Next Article in Journal
Pro- and Anti-Inflammatory Prostaglandins and Cytokines in Humans: A Mini Review
Previous Article in Journal
Bradykinin and Neurotensin Analogues as Potential Compounds in Colon Cancer Therapy
Previous Article in Special Issue
Mesoscale Simulations of Structure Formation in Polyacrylonitrile Nascent Fibers Induced by Binary Solvent Mixture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Energetic Polymer Possessing Furazan, 1,2,3-Triazole, and Nitramine Subunits

by
Pavel S. Gribov
1,
Natalia N. Kondakova
2,
Natalia N. Il’icheva
2,
Evgenia R. Stepanova
2,
Anatoly P. Denisyuk
2,
Vladimir A. Sizov
2,
Varvara D. Dotsenko
2,
Dmitry B. Vinogradov
1,
Pavel V. Bulatov
1,
Valery P. Sinditskii
2,
Kyrill Yu. Suponitsky
3,4,
Mikhail M. Il’in
3,
Mukhamed L. Keshtov
3 and
Aleksei B. Sheremetev
1,*
1
Zelinsky Institute of Organic Chemistry, Russian Academy of Sciences, 47 Leninsky Prosp., Moscow 119991, Russia
2
Mendeleev University of Chemical Technology, 9 Miusskaya pl., Moscow 125047, Russia
3
Institute of Organoelement Compounds, Russian Academy of Sciences, Moscow 119991, Russia
4
Basic Department of Chemistry of Innovative Materials and Technologies, Plekhanov Russian University of Economics, 36 Stremyannyi Line, Moscow 117997, Russia
*
Author to whom correspondence should be addressed.
Int. J. Mol. Sci. 2023, 24(11), 9645; https://doi.org/10.3390/ijms24119645
Submission received: 12 May 2023 / Revised: 28 May 2023 / Accepted: 30 May 2023 / Published: 1 June 2023

Abstract

:
A [3 + 2] cycloaddition reaction using dialkyne and diazide comonomers, both bearing explosophoric groups, to synthesize energetic polymers containing furazan and 1,2,3-triazole ring as well as nitramine group in the polymer chain have been described. The developed solvent- and catalyst-free approach is methodologically simple and effective, the comonomers used are easily available, and the resulting polymer does not need any purification. All this makes it a promising tool for the synthesis of energetic polymers. The protocol was utilized to generate multigram quantities of the target polymer, which has been comprehensively investigated. The resulting polymer was fully characterized by spectral and physico-chemical methods. Compatibility with energetic plasticizers, thermochemical characteristics, and combustion features indicate the prospects of this polymer as a binder base for energetic materials. The polymer of this study surpasses the benchmark energetic polymer, nitrocellulose (NC), in a number of properties.

1. Introduction

The need to improve efforts to synthesize components of energetic materials is obvious since the requirements for such materials that are promising for future use are increasing from year to year. Explosives, propellants, and powders are usually multicomponent energetic materials. For instance, a multicomponent system of solid composite propellants (SCPs) usually includes an oxidizer (50–80%, oxygen-rich inorganic and/or organic compounds), combustible (up to 25%, high-calorific metals or their alloys, metal hydrides, boron and its derivatives, high nitrogen compounds, etc.), binder (10–25%, plasticized polymers), catalysts, and processing aids [1,2,3,4,5,6,7]. All these components, complementing each other, should provide the required effect—the development of the maximum thrust and the maximum increase in the speed of the aircraft in the process of propellant burning, ensuring the delivery of the certain mass to the maximum possible distance. The extraordinary promise of SCPs for space, military, and civilian applications have led to a rapid growth of innovative research in this area.
The improvement of binder components for powders and SCP is an important area of research. Back in the 1970s, it was shown that higher propellant efficiency indicators can be achieved if active oxygen is delivered not only by an oxidizer, but also contained in a binder, in a polymer, or plasticizer of which, for example, bears nitroxy groups [8,9,10]. The expediency of introducing high-enthalpy moieties into polymers, such as an azide group and other explosophoric units also became obvious [11,12,13,14,15,16,17].
Binders created on the basis of a polymer and a plasticizer enriched with explosophoric groups [18] gave a significant increase in the specific impulse for metallized SCPs. Such components are even more in demand for propellants based on solid high-enthalpy high-density, but oxygen-poor fillers, CxHyNzOw, for which the oxygen coefficient, α = w/(2x + 0.5y), is significantly less than 1, as well as for propellants containing aluminum hydride [19] or boron-containing fuel [20].
When screening new binder components for modern powders and SCPs, their characteristics such as enthalpy of formation, ΔH0f, coefficient of supply with oxidizing elements, α, hydrogen content, %H, and density, ρ, are being analyzed. Of course, these four characteristics are interdependent and, usually, the improvement (growth) of one of them leads to the deterioration (decrease) of the others. For example, it is difficult to increase the proportion of oxygen in an oxidant without reducing its formation enthalpy.
Academic and industrial laboratories around the world are trying to clarify which components, in what combination and ratio, should be used to create effective powders and SCPs, and they are getting closer, step by step, to the cherished goal. However, there are still many unresolved issues.
The creation of a binder with an optimal chemical composition, which, in combination with an oxidizer and other components, would be suitable for processing both an uncured propellant mass and a hardened charge with the desired physical and mechanical properties, is a very difficult task.
As is known, to ensure sufficient rheological properties of the uncured propellant mass, where solid fillers are aluminum, inorganic and organic oxidizers, or gas-generating components, and to impart satisfactory physical and mechanical properties of the cured charge, the required minimum volume content of the binder should be ca. 19% [21]. Evidently, such a percentage of the binder has a significant impact on the energy performance of SCP. Energetic polymers are making a huge impact on development of new energetic materials due to their potential to improve performance.
While a variety of high-enthalpy heterocyclic building blocks have been widely used in the creation of low-molecular-weight energetic compounds in recent years, only tetrazole building blocks are used relatively widely in the construction of energetic polymers [22,23,24,25]. The use of other azoles as structural subunits of the polymer chain in the chemistry of energetic compounds is still used much less frequently.
It is obvious that the introduction of high-nitrogen subunits and explosophoric groups into the polymer can have a profound effect on the energy and physicochemical properties; therefore, synthesis methods and structural modifications of the polymer must be carefully designed to be effective.
Furazans (1,2,5-oxadiazoles) occupy a privileged position in energy compound chemistry and appear in multiple blockbuster low-molecular-weight energetic materials [26,27,28,29]. However, methods for the synthesis of energetic polymers that would incorporate the furazan subunit are rare, and only a few examples of polymerization processes leading to them have been reported by us [30,31] and others [32] previously. Scheme 1 shows known representative energetic polymers bearing the furazan subunit.
However, these methods are either difficult to scale, or the target polymers had an unsuitable complex of physicochemical properties. A detailed description of both the preparation and properties of polymers 1–3 has not been reported in the public literature.
At the same time, the incorporation of furazan ring in the plasticizer molecules had a significant impact on the development of energetic binders [33].
The introduction of both polar and nonpolar groups into the polymer favors the polymer–plasticizer and polymer–filler interactions, which are very important for regulating the properties of the solid composite propellant being created. Only with a wide range of polymers and plasticizers at hand, there is a chance to create a suitable binder.
Recently, while developing new effective components of energetic binders, we have reported a series of energetic plasticizers that combine azido and nitramino groups on a dialkyl ether backbone, which have some specific properties realized due to the synergistic effect of the included functional moieties.
In our ongoing research focused on the development of new approaches leading to the formation of energetic molecules with combined functionality, here we report on the simple syntheses and characterization of energetic polyether, incorporated furazan and 1,2,3-triazol subunits, and a nitramine bridge. This structure will ensure the compatibility of the final polymer with other powder and SCP components, such as nitramines (for example, ammonium dinitramide (ADN), octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine (HMX), or hexanitrohexaazaisowurtzitane (CL-20)) or modern high-nitrogen energetic compounds.
Tasked with developing a practical route for the preparation of a densely functionalized energetic polymer, we were attracted by the possibility of constructing a polymer chain via a simple azide–alkyne cycloaddition using available comonomers. A survey of the literature revealed a number of examples of the use of this [3 + 2] cycloaddition reaction for the construction of energetic polymers [34,35,36]. While for the preparation of the target 1.2.3-triazole-based polymer, the initial monomer may contain both azido and alkyne moieties, the use of two comonomers, one of which contains two azido groups and the other two alkyne moieties, is more preferable for controlling the polymerization process.
To the best of our knowledge, there are no reports describing the use of any furazan derivatives in azide–alkyne cycloaddition polymerization. Moreover, we have not found in the literature any mention of the possibility of using diazido- and dialkynyl-comonomers, so that both contain explosophoric groups or units. Recently, the azide–alkyne cycloaddition is most often realized under copper catalysis, CuAAC reaction [37,38,39,40,41]. However, energetic polymers are used to create propellants that are designed for the combustion process. Copper derivatives have a great influence on the combustion process [42,43,44,45]. Since the copper content may be different in different batches of the polymer, this may lead to an uncontrolled change in the ballistic characteristics of the powders and SCPs containing this polymer. Therefore, the use of CuAAC in the production of polymers is undesirable in this case.
Herein, a copper-free thermally promoted azide–alkyne cycloaddition has been exploited to form a new energetic polymer. Such a protocol could be a useful tool for the simple development of promising energy binder components to meet diverse applications.
The resulting triazole–furazan-based nitraminopolymer has been studied both in terms of its energy properties and for its ability to plasticize, giving acceptable physicochemical properties to the binder.

