Open Access Article
This Open Access Article is licensed under a
Creative Commons Attribution 3.0 Unported Licence

Synthesis and characterization of novel nitrofurazanyl ethers as potential energetic plasticizers

Patrick Liebera, Uwe Schaller*a and Thomas M. Klapötkeb
aFraunhofer Institute for Chemical Technology, Joseph-von-Fraunhofer-Str. 7, 76327 Pfinztal, Germany. E-mail: uwe.schaller@ict.fraunhofer.de
bDepartment of Chemistry, Ludwig-Maximilian-University of Munich, Butenandtstr. 5-13, 81377 Munich, Germany. E-mail: tmk@cup.uni-muenchen.de

Received 21st February 2025 , Accepted 10th April 2025

First published on 22nd April 2025


Abstract

Energetic plasticizers are used to improve the mechanical properties of advanced energetic formulations while increasing the overall energy content. Although nitro-1,2,5-oxadiazoles (nitrofurazans) possess excellent energetic properties such as a favorable oxygen balance and high heat of formation, their use as plasticizers has received little attention in the scientific literature. Four nitrofurazanyl ethers were synthesized by substitution of dinitrofurazan with linear alkoxides. The synthesized compounds were extensively analyzed by Fourier-transform infrared (FT-IR) spectroscopy, Raman spectroscopy, differential scanning calorimetry (DSC), thermogravimetric analysis (TGA), electrospray ionization (ESI) mass spectroscopy, mechanical sensitivity test, 1H nuclear magnetic resonance (NMR) spectroscopy and 13C NMR spectroscopy. They have lower mechanical sensitivity (>40 J) compared to modern energetic plasticizers in use, including 2,2-dinitropropyl formal/acetal (BDNPA/F), n-butylnitratoethylnitramine (BuNENA), and dinitrodiazaalkane (DNDA-57). In addition, the most promising compound 3-(2-(2-(2-azidoethoxy)ethoxy)ethoxy)-4-nitro-1,2,5-oxadiazole (NFPEG3N3) exhibits competitive thermal properties, with a lower glass transition temperature of −72 °C compared to BNDPA/F (−67 °C) and a higher thermal decomposition temperature of 179 °C compared to BuNENA (173 °C). The enthalpy of formation and heat of explosion of NFPEG3N3 were calculated to be −41.7 kJ mol−1 and 3421 J g−1, respectively. The impact of NFPEG3N3 on the glass transition temperature, viscosity and decomposition of the energetic binder glycidyl azide polymer (GAP)-diol was investigated and showed a remarkable decrease in viscosity (45.4%) and glass transition temperature (−3.3 °C) when compared to benchmark plasticizers in 10 wt% mixtures. These results demonstrate the potential of NFPEG3N3 as an insensitive and highly energetic plasticizer.


Introduction

Energetic plasticizers are critical to the development of advanced solid propellants and polymer-bonded explosives.1–4 Typically, these plasticizers are high-boiling organic liquids that improve the mechanical properties and lower the glass transition temperature of binders. They also help reduce the viscosity during processing and can modify the burn rate of energetic formulations. Unlike inert plasticizers, energetic plasticizers contain energetic groups such as nitro, nitrate ester, nitramino, and azido, which increase the overall energy of the formulation. Depending on the application, the ideal properties for energetic plasticizers can vary and sometimes be contradictory. In general, they should have high heat of formation, high oxygen balance, low glass transition temperature, low migration tendency, low viscosity, high thermal stability and low mechanical sensitivity.5 Investigation of the glass transition temperature depression, viscosity reduction and decomposition temperature shift of a liquid plasticizer-binder mixture compared to pure binder samples provides initial insight into the suitability of a new compound.6

Known energetic plasticizers such as 2,2-dinitropropyl formal/acetal7,8 (BDNPA/F), n-butylnitratoethylnitramine9 (BuNENA) and dinitrodiazaalkane10 (DNDA-57) primarily consist of a saturated, linear carbon backbone, augmented by heteroatoms, most of which form energetic side groups (Fig. 1). Studies on heterocycle-based energetic plasticizers remain limited in the literature although heterocyclic building blocks are energy-rich and have been extensively studied for secondary and primary explosives.11–13


image file: d5ra01282a-f1.tif
Fig. 1 Examples of energetic plasticizers in use.

In particular, oxadiazoles may be advantageous for plasticizers due to their higher oxygen content compared to triazoles or tetrazoles. Among the oxadiazole isomers, the 1,2,5-oxadiazole (furazan) ring has the highest heat of formation at 216 kJ mol−1.15–17 Notably, 3,4-dinitrofurazan (DNF) is a liquid at room temperature, with a melting point of −15 °C.18–20 Although its synthesis is challenging, we have recently improved its safety by using a flow process.21 Derivatization of DNF via nucleophilic substitution is a straightforward reaction, as demonstrated by Sheremetev et al.22,23 In this work we present novel energetic plasticizers that exploit the outstanding energetic properties of the nitrofurazan structure.

