin, J.F.E.; Zard, S.Z. Iminyl radicals: Part I. generation and intramolecular capture by an olefin. Tetrahedron 1994, 50, 1745–1756. [Google Scholar] [CrossRef]
  7. Xiao, T.; Huang, H.; Anandb, D.; Zhou, L. Iminyl-Radical-Triggered C-C Bond Cleavage of Cycloketone Oxime Derivatives: Generation of Distal Cyano-Substituted Alkyl Radicals and Their Functionalization. Synthesis 2020, 52, 1585–1601. [Google Scholar] [CrossRef]
  8. Kitamura, M.; Narasaka, K. Synthesis of Aza-Heterocycles from Oximes by Amino-Heck Reaction. Chem. Rec. 2002, 2, 268–277. [Google Scholar] [CrossRef]
  9. Kitamura, M.; Narasaka, K. Amination with Oximes. Eur. J. Org. Chem. 2005, 2005, 4505–4519. [Google Scholar] [CrossRef]
  10. Zard, S.Z. Recent progress in the generation and use of nitrogen-centred radicals. Chem. Soc. Rev. 2008, 37, 1603–1618. [Google Scholar] [CrossRef] [PubMed]
  11. Walton, J.C. The Oxime Portmanteau Motif: Released Heteroradicals Undergo Incisive EPR Interrogation and Deliver Diverse Heterocycles. Acc. Chem. Res. 2014, 47, 1406–1416. [Google Scholar] [CrossRef]
  12. Chen, J.-R.; Hu, X.-Q.; Lu, L.-Q.; Xiao, W.-J. Visible light photoredox-controlled reactions of N-radicals and radical ions. Chem. Soc. Rev. 2016, 45, 2044–2056. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Jackman, M.M.; Cai, Y.; Castle, S.L. Recent Advances in Iminyl Radical Cyclizations. Synthesis 2017, 49, 1785–1795. [Google Scholar] [CrossRef]
  14. Li, J.; Zhang, P.; Jiang, M.; Yang, H.; Zhao, Y.; Fu, H. Visible light as a sole requirement for intramolecular C(sp3)–H imination. Org. Lett. 2017, 19, 1994–1997. [Google Scholar] [CrossRef] [PubMed]
  15. Strieth, F.S.; Glorius, F. Triplet Energy Transfer Photocatalysis: Unlocking the Next Level. Chem. 2020, 6, 1888–1903. [Google Scholar] [CrossRef]
  16. Rykaczewski, K.A.; Wearing, E.R.; Blackmun, D.E.; Schindler, C.S. Reactivity of oximes for diverse methodologies and synthetic applications. Nat. Synth. 2022, 1, 24–36. [Google Scholar] [CrossRef]
  17. Walton, J.C. Synthetic Strategies for 5- and 6-Membered Ring Azaheterocycles Facilitated by Iminyl Radicals. Molecules 2016, 21, 660. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  18. Forrester, A.R.; Napier, R.J.; Thomson, R.H. Iminyls. Part 7. Intramolecular hydrogen abstraction; synthesis of heterocyclic analogues of α-tetralone. J. Chem. Soc. Perkin Trans. 1981, 1, 984–987. [Google Scholar] [CrossRef]
  19. Portela-Cubillo, F.; Scanlan, E.M.; Scott, J.S.; Walton, J.C. From dioxime oxalates to dihydropyrroles and phenanthridines viaiminyl radicals. Chem. Commun. 2008, 4189, 4189–4191. [Google Scholar] [CrossRef] [PubMed]
  20. Jochims, J.C.; Hehl, S.; Herzberger, S. Preparation and Beckmann Rearrangement of O-(Chlorooxalyl)oximes. Synthesis 1990, 1990, 1128–1133. [Google Scholar] [CrossRef]
  21. Chernykh, A.V.; Radchenko, D.S.; Chernykh, A.V.; Kondratov, I.S.; Tolmachova, N.A.; Datsenko, O.P.; Kurkunov, M.A.; Zozulya, S.X.; Kheylik, Y.P.; Bartels, K.; et al. Synthesis and Physicochemical Properties of 3-Fluorocyclobutylamines. Eur. J. Org. Chem. 2015, 2015, 6466–6471. [Google Scholar] [CrossRef]
  22. Ma, G.; Xu, Z.; Zhang, P.; Liu, J.; Hao, X.; Ouyang, J.; Liang, P.; You, S.; Jia, X. A Novel Synthesis of Rasagiline via a Chemoenzymatic Dynamic Kinetic Resolution. Org. Process Res. Dev. 2014, 18, 1169–1174. [Google Scholar] [CrossRef]
  23. Damljanović, I.; Vukićević, M.; Vukićević, D.R. A Simple Synthesis of Oximes. Monatsh. Chem. 2006, 137, 301–305. [Google Scholar] [CrossRef]
  24. Horáková, E.; Drabina, P.; Brůčková, L.; Štěpánková, Š.; Vorčáková, K.; Sedlák, M. Synthesis and in vitro evaluation of novel N-cycloalkylcarbamates as potential cholinesterase inhibitors. Monatsh. Chem. 2017, 148, 2143–2153. [Google Scholar] [CrossRef]
  25. Palatinus, L.; Chapuis, G. SUPERFLIP—A computer program for the solution of crystal structures by charge flipping in arbitrary dimensions. J. Appl. Crystallogr. 2007, 40, 786–790. [Google Scholar] [CrossRef]
