Next Article in Journal
Hepatic Pin1 Expression, Particularly in Nuclei, Is Increased in NASH Patients in Accordance with Evidence of the Role of Pin1 in Lipid Accumulation Shown in Hepatoma Cell Lines
Next Article in Special Issue
Asymmetrical Methylene-Bridge Linked Fully Iodinated Azoles as Energetic Biocidal Materials with Improved Thermal Stability
Previous Article in Journal
The High Level of RANKL Improves IκB/p65/Cyclin D1 Expression and Decreases p-Stat5 Expression in Firm Udder of Dairy Goats
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Communication

Enhanced Energetic Performance via the Combination of Furoxan and Oxa-[5,5]bicyclic Structures

1
School of Materials Science & Engineering, Beijing Institute of Technology, Beijing 100081, China
2
Experimental Center of Advanced Materials, School of Materials Science & Engineering, Beijing Institute of Technology, Beijing 100081, China
3
Chongqing Innovation Center, Beijing Institute of Technology, Chongqing 401120, China
*
Authors to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(10), 8846; https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24108846
Submission received: 23 April 2023 / Revised: 8 May 2023 / Accepted: 11 May 2023 / Published: 16 May 2023
(This article belongs to the Special Issue Molecular Research on Energetic Materials)

Abstract

:
Three new compounds based on the combination of furoxan (1,2,5-oxadiazole N-oxide) and oxa-[5,5]bicyclic ring were synthesized. Among them, the nitro compound showed satisfactory detonation properties (Dv, 8565 m s−1; P, 31.9 GPa), which is comparable to the performance of RDX (a classic high-energy secondary explosive). Additionally, the introduction of the N-oxide moiety and oxidation of the amino group more effectively improved the oxygen balance and density (d, 1.81 g cm−3; OB%, +2.8%) of the compounds compared to furazan analogues. Combined with good density and oxygen balance as well as moderate sensitivity, this type of furoxan and oxa-[5,5]bicyclic structure will open up a platform for the synthesis and design of new high-energy materials.

1. Introduction

Furoxan is a typical planar nitrogen heterocyclic skeleton containing N-oxide units [1,2,3,4]. It is commonly used as structural block of high-energy materials due to its “latent nitro” in the ring [5,6]. Benefitting from the dense molecular structure and high nitrogen and oxygen content (69.7%), furoxan could effectively improve the oxygen balance (OB%) and provide higher energy density and prominent enthalpy of formation (198.5 kJ mol−1) as an explosive group for energetic compounds [7]. The furoxan heterocyclic structure has the form of 2-oxide and 5-oxide tautomers, which can undergo tautomerism through the active cis-1,2-dinitrosoethylene intermediate under heating conditions. The existence of tautomers reduces the stability of furoxan skeleton and also causes difficulties in energetic synthesis (Figure 1) [8]. Owing to the good detonation properties, furoxan with attached heterocyclic energetic compounds has been extensively studied in the past decades [9,10]. However, the relatively high sensitivity and poor thermal stability remain constraints on its applications. Common energetic compounds containing furoxan include structures that multiple furoxans directly connect or through the NHCH2NH, N=N (O) bridge connection [11,12,13]. Another kind of structure is mainly based on the combination of the furoxan ring and five-membered heterocycles, such as furoxan connected with isoxazole [14], 1,2,4-oxadiazole [15,16], 1,3,4-oxadiazole [17,18], 1,2,5-oxadiazole [19,20], 1,2,4-triazole [21,22], tetrazole [23,24], etc. (Figure 1). Nevertheless, the synthetic reaction pattern of these compounds is relatively fixed and unitary, generally through the ring-closing reaction of amino-oxime, amino-hydrazone, and hydrazide furoxan with cyanogen bromide. Although successive progress has been made on the direct connection of furoxan with the five-membered ring, a structure connected with the bicyclic structure has not been reported so far.
Recent studies have shown that the introduction of a fused-ring skeleton can effectively increase the enthalpy of formation, thereby further enhancing detonation performance [25,26,27]. In order to simultaneously accomplish the oxygen balance boost, the oxa-[5,5]bicyclic ring is a preeminent candidate structure. Due to the high tension of the skeleton ring, the oxa-[5,5]bicyclic structure can store more chemical energy, giving it great application potential in the area of high-energy-density materials (HEDMs) [28]. Oxa-[5,5]bicyclic rings have a variety of atomic arrangements and may form a large number of different skeleton combinations. Although this class of structure has received more attention in recent years, only four kinds of skeleton structures have been applied in energetic compounds so far. The number of neutral energetic compounds is few, and we lack knowledge of their physical and chemical properties. In 2017, Tang et al. prepared pyrazolo [3,4-c]furazan N-oxide and several nitrogen-rich energetic salts that exhibited excellent detonation properties but had high impact and were friction-sensitive [29]. In 2020, 4H-[1,2,3]triazolo [4,5-c][1,2,5]oxadiazole 5-oxide and its salts, with moderate impact and friction sensitivity, were reported by Churakov [30]. Until 2022, our group has worked on constructing two new oxygen-containing [5,5]-fused skeletons with more modifiable sites, namely the [1,2,4]triazolo [1,5-d][1,2,4]oxadiazole and imidazo [1,2-d][1,2,4]oxadiazole ring system (Figure 1) [31,32]. However, attributing to the instability of N-O and C-N bonds in the oxa-[5,5]bicyclic ring, the two rings may have simultaneous cracking. Meanwhile, due to the limited synthesis methods and the difficulty of energetic derivation, the type and quantity of energetic compounds with the oxa-[5,5]bicyclic structure actually synthesized are very limited.
Based on our previous work, we propose starting from easily available and obtained 4-amino-3-cyano-1,2,5-oxadiazole 2-oxide to realize the synthesis of furoxan connected with oxa-[5,5]bicyclic ring energetic compounds, through conventional cyano functional group conversion. The introduction of the N-oxide moiety and amino oxidation can further improve the oxygen balance and energy density compared to furazan analogues (Figure 1) [33]. Considering that furoxan is relatively sensitive to alkaline conditions and high temperature, there will be problems regarding the stability of target products and the competitive reaction of chloroxime condensation under alkaline conditions in the reaction process. In this work, the oxa-[5,5]bicyclic ring energetic compounds of furoxan were obtained through the ring-closing reaction of amino-protected chloroxime furoxan and 2-chloro-4-nitro-1H-imidazole under alkaline conditions, followed by deprotection of the amino group. The oxidation products and azo products were also successfully prepared, and all compounds were deeply characterized using 1H, 13C, 14N NMR spectroscopy, high-resolution mass spectrometry (HRMS), infrared (IR), and elemental analysis. Among them, amino and nitro structures were clearly identified by X-ray single-crystal diffraction analysis. Overall, this new cyclization mode could construct furoxan and oxa-[5,5]bicyclic structures to enrich their types of energetic compounds by conducting cyclization reaction between furoxan containing a chloroxime group and N-α halogenated azole compounds.