Results and discussion

Synthesis

The starting material chosen for the synthesis was 3,4-diaminofurazan (DAF). Products were obtained by two-step syntheses (Scheme 1). First, DNF was obtained by complete amine oxidation of DAF with a mixture of hydrogen peroxide, sulfuric acid and sodium tungstate at 60 °C according to the flow chemistry procedure we reported.21 DNF is a sensitive primary explosive. However, in the context of the synthesis shown here, it can be safely handled and stored diluted in anhydrous dichloromethane solution. Side chains were introduced by nucleophilic substitution with sodium and lithium alkoxides, eliminating a nitro group as the corresponding nitrite. NFOEt was prepared based on literature methods.22 The purity of the longer-chain compounds NFPEG2Me and NFPEG3N3 was verified using HPLC (S6). Non-commercially available alkoxides were synthesized by deprotonation of the alcohols with n-butyllithium in tetrahydrofuran. The deprotonation of the alcohols and the substitution reaction were performed at −94 °C. The prepared alkoxides were then used in the substitution reaction without further purification.
image file: d5ra01282a-s1.tif
Scheme 1 Synthesis of nitrofurazanyl ethers starting from DAF.

Physicochemical properties

An overview of the properties of the synthesized nitrofurazanyl ethers is given in Table 1, in comparison to conventional energetic plasticizers. The heat of explosion and gas volume generated were determined using the ICT-thermodynamic code.24 The compounds show a heat of explosion ranging from 3004 J g−1 to 3976 J g−1 and a generated gas volume ranging from 928 cm3 g−1 to 956 cm3 g−1, which are similar to energetic plasticizers in use. However, significant differences are observed for the glass transition temperature.
Table 1 Comparison of conventional energetic plasticizers in use to synthesized nitrofurazanyl ethers
Compound Tga [°C] Tmb [°C] TTGAc [°C] Tdecd [°C] ISe [J] ρf [g cm−3] Qxg [J g−1] OBCO2h [%] Ni [%] Vxj [cm3 g−1]
a Glass transition temperature measured by DSC.b Melting point measured by DSC.c Temperature at maximum mass loss rate measured by TGA.d Thermal decomposition temperature (onset) measured by DSC in pressure-tight crucibles.e Impact sensitivity measured by BAM drop hammer.f Density.g Heat of explosion calculated by ICT-thermodynamic code (water liquid).h Oxygen balance calculated on CO2.i Nitrogen content.j Gas volume calculated by ICT-thermodynamic code without H2O at 25 °C.k Physical properties of conventional energetic plasticizers partially from Schaller et al.14
BDNPA/Fk −67   182 207 3 1.39 3469 −57.6 17.6 957
DNDA-57k −52   159 221 3 1.35 3848 −72.3 30.9 1078
BuNENAk −82   152 173 6 1.22 3573 −104.3 17.4 1045
NFOEt   8.7 65 233 >40 1.33 3976 −65.4 26.4 928
NFOBu   −6.2 88 191 >40 1.21 3379 −106.9 22.5 940
NFPEG2Me −69   135 156 >40 1.30 3004 −92.6 18.2 956
NFPEG3N3 −72   167 179 >40 1.34 3421 −88.8 29.2 948


The glass transition temperature of the ethylene glycol type side chain-containing ethers NFPEG2Me and NFPEG3N3 are −69 °C and −72 °C, respectively (Fig. 2a). These are slightly higher than the glass transition temperature of BuNENA (−82 °C), but lower than that of BNDPA/F (−67 °C). In contrast, NFOEt and NFOBu do not exhibit a glass transition, but rather melting points at 8.66 °C and −6.19 °C, respectively. The thermal stability of the compounds was analyzed by DSC in pressure-tight stainless steel crucibles and TGA under nitrogen flow. The alkyl ethers showed significantly higher thermal stability than the ethylene glycol ethers in DSC but evaporated at lower temperatures as shown in the TGA (Fig. 2b). The exothermic onsets for NFPEG2Me and NFPEG3N3 in DSC were found to be 156 °C and 179 °C, respectively. The complex DSC curves suggest a multi-stage decomposition processes (S4). Until now we have no information about the exact mechanism of the decomposition of these substances. Compared to the reference plasticizers, NFPEG2Me shows a lower thermal stability while NFPEG3N3 outperforms BuNENA. These properties suggest that, depending on the side chain, the new materials could potentially be used as plasticizers in energetic formulations such as solid rocket propellants. They could also potentially be used as high-explosive plasticizers or in modern gun propellant formulations. The ethylene glycol type ethers have an oxygen balance slightly higher than BuNENA and NFPEG3N3 has a high nitrogen content of 29.2%, comparable to DNDA-57. In general, a compound with a higher oxygen balance tends to be more explosive, powerful, or sensitive.25


image file: d5ra01282a-f2.tif
Fig. 2 (a) Glass transition temperature curves obtained by DSC (b) thermal mass loss curves obtained by TGA.

The densities of the synthesized compounds are between 1.30 g cm−3 and 1.34 g cm−3, except for NFOBu which has a much lower density of 1.21 g cm3. NFPEG3N3 has a density similar to that of GAP-diol (1.34 g cm3) which ensures a good miscibility. The impact sensitivity was determined by using standard BAM techniques.26 All nitrofurazanyl ethers have been found to be non-impact sensitive within UN regulations (>40 J). As a result, they exhibit superior mechanical sensitivity compared to conventional energetic plasticizers. NFPEG3N3 has a low viscosity of 32 mPa s at 20 °C, ensuring adequate processability over a wide temperature range.

