CHINESE JOURNAL OF ENERGETIC MATERIALS
+高级检索
参考文献 1
ZHANGQing‑hua, HEChun‑lin, YINPing, et al. Insensitive nitrogen‑rich materials incorporating the nitroguanidyl functionality[J]. Chemistry-an Asian Journal, 2014, 9(1): 212-217.
参考文献 2
GAOHai‑xiang, ShreeveJ M. Azole‑​based energetic salts[J]. Chemical. Reviews, 2011, 111(11): 7377-7436.
参考文献 3
WANGRui‑hu, XUHong‑yan, GUOYong, et al. Bis[3‑(5‑nitroimino‑1,2,4‑triazolate)]‑based energetic salts: synthesis and promising properties of a new family of high‑density insensitive materials[J]. Journal of the American Chemical Society, 2010, 132(34): 11904-11905.
参考文献 4
WANGKai, ParrishD A, ShreeveJ M. 3‑Azido‑N‑nitro‑1H‑1,2,4‑triazol‑5‑amine‑based energetic salts[J]. Chemistry‑A European Journal, 2011, 17(51): 14485-14492
参考文献 5
ZHANGJia‑heng, DharavathS, MitchellL A, et al. Energetic salts based on 3,5‑bis(dinitromethyl)‑1,2,4‑triazole monoanion and dianion: controllable preparation, characterization, and high performance[J]. Journal of the American Chemical Society, 2016, 138(24): 7500-7503
参考文献 6
KlapötkeT M, MayrN, StierstorferJ, et al. Maximum compaction of ionic organic explosives: bis(hydroxylammonium) 5,​5'‑dinitromethyl‑3,3′‑bis(1,2,4‑oxadiazolate) and its derivatives[J]. Chemistry‑A European Journal, 2014, 20(5): 1410-1417.
参考文献 7
DENGMu‑chong, ZHANGQing‑hua, WANGKang‑cai, et al. Synthesis and properties of 5,10‑bis(dinitromethyl)‑furazan[3,4‑e]bis([1,2,4]triazolo)[4,3‑a:3′,4′‑c] pyrazine and its energetic ion compounds[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2018, 26(2):144-149.
参考文献 8
LIYa‑nan, SHUYuan‑jie, ZHANGSheng‑yong, et al. Synthesis and thermal Properties of 4,4′,5,5′‑Tetranitro‑2,2′‑biimidazole and its energetic ion salts[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2017, 25(4): 298-303.
参考文献 9
WANGRui‑hu, GUOYong, ZENGZhuo, et al. Furazan‑functionalized tetrazolate‑based salts: a new family of insensitive energetic materials[J]. Chemistry‑A European Journal, 2009, 15(11): 2625-2634.
参考文献 10
WANGRui‑hu, GUOYong, Rong‑jianSA, et al. Nitroguanidine‑fused bicyclic guanidinium salts: a family of high‑density energetic materials[J]. Chemistry‑A European Journal, 2010, 16(28): 8522-8529
参考文献 11
TANGYong‑xing, ZHANGJia‑heng, MitchellL A, et al. Taming of 3,4‑di(nitramino)​furazan[J]. Journal of the American Chemical Society, 2015, 137(51):15984-15987
参考文献 12
TANGYong‑xing, GAOHai‑xiang, MitchellL A, et al. Syntheses and promising properties of dense energetic 5,5′‑dinitramino‑3,3'‑azo‑1,2,4‑oxadiazole and its salts[J]. Angewandte Chemie International Edition, 2016, 55(9): 3200-3203.
参考文献 13
HAOWei, HEChun‑lin, ZHANGJia‑heng, et al. Combination of 1,2,4‑​oxadiazole and 1,2,5‑oxadiazole moieties for the generation of high‑​performance energetic materials[J]. Angewandte Chemie International Edition, 2015, 54(32): 9367-9371.
参考文献 14
HEChun‑lin, TANGYong‑xing, MitchellL A, et al. N‑​oxides light up energetic performances: synthesis and characterization of dinitraminobisfuroxans and their salts[J]. Journal of Materials Chemistry A, 2016, 4(23): 8969-8973.
参考文献 15
WANGZuo‑quan, ZHANGHong, KillianB J, et al. Synthesis, characterization and energetic properties of 1,​3,​4‑​oxadiazoles[J]. European Journal of Organic Chemistry, 2015, 2015(23): 5183-5188.
参考文献 16
WUQiong, ZHUWei‑hua, XIAOHe‑ming. Molecular design of trinitromethyl‑substituted nitrogen‑rich heterocycle derivatives with good oxygen balance as high‑energy density compounds[J]. Structural Chemistry, 2013, 24(5):1725-1736.
参考文献 17
TIANJia‑wei, XIONGHua‑lin, LINQiu‑han, et al. Energetic compounds featuring bi(1,3,4‑oxadiazole)​: a new family of insensitive energetic materials[J]. New Journal of Chemistry, 2017, 41(5): 1918-1924.
参考文献 18
YUQiong, YINPing, ZHANGJia‑heng, et al. Pushing the limits of oxygen balance in 1,3,4‑oxadiazoles[J]. Journal of the American Chemical Society, 2017, 139(26): 8816-8819.
参考文献 19
HermannT S, KaraghiosoffK, KlapötkeT M, et al. Synthesis and characterization of 2,2′‑dinitramino‑5,5′‑bi(1‑oxa‑3,4‑diazole) and derivatives as economic and highly dense energetic materials[J]. Chemistry‑A European Journal, 2017, 23(50): 12092-12095.
参考文献 20
ZHANGWen‑quan, ZHANGJia‑heng, DENGMu‑chong, et al. A promising high‑energy‑density material[J]. Nature Communication, 2017, 8(1): 181-187.
参考文献 21
HariharanP C, PopleJ A. Influence of polarization functions on MO hydrogenation energies[J]. Theoretica Chimica Acta, 1973, 23(8): 213-216.
参考文献 22
JenkinsH D B, TudealD, GlasserL. Lattice potential energy estimation for complex ionic salts from density measurements[J]. Inorganic Chemistry, 2002, 41(9): 2364-2367.
参考文献 23
KlapötkeT M, PengerA, PflügerC, et al. Advanced open‑chain nitramines as energetic materials: heterocyclic‑substituted 1,3‑dichloro‑2‑nitrazapropane[J]. European Journal of Inorganic Chemistry, 2013, 2013(26): 4667-4678.