2. Results and Discussion

2.1. Synthesis and Characterization of Monomers and Polymer

When an energetic compound can be synthesized simply and safely, and the starting materials are available and inexpensive, it is attractive for practical application. With this in mind, designing such bimodal monomers, wherein two molecules that serve two different functions can react with each other without the use of a catalyst and a solvent, is an ideal approach for producing energetic polymers. This protocol has a high atom economy, since all comonomers atoms are entered into the target polymer.
As shown in Scheme 2, the diazide monomer, 3,4-di(azidomethyl)furazan (7), was generated in three steps from commercially available dimethylglyoxime (4). Compounds 5 and 6 were prepared based on our previous report [46]. The nucleophilic azidation of (halogenomethyl)furazans is well known [47,48], and thus, diazide 7 was formed in quantitative yield by using three equivalent of NaN3 in acetone with 6 at room temperature.
Our second task was to devise a practical approach toward a suitable alkynyl-monomer-bearing nitramine groups. To the best of our knowledge, there is only one report [49] on the synthesis of dipropargylic ether, the bridge group of which contained nitramino-units, namely 1,6-di(2-propyn-1-yloxy)-2,5-dinitro-2,5-diazahexane (9), which was selected as a starting point in the creation of an energetic polymer. However, neither synthetic details nor fully characterization for this dipropargylic ether are given in the patent [49]. We, therefore, have investigated the chlorine displacement in N,N′-bis(chloromethyl)nitramine 8 with propargyl alcohol. To establish optimal conditions, we used readily available 1,6-dichloro-2,5-dinitro-2,5-diazahexane (8a) [50] and propargyl alcohol as our exemplar reagent set (Scheme 3).
After an extensive survey of ratio of reactants, solvents, catalysts, and temperature, there are several important features to note. First, a solvent and a catalyst are not required for the successful course of this reaction. Secondly, propargyl alcohol serves both as a solvent and a reagent when using an excess of 20 equivalents. After the reaction was completed, the excess of propargyl alcohol was removed by vacuum distillation; thus, this alcohol was recycled and re-used. Thirdly, during the reaction, dry nitrogen should be slowly passed through the reaction mixture to remove the HCl released. Finally, we were able to reduce the reaction time to 1 h (vs. 21 h in the patent [49]) and improve the yield (85% vs. 68% in the patent [49]) at room temperature. The best reaction conditions found in this screening resulted in a higher yield of product 9 using 20 eq of propargyl alcohol, a reaction temperature of 20 °C, and a short reaction time (1 h), and scales well in kilogram quantities.
Under similar conditions, monoether 10 was prepared in a quantitative yield.
Dipropargyl ether 9 is colorless solid, and monoether 10 is slightly yellow oil. These ethers can be stored at room temperature in the absence of light for a long time.
All compounds are well characterized by IR, 1H, 13C, and 14N NMR spectroscopic data as well as CHN analysis (see Supplementary Materials). The structure of dipropargyl ether 9 was unambiguously confirmed by X-ray crystallography (Figure 1, Supplementary Materials: Figures S1 and S2, Tables S1 and S2). Compound 9 crystallizes in space group P21/c with a calculated density of 1.475 gcm−3 at 20 °C. An asymmetric unit cell contains half of the molecule which is located at the symmetry center. The nitro groups are oriented pseudo-trans relative to each other so that the pseudo-torsional angle N2-N1 … N1A-N2A is equal exactly to 180° (Figure 1). The propargyl groups are oriented “inside” the molecule and form an angle of 21.5(2)° with the planes of the corresponding nitramino groups. The C2 … N2 (C2A … N2A) interatomic separation (3.131(2)Å) is less than sum of van-der-Waals atomic radii (3.36Å [51]), which might imply π … π interaction between propargyl and nitramino groups.
To clarify this point, we optimized the isolated molecule at the M052X/def2tzvp approximation level (for more details, see Supplementary Materials). The optimized molecular conformation is almost the same as observed experimentally. Topological analysis of the calculated electron density (using AIM theory [52,53]) revealed a critical bond point between C2 and N2 (C2A and N2A) atoms, which indicates an attractive noncovalent interaction between propargyl and nitramino groups which contribute to the stabilization of the observed conformation.
Using combustion calorimetry, it was found that diazide 7 and dipropargyl ether 9 have enthalpy of formation +781.6 and 43.5 kJ/mol [54], respectively. Previously, only calculated values of the enthalpy of formation for any propargyl ethers bearing explosophoric groups were published [35,55].
Before investigating the dipolar cycloaddition polymerization process, we needed to establish the effect of explosophoric units, the furazan ring, and the nitramine group on the reactivity of our monomers. Toward proving that the above energetic monomers are suitable for alkyne–azide cycloaddition chemistry, we examined the model reaction of diazide 7 with monopropargyl ether 10 under different conditions, expecting to obtain molecules containing 1,2,3-triazole rings, which is demonstrated in Scheme 4. We started our study with copper(I)-catalysed cycloaddition of an alkyne and azide (CuAAC), as this is known selectively provides 1,4-substituted-1,2,3-triazoles [37,38,39,40,41]. The cycloaddition between diazide 7 and acetylene 10 (2 equiv.) in N,N-dimethylformamide (DMF) proceeded to completion within 6 h (1H NMR control) at room temperature upon the addition of a catalytic amount of CuSO4·5H2O (0.05 equiv.) and sodium ascorbate (0.1 equiv.) to give ditriazole 11a as the only the product in good isolated yield (analytically pure sample, 69%).
In the absence of a catalyst, the reaction between diazide 7 and acetylene 10 in refluxing 1,2-dichloroethane (DCE, 83 °C) required 90 h to achieve full consumption of the reagents. The catalyst-free reaction gave an inseparable mixture of three regioisomeric di-1,2,3-triazoles 11a, 11b, and 11c in a ratio of 5:6:2, as observed by the 1H NMR analysis of the crude product which displayed, for example, two distinctive peaks (8.26 and 8.24 ppm) for the 1,4-substituted isomers and one peak (7.83 ppm) for the 1,5-isomer (see, Figure 2), the chemical shifts being consistent with those reported [56]. The regiochemical assignments were based both on chemical shifts in the pure isomer 11a and on a combination of HSQC spectroscopy and NMR correlation (for more details, see Supplementary Materials).
Since the alkyne–azide cycloaddition chemistry is suitable for binding two small molecules both bearing explosophoric groups together to form a new compound with the required functionality, we tried to apply this approach in the field of energetic materials to synthesize a new energetic polymer.
With the bifunctional precursors 7 and 9 in hand, we proceeded to investigate conditions for the cycloaddition polymerization. In a typical experiment, diazide 7 undergo the reaction with one equivalent of diacetylene 9 only by heating at a certain temperature without using a solvent or catalyst (Scheme 5).
The 1,2,3-triazole rings containing products were characterized by gel permeation chromatography (GPC) and 1H NMR methods without any workup or further purification. Performing the reaction at 30 °C for 60 h led to very low conversion of the starting monomers. At 60 °C, as shown in Table 1, the molecular weight of the product depends on the reaction time; however, a further increase in the duration of the reaction to 100 h did not lead to a noticeable increase in the molecular weight of the polymer 12. The polydispersity (ÐM) values are in range from 1.42 to 2.44, increasing with the increase in molecular weight. Nevertheless, under these conditions, the average molecular weight remains low.
When the reaction was gradually heated to 80 °C and kept at this temperature for 37 h, a light yellow plastic product 12 in quantitative yield was obtained. Purification of the resulting polymer from the residues of low-molecular-weight impurities was achieved by double precipitation from the DMSO solution with methanol. A colorless powder of 12 was obtained in 89% yield. Relatively high molecular weight was achieved that possessed modest dispersity (Table 1, Entry 6, and Supplementary Materials: Figure S3).
The FTIR spectrum of polymer 12 was compared with these of the initial comonomers 7 and 9 and the model compound 11. Figure 3 clearly demonstrates the pertinent features and differences of these spectra. The spectrum of diazide 7 shows the presence of azido groups at ca. 2111 cm−1. In the FTIR spectrum of the diacetylene compound 9, two characteristic absorption bands at ca. 3290 and 2117 cm−1 can be attributed to C≡CH groups, whereas absorption bands for nitramino groups were observed at ca. 1550 and 1350 cm−1. The spectrum of model compound 11 showed the indicative loss of the aforementioned acetylene and azide bands, as a result of complete 1,3-dipolar cycloaddition. A similar trend was observed in the FTIR spectrum of polymer 12; here, however, there is a weak residual band at ca. 2115 due to the presence of terminal C≡CH and N3 groups.
Similar to the reaction that took place during the synthesis of model compound 11, the catalyst-free polymerization leads to the formation of a product including regioisomerically substituted 1,2,3-triazole subunits. As outlined in Figure 4, the 1H NMR spectrum of the resulting polymer 12 clearly showed the asymmetry that is present in the product. In particular, this spectrum displayed the appearance of three separate signals characteristic of a proton in the triazole ring at δ ca. 7.7–8.4, confirming polymer formation. This is noteworthy; the ratio of isomeric 1,2,3-triazole subunits is almost the same as in the model compound (see, Figure 2). The residual terminal alkyne proton is also clearly visible in the spectrum at δ ca. 3.6 ppm (for comparison, see Supplementary Materials, 1H MNR of 9).
The 13C NMR spectrum (see, Supplementary Materials) further confirms the conclusions made on the basis of FTIR and 1H NMR. This spectrum of polymer 12 displayed the disappearance of characteristic alkyne moiety signals at δ ca. 77.5 and 79.5. This gives reason to conclude that most of the azide and alkyne groups of monomers 7 and 9 have been successfully spent on the formation of a triazole ring. The 13C NMR spectrum confirmed the formation of the 1,2,3-triazole subunit with the appearance of the signal at 125.1 ppm (CH of the 1,5-isomer) and 133.9 ppm (CH of the 1,4-isomer) which is indicative of the heterocycle [56]. The 1H and 13C NMR shifts illustrated in Figure 2 and in Supplementary Materials also support the assignment of proton and carbon shifts in the polymer 12.
Taking into account the enthalpies of the formation (ΔHf0) for comonomers 7 and 9 (see above), as well as the fact that the cycloaddition reaction is very exothermic (between −209 and −272 kJ mol−1 [57]), the theoretical value for ΔHf0 of polymer 12 is +1.4 kJ g−1, indicating that 12 has a higher energy content than NG (ΔHf0 = −2.2 kJ g−1 [18]).