2. Results and Discussion

In Scheme 1, compound 1 was synthesized in good yield by oxidation of o-dioxime by (diacetoxyiodo)benzene [33]. Compound 4 was prepared according to the literature [34]. Then, a methanol suspension of 4 and 1.0 equivalent of 2-chloro-4-nitro-1H-imidazole with 2.0 eq triethylamine was stirred overnight at room temperature to afford target product 5 in 75% yield. In this process, 3,4-bis(4′-dimethylaminomethyleneaminofuroxano-3′)furoxan can be inevitably observed as the byproduct. Then, the amino group in compound 5 can easily split the amidine protection by reacting with the hydrochloric acid aqueous solution to give the amino compound 6. A mixture of 30% H2O2, Na2WO4, and H2SO4 was able to realize the oxidation of the amino group to obtain nitro product 7 in 82% yield. Notably, if the reaction time is prolonged to overnight, the [5,5]-fused ring will break to give compound 7a instead of 7, so the reaction time was controlled to 4 h. The structures of 57 and 7a were clarified by single-crystal X-ray diffraction (CCDC 2237350, 2237351, 2237352, and 2253262; Tables S1–S4). To construct an intramolecular azo bridge compound 8, oxidants trichloroisocyanuric acid (TCICA) and 10% sodium hypochlorite aqueous solution were tested. It was found that when TCICA was the oxidant, the target product could be obtained with a yield of 62%, but the stability of the product in the reaction system was relatively poor, so it required quenching immediately after the reaction finished. In addition, the target compound was difficult to separate and purify due to the presence of impurities in the reaction system. The target product can be observed in the oxidation system of 10% NaClO solution and sodium bicarbonate, but it decomposes quickly. Therefore, the combination of 10% NaClO solution and acetic acid was chosen as the most appropriate reaction conditions, and compound 8 was smoothly prepared with 79% yield.
We also tried to use 4-(chloro(hydroxyimino)methyl)-3-cyano-1,2,5-oxadiazole 2-oxide (9) as a raw material to introduce the cyano group for further derivatization (Scheme 2). Under similar reaction conditions, the target product 10 (CCDC 2252471; Table S5) can be obtained smoothly, but it is necessary to change the solvent to dichloromethane. Then, compound 10 underwent a cycloaddition reaction with sodium azide, but after acidification to pH 1~2 with an appropriate concentrated hydrochloric, the [5,5]-fused ring was degraded to afford compound 11 (CCDC 2252472; Table S6), indicating the instability of the oxa-[5,5]bicyclic structure.
Considering that 1,2,4-triazole has higher density, enthalpy of formation, and better detonation performance than imidazole, the similar ring-closing reaction of compound 4 with 5-bromo-3-nitro-1H-1,2,4-triazole was further attempted (Scheme 3). Fortunately, the target product 5a was successfully obtained with a yield of 65%, and its structure was confirmed by single-crystal X-ray diffraction (CCDC 2245807; Table S7). However, the amino product 6a could not be obtained smoothly during the next amino deprotection process, and only a small amount of the byproduct 4-amino-3-carbamoyl-1,2,5-oxadiazole 2-oxide 6b was obtained (CCDC 2245809; Table S8). In addition to hydrochloric acid aqueous solution, using other reaction conditions such as trifluoroacetic acid or ZnCl2, the target product was not acquired. As for the 4-(chloro(hydroxyimino)methyl)-3-cyano-1,2,5-oxadiazole 2-oxide, unexpectedly, a cyclization reaction occurred to form 1,4,2,5-dioxadiazine bridge-based furoxans instead of the expected reaction with 5-bromo-3-nitro-1H-1,2,4-triazole. We conducted a validation experiment, and under the condition of triethylamine, two molecules of compound 9 were cyclized to obtain 12 (CCDC 2252474; Table S9).
Crystals of 6 and 7 suitable for X-ray diffraction were obtained by slow evaporation of ethyl acetate and petroleum ether solutions at room temperature. Compound 6 crystallizes in the monoclinic space group P21/c with a calculated density of 1.807 g cm−3 at 298 K and a unit cell consisting of four molecules (Z = 4) (Figure 2a). All atoms in 6 are almost coplanar, with the dihedral angles of N5-C3-C2-C1, N7-C6-C5-N5, and N2-C2-C1-N3 as 177.6°, 179.1° and 176.3°, respectively. Compound 6 shows a wave-like stacking in the packing diagram (Figure 2b). Two typical intramolecular hydrogen bonds [N(3)–H(3B)/N(4) and N(12)–H(12B)/N(11)] and three intermolecular hydrogen bonds [N(3)–H(3A)/O(4)a, N(3)–H(3B)/O(9)b and N(12)–H(12A)/O(7)c, symmetry code a: 1 + x, y, z; b: −x, 1/2 + y, 3/2 − z; c: 1 + x, 1/2 − y, −1/2 + z] were observed in the crystal structure of compound 6 (Figure 2c). Compound 7 crystallized in the orthorhombic space group Pbca with eight molecular units (Z = 8) in the unit cell and a calculated density of 1.810 g cm−3 at 298 K (Figure 2d). After the oxidation of the amino group in compound 6 to a nitro group, the furoxan plane in the structure of compound 7 formed an angle of 45.21° with the plane of [5,5]-fused ring (Figure 2e), and compound 7 exhibited a mixed stacking in the packing diagram (Figure 2f). In addition, with the substitution of the nitro for the amino group, compound 7 lost favorable hydrogen bonding interactions due to the lack of hydrogen atoms, leading to an increase in friction and impact sensitivities.