    Abstract

    Using 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazole) as starting material, a serials of energetic salts were synthesized. All compounds were characterized by FT‑IR, multinuclear NMR spectroscopy and elemental analysis. The structures of di(3‑amino‑1,2,4‑triazolium) 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate)·2H2O (9·2H2O) and di(4‑amino‑1,2,4‑triazolium)5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate)(10) were further confirmed by single crystal X‑ray diffraction. Their thermal stabilities were determined by differential scanning calorimetry (DSC). The detonation performance was calculated with Explo5 v6.02 software. The results show that the temperature of the thermal decomposition reaction is in the range of 146.8-239.9 ℃. The calculated detonation velocity and pressure are higher than 7693 m·s-1 and 21.3 GPa, respectively. Their densities are from 1.683 to 1.941 g·cm-3. The measured impact sensitivities are between 10 J and 28 J and friction sensitivities are between 160 N and 360 N, which indicated that most salts of 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazole) are high energy‑density materials with good properties.

    摘要

    以5,5′‑二硝胺基‑2,2′‑联‑1,3,4‑噁二唑为原料合成了一系列含能盐,采用了红外(FT‑IR))、核磁(NMR)和元素分析进行了结构表征。并用X‑射线单晶衍射进一步确定了3‑氨基‑1,2,4‑三唑盐(9·2H2O)和4‑氨基‑1,2,4‑三唑盐(10)的结构,用差热扫描法(DSC)测定了它们的热分解温度,用Explos 5 v6.02计算了它们的爆轰性能。结果表明它们的热分解温度范围为146.8~239.9 ℃;计算爆速高于7693 m·s-1,爆压高于21.3 GPa;密度介于1.683~1.941 g·cm-3,实测撞击感度介于10~28 J,摩擦感度介于160~360 N, 表明5,5′‑二硝胺基‑2,2′‑联‑1,3,4‑噁二唑类含能盐是一类性能较好的高能量密度材料。

  • 1 Introduction

    1

    Over the past decade, the synthesis and development of new high‑energy density materials(HEDMs) have attracted increasing attention in the world[1,2] . New HEDMs not only focus on high density, high detonation velocity and pressure but also aim at the high positive heat of formation, high thermal stability, low sensitivity towards external forces such as impact and friction and environmental compatibility. However, among these characteristics, high performance and low sensitivity tend to be contradicting aspects, making the design and synthesis of energetic materials an interesting and challenging work[3,4] . Heterocyclic energetic salts with high nitrogen content is one of the favorite topics in the search for high‑performance energetic materials recently[5,6,7,8]. The energetic salts possess advantages over nonionic molecules since these salts tend to exhibit lower vapor pressures and higher thermal stability than their atomically similar nonionic analogues. This provides an efficient methodology for the design and synthesis of HEDMs[9,10].

    Oxadiazole is a favorite example of a nitrogen‑and oxygen‑rich ring that has been wildly used in the syntheses of drugs, ionic liquids and scintillators. The energetic materials based on 1,2,5‑oxadiazole (furazan) have been exhaustively investigated[11,12,13,14]. Compared to furazan, 1,3,4‑oxadiazoles have been studied very rarely as energetic materials. The energetic materials based on 1,3,4‑oxadiazole have become more and more attractive due to their higher thermal stability and lower sensitivity[15,16]. For example, di(4‑amino‑1,2,4‑triazolium) 5,5′‑dinitromethyl‑2,2′‑bis(1,3,4‑oxadiazolate) (A)[17] shows a detonation velocity of 8601 m·s-1 and a detonation pressure of 31.8 GPa; hydrazinium 2,5‑bis(trinitromethyl)‑1,3,4‑oxadiazolate (B)[16], has a high density of 1.90 g·cm-3 and excellent detonation performance (D=8900 m·s‑1, p=36.3 GPa). And both of them have good thermal stabilities (A: 205.8 ℃; B: 190 ℃) and acceptable sensitivity towards impact and friction (A: IS: 16 J, FS: 200 N; B: IS: 19 J, FS: 80 N).

    In our effort to seek more powerful, eco‑friendly and less sensitive explosives, we were interested in 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazole), because of its high thermal stability (decomposition temperature 210 ℃), high density (1.99 g·cm-3) and excellent explosive properties (D=9481 m·s-1, p=41.9 GPa)[18,19,20]. However, energetic salts based on 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazole) have not been fully investigated. Herein, we report the synthesis and characterization of energetic salts based on 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazole). All the salts are well characterized by IR and multinuclear NMR spectroscopy, differential scanning calorimetry (DSC) and elemental analysis, and some are further measured by single crystal X‑ray diffraction. Their key detonation performance and sensitivity towards impact and friction were determined by experimental and theoretical methods.

  • 2 Experiments

    2
  • 2.1 General Methods

    2.1

    1H and 13C spectra were recorded on a 500 MHz (Bruker AVANCE 500) nuclear magnetic resonance spectrometer operating at 500 and 125 MHz. The chemical shifts were given relative to external tetramethylsilane. IR spectra were recorded by using the attenuated total reflection mode for solids on a Thermo Scientific Nicolet is 10 spectrometer. The decomposition temperatures were determined on a differential scanning calorimeter (DSC823e instrument) at a heating rate of 5 ℃·min-1. Elemental analysis was performed with a Vario EL III instrument. Sensitivity towards impact and friction was determined using an HGZ‑1 drop hammer and a BAM friction tester. The densities of the compounds were determined at room temperature by employing a gas pycnometer.

  • 2.2 Synthetic Route

    2.2

    Synthesis of energetic salts based on 5,5′‑dinitroamino‑2,2'‑bi(1,3,4‑oxadiazole) are shown in Scheme 1.