2.2. Thermal Analysis

According to differential scanning calorimetric (DSC) and thermogravimetric analysis (TGA) measurements (scanning at 10 °C min−1, Figure 5 and Table 2) compound 9 melted with sharp endothermic peaks at 83 °C and began to decompose at 237 °C. The total energy of decomposition was 2212 J g−1. The temperatures of the beginning of the mass loss and the intensive decomposition of 9 coincide, indicating that the mass loss is due to decomposition, not evaporation. When heated, diazide 7 evaporates almost completely to a temperature of 200 °C, i.e., before decomposition begins (Table 2).
For polymer 12, DSC revealed a glass transition (Tg) near 48 °C, as well as a small endothermic peak (~6 J g−1) at 58 °C, associated with melting. However, the most striking observation from the DSC for this polymer was that a sample with Mw = 6747 g mol−1 and a sample with a higher molecular weight Mw = 42,400 g mol−1 have the same glass transition temperature. It is important to point out that polymer 12 has a significantly lower glass transition temperature than the benchmark energetic polymer, nitrocellulose (NC) (at 11.9% N, Tg = 160 °C [58]). The TGA plot in Figure 5b shows a mass loss due to decomposition commencing at 248 °C, a significant mass loss occurring as the temperature rises to Tpeak, and a loss of 60% of the original mass at 360 °C. The data in Table 2 demonstrate that the thermal stability of polymer 12 is higher than that of NC.
DSC thermograms for the decomposition of polymer 12 at different heating rates (from 5 to 20 °C min−1) are depicted in Figure 6. As indicated in Table 3, the apparent decomposition rate constants (k) of 12 that were obtained by the Kissinger method [59] are in accordance to Equation k = 4.66·1014·exp(−21,155/T), with activation energy Ea = 175.9 kJ mol−1.
The polymer 12 contains three units that can provoke the initial stage of decomposition, namely the nitroamine group, triazole, and furazan rings. It is possible to estimate which unit is responsible for the thermal stability of the polymer based on the kinetic parameters of the decomposition of these units.
For comonomer 9 bearing nitramino groups, thermal decomposition under isothermal conditions in the temperature range from 170 to 210 °C was carried out in thin-walled glass manometers of the compensation type, the glass Bourdon gauge (see Supplementary Materials, Table S4). This nitramine decomposed in the melt with weak acceleration. The gas emission curves can be described by a first-order model with linear autocatalysis [60]. This makes it possible to obtain (i) the initial decomposition rate constant (k1), characterizing the destruction of the nitramino group, and (ii) the acceleration rate constant (k2), which is probably due to the interaction of decomposition products (nitrogen oxides) with the initial 9. As can be seen from Figure 7 the decomposition rate constants of polymer 12 fall on a straight line, which is a continuation of the straight line describing the initial decomposition rate of nitramine 9 (k1).
The kinetic parameters of the decomposition of polymer 12 and 3,4-dimethylfurazan (DMF) [61] can be seen in Figure 7. As can be seen from the comparison, the decomposition rate of DMF is more than an order of magnitude slower than that of the polymer, which suggests that the furazan ring does not decompose under DSC conditions. At the same time, taking into account the published data [62,63], the thermal stability of 1,2,3-triazole ring is comparable to the stability of polymer 12. Summarizing the above, it can be assumed that the initial stage of decomposition of polymer 12 includes the decomposition of both 1,2,3-triazole units and nitramine groups.
Comparison of kinetic data of decomposition of NC (13.9% N) [64] and polymer 12 (Figure 7), shows that the decomposition rate of 12 is almost three orders of magnitude less than NC.

2.3. Plastification

It is well known that the introduction of a plasticizer into a polymer leads to a decrease in the glass transition temperature and an improvement in technological characteristics. Figure 8 demonstrates the appearance of the resulting polymer 12 before and after thermostating at ~60 °C. At room temperature, the polymer is a solid and after warming up, it turns into a viscous resin.
Nitroglycerin (NG) is usually used as a traditional plasticizer for NC. Both of these components and compositions based on them were benchmarked in this study. Figure 9 shows a DSC plot for a mixture of NC (12.2% N) with NG in a 50:50 ratio. The glass transition temperature (Tg) of this mixture is −52 °C, and there are no signs of phase decay.
Polymer 12 is well compatible with NG. In Figure 10, the DSC data indicate that Tg of the mixtures of polymer 12 with NG range from −52 to −12 °C depending on the ratio of components; thus, only at 70% NG in the mixture it is possible to reach the glass transition temperature inherent to the NC/NG mixture with a 50:50 ratio. The limit of compatibility of polymer 12 with diethylene glycol dinitrate (DEGDN) is 20% (12: DEGDN = 80:20, Tg = 2 °C, Figure 11).
Not only nitroesters such as NG and DEGDN, but also organic azides [2,65,66] such as Z1, Z3, Z8, and Z12 (Figure 12) were used to plasticize polymer 12. The use of azide plasticizers allows to reduce the combustion temperature of the binder including them in comparison with compositions based on nitro ester plasticizers.
Azide plasticizers Z1, Z3, and Z8 are well compatible with polymer 12 in a 50:50 ratio (Figure 13). Only one relaxation transition associated with the glass transition process of the plasticized composition was recorded on all profiles. At the same time, polymer 12 is practically incompatible with azide Z12; after 5 days of thermostating at +70 °C of the polymer–plasticizer mixture, the mass of polymer 12 increased by only 3.67%.
For all of the above mixtures, only the relaxation transition caused by the glass transition process of plasticized compositions was recorded on thermograms. The results are summarized in Table 4. It should be noted that the glass transition temperatures of mixtures 12/Z1 (50:50) and NC/NG (50:50) are practically equal. The glass transition temperatures of mixtures 12/Z3 and 12/Z8 are 13 and 9 °C, respectively, higher than that of the NC/NG composition (50:50).
The polymer composition 12/NG (60:40) is practically unable to flow at room temperature; the polymer 12, containing up to 60% NG, is a highly viscous resin, and when the NG content increases to 70%, a viscous liquid is formed (Figure 14)
As expected, with an increase in the plasticizer content in the polymer, the viscosity of the composition decreases (Figure 15, Table 5). The viscosity of polymer 12 containing 60 and 70% NG practically does not depend on the shear stress, which indicates the destruction of intermolecular interactions between polymer chains. With a content of 40% plasticizer in the composition, the interaction between the polymer macromolecules is still preserved: the viscosity decreases with increasing shear stress.

2.4. Combustion

The burning rate (rb) and its dependence on pressure are key parameters that need to be available when creating new rocket propellants for various purposes [1,3,67,68]. The accumulation and analysis of experimental data on burning rates and temperature distribution [69] makes it possible to evaluate and, in the long term, influence the kinetics of decomposition of propellant components burning in the condensed phase.
Burning rate was determined in a constant pressure bomb with a volume of 2 l in a nitrogen atmosphere. The process was recorded using a pressure strain gauge and a digital oscilloscope. Both a high-molecular polymer (Mw = 42,400 g mol−1) and its low-molecular analogue (Mw = 6747 g mol−1) were used for the study. Charges for determining the burning rate were prepared by pouring the polymer 12 presoftened at 75–80 °C into transparent polycarbonate tubes with an inner diameter of 7.5 mm, an outer diameter of 11 mm. As comparison samples, charges from both comonomer 9 and NC were prepared by blind pressing at a pressure of 450 MPa.
The results of these combustion experiments are shown in Figure 16, and data on the laws of combustion are summarized in Table 6. A cursory inspection of the data presented in Figure 16 and Table 6 demonstrates that the burning rates of high-molecular polymer 12 and NC is almost the same. Polymer 12 is a non-volatile material with a relatively low combustion temperature and burning rate is undoubtedly determined by the heat release in the condensed phase.
At a pressure of 10MPa, the burning rate of the comonomer 9 turned out to be very low. This is typical for the combustion of aliphatic nitramines [70], which is explained by their high thermal stability. Low burning rates imply a long residence time of the compounds at boiling points and, accordingly, a large contribution of the heat release in the molten phase in the control of the combustion. In this case, pressure exponent in the burning rate law is determined by the ratio of the activation energy of the leading combustion reaction and the enthalpy of evaporation, which is responsible for the temperature change with increasing pressure in this zone. The higher burning rate of the polymer 12 compared to that of the comonomer 9 is undoubtedly due to the higher surface temperature of the polymer, which is most likely due to less evaporation. This assumption is supported by a lower burning rate of a low-molecular polymer (Figure 16 and Table 6).
From a practical point of view, reducing the pressure exponent n is highly desirable; the lower n, the more stable the operation of the rocket engine is in relation to pressure fluctuations. Table 6 shows clearly that plasticization of polymer 12 with NG reduces the pressure exponent n from 0.93 to 0.6. More significantly, the effect of NG on polymer 12 is significantly higher than on NC (Figure 17). The burning rate of 12/NG is higher than that of NC/NG.