Infrared spectroscopy

The infrared spectra of the nitrofurazanyl ethers and DNF were analyzed, and specific group frequencies were identified (Fig. 3).27 While the 1,2,5-oxadiazole ring vibration is found at 1570 cm−1 for DNF, it shifts to 1560 cm−1 for the ethers. The asymmetric stretching of the nitro group also shifts from 1538 cm−1 to 1548–1544 cm−1. In contrast the symmetric stretching of NO2 shows no shift and is found at 1350 cm−1 for all compounds. Specific frequencies for an allylic ether occur at 1203 cm−1 and 828 cm−1 indicating the replacement of a nitro group by the alkoxy substituents. The characteristic N[double bond, length as m-dash]N stretching of the azide group is found for NFPEG3N3 at 2106 cm−1. The analysis shows that the IR spectroscopy allows rapid identification and reaction control in this case.
image file: d5ra01282a-f3.tif
Fig. 3 Infrared spectra of synthesized nitrofurazanyl ethers and 3,4-dinitrofurazan.

Calculation of the formation enthalpy

The standard enthalpy of formation of the synthesized nitrofurazanyl ethers in liquid-state was calculated from the gas-phase enthalpy of formation and the vaporization enthalpy using density functional theory calculations within the Gaussian 16 software package.28 The vaporization enthalpies were estimated using the semi-empirical methods of Politzer,29 Rice30 and improved Rice31 (Table 2). In our experience, for molecules without any enthalpy data in literature, an average of all three methods gives the best results as it provides the broadest empirical base. The methods are based on geometry optimizations and calculations of the ground state energy with the DFT functional/basis set combinations B3PW91/6-31G** and B3LYP/6-31G*. The improved method of Rice utilizes B3LYP/6-311++G(2df,2p) for calculation of the electronic energy of the ground state.
Table 2 Vaporization enthalpies calculated by the methods of Rice and Politzer
ΔHvap. [kJ mol−1] Politzer Rice Rice II Average
NFOEt 53.5 58.5 62.6 58.2
NFOBu 60.8 68.7 74.9 68.1
NFPEG2Me 69.7 80.5 88.4 79.6
NFPEG3N3 78.6 92.5 103.1 91.4


The functional/basis set combinations B3PW91/aug-cc-pVTZ and B3LYP/aug-cc-pVTZ were also used to find and validate the energetic ground state. Since the accuracy of the DFT methods is not sufficient for the calculations of the gas phase enthalpy of formation with the atomization method, geometry optimizations and ground state energy calculations were performed with the so-called composite methods CBS-QB3,32,33 G4 (ref. 34 and 35) and G4MP2 (ref. 35) (Table 3 and Fig. 4). Finally, the standard enthalpy of formation of the liquid phase was determined by combining results of the precise G4 calculations and average value of the three semi-empirical values for the vaporization enthalpy (Table 4). The standard enthalpy of formation of the liquid state for NFOEt, NFOBu, NFPEG2Me, and NFPEG3N3 were found to be −13.8 kJ mol−1, −72.7 kJ mol−1, −339.5 kJ mol−1, and −41.7 kJ mol−1, respectively.

Table 3 Gas-phase heat of formation calculated by composite methods
ΔfHm [kJ mol−1] CBS-QB3 G4 G4MP2
NFOEt 29.2 44.4 56.7
NFOBu −17.7 −4.5 7.7
NFPEG2Me −285.2 −260.0 −244.0
NFPEG3N3 26.8 49.7 70.0



image file: d5ra01282a-f4.tif
Fig. 4 Molecular structures of the synthesized nitrofurazanyl ethers geometry optimized using the G4 composite method.
Table 4 Standard heat of formation in the liquid phase
  ΔfHm(l) [kJ mol−1] ΔfHm(l) [kcal mol−1]
NFOEt −13.8 −3.31
NFOBu −72.7 −17.36
NFPEG2Me −339.5 −81.14
NFPEG3N3 −41.7 −9.98


Impact of NFPEG3N3 in mixtures with GAP-diol

In addition to the mechanical properties of the cured formulation, plasticizers are also critical to the processability of the uncured formulation. Therefore, we studied the viscosity change of mixtures of GAP-diol and NFPEG3N3 with ratios from pure GAP-diol to 40 wt% at 10 °C to 100 °C (Fig. 5a).
image file: d5ra01282a-f5.tif
Fig. 5 (a) Logarithmic representation of temperature and NFPEG3N3 concentration impact on the viscosity of GAP-diol. (b) Viscosity and glass transition temperature of GAP-diol with 10 wt% of NFPEG3N3 and selected plasticizers in use.