The physical properties of 68 compared with the properties of RDX are summarized in Table 1. Thermal stability determines the adaptability of energetic materials to practical applications and is one of the important physical properties for evaluating their safety. The thermal stabilities of the new compounds were determined by differential scanning calorimetry (DSC) with a heating rate of 5 °C min−1. These compounds exhibited good thermal stability, while azo compound 8 was the least thermally stable compound, decomposing at 158.8 °C. The decomposition temperature (onset) of compound 6 was 163.6 °C, and the thermal stability of the compound 7 was improved, with a decomposition temperature of 170.3 °C. To further investigate the thermal stability of compounds 6 and 7 from the perspective of molecular structure, the multicenter bond orders were calculated by Multiwfn (Figure 3) [35,36]. The results demonstrate that the furoxan ring and oxa-[5,5]bicyclic ring of 7 displays higher multicenter bond orders than that of 6 (6: BM = 0.0093/0.0177/0.0474; 7: BM = 0.0167/0.0201/0.0481), indicating that 7 exhibits stronger aromaticity.
The densities of compounds 68 were measured using a gas pycnometer at 25 °C, and their experimental densities ranged from 1.78 to 1.82 g cm−3. The densities of compounds 6 and 7 were slightly higher than that of classic explosive RDX (1.80 g cm−3). The heat of formation (HOF) values of compounds 68 were calculated by the Gaussian 09 (Revision E.01) program using the isodesmic reactions shown in the ESI (Figure S1, Table S10). The presence of the high enthalpies of furoxan and [5,5]-fused-ring skeletons caused the three prepared compounds to feature high positive HOF, much higher than RDX (70.3 kJ mol−1), ranging from 370.7 kJ mol−1 to 1212.7 kJ mol−1. As shown in Table 1, among these compounds, the azo compound 8 demonstrated the highest positive HOF (1212.7 kJ mol−1), which is attributed to the introduction of the azo group adding more high-energy N=N and C-N bonds.
Together with the measured densities, some important energetic performance measurements, such as the energy and temperature of the explosion, were evaluated by using the EXPLO5 6.05 program [37]. Compounds 68 showed good detonation properties, and compound 7 especially possesses excellent detonation velocity and pressure. The calculated detonation velocity and detonation pressure of nitro product 7 were 8565 m s−1 and 31.9 GPa respectively, which are comparable to RDX (Dv = 8795 m s−1, P = 34.9 GPa), elucidating its potential as an RDX replacement candidate. The heat of detonation and detonation temperature were also evaluated and compared with RDX. Similar to other detonation properties, compound 7 has the highest heat of detonation (5812 kJ kg−1) and temperature of detonation (4490 K), both higher than that of RDX. Impact and friction sensitivities were recorded by BAM fall-hammer and BAM friction-sensitivity tester. Compounds 68 have similar moderate sensitivity, as shown in Table 1. Because of the existing furoxan, it is not surprising that these compounds displayed relatively high sensitivity (IS: 3~5 J, FS: 80~84 N). Additionally, due to the introduction of the oxygen-containing heterocyclic and furoxan skeleton, the nitrogen and oxygen contents of the three compounds were more than 70%, and the oxygen balance of compound 7 was also significantly improved.
The calculated electrostatic potentials (ESP) of the molecular surfaces are closely related to sensitivities [38]. In general, the greater positive ESP values and more extensive electropositive regions (red areas) usually lead to more sensitive energetic compounds. As shown in Figure 4, compound 7 has higher ESP maximum values (+65.80 kcal mol−1 for 7, +57.69 kcal mol−1 for 6) and a larger electrostatic potential difference (103.23 kcal mol−1 for 7, 98.75 kcal mol−1 for 6) compared to compound 6, which agrees with the experimental results.
The two-dimensional (2D) fingerprint and associated Hirshfeld surfaces of compounds 6 and 7 were investigated to gain a further understanding of the intermolecular interactions (Figure 5) [39]. It can be intuitively noted that compound 6 is nearly plate-shaped, with most of the red dots being distributed on the surface edges. More hydrogen bonding interactions (N···H, 9.6%; O···H 24.7%) were observed in compound 6 than 7 (N···H, 2.7%; O···H 10.7%). Meanwhile, the π–π interactions (N···O/O···N, N···N contact) accounted for 29.2% in compound 6 and 34.1% in compound 7, demonstrating that they also play an important role in reducing the sensitivity of the compounds. Other than the favorable contact, the unfavorable O···O contact of compound 7 is evidently higher than that of compound 6, which results in a relatively high sensitivity.
The non-covalent interaction (NCI) plots of compounds 6 and 7 were studied to further investigate their inter- and intra-molecular interactions (Figure 6) [40]. The existence of green isosurfaces in 6 is greater than that in 7, indicating the presence of extensive π–π and hydrogen bonding interactions for compound 6. Therefore, it can be inferred from these calculations that compound 6 had a lower sensitivity than 7.