2.5. Gunpowder

The resulting polymer 12, as a potential heat-resistant replacement for NC, for example, can be used in gunpowders. The gauge for rating the efficiency of gunpowders is powder force (F). F is the work that could be conducted by gaseous products formed at the consumption of one kilogram of gunpowder, freely expanding at atmospheric pressure as a result of their heating from 273 K to the combustion temperature (Tc) [3,71,72]. It is important to note that the higher the combustion temperature, the more destructive the effect of outgoing hot gases on metal equipment, whether it is a gun barrel, a rocket nozzle, etc. Calculated parameters of NC/NG and 12/NG compositions are summarized in Table 7.
As can be seen from Table 7, according to the thermodynamic calculation, the compositions 12/NG (50/50) and NC/NG (70/30) are comparable in magnitude to F, whereas the combustion temperature of 12/NG is better (755 K lower) compared NC/NG. A more favorable Tc is provided by a higher nitrogen content in polymer 12 compared to NC.

3. Materials and Methods

IR spectra were recorded on a BrukerALPHA instrument in KBr pellets. The 1H, 13C, and 14N spectra were recorded on a Bruker AM-300 instrument (300.13, 75.47, and 21.69 MHz, respectively) at 299 K. The chemical shifts of 1H and 13C nuclei were reported relative to TMS, for 14Nrelative to MeNO2, high-filed chemical shifts are given with a minus sign. Elemental analysis was performed on a PerkinElmer 2400 Series II instrument. Analytical TLC was performed using commercially pre-coated silica gel plates (Kieselgel 60 F254), and visualization was affected with short-wavelength UV light.
Gel permeation chromatography (GPC) measurements of molecular weights were performed in N-methylpyrrolidone using a Smartline HPLC Series KNAUER system and calibrated with polystyrene standards.
Thermal stability, relaxation, and phase transitions were studied by differential scanning calorimetry (DSC) using a Mettler Toledo DSC 822e module. Approximately 2 mg of compounds were weighed and placed into in a 40 µL aluminum crucible, sealed under air with the appropriate sample press, and then pierced with a needle to leave one hole of approximately 1 mm in diameter. The decomposition of a sample was carried out in a nitrogen atmosphere at a purge rate of 50 μL min−1. The temperature of the onset of intense decomposition (Tonset) was taken as the temperature determining thermal stability. To study relaxation and phase transitions, the samples were uncontrollably cooled to −130 °C and then heated at a rate of 10 °C min−1. The temperature of the midpoint of the relaxation transition was taken as the glass transition temperature (Tg). The glass transition process is accompanied by a change in the heat capacity of the sample ΔCp, which was also measured. The melting point (Tm) of the individual compounds was determined as the temperature of the melting effect start point.
Thermal degradation was quantified using a TGA 822e Mettler Toledo (TA instruments). Thermograms were recorded under an N2 atmosphere at a heating rate of 10 K min−1 from 25 to 400 °C.
The burning rate was determined in a constant pressure device (Crawford bomb) with a volume of 2 L in a nitrogen atmosphere. Liquid compounds were mixed with 4% nitrocellulose (colloxylin 12% N). The dissolution of nitrocellulose was carried out from 1 to 2 h at 50–60 °C into transparent acrylic tubes of 7 mm i.d. The combustion process of the sample was recorded using a pressure strain gauge, which transmitted the signal to a digital oscilloscope. The start and end times of combustion were determined from oscillograms. The burning rate was calculated by dividing the sample height by the burning time and was related to the mean integral pressure during the experiment. The error in determining the burning rate does not exceed 3%.
Most of the reagents and starting materials were purchased from commercial sources and used without additional purification. The starting compound 6 [46] and 8a,b [50] were synthesized by using previously reported procedures.
Caution! Although we have encountered no difficulties during preparation and handling of these compounds, they are potentially explosive energetic materials. Manipulations must be carried out by using appropriate standard safety precautions.
3,4-Bis(azidomethyl)furazan (7). To a solution of 3,4-bis(bromomethyl)furazan (6) (2.56 g, 10 mmol) in acetone (10 mL) was added NaN3 (1.95 g, 30 mmol) at room temperature to give a suspension. The reaction was monitored by TLC and completed in about 24 h. Solvent was removed under reduced pressure, and the residue was extracted with Et2O (3 × 15 mL). After evaporation of diethyl ether, the crude product was purified by flash chromatography (CCl4) on silica gel to afford 7 as a light yellow liquid (1.76 g, 98%). Rf = 0.7 (CH2Cl2). 1H NMR (DMSO-d6) δ 4.86 (s, 2H, CH2). 13C NMR (DMSO- d6) δ 42.2, 151.1. IR (KBr): 2967, 2940, 2107, 1443, 1339, 1279, 1185, 1027, 890, 791 cm−1. Anal. Calcd. for C4H4N8O (180.13): C, 26.67; H, 2.24; N, 62.21. Found: C, 26.72; H, 2.27; N, 62.13.