Newtonian behavior was observed for all measured compounds and mixtures. With increasing plasticizer concentration, the viscosity decreases at all measured temperatures, but each increase in plasticizer concentration has a smaller effect than the previous one. In direct comparison with other plasticizers at 10 wt% NFPEG3N3 shows a superior viscosity reduction of GAP-diol from 4560 mPa s to 2490 mPa s at 20 °C. This result is slightly better than BuNENA, which reduced the viscosity of GAP-diol to 2560 mPa s in our measurement (Fig. 5b). NFPEG3N3 also shows a competitive lowering of the glass transition temperature from −49.4 °C to −52.7 °C. At concentrations up to 40 wt% a linear decrease of the glass transition temperature compared to pure GAP-diol was observed (Fig. 6a). First results on the compatibility of GAP and NFPEG3N3 could be received by TGA measurements of mixtures and pure substances (Fig. 6b). Two separate decomposition peaks were found for all investigated mixtures. The first decomposition event can be attributed to NFPEG3N3 which shows a mass loss in TGA at 167 °C as a pure substance and in a 10 wt% mixture in GAP-diol. The second decomposition event could be attributed to the GAP diol and took place constantly at 231 °C without any influence of the presence or content of NFPEG3N3.


image file: d5ra01282a-f6.tif
Fig. 6 (a) Glass transition temperatures of pure GAP-diol and mixtures with NFPEG3N3 obtained by DSC. (b) TGA-curves of pure GAP-diol and mixtures with NFPEG3N3.

Conclusion

In this study, three new nitrofurazanyl ethers were synthesized in high yields. Theoretical calculations show that nitrofurazanyl ethers with ethylene glycol-type side chains have promising energetic properties, including high heat of explosion and large generated gas volume. All synthesized ethers were found to be non-impact sensitive within UN regulations.26 Among the synthesized compounds, NFPEG3N3 emerged as the most promising candidate for use as an energetic plasticizer. Remarkably, NFPEG3N3 outshines BuNENA with its superior oxygen balance, nitrogen content and decomposition temperature, while maintaining a low glass transition temperature of −72 °C and viscosity of 32 mPa s at 20 °C. In addition, NFPEG3N3 demonstrated a significant ability to reduce viscosity and glass transition temperature when formulated with GAP diol. NFPEG3N3 does not adversely affect the decomposition temperature of GAP-diol, even at concentrations up to 40 wt%. These results demonstrate for the first time the potential of nitrofurazanyl ethylene glycol ethers as advanced energetic plasticizers.

Experimental

Materials and instruments

BuNENA and BDNPA/F were purchased from Chemring. DNDA-57 was obtained from N. D. Zelinsky Institute of Organic Chemistry (Moscow, Russia). DAF was purchased from Chemicalpoint (Deisenhofen, Germany). 2-(2-(2-Azidoethoxy)ethoxy)ethan-1-ol (PEG3N3) was purchased from Abcr (Karlsruhe, Germany). GAP-diol charge 03S19 (MW 1814 g mol−1, EQ 1228 g mol−1, functionality 1.5) was supplied from Eurenco. Other chemicals were purchased from Sigma Aldrich, Merck or Carl Roth. Unless otherwise stated, all chemicals were used without further purification. A cooling bath from liquid nitrogen and acetone was used for low temperature reactions. 1H-NMR and 13C-NMR spectra were recorded on a 400 MHz Bruker AV-400 spectrometer. The melting point and glass transition temperature were recorded on a TA instruments Q1000 differential scanning calorimeter (DSC) at a heat rate of 10 °C min−1 using pierced aluminum crucibles. Glass transition temperatures were measured from heating up after cooling to −90 °C. Reported glass transition temperatures are inflection point temperatures. Reported melting points are peak temperatures. Measurements were carried out under nitrogen flux (25 mL min−1). The thermal decomposition temperature was measured on the same apparatus at a heat rate of 5 °C min−1 in pressure-tight steel crucibles (F20) purchased from the Swiss Institute for the Promotion of Safety. Thermogravimetric analysis (TGA) was performed on a TA Q500 apparatus with a heating rate of 5 °C min−1 in a platinum 100 μL pan under nitrogen flux (25 mL min−1). Reported values are the central points, according to DIN ENISO 11358. Infrared (IR) spectra were recorded on a Thermo Scientific iS 50 FTIR spectrometer in attenuated total reflection (ATR) mode. Raman spectra were recorded on a Bruker FT MultiRAM spectrometer at 1064 nm (Neodymium-YAG laser). Densities were determined by scale and volumetric flask. Elemental analyses were performed on a Thermo Flash EA with helium as carrier gas. Carbon, hydrogen and nitrogen were determined at a combustion temperature of 900 °C using a tin sleeve and oxygen at 1090 °C using a silver sleeve. High resolution mass spectra were recorded on a Thermo Fisher QExactive Plus spectrometer using electrospray ionization (ESI). Mechanical sensitivity was measured on a BAM drop hammer. A high-pressure liquid chromatography system Agilent 1200 equipped with a ZorbaxBonus RP 4.6 × 2500 mm column and a diode array detector was used to verify the purity of NFPEG2Me and NFPEG3N3. Gradient determinations were performed by injecting 5 μL acetonitrile diluted samples of the products. The mobile phase was acetonitrile/water with 0.1% trifluoroacetic acid (gradient from 20[thin space (1/6-em)]:[thin space (1/6-em)]80 to pure acetonitrile in 15 minutes) the flow rate was 1.0 mL min−1.