3. Materials and Methods

All reagents were purchased from commercial sources (Energy Chemical (Energy Chemical, Shanghai, China), Adamas-beta® (Titan, Shanghai, China), J&K Scientific (J&K Scientific, Shanghai, China), Sigma-Aldrich (Merck, Shanghai, China)) and used without purification unless otherwise mentioned. The products were purified by column chromatography over silica gel (200–300 size). 1H, 13C, and 14N nuclear magnetic resonance (NMR) spectra were recorded at 25 °C on a Bruker 400 MHz, 100 MHz; TMS was used as internal standard. Infrared spectra (IR) were obtained on a PerkinElmer Spectrum BX FT-IR instrument equipped with an ATR unit at 25 °C. Elemental analyses of C/H/N were investigated on a Vario EL III Analyzer. High-resolution mass spectra (HRMS) were recorded on Thermo Scientific LTQ Orbitrap XL and Thermo Scientific Q Exactive by using ESI method. The onset decomposition temperature was recorded on TA Discovery DSC 25 or METT-FT 900 at a heating rate of 5 °C min−1 under a dry nitrogen atmosphere. Impact and friction sensitivities were measured with a BAM fall-hammer and friction tester. Densities were calculated at 298K based on crystal structure data and determined at room temperature by employing a Micromeritics AccuPyc II 1345 gas pycnometer. All quantum chemical calculations were carried out using the Gaussian 09 program package and visualized by GaussView 5.0. The geometric optimization and frequency analyses of the structures were carried out using the B3LYP functional with 6-311 + g(d,p) basis set, and single energy points were calculated at the M062X/6-311g(d,p) level. All of the optimized structures were characterized to be true local energy minima on the potential energy surface without imaginary frequencies.
5-Bromo-3-nitro-1H-1,2,4-triazole
Prepared according to the literature [41]. Yield 89%.
2-Chloro-4-nitro-1H-imidazole
Commercially available reagent.
4-Amino-3-cyano-1,2,5-oxadiazole 2-oxide (1)
Prepared according to the literature [33]. Yield 70%.
3-Chlorohydroximoyl-4-dimethylaminomethyleneaminofuroxan (4)
Prepared according to the literature [34]. Yield 87%.
4-(((Dimethylamino)methylene)amino)-3-(6-nitroimidazo [1,2-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide (5): Compound 4 (2.2 mmol, 514 mg) and 2-chloro-4-nitroimidazole (2 mmol, 295 mg) were added to 10 mL MeOH, and the triethylamine (4 mmol, 405 mg) was slowly added. The resulting mixture was stirred overnight at room temperature until the substrate disappeared, and then, the MeOH was removed under reduced pressure. Column chromatography of the residue on silica gel eluting with PE/EA (3:1) gave compound 5 as a pale-yellow solid, yield 450 mg (75%). Tdec: 151.1 °C; 1H NMR (400 MHz, DMSO) δ 8.74 (s, 1H), 8.49 (s, 1H), 3.23 (s, 3H), 3.12 (s, 3H). 13C NMR (100 MHz, DMSO) δ 160.0, 158.0, 157.7, 150.1, 143.3, 112.3, 103.3, 41.2, 35.3. IR (cm−1) 3184, 2925, 1719, 1636, 1610, 1539, 1516, 1434, 1392, 1331, 1294, 1274, 1213, 1127, 1110, 989, 896, 873, 849, 786, 732, 599, 587, 416. Elemental analysis (%) for C9H8N8O5 (308.21): calcd: C 35.07, H 2.62, N 36.36. Found: C 34.65, H 2.45, N 35.67. HRMS (ESI) m/z calcd for C9H9N8O5+ (M+H)+ 309.06904, found 309.06897 (Figures S2, S3 and S23).
4-Amino-3-(6-nitroimidazo [1,2-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide (6): Compound 5 (308 mg, 1.0 mmol) was added to a mixture of concentrated HCl (2 mL) and water (6 mL), and the suspension was stirred at 25 °C for 48 h. The solid was collected by filtration and washed with ice water (10 mL) to give compound 6 as a yellow solid, yield 182 mg (72%). Tdec: 163.6 °C; 1H NMR (400 MHz, DMSO) δ 8.41 (s, 1H), 6.86 (s, 2H). 13C NMR (100 MHz, DMSO) δ 157.9, 155.5, 150.4, 143.2, 111.6, 101.1. IR (cm−1) 3466, 3365, 3164, 2923, 1629, 1590, 1537, 1484, 1366, 1285, 1214, 1046, 978, 963, 857, 804, 789, 767, 745, 721, 708, 648. Elemental analysis (%) for C6H3N7O5 (253.13): calcd: C 28.47, H 1.19, N 38.73. Found: C 28.02, H 1.34, N 38.27. HRMS (ESI) m/z calcd for C6H4N7O5+ (M+H)+ 254.02684, found 254.02690 (Figures S4, S5 and S20).
3-(5-Amino-1,2,4-oxadiazol-3-yl)-4-nitro-1,2,5-oxadiazole 2-oxide (7a): Compound 6 (253 mg, 1.0 mmol) was dispersed in 30% H2O2 (3.6 g), and 98% H2SO4 (4.5 mL) was added dropwise at 0 °C. Then, Na2WO4 (330 mg, 1 mmol) was added in portions, and the reaction mixture was stirred overnight at room temperature. After the reaction was completed by TLC monitoring, the reaction mixture was quenched with 15 mL ice water and extracted with EA. The organic phase was combined and washed with water and then dried with Na2SO4. After the organic solvent was removed, the residue was purified by flash column chromatography on silica gel eluting with PE/EA (4:1) to give compound 7a as a pale-yellow solid, yield 150 mg (70%). Tdec: 176.8 °C; 1H NMR (400 MHz, DMSO) δ 8.48 (s, 1H). 13C NMR (100 MHz, DMSO) δ 173.1, 158.0, 155.6, 103.4. IR (cm−1) 3441, 3140, 1678, 1633, 1572, 1537, 15034, 1386, 1313, 1229, 1097, 1070, 1049, 991, 960, 9078, 847, 791, 760, 665, 464, 418. Elemental analysis (%) for C4H2N6O5 (214.10): calcd: C 22.44, H 0.94, N 39.25. Found: C 22.75, H 1.00, N 38.82 (Figures S6 and S7).