    Scheme 1 Synthesis of energetic salts based on 5,5′‑dinitroamino‑2,2'‑bi(1,3,4‑oxadiazole)

    The starting material 5,5′‑diamino‑2,2′‑bi(1,3,4‑oxadiazole)(1) was prepared by treating oxalyl dihydrazide with cyanogen bromide in anhydrous ethanol for 24 h. Nitramino substituted compound 2 was obtained by direct nitration of 5,5′‑diamino‑2,2'‑bi(1,3,4‑oxadiazole) using 100% nitric acid at 0 ℃. The energetic salt derivatives of compound 2 were readily synthesized by the acid‑base neutralization reaction in which compound 2 reacted with such as ammonia, hydroxylamine, hydrazine, guanidine, aminoguanidine, 1,2,4‑triazole, 3‑amino‑1,2,4‑triazole, 4‑amino‑1,2,4‑triazole and carbohydrazide.

    General procedures for compounds 3-11: NH3·H2O(2.0 mmol), hydroxylamine solution (2.0 mmol), hydrazine hydrate (2.0 mmol), guanidine carbonate(1.0 mmol), aminoguanidine bicarbonate(2.0 mmol), 1,2,4‑triazole(2.0 mmol), 3‑amino‑1,2,4‑triazole(2.0 mmol) and 4‑amino‑1,2,4‑triazole(2.0 mmol), carbohydrazide(2.0 mmol) were, respectively, added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield.

  • 2.3 Characterization

    2.3

    Ammonium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate)(3)

    NH3·H2O(2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.237 g of compound 3 was obtained as a yellow solid in a yield of 81.1%. 1H NMR(500 MHz, DMSO‑d 6)δ:7.17 (s, 8H). 13C NMR(125 MHz, DMSO‑d 6)δ: 166.15 (s), 148.32 (s). IR (KBr, ν/cm-1): 3448, 3372, 3216, 3050, 1524, 1488, 1447, 1424, 1430, 1311, 1279, 1150, 1069, 1022, 963, 728. Anal. calcd for (%) for C4H8N10O6(292.17): C 16.44, H 2.76, N 47.94; found: C 16.52, H 2.53, N 47.52.

    Hydroxylammonium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑ oxadiazolate)(4)

    Hydroxylamine solution (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.279 g of compound 4 was obtained as a yellow solid in a yield of 86.1%. 1H NMR (500 MHz, DMSO‑d 6)δ: 7.09. 13C NMR (125 MHz, D2O)δ: 167.41 (s), 149.97 (s). IR (KBr, ν/cm-1): 3448, 3138, 2718, 1582, 1523, 1489, 1417, 1339, 1289, 1159, 1078, 1017, 987, 775, 728. Anal. calcd for (%) for C4H8N10O8(324.17): C 14.82, H 2.49, N 43.21; found C 14.71, H 2.38, N 43.33.

    Hydrazinium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑ oxadiazolate)(5)

    Hydrazine hydrate (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.254 g of compound 5 was obtained as a yellow solid in a yield of 78.9%. 1H NMR(500 MHz, DMSO‑d 6)δ: 6.16(br.). 13C NMR(125 MHz, DMSO‑d 6)δ: 166.11 (s), 148.34 (s). IR(KBr, ν/cm-1): 3330, 3303, 3222, 3161, 1516, 1490, 1435, 1301, 1280, 1158, 1110, 1071, 1014, 954, 786, 731. Anal. calcd for (%) for C4H10N12O6 (322.20): C 14.91, H 3.13, N 52.17; found C 14.78, H 3.02, N 52.32.

    Diaminomethaniminium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (6)

    Guanidine carbonate (1.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.312 g of compound 6 was obtained as a yellow solid in a yield of 82.9%. 1H NMR (500 MHz, DMSO‑d 6)δ: 6.93 (s, 12H). 13C NMR (125 MHz, DMSO‑d 6)δ: 166.05 (s), 157.91 (s), 148.36 (s). IR (KBr, ν/cm-1): 3386, 3270, 3205, 1681, 1657, 1525, 1494, 1407, 1297, 1155, 1074, 635. Anal. calcd for (%) for C6H12N14O6 (376.25): C 19.15, H 3.21, N 52.12; found: C 19.03, H 3.42, N 52.24.

    Amino(hydrazinyl)methaniminium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (7)

    Aminoguanidine bicarbonate (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.352 g of compound 7 was obtained as a yellow solid in a yield of 86.6%. 1H NMR (500 MHz, DMSO‑d 6)δ: 8.56 (s, 2H)δ: 7.02 (s, 8H) δ: 5.36 (s, 4H). 13C NMR (125 MHz, DMSO‑d 6)δ: 164.65(s), 158.76(s), 146.85(s). IR(KBr, ν/cm-1): 3385, 3310, 3164, 1664, 1590, 1522, 1488, 1403, 1277, 1153, 1073, 1016, 956, 776, 660. Anal. calcd for (%) for C6H14N16O6(406.28): C 17.74, H 3.47, N 55.16; found: C 17.68, H 3.52, N 55.22.

    Di(1,2,4‑triazolium) 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (8)

    1,2,4‑ Triazole (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.347 g of compound 8 was obtained as a white solid in a yield of 87.6%. 1H NMR (500 MHz, DMSO‑d 6)δ:10.06 (br.). 8.97(s, 4H). 13C NMR (125 MHz, DMSO‑d 6)δ: 164.33(s), 146.48(s), 144.53(s). IR(KBr, ν/cm-1): 3105, 1568, 1526, 1491, 1451, 1322, 1295, 1264, 1159, 1077, 1018, 945, 735, 625. Anal.calcd for(%) for C8H8N14O6(396.24): C 24.25, H 2.04, N 49.49; found: C 24.20, H 2.11, N 49.53.

    Di(3‑amino‑1,2,4‑triazolium) 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (9)

    3‑Amino‑1,2,4‑triazole (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.362 g of compound 9 was obtained as a white solid in a yield of 84.9%. 1H NMR (500 MHz, DMSO‑d 6)δ: 8.27(s, 2H)δ: 8.05 (br.). 13C NMR (125 MHz, DMSO‑d 6)δ: 165.94 (s), 151.11 (s), 148.11 (s), 139.64 (s). IR (KBr, ν/cm‑1): 3405, 3149, 2738, 1691, 1572, 1548, 1503, 1456, 1323, 1281, 1161, 1135, 1084, 1024, 946, 914, 887, 774,728. Anal. calcd for (%) for C8H10N16O6(426.27): C 22.54, H 2.36, N 52.57; found: C 22.21, H 2.39, N 52.49.