1,6-Di(2-propyn-1-yloxy)-2,5-dinitro-2,5-diazahexane (9). 1,6-Dichloro-2,5-dinitro-2,5-diazahexane (8a, 2.47 g, 10.0 mmol) was dissolved in dry propargyl alcohol (10.8 mL, 200.0 mmol) and stirred at room temperature, passing dry N2 through the solution for 1 h. The excess of propargyl alcohol was removed by vacuum distillation at 30 °C to afford a yellow residue. After addition of H2O (30 mL), the mixture was stirred for 10 min. The flaxen precipitate was filtered, washed with water (3 × 30 mL), and dried. The crude product was recrystallized from CHCl3 to give 9 (2.43 g, 85%) as a white solid, mp 83-84 °C. 1H NMR (DMSO-d6) δ 3.47 (s, 1H, CH), 4.08 (s, 2H, CH2CH2), 4.25 (d, 2H, J = 1.9 Hz, CH2C), 5.21 (s, 2H, NCH2O). 13C NMR (DMSO-d6) δ 47.9, 56.5, 77.6, 78.7, 79.5. 14N NMR (DMSO-d6) δ −29.9 (NO2). IR (KBr): 3293, 2928, 2120, 1527, 1438, 1290, 1269, 1113, 1069, 1026, 981, 891, 891, 844, 663, 606 cm−1. Anal. calcd. for C10H14N4O6 (286.24): C 41.96, H 4.93, N 19.57. Found: C 42.02, H 4.96, N 19.61.
1-(2-Propyn-1-yloxy)-2-nitro-2-azapropane (10). The procedure is the same as for 9. After evaporation of an excess of propargyl alcohol, the crude product was purified by flash chromatography (CH2Cl2, Rf = 0.60) on silica gel. The title compound 10 (97%) is a light yellow liquid; 1H NMR (DMSO-d6) δ 3.35 (s, 3H, CH3), 3.49 (t, 1H, J = 2.2 Hz, CH), 4.26 (d, 2H, J = 2.2 Hz, CH2C), 5.22 (s, 2H, NCH2O). 13C NMR (DMSO-d6) δ 37.9, 56.5, 77.5, 79.3, 79.6. 14N NMR (DMSO-d6) δ −28.2 (NO2). IR (KBr): 3288, 2955, 2930, 2864, 2119, 1532, 1473, 1435, 1299, 1250, 1080, 1048, 995, 978 cm−1. Anal. calcd. for C5H8N2O3 (144.13): C 41.67, H 5.59, N 19.44. Found: C 41.71, H 5.61, N 19.39.
3,4-Bis((4-(2-methoxy-2-nitro-2-azapropane)-1H-1,2,3-triazol-1-yl)methyl)furazan (11a). To a solution of compound 7 (0.18 g, 1 mmol) and 10 (0.288 g, 2 mmol) in DMF (5 mL) was added CuSO4·5H2O (0.013 g, 0.05 mmol) followed by sodium ascorbate (0.019 g, 0.1 mmol) under argon. The reaction mixture was stirred at room temperature for 6 h and then diluted with H2O (50 mL) and extracted using CH2Cl2 (5 × 20 mL). Combined organic layers were filtered through Celite, dried over MgSO4, filtered, concentrated on a rotary evaporator, and recrystallized from dichloroethane to give colorless solid (0.323 g, 69%), mp 117–119 °C. Rf = 0.25 (MeCN:CCl4 1:3). 1H NMR (DMSO-d6) δ 3.37 (s, 3H, CH3), 4.68 (s, 2H, OCH2), 5.24 (s, 2H, NCH2O), 6.01 (s, 2H, Het-CH2-Het’), 8.26 (s, 1H, CH). 13C NMR (DMSO-d6) δ 37.4, 41.5, 61.5, 79.2, 124.6, 143.2, 150.3. IR (KBr): 3141, 3101, 2975, 2953, 1537, 1519, 1475, 1430, 1318, 1293, 1252, 1311, 1106, 1073, 1056, 1000, 938, 800 cm−1. Anal. Calcd. for C14H20N12O7 (468.39): C, 35.90; H, 4.30; N, 35.89. Found: C, 36.01; H, 4.24; N, 35.77.
Mixture of isomers 11a–11c (Catalyst free conditions). To a solution of compound 7 (0.18 g, 1 mmol) and 10 (0.288 g, 2 mmol) in DCE (5 mL) was heated under reflux until TLC analysis (3:2 hexanes–EtOAc) indicated complete consumption of starting materials (~90 h). Then DCE was evaporated in vacuo to dryness. Flash chromatography (CHCl3) furnished the product as an inseparable mixture of isomers 11a–11c in 60–65% yield. The ratio of isomers was determined by 1H NMR spectroscopy of the crude and the purified isomer mixture (the ratio is invariable).
Isomer 11a: 1H NMR (DMSO-d6) δ 3.37 (s, 3H, CH3), 4.69 (s, 2H, OCH2), 5.25 (s, 2H, NCH2O), 6.01 (s, 2H, Het1,4-CH2-Het’), 8.26 (s, 1H, CH). 13C NMR (DMSO-d6) δ 37.9, 42.1, 62.1, 79.8, 125.2, 143.7, 150.8. IR (KBr): 3141, 3101, 2975, 2953, 1537, 1519, 1475, 1430, 1318, 1293, 1252, 1311, 1106, 1073, 1056, 1000, 938, 800 cm−1.
Isomer 11b: 1H NMR (DMSO-d6) δ 3.37 (s, 3H, CH3), 4.81 (s, 2H, OCH2), 5.18 (s, 2H, NCH2O), 5.94 (s, 2H, Het1,5-CH2-Het’), 7.83 (s, 1H, CH). 13C NMR (DMSO-d6) δ 38.0, 40.8, 58.7, 79.8, 134.1, 143.7, 150.6.
Isomer 11c: 1H NMR (DMSO-d6) δ 3.37 (s, 3H, CH3), 4.69 (s, 2H, OCH2), 4.82 (s, 2H, OCH2), 5.18 (s, 2H, NCH2O), 5.25 (s, 2H, NCH2O), 5.97 (s, 2H, Het1,5-CH2-Het’), 5.98 (s, 2H, Het1,4-CH2-Het’), 7.83 (s, 1H, CH), 8.24 (s, 1H, CH). 13C NMR (DMSO-d6) δ 37.9, 40.8, 42.2, 58.7, 62.1, 79.8, 125.2, 134.0, 134.1, 143.7, 150.6, 150.8.
Polymer 12. To a 50 mL beaker immersed in a water bath was added diazide 7 (10 g, 55.56 mmol). After the diazide was heated to 45–50 °C, diacetylene 9 (15.9 g, 55.56 mmol) was added portion-wise maintaining 45–50 °C. The oily reaction mixture was then stirred at 50 °C overnight, after which it was transferred to a Teflon mold and kept at 80 °C in a thermostat at reduced pressure for 37 h. The reaction mixture was allowed to cool to ambient temperature. The resulting target polymer 12, which is a light yellow transparent plastic, was used in subsequent investigations without additional purification. An analytical sample was prepared by double precipitation from the DMSO solution with methanol, after which the precipitate was filtered, washed with methanol, and dried in vacuum for a day at 60 °C, which gave a colorless powder (89%). 1H NMR (DMSO-d6) δ 3.52, 4.09 (s, 2H, 1,4-triazole–N(NO)2CH2CH2), 4.11 (s, 2H, N(NO)2CH2CH2–Het1,5), 4.27 (s, 2H, CH2CCH), 4.68 (s, 2H, OCH2– Het1,5), 4.81 (s, 2H, OCH2–Het1,4), 5.19 (s, 2H, N(NO)2CH2OCH2–Het1,5), 5.24 (s, 2H, N(NO)2CH2OCH2–Het1,4), 5.93 (s, 2H, Het1,5–CH2–furazan–CH2–Het1,5), 5.96 (s, 2H, Het1,4–CH2–furazan–CH2–Het1,5), 5.97 (s, 2H, Het1,4–CH2–furazan–CH2–Het1,5), 6.01 (s, 2H, Het1,4–CH2–furazan–CH2–Het1,4), 7.80 (s, 1H, Het1,5–CH), 8.22 (s, 1H, (Het1,4–CH)–CH2–furazan–CH2–(Het1,5–CH)), 8.24 (s, 1H, Het1,4–CH). 13C NMR (DMSO-d6) δ 42.5, 42.6, 48.4, 56.9, 59.1, 62.5, 78.1, 79.6, 125.6, 134.5, 144.1, 150.9, 151.3. IR (KBr): 3293, 3144, 2963, 2113, 1528, 1461, 1437, 1290, 1270, 1111, 1069, 1025, 981. Anal. Calcd. for C14H18N12O7 (466.39): C, 36.06; H, 3.89; N, 36.04. Found: C, 36.12; H, 3.92; N, 35.99.