Scheme 1. (a) addition to π-systems, (b) iminyl radicals to C-centered radicals, and (c) synthesis of nitrogen heterocycles.
Scheme 1. (a) addition to π-systems, (b) iminyl radicals to C-centered radicals, and (c) synthesis of nitrogen heterocycles.
Molbank 2022 m1473 sch001
Scheme 2. (a) iminyl radical formation via SET and EnT; (b) iminyl radical from dioxime oxalates; and (c) dioxime oxalate typical synthesis.
Scheme 2. (a) iminyl radical formation via SET and EnT; (b) iminyl radical from dioxime oxalates; and (c) dioxime oxalate typical synthesis.
Molbank 2022 m1473 sch002
Scheme 3. Formation of 2a in MeCN and DCM.
Scheme 3. Formation of 2a in MeCN and DCM.
Molbank 2022 m1473 sch003
Figure 1. (a) Molecular structure of cyclopentanone dioxime oxalate, with anisotropic thermal vibration ellipsoids drawn at the 50% probability level. The hydrogen atoms are shown as spheres of arbitrary radius. (b) Dihedral angle between cyclopentane moieties and dioxime oxalate fragment. (c) Molecular sheets are formed by C-H···O hydrogen interactions. (d) Packing showing dioxime oxalate fragments oriented along the [100] direction.
Figure 1. (a) Molecular structure of cyclopentanone dioxime oxalate, with anisotropic thermal vibration ellipsoids drawn at the 50% probability level. The hydrogen atoms are shown as spheres of arbitrary radius. (b) Dihedral angle between cyclopentane moieties and dioxime oxalate fragment. (c) Molecular sheets are formed by C-H···O hydrogen interactions. (d) Packing showing dioxime oxalate fragments oriented along the [100] direction.
Molbank 2022 m1473 g001
Figure 2. Hirshfeld surface mapped over dnorm and 2D-fingerprint plots, including relative contributions (%) to the Hirshfeld surface area for the various close intermolecular contacts.
Figure 2. Hirshfeld surface mapped over dnorm and 2D-fingerprint plots, including relative contributions (%) to the Hirshfeld surface area for the various close intermolecular contacts.
Molbank 2022 m1473 g002
Table 1. Reaction with other oximes.
Table 1. Reaction with other oximes.
Molbank 2022 m1473 i001
EntryOxime (1)Dioxime Oxalate (2), Yield (%) 1
1Molbank 2022 m1473 i002(2b), 60
2Molbank 2022 m1473 i003(2c), 60 2
3Molbank 2022 m1473 i004(2d), 89
4Molbank 2022 m1473 i005(2e), 94
1 Isolated yield after chromatographic purification. 2 NMR data agreed with previous descriptions.
Table 2. Crystallographic data of cyclopentanone dioxime oxalate.
Table 2. Crystallographic data of cyclopentanone dioxime oxalate.
Crystal Data
Chemical FormulaC12H16N2O4
Mr252.27
Crystalline system, space groupMonoclinic, I2/c
a, b, c (Å)8.9695 (11), 11.2015 (11), 13.3635 (15)
α, β, γ (°)90, 109.087 (13), 90
Volume, (Å3)1268.8 (3)
ρ, kg m−31.321
Z4
Temperature, (K)298 (2)
Radiation typeCu Kα
μ (mm−1)0.84
Theta range for data collection5.278° < 2θ < 76.337°
Index range−11 ≤ h ≤ 11,
−13 ≤ k ≤ 14,
−15 ≤ l ≤ 16
Data collection
DiffractometerSuperNova, Dual, Cu at zero, Atlas
Absorption correctionMulti-scan method CrysAlis PRO 1.171.41.119a (Rigaku Oxford Diffraction, 2021)
Tmin, Tmax0.908, 1.000
No. of measured, independent, and observed reflections [I > 2σ(I)]5200, 1314, 1156
Rint0.034
(sin θ/λ)max (Å−1)0.630
Refinement
R[F2 > 2σ(F2)], wR(F2), S0.048, 0.142, 1.09
No. of reflections1314
Refined parameters83
H-atoms treatmentH-atom parameters constrained
Δρmax, Δρmin (e Å−3)0.23, −0.18
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Adarve-Cardona, L.; Ezenarro-Salcedo, D.; Macías, M.A.; Gamba-Sánchez, D. One-Pot Synthesis of Dioxime Oxalates. Molbank 2022, 2022, M1473. https://doi.org/10.3390/M1473

AMA Style

Adarve-Cardona L, Ezenarro-Salcedo D, Macías MA, Gamba-Sánchez D. One-Pot Synthesis of Dioxime Oxalates. Molbank. 2022; 2022(4):M1473. https://doi.org/10.3390/M1473

Chicago/Turabian Style

Adarve-Cardona, Laura, David Ezenarro-Salcedo, Mario A. Macías, and Diego Gamba-Sánchez. 2022. "One-Pot Synthesis of Dioxime Oxalates" Molbank 2022, no. 4: M1473. https://doi.org/10.3390/M1473

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