4-Nitro-3-(6-nitroimidazo [1,2-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide (7): Compound 6 (253 mg, 1.0 mmol) was dispersed in 30% H2O2 (3 mL), and 98% concentrated sulfuric acid (3.5 mL) was added dropwise at 0 °C. Then, Na2WO4 (330 mg, 1 mmol) was added in batches, and the reaction mixture was slowly raised to room temperature and stirred for 4 h. After the reaction was completed, the reaction mixture was poured into 15 mL ice water and extracted with ethyl acetate. The organic phase was combined and washed with water and then dried with anhydrous sodium sulfate. After the organic solvent was removed, the solid residue was purified by flash column chromatography on silica gel eluting with petroleum ether/ethyl acetate (3:1) to give compound 7 as a pale-yellow solid, yield 232 mg (82%). Tdec: 170.3 °C; 1H NMR (400 MHz, DMSO) δ 8.74 (s, 1H). 13C NMR (100 MHz, DMSO) δ 158.2, 157.7, 150.7, 140.6, 111.4, 100.1. 14N NMR (40 MHz, DMSO) δ -35.53. IR (cm−1) 3187, 2923, 2853, 1640, 1604, 1587, 1563, 1517, 1450, 1488, 1347, 1310, 1209, 1141, 1092, 1067, 987, 965, 890, 855, 841, 801, 761, 704, 666, 613, 494. Elemental analysis (%) for C6HN7O7 (283.12): calcd: C 25.45, H 0.36, N 34.63. Found: C 25.43, H 0.44, N 33.73. HRMS (ESI) m/z calcd for C6H2N7O7+ (M+H)+ 284.00102, found 284.00015 (Figures S8–S10 and S21).
4,4′-(Diazene-1,2-diyl)bis(3-(6-nitroimidazo [1,2-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide) (8): To a solution of pure compound 6 (253 mg, 1.0 mmol) in acetic acid (2.5 mL), aqueous sodium hypochlorite solution (6–14% available chlorine, 1.2 g) was added dropwise and stirred at 25 °C for 5 min. The reaction solution was quenched with a small amount of water, resulting in solid precipitation. Subsequently, the obtained solid was filtered, washed with water, and dried to obtain a crude product. Then, the solid residue was purified by flash column chromatography on silica gel eluting with petroleum ether/ethyl acetate/AcOH (5:1:0.025) to give compound 8 as a yellow solid, yield 198 mg (79%). Tdec: 158.8 °C; 1H NMR (400 MHz, CD3CN) δ 8.09 (s, 2H). 13C NMR (100 MHz, CD3CN) δ 161.5, 159.1, 152.2, 141.7, 109.9, 99.8. IR (cm−1) 3158, 2924, 1634, 1601, 1539, 1487, 1362, 1278, 1123, 1068, 1005, 968, 900, 884, 855, 792, 777, 755, 714, 644, 611, 552, 510, 450. HRMS (ESI) m/z calcd for C12H3N14O10+ (M+H)+ 503.01511, found 503.01480 (Figures S11, S12 and S22).
4-(Chloro(hydroxyimino)methyl)-3-cyano-1,2,5-oxadiazole 2-oxide (9)
Prepared according to the literature [6]. Yield 74%.
4-Cyano-3-(6-nitroimidazo [1,2-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide (10): 4-(chloro(hydroxyimino)methyl)-3-cyano-1,2,5-oxadiazole 2-oxide (1.1 mmol, 188 mg) and 2-chloro-4-nitroimidazole (1 mmol, 148 mg) were added to 10 mL dichloromethane, and the triethylamine (2 mmol, 203 mg) diluted with 1.5 mL of dichloromethane was slowly added at 25 °C. The resulting mixture was stirred overnight at room temperature until the substrate disappeared, and then, the dichloromethane was removed under reduced pressure. Column chromatography of the residue on silica gel eluting with PE/EA (5:1) gave the compound 10 as a white solid, yield 184 mg (70%). Tdec: 181.8 °C; 1H NMR (400 MHz, DMSO) δ 9.01 (s, 1H). 13C NMR (100 MHz, DMSO) δ 158.3, 151.8, 142.9, 142.4, 110.2, 105.7, 96.7. IR (cm−1) 3137, 2254, 1630, 1612, 1594, 1538, 1493, 1468, 1429, 1366, 1283, 1240, 1105, 1042, 991, 962, 885, 855, 830, 792, 753, 722, 693, 574, 502, 470, 427. HRMS (ESI) m/z calcd for C7N7O5 (M-H) 261.99664, found 261.99609 (Figures S13, S14 and S24).
4-(5-Amino-1,2,4-oxadiazol-3-yl)-3-(1H-tetrazol-5-yl)-1,2,5-oxadiazole 2-oxide (11): Compound 10 (263 mg, 1.0 mmol) was dissolved in water (10 mL), and sodium azide (130 mg, 2.0 mmol) and zinc chloride (204 mg, 1.5 mmol) were added. The solution was heated to 50 °C and stirred for overnight. The reaction mixture was cooled to ambient temperature and acidified with 2 N HCl to pH 1~2. The mixtures were extracted with ethyl acetate. After the combined organic phase was evaporated, the obtained solid was recrystallized to afford compound 11, yield 159 mg (52%). Tdec: 251.0 °C; 1H NMR (400 MHz, DMSO) δ 13.73 (s, 1H), 8.41 (s, 1H). 13C NMR (100 MHz, DMSO) δ 173.2, 170.9, 158.9, 147.2, 145.5, 106.6. IR (cm−1) 3412, 3288, 3211, 3057, 3012, 1785, 1677, 1629, 1522, 1371, 1287, 1220, 1109, 1056, 993, 830, 541. HRMS (ESI) m/z calcd for C5H2N9O3 (M-H) 236.02806, found 236.02824 (Figures S15 and S16).
4-(((Dimethylamino)methylene)amino)-3-(6-nitro-[1,2,4]triazolo [1,5-d][1,2,4]oxadiazol-3-yl)-1,2,5-oxadiazole 2-oxide (5a): Compound 4 (2.2 mmol, 514 mg) and 5-bromo-3-nitro-1H-1,2,4-triazole (2 mmol, 386 mg) were added to 10 mL dichloromethane, and the triethylamine (4 mmol, 405 mg) was slowly added at 25 °C. The resulting mixture was stirred overnight at room temperature until the substrate disappeared, and then, the solvent was removed under reduced pressure. Column chromatography of the residue on silica gel eluting with PE/EA (4:1) gave the compound 5a as a pale-yellow solid, yield 278 mg (45%). Tdec: 116.4 °C; 1H NMR (400 MHz, CD3CN) δ 8.34 (s, 1H), 3.14 (s, 3H), 3.11 (s, 3H). 13C NMR (100 MHz, CD3CN) δ 167.8, 167.3, 160.7, 157.3, 143.6, 102.2, 41.31, 35.1. IR (cm−1) 2930, 1637, 1607, 1587, 1552, 1525, 1434, 1399, 1298, 1268, 1112, 1062, 964, 870, 848, 837, 753, 737, 633, 591, 473. HRMS (ESI) m/z calcd for C8H8N9O5+ (M+H)+ 310.06429, found 310.06470 (Figures S17, S18 and S25).
4,4′-(1,4,2,5-Dioxadiazine-3,6-diyl)bis(3-cyano-1,2,5-oxadiazole 2-oxide) (12): 3-(chloro(hydroxyimino)methyl)-4-cyano-1,2,5-oxadiazole 2-oxide (1.1 mmol, 188 mg) were added to 10 mL dichloromethane, and the triethylamine (2 mmol, 203 mg) diluted with 1.5 mL of dichloromethane was slowly added at 25 °C. The resulting mixture was stirred overnight at room temperature until the substrate disappeared, and then, the solvent was removed under reduced pressure. Column chromatography of the residue on silica gel eluting with PE/EA (10:1) gave compound 12 as a pale-yellow solid, yield 94 mg (62%). Tdec: 170.4 °C; 13C NMR (100 MHz, DMSO) δ 152.1, 145.1, 105.9, 97.2. IR (cm−1) 2923, 2852, 2258, 1747, 1626, 1522, 1483, 1347, 1228, 1046, 1025, 1006, 964, 896, 838, 739, 633, 577, 538, 507, 439, 406. HRMS (ESI) m/z calcd for C8HN8O6+ (M+H)+ 305.00190, found 305.00146 (Figures S19 and S26).