    Di(4‑amino‑1,2,4‑triazolium) 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (10)

    4‑Amino‑1,2,4‑triazole (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.356 g of compound 10 was obtained as a white solid in a yield of 83.5%. 1H NMR (500 MHz, DMSO‑d 6)δ: 9.25 (s, 4H), 8.43 (br). 13C NMR (125 MHz, DMSO‑d 6)δ: 164.86 (s), 147.03 (s), 144.07 (s). IR (KBr, ν/cm-1): 3312, 3120, 1571, 1555, 1537, 1493, 1495, 1322, 1277, 1161, 1086, 1040, 1019, 984, 798, 784, 622. Anal. calcd for (%) for C8H10N16O6 (426.27): C 22.54, H 2.36, N 52.57; found: C 22.37, H 2.48, N 52.31.

    Didiaminouronium 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) (11)

    Carbohydrazide (2.0 mmol) was added to a solution of compound 2 (0.258 g, 1.0 mmol) in 30 mL of acetonitrile and the mixture was stirred at room temperature. After 2 h the precipitate was filtered, washed with ether and air‑dried to yield. 0.392 g of compound 11 was obtained as a white solid in a yield of 89.5%. 1H NMR (500 MHz, DMSO‑d 6)δ: 8.50. 13C NMR (125 MHz, DMSO‑d 6)δ: 166.11 (s), 159.14 (s), 148.38 (s), IR (KBr, ν/cm-1): 3320, 3275, 3213, 1697, 1623, 1602, 1555, 1529, 1492, 1429, 1350, 1316, 1281, 1220, 1168, 1120, 1083, 1019, 843, 776, 690. Anal. calcd for (%) for C6H14N16O8 (438.28): C 16.44, H 3.22, N 51.13; found: C 16.40, H 3.54, N 50.28.

  • 3 Results and Discussion

    3
  • 3.1 Single Crystal Structure Analysis

    3.1

    A solvent evaporation method was applied to grow good quality crystals of 9·2H2O and compound 10 for X‑ray diffraction. The crystals of 9·2H2O and compound 10 were grown from a solution of water/CH3OH. Their structures are shown in Fig.1 to 4 and the crystallographic and structural refinement data are listed in Table .1

    Compound 9·2H2O crystallizes in the triclinic space group Pī with one formula unit in the unit cell and a calculated density of 1.761 g·cm-3 at 172(2) K. The asymmetric unit consists of one 3‑amino‑1,2,4‑triazole cation, one and a half 5,5′‑dinitroamino‑2,2′‑bi(1,3,4‑oxadiazolate) anion and a lattice water molecule. As shown in Fig.,1, the anion exhibits a central symmetric structure where an inversion centre lies. The packing diagram of 9·2H2O is built up by hydrogen bonds(Fig.2). The cations connect to anions by the classical hydrogen bonds (N(7)—H(7)…O(2), 2.800(2) Å; N(7)—H(7)…N(3), 3.051(2) Å; N(7)—H(7)…N(4), 3.353(2) Å; N(8)—H(8)…N(2), 3.021(2) Å) and the extensive hydrogen‑bonding in the structure may also contribute to its good thermal stability.

    Fig.1
                            Molecular structure of 9·2H2O

    Fig.1 Molecular structure of 9·2H2O

    Fig.2
                            Ball and stick packing diagram of 9·2H2O

    Fig.2 Ball and stick packing diagram of 9·2H2O

    Table 1 Crystallographic data for 9·2H2O and 10

    crystal9·2H2O10
    CCDC15442541543503
    formulaC8H10N16O6·2H2OC8H10N16O6
    formula mass462.35426.32
    temperature /K172(2)171(2)
    crystal systemtriclinicmonoclinic
    space group P‑1 P21/c
    volume /Å3 435.87(6)812.65(14)
    a6.0458(5)6.9450(6)
    b6.6277(6)6.3230(6)
    c11.3541(9)18.576(2)
    α/(°)97.206(3)90
    β/(°)101.955(3)94.983(4)
    γ/(°)97.438(3)90
    Z 12
    ρ/g·cm-3 1.7611.742
    F(000)238.0436
    F 2 1.0501.031
    R 1,wR 2 [all data]0.0612, 0.08870.0851, 0.1087
    R 1,wR 2[I>=2σ(I)]0.0413, 0.08160.0511, 0.0973

    Compound 10 crystallizes in the monoclinic space group P21/c with Z=2 and a cell volume of 812.65(14) Å3. A density of 1.742 g·cm-3 was determined at 171 K from the X‑ray diffraction analysis. As presented in Fig.,3, the structure of anion is similar to that of compound 10 and there are little differences in bond lengths and angles. The anion within the asymmetric unit (—CNNOC—N—NO2) is nearly coplanar with an root mean square deviation value of 0.0237 Å. As can be seen from Fig.,4, large amount of hydrogen bonds can be found between the anions and cations (N(1)—H(1)…O(3), 2.888(3) Å; N(1)—H(1)…N(5), 2.726(3) Å; N(4)—H(4)…N(7), 2.984(4) Å). The oxygen atoms (O(2), O(3)) in the anions are involved in the formation of non‑classical hydrogen bonds with the CH fragments (C(2)—H(2)…O(3), 2.950(3) Å; C(2)—H(2)…O(2), 3.024(4) Å).

    Fig.3
                            Molecular structure of compound 10

    Fig.3 Molecular structure of compound 10

    Fig.4
                            Ball and stick packing diagram of compound 10

    Fig.4 Ball and stick packing diagram of compound 10

  • 3.2 Computational Details

    3.2

    Computations were performed by using the Gaussian 09 suite of programs. The elementary geometric optimization and the frequency analysis were performed at the level of the Becke three parameter, Lee‑Yan‑Parr (B3LYP) functional with the 6‑311+G** basis set[21]. All the optimized structures were characterized to be local energy minima on the potential surface without imaginary frequencies. Atomization energies were calculated by the CBS‑4M. The lattice energy of the trinitroethyl derivatives were predicted by the formula suggested by Jenkins et al.

    The predictions of heats of formation (HOF) used the hybrid DFT/B3LYP methods with the 6‑311+G** basis set through designed isodesmic reactions. The isodesmic reaction processes, that is, the number of each kind of formal bond is conserved, were used with the application of the bond separation reaction (BSR) rules. The molecule was broken down into a set of two heavy‑atom molecules containing the same component bonds. The isodesmic reactions used to derive the HOF of compounds 3-11 are shown in Scheme 2.