Quantum and thermochemical calculations

Quantum chemical calculations were performed with the software Gaussian 16.28 The molecular structures were calculated as singlets in their electronic ground states and verified as true minima by frequency calculations (no imaginary frequencies). The electrostatic potential of the optimized structures was investigated using the software MultiFWN.36,37 Thermochemical data were calculated using ICT-thermodynamic code.24

General procedure for the preparation of nitrofurazanyl ethers

General procedure 1: preparation of the alkoxide solution. To a solution of the alcohol in THF, 2.5 M n-butyllithium in hexane was added to that at −94 °C and under protective atmosphere. The reaction mixture was slowly warmed up to ambient temperature. The alkoxide solution was subsequently used for the substitution without further purification.
General procedure 2: substitution. To a solution of DNF in DCM, the alkoxide solution was added to that at −94 °C under protective atmosphere. The reaction mixture was slowly warmed up to ambient temperature. Then it was washed with water (2 × 300 mL) and the organic phase was dried over Na2SO4. The solvent was removed, and the residue was purified by basic alumina column chromatography using 25 vol% ethyl acetate–petroleum ether to deliver a pure product.
Synthesis of 3,4-dinitro-1,2,5-oxadiazole (DNF). DNF was prepared according to the literature.21 It was received as a colorless solution in dichloromethane (21.74 g L−1, 50%). IR (ATR, cm−1): 1570, 1538, 1451, 1351, 1325, 1137, 1031, 902, 839, 803, 769, 734, 615, 475. 13C NMR (400 MHz, CDCl3) δ [ppm] = 152.72 (t, J = 20.2 Hz).
Synthesis of 3-ethoxy-4-nitro-1,2,5-oxadiazole (NFOEt). NFOEt (440 mg, 73%) was prepared according to the literature and received as a colorless oil.22 Mp 7.1 °C. TGA 65 °C. Tdec 233 °C. Found: C, 30.15; H, 3.25; N, 26.4; O, 40.2%; M (mass spectrum), 159.0275. C4H5N3O4 requires C, 30.2; H, 3.2; N, 26.4; O, 40.2%; M, 159.0280. IR (ATR, cm−1): 2992, 1597 (furazan ring), 1545, 1498, 1471, 1446, 1425, 1393, 1353, 1298, 1204, 1170, 1156, 1114, 1093, 1033, 1011, 928, 889, 871, 828, 765, 741, 696, 593. Raman (1064 nm, cm−1): 2991, 2945, 2898, 2877, 2768, 2721, 1728, 1605, 1550, 1501, 1454, 1426, 1369, 1355, 1279, 1207, 1116, 1094, 1016, 928, 872, 831, 767, 742, 696, 598, 448. 1H NMR (400 MHz, CDCl3) δ [ppm] = 4.55 (q, J = 7.1 Hz, 2H), 1.55 (t, J = 7.1 Hz, 3H) ppm. 13C NMR (101 MHz, CDCl3) δ [ppm] = 158.19, 151.62 (t, J = 17.5 Hz), 70.80, 14.29. Impact sensitivity > 40 J. Friction sensitivity > 360 N. ρ 1.33 ± 0.02 g cm−3.
Synthesis of 3-butoxy-4-nitro-1,2,5-oxadiazole (NFOBu). The reaction has been performed by following general procedure 2.

20 wt% sodium butoxide in butanol (10.65 g, 22.11 mmol) and DNF (3.37 g, 21.06 mmol) in dichloromethane (180 mL) were used. The product was received as colorless oil (3.11 g, 79%). Mp −7.2 °C. TGA 86 °C. Tdec 191 °C. Found: C, 38.5; H, 4.9; N, 22.3; O, 34.1%; M (mass spectrum), 187.0591. C6H9N3O4 requires C, 38.5; H, 4.85; N, 22.45; O, 34.2%; M, 187.0593. IR (ATR, cm−1): 2964, 2938, 2877, 1598 (furazan ring), 1546, 1498, 1464, 1424, 1383, 1353, 1203, 1057, 1034, 1017, 994, 960, 942, 872, 857, 827, 765, 746, 592, 496. Raman (1064 nm, cm−1): 2940, 2918, 2877, 2743, 1603, 1550, 1500, 1453, 1426, 1356, 1303, 1263, 1232, 1204, 1147, 1127, 1059, 1036, 994, 961, 944, 906, 871, 832, 766, 696, 596, 438. 1H NMR (400 MHz, CDCl3) δ [ppm] = 4.48 (t, J = 6.5 Hz, 2H), 1.94–1.81 (m, 2H), 1.59–1.43 (m, 2H), 0.99 (t, J = 7.4 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ [ppm] = 158.39, 151.63, 74.59, 30.61, 18.90, 13.68. Impact sensitivity > 40 J. Friction sensitivity > 360 N. ρ 1.21 ± 0.02 g cm−3.

Synthesis of 3-(2-(2-methoxyethoxy)ethoxy)-4-nitro-1,2,5-oxadiazole (NFPEG2Me). The reaction has been performed by following the general procedures 1 and 2.