4. Conclusions

In summary, three newly designed energetic compounds based on the combination of furoxan and oxa-[5,5]bicyclic ring were synthesized and deeply characterized. Compound 7 possesses a high density of 1.810 g cm−3 and is thermally stable up to 170.3 °C. Because of the introduction of nitrofuroxan, the synthesized nitro compound 7 exhibits good oxygen balances and detonation properties (OB, 2.83%; Dv, 8565 m s−1; P, 31.9 GPa) comparable to that of RDX. These novel structures represent the first examples of a combination of the furoxan and oxa-[5,5]bicyclic structures, which would open a new avenue for the construction of new high-energy-density material.

Supplementary Materials

The following supporting information can be downloaded at: https://0-www-mdpi-com.brum.beds.ac.uk/article/10.3390/ijms24108846/s1.

Author Contributions

Conceptualization, Q.Z. and C.H.; methodology, Q.Z.; formal analysis, Q.Z. and X.Z.; writing—original draft preparation, Q.Z.; writing—review and editing, C.H. and S.P.; supervision, C.H. and S.P.; funding acquisition, C.H. and S.P. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported by the National Natural Science Foundation of China (No. 22235003).

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.

References

  1. Pasinszki, T.; Havasi, B.; Hajgató, B.; Westwood, N.P.C. Synthesis, spectroscopy and structure of the parent furoxan (HCNO)2. J. Phys. Chem. A 2009, 113, 170–176. [Google Scholar] [CrossRef]
  2. Larin, A.A.; Fershtat, L.L. Energetic heterocyclic N-oxides: Synthesis and performance. Mendeleev Commun. 2022, 32, 703–713. [Google Scholar] [CrossRef]
  3. Sheremetev, A.B.; Makhova, N.N.; Friedrichsen, W. Monocyclic furazans and furoxans. Adv. Heterocyclic Chem. 2001, 78, 65–188. [Google Scholar]
  4. Stephens, D.E.; Larionov, O.V. Energetic Heterocyclic N-Oxides. Top. Heterocycl. Chem. 2017, 53, 1–28. [Google Scholar]
  5. Makhova, N.N.; Kulikov, A.S. Advances in the chemistry of monocyclic amino- and nitrofuroxans. Russ. Chem. Rev. 2013, 82, 1007–1033. [Google Scholar] [CrossRef]
  6. Liu, Y.; He, C.; Tang, Y.; Imler, G.H.; Parrish, D.A.; Shreeve, J.M. Asymmetric nitrogen-rich energetic materials resulting from the combination of tetrazolyl, dinitromethyl and (1,2,4-oxadiazol-5-yl)nitroamino groups with furoxan. Dalton Trans. 2018, 47, 16558–16566. [Google Scholar] [CrossRef] [PubMed]
  7. Larin, A.A.; Shaferov, A.V.; Epishina, M.A.; Melnikov, I.N.; Muravyev, N.V.; Ananyev, I.V.; Fershtat, L.L.; Makhova, N.N. Pushing the energy-sensitivity balance with high-performance bifuroxans. ACS Appl. Energy Mater. 2020, 3, 7764–7771. [Google Scholar] [CrossRef]
  8. Zhang, J.; Zhou, J.; Bi, F.; Wang, B. Energetic materials based on poly furazan and furoxan structures. Chin. Chem. Lett. 2020, 31, 2375–2394. [Google Scholar] [CrossRef]
  9. Makhova, N.N.; Fershtat, L.L. Recent advances in the synthesis and functionalization of 1,2,5-oxadiazole 2-oxides. Tetrahedron Lett. 2018, 59, 2317–2326. [Google Scholar] [CrossRef]
  10. Wang, L.; Zhai, L.; She, W.; Wang, M.; Zhang, J.; Wang, B. Synthetic strategies toward nitrogen-rich energetic compounds via the reaction characteristics of cyanofurazan/furoxan. Front. Chem. 2022, 10, 871684–871706. [Google Scholar] [CrossRef] [PubMed]
  11. Finogenov, A.O.; Epishina, M.A.; Kulikov, A.S.; Makhova, N.N.; Anan’ev, I.V.; Tartakovsky, V.A. Synthesis and nitration of N,N′-bis(3-R-furoxan-4-yl)methylenediamines. Russ. Chem. Bull. 2010, 59, 2108–2113. [Google Scholar] [CrossRef]
  12. Chen, P.; Dou, H.; Zhang, J.; He, C.; Pang, S. Trinitromethyl Energetic Groups Enhance High Heats of Detonation. ACS Appl. Mater. Interfaces 2023, 15, 4144–4151. [Google Scholar] [CrossRef] [PubMed]
  13. Larin, A.A.; Bystrov, D.M.; Fershtat, L.L.; Konnov, A.A.; Makhova, N.N.; Monogarov, K.A.; Meerov, D.B.; Melnikov, I.N.; Pivkina, A.N.; Kiselev, V.G.; et al. Nitro-, cyano-, and methylfuroxans, and their bis-derivatives: From green primary to melt-cast explosives. Molecules 2020, 25, 5836. [Google Scholar] [CrossRef]
  14. Johnson, E.C.; Sabatini, J.J.; Chavez, D.E.; Wells, L.A.; Banning, J.E.; Sausa, R.C.; Byrd, E.F.C.; Orlicki, J.A. Bis(nitroxymethylisoxazolyl) furoxan: A promising standalone melt-castable explosive. Chempluschem 2020, 85, 237–239. [Google Scholar] [CrossRef]
  15. Johnson, E.C.; Bukowski, E.J.; Sabatini, J.J.; Sausa, R.C.; Byrd, E.F.C.; Garner, M.A.; Chavez, D.E. Bis(1,2,4-oxadiazolyl) furoxan: A promising melt-castable eutectic material of low sensitivity. Chempluschem 2019, 84, 319–322. [Google Scholar] [CrossRef]
  16. Hermann, T.S.; Klapötke, T.M.; Krumm, B.; Stierstorfer, J. Synthesis, characterization and properties of ureido-furazan derivatives. J. Heterocycl. Chem. 2018, 55, 852–862. [Google Scholar] [CrossRef]