    Scheme 2 Isodemic reactions of N,N'‑([2,2'‑bi(1,3,4‑oxadiazole)]‑5,5'‑diyl)dinitramide anion

    The change of enthalpy for the reactions at 298K can be expressed by Equ.(1):

    ΔH 298 KΔf H P–ΣΔf H R (1)

    Where ΣΔf H P and ΣΔf H R are the HOF of the reactants and products at 298 K, respectively, and ΔH 298 K can be calculated from the following expression in Equ.(2):

    ΔH 298 KE 298 K+Δ(pV)=ΔE 0+ΔZPE+ΔH TnRT (2)

    where ΔE 0 is the change in total energy between the products and the reactants at 0 K; ΔZPE is the difference between the zero‑point energies (ZPE) of the products and the reactants at 0 K. ΔH T is the thermal correction from 0 to 298 K. The Δ(pV) value in Equ.(2) is the pV work term. It equals ΔnRT for the reactions of an ideal gas. For the isodesmic reactions Δn=0, so Δ(pV)=0. On the left side of Equ.(2) apart from target compound all the others are called reference compounds. The HOF of reference compounds are available either from experiments or from the high level computing such as CBS‑4M.

    Based on a Born‑Haber energy cycle (Scheme 3), the heat of formation of a salt can be simplified by Equ.(3):

    ΔH f 0(ionic salt, 298 K)=ΔH f 0(cation, 298 K)+ ΔH f 0 (anion, 298 K)-ΔH L (3)

    where ΔH L is the lattice energy of the salt which could be predicted by the formula suggested by Jenkins et al[22]. as given in Equ.(4):

    ΔHL =UPOT+[p(n M/2‑2)+q(n X/2‑2)]RT (4)

    where nM and nX depend on the nature of the ions Mp+ and Xq-, respectively, and are Equ.(3) for monatomic ions, Equ.(5) for linear polyatomic ions, and 6 for nonlinear polyatomic ions. The equation for lattice potential energy U POT takes the form of Equ.(5):

    U POT=γ(ρm/Mm )1/3+δ (5)

    where ρm is the density ,g·cm-3. M mis the chemical formula mass of the ionic material and the coefficients γ , kJ·mol-1 ·cm .δ are 8375.6 and -178.8 kJ·mol-1, respectively.

    Scheme 3 Born‑Haber cycle for the formation of energetic salts

  • 3.3 Physicochemical Properties

    3.3

    Thermal stability is one of the most important physicochemical properties of energetic materials. The thermal stabilities of compounds 3-11 were determined by differential scanning calorimetry (DSC) at a heating rate of 5 ℃·min-1 using dry nitrogen in Al pans. For all compounds, the decomposition temperatures were determined by the decomposition onset temperatures. As show in Fig.,5, all the compounds decompose directly without melting, and except for compound 3, that melts at 73.5 ℃. In particular, the decomposition process of 4-6, and 9 possess two steps with the onset temperatures. The decomposition onset temperatures of the energetic compounds are observed in a range from 146.7(4) to 239.9 ℃(6). Compounds 6-10 are observed in a range from 181.7(9) to 239.9 ℃(6), As is known, thermal stability above 180 ℃ is an essential requirement for energetic compounds for adaptation for practical use[22]. The guanidinium salt 6 is the most thermally stable one at 239.9 ℃. Both compound 6 and 9 have higher decomposition temperatures than RDX (204°C).

    Fig.5
                            DSC curves of compounds 3-11

    Fig.5 DSC curves of compounds 3-11

  • 3.4 Energetic Performance

    3.4

    Impact and friction sensitivities are high priorities for secondary explosives. Low sensitivities of energetic materials can reduce the risk of serious and fatal accidents during the applications. Compounds 3-11 were tested for their sensitivity towards friction and impact using the BAM methods. The impact and friction sensitivities of compounds 3-11 fall in the range of 10-28 J or 120-280 N. The energetic salts are well stabilized and exhibit better impact and friction sensitivities (10-28 J for IS and 160-360 N for FS) than their parent compounds. All the compounds are less sensitive than RDX(7.5 J), the IS values of 3-4, 6-11 are higher than that of TNT (15 J). In the terms of the friction sensitivities, all of the energetic salts are much higher than RDX (120 N). By comparing with nitrogen‑rich cations, the salts have additional ionic and hydrogen bonding, and thus exhibit lower impact and friction sensitivities.

    The enthalpy of formation (HOF) is essential for calculating the detonation performance. The heats of formation of the synthesized compounds 3-11 were calculated based on appropriate isodesmic reactions. Calculations were carried out using the Gaussian 09 program suite. The geometry optimization of the structures and frequency analyses were carried out using the B3LYP functional with the 6‑311+G** basis set. All of the optimized structures were characterized by true local energy minima on the potential energy surface without imaginary frequencies. The results are summarized in Table 2, the enthalpy of formation ranging from -0.671(6) kJ·g-1 to 1.143 kJ·g-1 (10). With the exception of 3, 4 and 6, the energetic salts possess positive heats of formation, which are higher than TNT (-1.30 kJ·g-1) and RDX (0.32 kJ·g-1) due to a large number of C—N and N—N bonds, and compound 10 has the highest value of 1.143 kJ·g-1.

    Table 2 The physicochemical properties of 3-11 compared with trinitrotoluene (TNT), 1,3,5‑trinitroperhydro‑1,3,5‑triazine (RDX), octahydro‑1,3,5,7‑tetranitro‑1,3,5,7‑tetrazocine(HMX)

    compounds T m / °C T d / °COB / % ρ / g·cm-3 ΔH f / kJ·mol-1 D / m·s-1 p / GPaIS / JFS / N
    373.5172.2-32.851.941-125.1913634.220180
    4_146.8-19.741.926-121.55907833.916160
    5_151.3-34.761.937113.40901234.110160
    6_239.9-51.031.711-252.57769321.328280
    7_188.6-55.131.692155.16810024.124240
    8_188.4-56.531.727382.29788623.926280
    9_219.6-56.311.751303.40798424.025360
    10_181.7-56.311.739602.68818125.822360
    11_155.9-40.151.683106.21814725.226360
    TNT[4] 80.4295-73.81.654-295688119.515_
    RDX[23] _204-21.61.8070.3879534.97.5120
    HMX[18] _280-21.61.91106.63932039.57.4120