4. Conclusions

It has been demonstrated that azide–alkyne cycloaddition copolymerization under solvent- and catalyst-free conditions can enable the synthesis of a hybrid polymer bearing nitramine, furazan, and 1,2,3-triazole subunits in the polymer backbone. The method described herein requires no purification steps of the synthesized polymer. Alkyne and azide groups can be easily introduced into the required difunctional readily available precursors, both incorporating explosophoric units. This ring-closure copolymerization allowed for efficient diazide and dialkyne comonomers coupling, yielding an energetic polymer with high yield and sufficiently narrow polydispersity. The formation of the copolymer was confirmed by NMR, FT-IR, and SEC. Due to the presence of the nitramine, furazan, and 1,2,3-triazole subunits within the polymer repeating unit, this polymer is better than the benchmark energetic polymer, nitrocellulose (NC), in terms of glass transition temperature, thermal stability, and enthalpy of formation, whereas their burning rates are close. The polymer of this study is well plasticized by energetic plasticizers, which opens up wide opportunities for its use in the compositions of gunpowders and rocket propellants.
To the best of our knowledge, this is the first example of combining nitramino-, furazan-, and triazole-subunits in a polymer architecture designed by the Hüisgen 1,3-dipolar cycloaddition reaction of diazide and dialkyne comonomers, both of which contain explosophoric groups. Therefore, it represents a new way to produce synthetic energetic polymers endowed with a number of useful properties. Due to the reliable, easy-to-use and reproducible result of this azide–alkyne cycloaddition copolymerization, as well as the ability to tune the structure of both initial comonomers, we are confident that this simple and modular approach will find utility in the creation of new nitrogen- and oxygen-rich polymers for various applications related to energy materials. Further investigations on the synthesis and study of the structure-properties relationships of new energetic polymers are continuing in our laboratory.

Supplementary Materials

The supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/ijms24119645/s1. References [73,74,75,76,77,78] are cited in this section.

Author Contributions

Conceptualization, A.B.S.; methodology, P.S.G. and A.B.S.; investigation, P.S.G., N.N.K., N.N.I., E.R.S., A.P.D., V.A.S., V.D.D., D.B.V., P.V.B., K.Y.S., M.M.I. and M.L.K.; writing—original draft, N.N.K., N.N.I., A.P.D., V.A.S. and V.P.S.; writing—review and editing, V.P.S. and A.B.S.; project administration, A.B.S.; funding acquisition, A.B.S. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by The Ministry of Science and Higher Education of the Russian Federation (Agreement No 075-15-2020-803).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available under request.

Conflicts of Interest

The authors declare no conflict of interest. The funders had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript; or in the decision to publish the results.