  17. Finogenov, A.O.; Kulikov, A.S.; Epishina, M.A.; Ovchinnikov, I.V.; Nelyubina, Y.V.; Makhova, N.N. The first synthesis of furoxan and 1,3,4-oxadiazole ring ensembles. J. Heterocycl. Chem. 2013, 50, 135–140. [Google Scholar] [CrossRef]
  18. Fershtat, L.L.; Kulikov, A.S.; Ananyev, I.V.; Struchkova, M.I.; Makhova, N.N. New method for the synthesis and reactivity of (5-R-1,3,4-Oxadiazol-2-yl)furoxans. J. Heterocycl. Chem. 2016, 53, 102–108. [Google Scholar] [CrossRef]
  19. Stepanov, A.I.; Astrat’ev, A.A.; Dashko, D.V.; Spiridonova, N.P.; Mel’nikova, S.F.; Tselinskii, I.V. Synthesis of linear and cyclic compounds containing the 3,4-bis(furazan-3-yl)furoxan fragment. Russ. Chem. Bull. 2012, 61, 1024–1040. [Google Scholar] [CrossRef]
  20. Yu, Q.; Chinnam, A.K.; Yin, P.; Imler, G.H.; Parrish, D.A.; Shreeve, J.M. Finding furoxan rings. J. Mater. Chem. A 2020, 8, 5859–5864. [Google Scholar] [CrossRef]
  21. Larin, A.A.; Ananyev, I.V.; Dubasova, E.V.; Teslenko, F.E.; Monogarov, K.A.; Khakimov, D.V.; He, C.-L.; Pang, S.-P.; Gazieva, G.A.; Fershtat, L.L. Simple and energetic: Novel combination of furoxan and 1,2,4-triazole rings in the synthesis of energetic materials. Energ. Mater. Front. 2022, 3, 146–153. [Google Scholar] [CrossRef]
  22. Larin, A.A.; Pivkina, A.N.; Ananyev, I.V.; Khakimov, D.V.; Fershtat, L.L. Novel family of nitrogen-rich energetic (1,2,4-triazolyl) furoxan salts with balanced performance. Front. Chem. 2022, 10, 1012605–1012616. [Google Scholar] [CrossRef] [PubMed]
  23. Larin, A.A.; Muravyev, N.V.; Pivkina, A.N.; Suponitsky, K.Y.; Ananyev, I.V.; Khakimov, D.V.; Fershtat, L.L.; Makhova, N.N. Assembly of tetrazolylfuroxan organic salts: Multipurpose green energetic materials with high enthalpies of formation and excellent detonation performance. Chem. Eur. J. 2019, 25, 4225–4233. [Google Scholar] [CrossRef] [PubMed]
  24. Fershtat, L.L.; Epishina, M.A.; Kulikov, A.S.; Ovchinnikov, I.V.; Ananyev, I.V.; Makhova, N.N. An efficient access to (1H-tetrazol-5-yl)furoxan ammonium salts via a two-step dehydration/[3+2]-cycloaddition strategy. Tetrahedron 2015, 71, 6764–6775. [Google Scholar] [CrossRef]
  25. Chinnam, A.K.; Tang, Y.; Staples, R.J.; Shreeve, J.M. Effects of nitric acid concentration for nitration of fused [1,2,5]oxadiazolo[3,4-d]pyrimidine-5,7-diamine. Dalton Trans. 2022, 51, 17987–17993. [Google Scholar] [CrossRef] [PubMed]
  26. Rudakov, G.F.; Sinditskii, V.P.; Andreeva, I.A.; Botnikova, A.I.; Veselkina, P.R.; Kostanyan, S.K.; Yudin, N.V.; Serushkin, V.V.; Cherkaev, G.V.; Dorofeeva, O.V. Energetic compounds based on a new fused Bis[1,2,4]Triazolo[1,5-b;5′,1′-f]-1,2,4,5-Tetrazine. Chem. Eng. J. 2022, 450, 138073–138086. [Google Scholar] [CrossRef]
  27. Banik, S.; Ghule, V.D.; Dharavath, S. Synthesis, characterization, testing, and detonation performance studies of fused pyrazole-based fluorescent energetic materials. Mater. Chem. Phys. 2023, 301, 127678–127686. [Google Scholar] [CrossRef]
  28. Sheremetev, A.B. Chemistry of furazans fused to five-membered rings. J. Heterocycl. Chem. 1995, 32, 371–385. [Google Scholar] [CrossRef]
  29. Tang, Y.; He, C.; Shreeve, J.M. A furazan-fused pyrazole N-oxide via unusual cyclization. J. Mater. Chem. A 2017, 5, 4314–4319. [Google Scholar] [CrossRef]
  30. Voronin, A.A.; Fedyanin, I.V.; Churakov, A.M.; Pivkina, A.N.; Muravyev, N.V.; Strelenko, Y.A.; Klenov, M.S.; Lempert, D.B.; Tartakovsky, V.A. 4H-[1,2,3]Triazolo[4,5-c][1,2,5]oxadiazole 5-oxide and its salts: Promising multipurpose energetic materials. ACS Appl. Energy Mater. 2020, 3, 9401–9407. [Google Scholar] [CrossRef]
  31. Dou, H.; Chen, P.; Hu, L.; He, C.; Pang, S. [1,2,4]Triazolo[1,5-d][1,2,4]oxadiazole ring system-a novel building block for creating energetic compounds. Chem. Eng. J. 2022, 444, 136708–136713. [Google Scholar] [CrossRef]
  32. Dou, H.; Chen, P.; He, C.-L.; Pang, S.-P. Synthesis and characterization of oxygen-containing imidazo[1,2-d][1,2,4]oxadiazole fused-ring energetic compounds. Energ. Mater. Front. 2022, 3, 154–160. [Google Scholar] [CrossRef]
  33. Zhang, Q.; Zhao, C.; Zhang, X.; He, C.; Pang, S. Oxidation of o-dioxime by (diacetoxyiodo)benzene: Green and mild access to furoxans. New J. Chem. 2022, 46, 1489–1493. [Google Scholar] [CrossRef]
  34. He, C.; Gao, H.; Imler, G.H.; Parrish, D.A.; Shreeve, J.M. Boosting energetic performance by trimerizing furoxan. J. Mater. Chem. A 2018, 6, 9391–9396. [Google Scholar] [CrossRef]
  35. Lu, T.; Chen, F. Multiwfn: A multifunctional wavefunction analyzer. J. Comput. Chem. 2012, 33, 580–592. [Google Scholar] [CrossRef] [PubMed]
  36. Lu, T.; Chen, Q. A simple method of identifying π orbitals for non-planar systems and a protocol of studying π electronic structure. Theor. Chem. Acc. 2020, 139, 25–36. [Google Scholar] [CrossRef]
  37. Sućeska, M. EXPLO5 6.05; Brodarski Institute: Zagreb, Croatia, 2020. [Google Scholar]