    T m is melting point. T d is decomposition temperature from DSC (5 ℃·min-1). OB is oxygen balance (%) for CaHbNcOd, and OB%=1600×(d‑2a‑b/2)/Mw (based on carbon dioxide). ρ is density measured using a gas pycnometer (25 ℃). ΔH f is calculated molar enthalpy of formation. D is calculated detonation velocity. p is calculated detonation pressure. IS is impact sensitivity. FS is friction sensitivity

    By using the calculated heats of formation and the experimental densities (gas pycnometer) of the new energetic compounds, the detonation pressures (p) and detonation velocities (D) were calculated by using the EXPLO5 v6.02 program. As is shown in Table 2, the calculated detonation velocities fall in the range between 7693 m·s-1(6) and 9136 m·s-1(3), and detonation pressures range from 21.3 GPa(6) to 34.2 GPa(3). Compound 3-5 show excellent detonation properties which are much higher than those of RDX (8795 m·s-1, 34.9 GPa), and to those of HMX (9320 m·s-1, 39.6 GPa).

  • 4 Conclusions

    4

    A new family of energetic salts featuring bi(1,3,4‑oxadiazole) were prepared and fully characterized. In addition, the structures of 9·2H2O and 10 were confirmed by single‑crystal X‑ray diffraction analysis. According to the DSC results, most of the energetic salts showed acceptable thermal stabilities, with the decomposition temperatures ranging from 146.8 to 239.9 ℃. The calculated detonation velocities lie in the range between 7693(6) and 9136(3) m·s-1, and the detonation pressures range from 21.3 to 34.2 GPa, which are larger than those of TNT. The detonation velocities and pressures of compounds 3-5 are comparable to those of RDX. Based on the impact and friction tests, all the salts are less sensitive than RDX. Ammonium salt which has excellent detonation properties (D=9136 m·s-1, p=34.2 GPa) and acceptable sensitivities (IS=20 J, FS=180 N) is a promising potential candidate for energetic materials.

  • Reference

    • 1

      ZHANG Qing‑hua, HE Chun‑lin, YIN Ping, et al. Insensitive nitrogen‑rich materials incorporating the nitroguanidyl functionality[J]. Chemistry-an Asian Journal, 2014, 9(1): 212-217.

    • 2

      GAO Hai‑xiang, Shreeve J M. Azole‑​based energetic salts[J]. Chemical. Reviews, 2011, 111(11): 7377-7436.

    • 3

      WANG Rui‑hu, XU Hong‑yan, GUO Yong, et al. Bis[3‑(5‑nitroimino‑1,2,4‑triazolate)]‑based energetic salts: synthesis and promising properties of a new family of high‑density insensitive materials[J]. Journal of the American Chemical Society, 2010, 132(34): 11904-11905.

    • 4

      WANG Kai, Parrish D A, Shreeve J M. 3‑Azido‑N‑nitro‑1H‑1,2,4‑triazol‑5‑amine‑based energetic salts[J]. Chemistry‑A European Journal, 2011, 17(51): 14485-14492

    • 5

      ZHANG Jia‑heng, Dharavath S, Mitchell L A, et al. Energetic salts based on 3,5‑bis(dinitromethyl)‑1,2,4‑triazole monoanion and dianion: controllable preparation, characterization, and high performance[J]. Journal of the American Chemical Society, 2016, 138(24): 7500-7503

    • 6

      Klapötke T M, Mayr N, Stierstorfer J, et al. Maximum compaction of ionic organic explosives: bis(hydroxylammonium) 5,​5'‑dinitromethyl‑3,3′‑bis(1,2,4‑oxadiazolate) and its derivatives[J]. Chemistry‑A European Journal, 2014, 20(5): 1410-1417.

    • 7

      DENG Mu‑chong, ZHANG Qing‑hua, WANG Kang‑cai, et al. Synthesis and properties of 5,10‑bis(dinitromethyl)‑furazan[3,4‑e]bis([1,2,4]triazolo)[4,3‑a:3′,4′‑c] pyrazine and its energetic ion compounds[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2018, 26(2):144-149.

    • 8

      LI Ya‑nan, SHU Yuan‑jie, ZHANG Sheng‑yong, et al. Synthesis and thermal Properties of 4,4′,5,5′‑Tetranitro‑2,2′‑biimidazole and its energetic ion salts[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2017, 25(4): 298-303.

    • 9

      WANG Rui‑hu, GUO Yong, ZENG Zhuo, et al. Furazan‑functionalized tetrazolate‑based salts: a new family of insensitive energetic materials[J]. Chemistry‑A European Journal, 2009, 15(11): 2625-2634.

    • 10

      WANG Rui‑hu, GUO Yong, SA Rong‑jian, et al. Nitroguanidine‑fused bicyclic guanidinium salts: a family of high‑density energetic materials[J]. Chemistry‑A European Journal, 2010, 16(28): 8522-8529

    • 11

      TANG Yong‑xing, ZHANG Jia‑heng, Mitchell L A, et al. Taming of 3,4‑di(nitramino)​furazan[J]. Journal of the American Chemical Society, 2015, 137(51):15984-15987

    • 12

      TANG Yong‑xing, GAO Hai‑xiang, Mitchell L A, et al. Syntheses and promising properties of dense energetic 5,5′‑dinitramino‑3,3'‑azo‑1,2,4‑oxadiazole and its salts[J]. Angewandte Chemie International Edition, 2016, 55(9): 3200-3203.

    • 13

      HAO Wei, HE Chun‑lin, ZHANG Jia‑heng, et al. Combination of 1,2,4‑​oxadiazole and 1,2,5‑oxadiazole moieties for the generation of high‑​performance energetic materials[J]. Angewandte Chemie International Edition, 2015, 54(32): 9367-9371.

    • 14

      HE Chun‑lin, TANG Yong‑xing, Mitchell L A, et al. N‑​oxides light up energetic performances: synthesis and characterization of dinitraminobisfuroxans and their salts[J]. Journal of Materials Chemistry A, 2016, 4(23): 8969-8973.