Diethylene glycol monomethyl ether (1.44 mL, 1.47 g,12.24 mmol), THF (5 mL), 2.5 M n-butyllithium in hexane (5 mL, 12.48 mmol) and DNF (1.96 g, 12.24 mmol) in DCM (90 mL) were used. The product was received as pale yellow liquid (2.31 g, 81%). Tg −69.2 °C. TGA 135 °C. Tdec 156 °C. Found: C, 34.95; H, 4.8; N, 17.9; O, 41.2%; [M + H]+ (mass spectrum), 234.0710. C7H11N3O6 requires C, 36.1; H, 4.8; N, 18.0; O, 41.2%; [M + H]+, 234.0726. IR (ATR, cm−1): 2883, 1599 (furazan ring), 1547, 1499, 1448, 1432, 1351, 1300, 1248, 1202, 1107, 1031, 983, 924, 892, 872, 828, 765, 743, 698, 591, 497. Raman (1064 nm, cm−1): 2947, 2895, 2830, 2742, 1604, 1550, 1501, 1472, 1446, 1431, 1368, 1287, 1243, 1205, 1132, 1032, 926, 872, 831, 766, 743, 699, 594, 441. 1H NMR (400 MHz, CDCl3) δ [ppm] = 4.63 (ddd, J = 5.8, 3.3, 1.2 Hz, 2H), 3.93 (ddd, J = 5.8, 3.4, 1.3 Hz, 2H), 3.74–3.67 (m, 2H), 3.57–3.51 (m, 2H), 3.36 (d, J = 1.2 Hz, 3H). 13C NMR (101 MHz, CDCl3) δ [ppm] = 158.40, 151.58, 73.75, 72.00, 71.03, 68.63, 59.16 ppm. Impact sensitivity > 40 J. Friction sensitivity > 360 N ρ 1.30 ± 0.02 g cm−3.

Synthesis of 3-(2-(2-(2-azidoethoxy)ethoxy)ethoxy)-4-nitro-1,2,5-oxadiazole (NFPEG3N3). The reaction has been performed by following the general procedures 1 and 2.

2-(2-(2-Azidoethoxy)ethoxy)ethan-1-ol (2.65 g, 15.15 mmol), THF (10 mL), 2.5 M n-butyllithium in hexane (6 mL, 15.2 mmol) and DNF (2.42 g, 15.15 mmol) in DCM (140 mL) were used. The product was received as yellow liquid (3.50 g, 80%). Tg −72.4 °C. TGA 167 °C. Tdec 179 °C. Found: C, 32.9; H, 4.2; N, 28.0; O, 33.45%; [M + H]+ (mass spectrum), 289.0882. C8H12N6O6 requires C, 33.3; H, 4.2; N, 29.2; O, 33.3%; [M + H]+, 289.0897. IR (ATR, cm−1): 2872, 2100 (–N3), 1599 (furazan ring), 1547, 1499, 1446, 1351, 1284, 1205, 1119, 1032, 923, 891, 871, 852, 827, 765, 743, 697, 645, 591, 557, 505. Raman (1064 nm, cm−1): 2944, 2875, 2096, 1604, 1549, 1501, 1472, 1446, 1430, 1368, 1286, 1247, 1126, 1033, 992, 923, 872, 831, 165, 742, 699, 645, 593, 440. 1H NMR (400 MHz, CDCl3) δ [ppm] = 4.61–4.54 (m, 2H), 3.93–3.86 (m, 2H), 3.71–3.64 (m, 2H), 3.68–3.58 (m, 4H), 3.31 (dq, J = 4.7, 2.4 Hz, 2H). 13C NMR (101 MHz, CDCl3) δ [ppm] = 158.29, 73.62, 71.03, 70.72, 70.17, 68.58, 50.69. Impact sensitivity > 40 J. Friction sensitivity > 360 N. ρ 1.34 ± 0.02 g cm−3.

Data availability

The data supporting this article, including infrared spectra, NMR spectra, mass spectra, DSC curves and HPLC chromatographs have been included as part of the ESI (S1–S6).

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

The authors are thankful to Dr D. Boskovic, Dr V. Gettwert, Dr T. Keicher and Dr A. Omlor for guidance and support throughout the study. We thank H. Schuppler for TGA and DSC analysis, Y. Kasimir for elemental analysis, S. Huber (LMU Munich) for mechanical sensitivity tests, U. Förther-Barth for rheological measurements, W. Schweikert for Raman and IR measurements, L. Hirsch (KIT Karlsruhe) for mass spectroscopy, T. Ohmer-Scherrer (KIT Karlsruhe) for NMR spectroscopy and M. Schwarzer for HPLC method development. The authors also would like to acknowledge the German Ministry of Defense for funding this work, and support provided by the WTD91.