References

  1. Kubota, N. Propellants and Explosives; Wiley-VCH: Weinheim, Germany, 2002. [Google Scholar]
  2. Davenas, A. Development of modern solid propellants. J. Propuls. Power 2003, 19, 1108–1128. [Google Scholar] [CrossRef]
  3. Pavlovets, G.A.; Tsutsuran, V.I. Physicochemical Properties of Powders and Propellants; Russian Ministry of Defense Publishing House: Moscow, Russia, 2009. (In Russian)
  4. DeLuca, L.T.; Shimada, T.; Sinditskii, V.P.; Calabro, M. (Eds.) Chemical Rocket Propulsion: A Comprehensive Survey of Energetic Materials; Springer: Berlin/Heidelberg, Germany, 2017. [Google Scholar]
  5. Singh, H.; Shekhar, H. Solid Rocket Propellants: Science and Technology Challenges; Royal Society of Chemistry: Cambridge, UK, 2017. [Google Scholar]
  6. Lysien, K.; Stolarczyk, A.; Jarosz, T. Solid Propellant Formulations: A Review of Recent Progress and Utilized Components. Materials 2021, 14, 6657. [Google Scholar] [CrossRef] [PubMed]
  7. Jin, Y.; Zhang, W.; Zhou, Z.; Liu, T.; Xia, H.; Huang, S.; Zhang, Q. Recent advances in hypergolic ionic liquids with broad potential for propellant applications. FirePhysChem 2022, 2, 236–252. [Google Scholar] [CrossRef]
  8. Gayathri, S.; Reshmi, S. Nitrato Functionalized Polymers for High Energy Propellants and Explosives: Recent Advances. Polym. Adv. Technol. 2017, 28, 1539–1550. [Google Scholar] [CrossRef]
  9. Liu, J. Nitrate Esters Chemistry and Technology; Springer: Berlin/Heidelberg, Germany, 2019. [Google Scholar] [CrossRef]
  10. Dou, J.; Xu, M.; Tan, B.; Lu, X.; Mo, H.; Wang, B.; Liu, N. Research progress of nitrate ester binders. FirePhysChem 2023, 3, 54–77. [Google Scholar] [CrossRef]
  11. Ang, H.G.; Pisharath, S. Energetic Polymers: Binders and Plasticizers for Enhancing Performance, 1st ed.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2012. [Google Scholar]
  12. Kumari, D.; Balakshe, R.; Banerjee, S.; Singh, H. Energetic Plasticizers for Gun & Rocket Propellants. Rev. J. Chem. 2012, 2, 240–262. [Google Scholar] [CrossRef]
  13. Pasquinet, E. Nitrogen-Rich Polymers as Candidates for Energetic Applications. In New Polymers for Special Applications; INTECH: London, UK, 2012; Chapter 10; pp. 313–338. [Google Scholar] [CrossRef] [Green Version]
  14. Paraskos, A.J. Energetic Polymers: Synthesis and Applications. In Energetic Materials; Shukla, M.K., Boddu, V., Steevens, J., Damavarapu, R., Leszczynski, J., Eds.; Springer: Berlin/Heidelberg, Germany, 2017; Chapter 4; pp. 91–134. [Google Scholar]
  15. Jarosz, T.; Stolarczyk, A.; Wawrzkiewicz-Jalowiecka, A.; Pawlus, K.; Miszczyszyn, K. Glycidyl Azide Polymer and its Derivatives-Versatile Binders for Explosives and Pyrotechnics: Tutorial Review of Recent Progress. Molecules 2019, 24, 4475. [Google Scholar] [CrossRef] [Green Version]
  16. Lysien, K.; Stolarczyk, A.; Jarosz, T. Energetic Polyoxetanes as High-Performance Binders for Energetic Composites: A Critical Review. Polymers 2022, N14, 4651. [Google Scholar] [CrossRef]
  17. Gozin, M.; Fershtat, L.L. Recent Advances in Chemistry of Nitrogen-Rich Energetic Polymers and Plasticizers. In Nitrogen-Rich Energetic Materials; Gozin, M., Fershtat, L.L., Eds.; Wiley-VCH GmbH: Weinheim, Germany, 2023; pp. 189–238. [Google Scholar] [CrossRef]
  18. Lempert, D.B.; Kazakov, A.I.; Sheremetev, A.B. Comparative ballistic efficiency of solid composite propellants: Which plasticizer/polymer combination is the energetically preferred binder? Mendeleev Commun. 2022, 32, 601–603. [Google Scholar] [CrossRef]
  19. Lempert, D.B.; Sheremetev, A.B. The energetic potential of azo- and azoxyfurazan nitro derivatives as components of composite rocket propellants. Chem. Heterocycl. Compd. 2016, 52, 1070–1077. [Google Scholar] [CrossRef]
  20. Lempert, D.B.; Dorofeenko, E.M. Reasons for the anomalous dependence of the specific impulse of rocket propellants on the content of borohydride. Combust. Explos. Shock. Waves 2013, 49, 472–477. [Google Scholar] [CrossRef]
  21. Dorofeenko, E.M.; Soglasnova, S.I.; Nechiporenko, G.N.; Lempert, D.B. Optimization of the Binder Formulation to Increase the Energetic Performance of Polynitrogen Oxidizers in Metal-Free Compositions. Combust. Explos. Shock. Waves 2018, 54, 698–703. [Google Scholar] [CrossRef]
  22. Kizhnyaev, V.N.; Vereshchagin, L.I. Vinyltetrazoles: Synthesis and properties. Russ. Chem. Rev. 2003, 72, 143–164. [Google Scholar] [CrossRef]
  23. Kizhnyaev, V.N.; Pokatilov, F.A.; Vereshchagin, L.I. Carbochain polymers with oxadiazole, triazole, and tetrazole cycles. Polym. Sci. Ser. C 2008, 50, 1–21. [Google Scholar] [CrossRef]