    • 15

      WANG Zuo‑quan, ZHANG Hong, Killian B J, et al. Synthesis, characterization and energetic properties of 1,​3,​4‑​oxadiazoles[J]. European Journal of Organic Chemistry, 2015, 2015(23): 5183-5188.

    • 16

      WU Qiong, ZHU Wei‑hua, XIAO He‑ming. Molecular design of trinitromethyl‑substituted nitrogen‑rich heterocycle derivatives with good oxygen balance as high‑energy density compounds[J]. Structural Chemistry, 2013, 24(5):1725-1736.

    • 17

      TIAN Jia‑wei, XIONG Hua‑lin, LIN Qiu‑han, et al. Energetic compounds featuring bi(1,3,4‑oxadiazole)​: a new family of insensitive energetic materials[J]. New Journal of Chemistry, 2017, 41(5): 1918-1924.

    • 18

      YU Qiong, YIN Ping, ZHANG Jia‑heng, et al. Pushing the limits of oxygen balance in 1,3,4‑oxadiazoles[J]. Journal of the American Chemical Society, 2017, 139(26): 8816-8819.

    • 19

      Hermann T S, Karaghiosoff K, Klapötke T M, et al. Synthesis and characterization of 2,2′‑dinitramino‑5,5′‑bi(1‑oxa‑3,4‑diazole) and derivatives as economic and highly dense energetic materials[J]. Chemistry‑A European Journal, 2017, 23(50): 12092-12095.

    • 20

      ZHANG Wen‑quan, ZHANG Jia‑heng, DENG Mu‑chong, et al. A promising high‑energy‑density material[J]. Nature Communication, 2017, 8(1): 181-187.

    • 21

      Hariharan P C, Pople J A. Influence of polarization functions on MO hydrogenation energies[J]. Theoretica Chimica Acta, 1973, 23(8): 213-216.

    • 22

      Jenkins H D B, Tudeal D, Glasser L. Lattice potential energy estimation for complex ionic salts from density measurements[J]. Inorganic Chemistry, 2002, 41(9): 2364-2367.

    • 23

      Klapötke T M, Penger A, Pflüger C, et al. Advanced open‑chain nitramines as energetic materials: heterocyclic‑substituted 1,3‑dichloro‑2‑nitrazapropane[J]. European Journal of Inorganic Chemistry, 2013, 2013(26): 4667-4678.

XIONGHua‑lin

机 构: 南京理工大学化工学院, 江苏 南京 210094

Affiliation: School of Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China

邮 箱:18100619214@163.com

Biography:XIONG Hua‑lin(1992-), male, synthesis of energetic materials. e‑mail:18100619214@163.com

YANGHong‑wei

机 构: 南京理工大学化工学院, 江苏 南京 210094

Affiliation: School of Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China

CHENGGuang‑bin

机 构: 南京理工大学化工学院, 江苏 南京 210094

Affiliation: School of Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China

角 色:通讯作者

Role: Corresponding author

邮 箱:gcheng@mail.njust.edu.cn.

Biography: CHENG Guang‑bin, male, professor, synthesis of energetic materials. e‑mail:gcheng@mail.njust.edu.cn.

html/hncl/CJEM2018142/media/e7437d06-4e4b-49c0-9066-b260cc6dc5eb-image002.png
html/hncl/CJEM2018142/media/e7437d06-4e4b-49c0-9066-b260cc6dc5eb-image003.png
crystal9·2H2O10
CCDC15442541543503
formulaC8H10N16O6·2H2OC8H10N16O6
formula mass462.35426.32
temperature /K172(2)171(2)
crystal systemtriclinicmonoclinic
space group P‑1 P21/c
volume /Å3 435.87(6)812.65(14)
a6.0458(5)6.9450(6)
b6.6277(6)6.3230(6)
c11.3541(9)18.576(2)
α/(°)97.206(3)90
β/(°)101.955(3)94.983(4)
γ/(°)97.438(3)90
Z 12
ρ/g·cm-3 1.7611.742
F(000)238.0436
F 2 1.0501.031
R 1,wR 2 [all data]0.0612, 0.08870.0851, 0.1087
R 1,wR 2[I>=2σ(I)]0.0413, 0.08160.0511, 0.0973
html/hncl/CJEM2018142/media/e7437d06-4e4b-49c0-9066-b260cc6dc5eb-image004.png
html/hncl/CJEM2018142/media/e7437d06-4e4b-49c0-9066-b260cc6dc5eb-image005.png
html/hncl/CJEM2018142/alternativeImage/e7437d06-4e4b-49c0-9066-b260cc6dc5eb-F008.jpg
compounds T m / °C T d / °COB / % ρ / g·cm-3 ΔH f / kJ·mol-1 D / m·s-1 p / GPaIS / JFS / N
373.5172.2-32.851.941-125.1913634.220180
4_146.8-19.741.926-121.55907833.916160
5_151.3-34.761.937113.40901234.110160
6_239.9-51.031.711-252.57769321.328280
7_188.6-55.131.692155.16810024.124240
8_188.4-56.531.727382.29788623.926280
9_219.6-56.311.751303.40798424.025360
10_181.7-56.311.739602.68818125.822360
11_155.9-40.151.683106.21814725.226360
TNT[4] 80.4295-73.81.654-295688119.515_
RDX[23] _204-21.61.8070.3879534.97.5120
HMX[18] _280-21.61.91106.63932039.57.4120

Fig.1 Molecular structure of 9·2H2O

Fig.2 Ball and stick packing diagram of 9·2H2O

Fig.3 Molecular structure of compound 10

Fig.4 Ball and stick packing diagram of compound 10

Fig.5 DSC curves of compounds 3-11

image /

无注解

无注解

无注解

无注解

无注解

无注解

T m is melting point. T d is decomposition temperature from DSC (5 ℃·min-1). OB is oxygen balance (%) for CaHbNcOd, and OB%=1600×(d‑2a‑b/2)/Mw (based on carbon dioxide). ρ is density measured using a gas pycnometer (25 ℃). ΔH f is calculated molar enthalpy of formation. D is calculated detonation velocity. p is calculated detonation pressure. IS is impact sensitivity. FS is friction sensitivity

  • Reference

    • 1

      ZHANG Qing‑hua, HE Chun‑lin, YIN Ping, et al. Insensitive nitrogen‑rich materials incorporating the nitroguanidyl functionality[J]. Chemistry-an Asian Journal, 2014, 9(1): 212-217.