References

  1. D. Kumari, K. D. B. Yamajala, H. Singh, R. R. Sanghavi, S. N. Asthana, K. Raju and S. Banerjee, Application of Azido Esters as Energetic Plasticizers for LOVA Propellant Formulations, Propellants, Explos., Pyrotech., 2013, 38, 805–809 CrossRef CAS.
  2. L. A. Wingard, E. C. Johnson, P. E. Guzmán, J. J. Sabatini, G. W. Drake, E. F. C. Byrd and R. C. Sausa, Synthesis of Biisoxazoletetrakis(methyl nitrate): A Potential Nitrate Plasticizer and Highly Explosive Material, Eur. J. Org Chem., 2017, 2017, 1765–1768 CrossRef CAS.
  3. R. S. Damse and A. Singh, Evaluation of Energetic Plasticisers for Solid Gun Propellant, Def. Sci. J., 2008, 58, 86–93 CrossRef CAS.
  4. S. Hafner, V. A. Hartdegen, M. S. Hofmayer and T. M. Klapötke, Potential Energetic Plasticizers on the Basis of 2,2-Dinitropropane-1,3-diol and 2,2-Bis(azidomethyl)propane-1,3-diol, Propellants, Explos., Pyrotech., 2016, 41, 806–813 CrossRef CAS.
  5. H. G. Ang and S. Pisharath, Energetic Polymers. Binders and Plasticizers for Enhancing Performance, Wiley-VCH-Verl, Weinheim, 1st edn, 2012 Search PubMed.
  6. S. K. Shee, S. T. Reddy, J. Athar, A. K. Sikder, M. B. Talawar, S. Banerjee and M. A. Shafeeuulla Khan, Probing the compatibility of energetic binder poly-glycidyl nitrate with energetic plasticizers: thermal, rheological and DFT studies, RSC Adv., 2015, 5, 101297–101308 RSC.
  7. V. Grakauskas, Polynitroalykyl ethers, J. Org. Chem., 1973, 38, 2999–3004 CrossRef CAS.
  8. K. Selim, S. Özkar and L. Yilmaz, Thermal characterization of glycidyl azide polymer (GAP) and GAP-based binders for composite propellants, J. Appl. Polym. Sci., 2000, 77, 538–546 CrossRef CAS.
  9. K. P. C. Rao, A. K. Sikder, M. A. Kulkarni, M. M. Bhalerao and B. R. Gandhe, Studies on n-Butyl Nitroxyethylnitramine (n-BuNENA): Synthesis, Characterization and Propellant Evaluations, Propellants, Explos., Pyrotech., 2004, 29, 93–98 CrossRef CAS.
  10. R. Vijayalakshmi, N. H. Naik, G. M. Gore and A. K. Sikder, Linear Nitramine (DNDA-57): Synthesis, Scale-Up, Characterization, and Quantitative Estimation by GC/MS, J. Energ. Mater., 2015, 33, 1–16 CrossRef CAS.
  11. Z. Xu, G. Cheng, H. Yang, X. Ju, P. Yin, J. Zhang and J. M. Shreeve, A Facile and Versatile Synthesis of Energetic Furazan-Functionalized 5-Nitroimino-1,2,4-Triazoles, Angew. Chem., Int. Ed. Engl., 2017, 56, 5877–5881 CrossRef CAS PubMed.
  12. M. A. Roux, M. Zeller, E. F. C. Byrd and D. G. Piercey, Synthesis and characterization of the energetic 5,7-diamino-2-nitro-1,2,4-triazolo[1,5-a]-1,3,5-triazine, Propellants, Explos., Pyrotech., 2023, 48(10), e20230012 CrossRef.
  13. J. Li, Y. Liu, W. Ma, T. Fei, C. He and S. Pang, Tri-explosophoric groups driven fused energetic heterocycles featuring superior energetic and safety performances outperforms HMX, Nat. Commun., 2022, 13, 5697 CrossRef CAS PubMed.
  14. U. Schaller, J. Lang, V. Gettwert, J. Hürttlen, V. Weiser and T. Keicher, 4-Amino-1-Butyl-1,2,4-Triazolium Dinitramide – Synthesis, Characterization and Combustion of a Low-Temperature Dinitramide-Based Energetic Ionic Liquid (EIL), Propellants, Explos., Pyrotech., 2022, 47(7), e202200054 CrossRef CAS.
  15. L. L. Fershtat and N. N. Makhova, 1,2,5-Oxadiazole-Based High-Energy-Density Materials: Synthesis and Performance, ChemPlusChem, 2020, 85, 13–42 CrossRef CAS.
  16. W.-L. Cao, Q.-N. Tariq, Z.-M. Li, J.-Q. Yang and J.-G. Zhang, Recent advances on the nitrogen-rich 1,2,4-oxadiazole-azoles-based energetic materials, Def. Technol., 2022, 18, 344–367 CrossRef.
  17. T. S. Hermann, T. M. Klapötke, B. Krumm and J. Stierstorfer, Synthesis, Characterization, and Properties of Di- and Trinitromethyl-1,2,4-Oxadiazoles and Salts, Asian J. Org. Chem., 2018, 7, 739–750 CrossRef CAS.
  18. A. B. Sheremetev, Efficient synthesis of nitrofurazans using HOF • MeCN, Russ. Chem. Bull., 2022, 71, 1818–1820 CrossRef CAS.
  19. T. S. Novikova, T. M. Mel’nikova, O. V. Kharitonova, V. O. Kulagina, N. S. Aleksandrova, A. B. Sheremetev, T. S. Pivina, L. I. Khmel’nitskii and S. S. Novikov, An Effective Method for the Oxidation of Aminofurazans to Nitrofurazans, Mendeleev Commun., 1994, 4, 138–140 CrossRef.
  20. K. Y. Suponitsky, A. F. Smol'yakov, I. V. Ananyev, A. V. Khakhalev, A. A. Gidaspov and A. B. Sheremetev, 3,4-Dinitrofurazan: Structural Nonequivalence of ortho -Nitro Groups as a Key Feature of the Crystal Structure and Density, ChemistrySelect, 2020, 5, 14543–14548 CrossRef CAS.