  38. Rice, B.M.; Hare, J.J. A Quantum mechanical investigation of the relation between impact sensitivity and the charge distribution in energetic molecules. J. Phys. Chem. A 2002, 106, 1770–1783. [Google Scholar] [CrossRef]
  39. Spackman, M.A.; Jayatilaka, D. Hirshfeld surface analysis. CrystEngComm 2009, 11, 19–32. [Google Scholar] [CrossRef]
  40. Johnson, E.R.; Keinan, S.; Mori-Sánchez, P.; Contreras-García, J.; Cohen, A.J.; Yang, W. Revealing noncovalent interactions. J. Am. Chem. Soc. 2010, 132, 6498–6506. [Google Scholar] [CrossRef]
  41. Witkowski, J.T.; Robins, R.K. Chemical synthesis of the 1,2,4-triazole nucleosides related to uridine, 2′-deoxyuridine, thymidine, and cytidine. J. Org. Chem. 1970, 35, 2635–2641. [Google Scholar] [CrossRef]
Figure 1. Furoxan and oxa-[5,5]bicyclic skeleton in energetic compounds.
Figure 1. Furoxan and oxa-[5,5]bicyclic skeleton in energetic compounds.
Ijms 24 08846 g001
Scheme 1. Synthetic routes for preparing compounds 68.
Scheme 1. Synthetic routes for preparing compounds 68.
Ijms 24 08846 sch001
Scheme 2. Synthetic routes of compounds 1011.
Scheme 2. Synthetic routes of compounds 1011.
Ijms 24 08846 sch002
Scheme 3. Synthetic procedures for preparing compounds 5a and 12.
Scheme 3. Synthetic procedures for preparing compounds 5a and 12.
Ijms 24 08846 sch003
Figure 2. (ac) X-ray single-crystal diffraction and crystal packing of compound 6; (df) X-ray single-crystal diffraction and crystal packing of compound 7.
Figure 2. (ac) X-ray single-crystal diffraction and crystal packing of compound 6; (df) X-ray single-crystal diffraction and crystal packing of compound 7.
Ijms 24 08846 g002
Figure 3. (a) The multicenter bond orders of 6; (b) the multicenter bond orders of 7.
Figure 3. (a) The multicenter bond orders of 6; (b) the multicenter bond orders of 7.
Ijms 24 08846 g003
Figure 4. (a) Electrostatic potential surfaces (ESP) for compound 6; (b) electrostatic potential surfaces (ESP) for compound 7. The blue and red spheres in the diagram represent the surface local minima and maximum of ESP, respectively.
Figure 4. (a) Electrostatic potential surfaces (ESP) for compound 6; (b) electrostatic potential surfaces (ESP) for compound 7. The blue and red spheres in the diagram represent the surface local minima and maximum of ESP, respectively.
Ijms 24 08846 g004
Figure 5. Hirshfeld surfaces (inside) and pie graph showing the percentage contributions of the individual atomic contacts to the Hirshfeld surfaces of 6 (a) and 7 (c). Fingerprint plots of 6 (b) and 7 (d).
Figure 5. Hirshfeld surfaces (inside) and pie graph showing the percentage contributions of the individual atomic contacts to the Hirshfeld surfaces of 6 (a) and 7 (c). Fingerprint plots of 6 (b) and 7 (d).
Ijms 24 08846 g005
Figure 6. Noncovalent interaction (NCI) analysis for compounds 6 (a) and 7 (b).
Figure 6. Noncovalent interaction (NCI) analysis for compounds 6 (a) and 7 (b).
Ijms 24 08846 g006
Table 1. Physicochemical and energetic properties of compounds 6, 7, and 8.
Table 1. Physicochemical and energetic properties of compounds 6, 7, and 8.
CompoundTd a
(°C)
d b
(g cm−3)
ΔHf c
(kJ mol−1)
D d
(m s−1)
P e
(Gpa)
Q f
(kJ kg−1)
Detonation Temperature g (K)IS h
(J)
FS i
(N)
OB j
(%)
[N+O] k
(%)
6163.61.807370.7796026.148713597580−15.870.3
7170.31.810454.9856531.9581244903842.8374.2
8158.81.7891212.7824128.655744292384−9.5670.9
RDX2041.80670.3879534.9574237397.5120081.1
a Decomposition temperature from DSC (5 °C min−1). b Crystal density (298K). c Calculated molar enthalpy of formation in solid state. d Detonation velocity calculated with EXPLO5 V6.05. e Detonation pressure calculated with EXPLO5 V6.05. f Heat of detonation calculated with EXPLO5 V6.05. g Detonation temperature calculated with EXPLO5 V6.05. h Impact sensitivity. i Friction sensitivity. j Oxygen balances based on CO as the product for CaHbOcNd, ΩCO (%) = 1600 (c–a–b/2)/Mw; Mw = molecular weight. k Nitrogen–oxygen content.
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

Zhang, Q.; Zhang, X.; Pang, S.; He, C. Enhanced Energetic Performance via the Combination of Furoxan and Oxa-[5,5]bicyclic Structures. Int. J. Mol. Sci. 2023, 24, 8846. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24108846

AMA Style

Zhang Q, Zhang X, Pang S, He C. Enhanced Energetic Performance via the Combination of Furoxan and Oxa-[5,5]bicyclic Structures. International Journal of Molecular Sciences. 2023; 24(10):8846. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24108846

Chicago/Turabian Style

Zhang, Qi, Xun Zhang, Siping Pang, and Chunlin He. 2023. "Enhanced Energetic Performance via the Combination of Furoxan and Oxa-[5,5]bicyclic Structures" International Journal of Molecular Sciences 24, no. 10: 8846. https://0-doi-org.brum.beds.ac.uk/10.3390/ijms24108846

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