    • 2

      GAO Hai‑xiang, Shreeve J M. Azole‑​based energetic salts[J]. Chemical. Reviews, 2011, 111(11): 7377-7436.

    • 3

      WANG Rui‑hu, XU Hong‑yan, GUO Yong, et al. Bis[3‑(5‑nitroimino‑1,2,4‑triazolate)]‑based energetic salts: synthesis and promising properties of a new family of high‑density insensitive materials[J]. Journal of the American Chemical Society, 2010, 132(34): 11904-11905.

    • 4

      WANG Kai, Parrish D A, Shreeve J M. 3‑Azido‑N‑nitro‑1H‑1,2,4‑triazol‑5‑amine‑based energetic salts[J]. Chemistry‑A European Journal, 2011, 17(51): 14485-14492

    • 5

      ZHANG Jia‑heng, Dharavath S, Mitchell L A, et al. Energetic salts based on 3,5‑bis(dinitromethyl)‑1,2,4‑triazole monoanion and dianion: controllable preparation, characterization, and high performance[J]. Journal of the American Chemical Society, 2016, 138(24): 7500-7503

    • 6

      Klapötke T M, Mayr N, Stierstorfer J, et al. Maximum compaction of ionic organic explosives: bis(hydroxylammonium) 5,​5'‑dinitromethyl‑3,3′‑bis(1,2,4‑oxadiazolate) and its derivatives[J]. Chemistry‑A European Journal, 2014, 20(5): 1410-1417.

    • 7

      DENG Mu‑chong, ZHANG Qing‑hua, WANG Kang‑cai, et al. Synthesis and properties of 5,10‑bis(dinitromethyl)‑furazan[3,4‑e]bis([1,2,4]triazolo)[4,3‑a:3′,4′‑c] pyrazine and its energetic ion compounds[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2018, 26(2):144-149.

    • 8

      LI Ya‑nan, SHU Yuan‑jie, ZHANG Sheng‑yong, et al. Synthesis and thermal Properties of 4,4′,5,5′‑Tetranitro‑2,2′‑biimidazole and its energetic ion salts[J]. Chinese Journal of Energetic Materials(Hanneng Cailiao), 2017, 25(4): 298-303.

    • 9

      WANG Rui‑hu, GUO Yong, ZENG Zhuo, et al. Furazan‑functionalized tetrazolate‑based salts: a new family of insensitive energetic materials[J]. Chemistry‑A European Journal, 2009, 15(11): 2625-2634.

    • 10

      WANG Rui‑hu, GUO Yong, SA Rong‑jian, et al. Nitroguanidine‑fused bicyclic guanidinium salts: a family of high‑density energetic materials[J]. Chemistry‑A European Journal, 2010, 16(28): 8522-8529

    • 11

      TANG Yong‑xing, ZHANG Jia‑heng, Mitchell L A, et al. Taming of 3,4‑di(nitramino)​furazan[J]. Journal of the American Chemical Society, 2015, 137(51):15984-15987

    • 12

      TANG Yong‑xing, GAO Hai‑xiang, Mitchell L A, et al. Syntheses and promising properties of dense energetic 5,5′‑dinitramino‑3,3'‑azo‑1,2,4‑oxadiazole and its salts[J]. Angewandte Chemie International Edition, 2016, 55(9): 3200-3203.

    • 13

      HAO Wei, HE Chun‑lin, ZHANG Jia‑heng, et al. Combination of 1,2,4‑​oxadiazole and 1,2,5‑oxadiazole moieties for the generation of high‑​performance energetic materials[J]. Angewandte Chemie International Edition, 2015, 54(32): 9367-9371.

    • 14

      HE Chun‑lin, TANG Yong‑xing, Mitchell L A, et al. N‑​oxides light up energetic performances: synthesis and characterization of dinitraminobisfuroxans and their salts[J]. Journal of Materials Chemistry A, 2016, 4(23): 8969-8973.

    • 15

      WANG Zuo‑quan, ZHANG Hong, Killian B J, et al. Synthesis, characterization and energetic properties of 1,​3,​4‑​oxadiazoles[J]. European Journal of Organic Chemistry, 2015, 2015(23): 5183-5188.

    • 16

      WU Qiong, ZHU Wei‑hua, XIAO He‑ming. Molecular design of trinitromethyl‑substituted nitrogen‑rich heterocycle derivatives with good oxygen balance as high‑energy density compounds[J]. Structural Chemistry, 2013, 24(5):1725-1736.

    • 17

      TIAN Jia‑wei, XIONG Hua‑lin, LIN Qiu‑han, et al. Energetic compounds featuring bi(1,3,4‑oxadiazole)​: a new family of insensitive energetic materials[J]. New Journal of Chemistry, 2017, 41(5): 1918-1924.

    • 18

      YU Qiong, YIN Ping, ZHANG Jia‑heng, et al. Pushing the limits of oxygen balance in 1,3,4‑oxadiazoles[J]. Journal of the American Chemical Society, 2017, 139(26): 8816-8819.

    • 19

      Hermann T S, Karaghiosoff K, Klapötke T M, et al. Synthesis and characterization of 2,2′‑dinitramino‑5,5′‑bi(1‑oxa‑3,4‑diazole) and derivatives as economic and highly dense energetic materials[J]. Chemistry‑A European Journal, 2017, 23(50): 12092-12095.

    • 20

      ZHANG Wen‑quan, ZHANG Jia‑heng, DENG Mu‑chong, et al. A promising high‑energy‑density material[J]. Nature Communication, 2017, 8(1): 181-187.

    • 21

      Hariharan P C, Pople J A. Influence of polarization functions on MO hydrogenation energies[J]. Theoretica Chimica Acta, 1973, 23(8): 213-216.

    • 22

      Jenkins H D B, Tudeal D, Glasser L. Lattice potential energy estimation for complex ionic salts from density measurements[J]. Inorganic Chemistry, 2002, 41(9): 2364-2367.

    • 23

      Klapötke T M, Penger A, Pflüger C, et al. Advanced open‑chain nitramines as energetic materials: heterocyclic‑substituted 1,3‑dichloro‑2‑nitrazapropane[J]. European Journal of Inorganic Chemistry, 2013, 2013(26): 4667-4678.