  21. P. Lieber, U. Schaller and T. M. Klapötke, in Proceedings of the 26th Seminar on New Trends in Research of Energetic Materials (NTREM), Pardubice, Czech Republic, 2024 Search PubMed.
  22. A. B. Sheremetev, O. V. Kharitonova, E. V. Mantseva, V. O. Kulagina, E. V. Shatunova, N. S. Aleksandrova and L. I. Khmel'nitskii, Nucleophilic substitution in the furazan series. Reactions with O-nucleophiles, Russ. J. Org. Chem., 1999, 35, 1525–1537 CAS.
  23. A. B. Sheremetev, V. O. Kulagina, I. A. Kryazhevskikh, T. M. Melnikova and N. S. Aleksandrova, Nucleophilic substitution in the furazan series. Reactions of nitrofurazans with ammonia, Russ. Chem. Bull., 2002, 51, 1533–1539 CrossRef CAS.
  24. S. Kelzenberg, P.-B. Kempa, S. Wurster, M. Herrmann and T. Fischer, New version of the ICT-thermodynamic code, 45th International Annual Conference of ICT – Energetic Materials Particles, Processing, Applications, Karlsruhe, 2014 Search PubMed.
  25. R. Haiges and K. O. Christe, Energetic high-nitrogen compounds: 5-(trinitromethyl)-2H-tetrazole and -tetrazolates, preparation, characterization, and conversion into 5-(dinitromethyl)tetrazoles, Inorg. Chem., 2013, 52, 7249–7260 CrossRef CAS PubMed.
  26. United Nations Economic Commission for Europe, Manual of Tests and Criteria, United Nations, New York, 8th rev edn, 2023 Search PubMed.
  27. G. Socrates, Infrared and Raman Characteristic Group Frequencies. Tables and Charts, Wiley, Chichester, 3rd edn, 2010 Search PubMed.
  28. M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R. Cheeseman, G. Scalmani, V. Barone, G. A. Petersson, H. Nakatsuji, X. Li, M. Caricato, A. V. Marenich, J. Bloino, B. G. Janesko, R. Gomperts, B. Mennucci, H. P. Hratchian, J. V. Ortiz, A. F. Izmaylov, J. L. Sonnenberg, D. Williams-Young, F. Ding, F. Lipparini, F. Egidi, J. Goings, B. Peng, A. Petrone, T. Henderson, D. Ranasinghe, V. G. Zakrzewski, J. Gao, N. Rega, G. Zheng, W. Liang, M. Hada, M. Ehara, K. Toyota, R. Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, T. Vreven, K. Throssell, J. A. Montgomery Jr, J. E. Peralta, F. Ogliaro, M. J. Bearpark, J. J. Heyd, E. N. Brothers, K. N. Kudin, V. N. Staroverov, T. A. Keith, R. Kobayashi, J. Normand, K. Raghavachari, A. P. Rendell, J. C. Burant, S. S. Iyengar, J. Tomasi, M. Cossi, J. M. Millam, M. Klene, C. Adamo, R. Cammi, J. W. Ochterski, R. L. Martin, K. Morokuma, O. Farkas, J. B. Foresman and D. J. Fox, Gaussian 16 Revision C.01, 2016 Search PubMed.
  29. P. Politzer, Y. Ma, P. Lane and M. C. Concha, Computational prediction of standard gas, liquid, and solid-phase heats of formation and heats of vaporization and sublimation, Int. J. Quantum Chem., 2005, 105, 341–347 CrossRef CAS.
  30. B. M. Rice, S. V. Pai and J. Hare, Predicting heats of formation of energetic materials using quantum mechanical calculations, Combust. Flame, 1999, 118, 445–458 CrossRef CAS.
  31. E. F. C. Byrd and B. M. Rice, Improved prediction of heats of formation of energetic materials using quantum mechanical calculations, J. Phys. Chem. A, 2006, 110, 1005–1013 CrossRef CAS PubMed.
  32. J. M. Simmie and K. P. Somers, Benchmarking Compound Methods (CBS-QB3, CBS-APNO, G3, G4, W1BD) against the Active Thermochemical Tables: A Litmus Test for Cost-Effective Molecular Formation Enthalpies, J. Phys. Chem. A, 2015, 119, 7235–7246 CrossRef CAS PubMed.
  33. G. A. Petersson and M. A. Al-Laham, A complete basis set model chemistry. II. Open-shell systems and the total energies of the first-row atoms, J. Chem. Phys., 1991, 94, 6081–6090 CrossRef CAS.
  34. S. Rayne and K. Forest, A G4MP2 and G4 Theoretical Study into the Thermochemical Properties of Explosophore Substituted Tetrahedranes and Cubanes, Propellants, Explos., Pyrotech., 2011, 36, 410–415 CrossRef CAS.
  35. L. A. Curtiss, P. C. Redfern and K. Raghavachari, Gaussian-4 theory using reduced order perturbation theory, J. Chem. Phys., 2007, 127, 124105 CrossRef PubMed.
  36. T. Lu and F. Chen, Multiwfn: a multifunctional wavefunction analyzer, J. Comput. Chem., 2012, 33, 580–592 CrossRef CAS PubMed.
  37. T. Lu, A comprehensive electron wavefunction analysis toolbox for chemists, Multiwfn, J. Chem. Phys., 2024, 161(8), 082503 CrossRef CAS PubMed.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra01282a

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.