  24. Kizhnyaev, V.N.; Pokatilov, F.A.; Vereshchagin, L.I.; Verkhozina, O.N.; Petrova, T.L.; Prodaikov, A.G.; Ratovskii, G.V.; Tyukalova, O.V. Synthesis of energetic polynuclear and polymeric nitroazole systems. Russ. J. Appl. Chem. 2009, 82, 1769–1775. [Google Scholar] [CrossRef]
  25. Kizhnyaev, V.N.; Golobokova, T.V.; Pokatilov, F.A.; Vereshchagin, L.I.; Estrin, Y.I. Synthesis of energetic triazole- and tetrazole-containing oligomers and polymers. Chem. Heterocycl. Compd. 2017, 53, 682–692. [Google Scholar] [CrossRef]
  26. Fershtat, L.L.; Makhova, N.N. 1,2,5-Oxadiazole-Based High-Energy-Density Materials: Synthesis and Performance. ChemPlusChem 2020, 85, 13–42. [Google Scholar] [CrossRef] [Green Version]
  27. 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]
  28. Zlotin, S.G.; Churakov, A.M.; Egorov, M.P.; Fershtat, L.L.; Klenov, M.S.; Kuchurov, I.V.; Makhova, N.N.; Smirnov, G.A.; Tomilov, Y.V.; Tartakovsky, V.A. Advanced energetic materials: Novel strategies and versatile applications. Mendeleev Commun. 2021, 31, 731–749. [Google Scholar] [CrossRef]
  29. Zhou, J.; Zhang, J.; Wang, B.; Qiu, L.; Xu, R.; Sheremetev, A.B. Recent synthetic efforts towards high energy density materials: How to design high-performance energetic structures? FirePhysChem 2022, 2, 83–139. [Google Scholar] [CrossRef]
  30. Sheremetev, A.B.; Shatunova, E.V. Furazanyl Ethers of Pentaerythritol Derivates. In Proceedings of the 28th International Annual ICT-Conference, Combustion and Detonation, Karlsruhe, Germany, 24–27 June 1997. FRG, 94/1-8. [Google Scholar]
  31. Sheremetev, A.B.; Yudin, I.L. Unusual Oxidation of 4-amino-4H,8H-bisfurazano-[3,4-b;3′,4′-e]pyrazines. Mendeleev Commun. 2002, 12, 66–67. [Google Scholar] [CrossRef]
  32. Willer, R.L. Calculation of the density and detonation properties of C, H, N, O and F compounds: Use in the design and synthesis of new energetic materials. J. Mex. Chem. Soc. 2009, 53, 108–119. [Google Scholar] [CrossRef]
  33. Lotmentsev, Y.M.; Kondakova, N.N.; Bakeshko, A.V.; Kozeev, A.M.; Sheremetev, A.B. 3-Alkyl-4-nitrofurazansPlasticizers for polymers. Chem. Heterocycl. Comp. 2017, 53, 740–745. [Google Scholar] [CrossRef]
  34. Gorman, I.E.; Willer, R.L.; Kemp, L.K.; Storey, R.F. Development of a triazole-cure resin system for composites: Evaluation of alkyne curatives. Polymer 2012, 53, 2548–2558. [Google Scholar] [CrossRef]
  35. Chavez, D.E.; Tappan, B.C.; Kuehl, V.A.; Schmalzer, A.M.; Leonard, P.W.; Wu, R.; Imler, G.H.; Parrish, D.A. [3+2] Click chemistry approach to tetrazine containing polymers. RSC Adv. 2022, 12, 28490–28493. [Google Scholar] [CrossRef]
  36. Petrov, A.O.; Karpov, S.V.; Malkov, G.V.; Shastin, A.V.; Badamshina, E.R. New non-symmetric azido-diacetylenic s-triazine monomer for polycycloaddition. Mendeleev Commun. 2022, 32, 464–466. [Google Scholar] [CrossRef]
  37. Johnson, J.A.; Finn, M.G.; Koberstein, J.T.; Turro, N.J. Construction of Linear Polymers, Dendrimers, Networks, and Other Polymeric Architectures by Copper-Catalyzed Azide-Alkyne Cycloaddition ‘‘Click’’ Chemistry. Macromol. Rapid Commun. 2008, 29, 1052–1072. [Google Scholar] [CrossRef]
  38. Singh, M.S.; Chowdhury, S.; Koley, S. Advances of azide-alkyne cycloaddition-click chemistry over the recent decade. Tetrahedron 2016, 72, 5257–5283. [Google Scholar] [CrossRef]
  39. Santos, C.S.; Oliveira, R.J.; Oliveira, R.N.; Freitas, J.C.R. 1,2,3-Triazoles: General and key synthetic strategies. Arkivoc 2020, i, 219–271. [Google Scholar] [CrossRef]
  40. Aflak, N.; Ayouchia, H.B.E.; Bahsis, L.; Anane, H.; Julve, M.; Stiriba, S.-E. Recent Advances in Copper-Based Solid Heterogeneous Catalysts for Azide–Alkyne Cycloaddition Reactions. Int. J. Mol. Sci. 2022, 23, 2383. [Google Scholar] [CrossRef]
  41. Bakulev, V.A.; Shafran, Y.M.; Beliaev, N.A.; Beryozkina, T.V.; Volkova, N.N.; Joy, M.N.; Fan, Z. Heterocyclic azides: Advances in their chemistry. Russ. Chem. Rev. 2022, 91, RCR5042. [Google Scholar] [CrossRef]
  42. Androsov, A.S.; Denisyuk, A.P.; Tokarev, N.P.; Fominov, K.G. Role of individual components in the catalysis of the combustion of ballistic powders. Combust. Explos. Shock. Waves 1975, 11, 14–20. [Google Scholar] [CrossRef]
  43. Androsov, A.S.; Denisyuk, A.P.; Tokarev, N.P. Mechanism of the effect of composite lead-copper catalysts on powder Combustion. Combust. Explos. Shock. Waves 1978, 14, 184–187. [Google Scholar] [CrossRef]
  44. Abdel-Ghani, N.; Elbeih, A.; Helal, F. The effect of different copper salts on the mechanical and ballistic characteristics of double base rocket propellants. Cent. Eur. J. Energetic Mater. 2016, 13, 469–482. [Google Scholar] [CrossRef]
  45. Zhou, X.; Xu, R.; Nie, H.; Yan, Q.; Liu, J.; Sun, Y. Insight into the precise catalytic mechanism of CuO on the decomposition and combustion of core–shell Al@AP particles. Fuel 2023, 346, 128294. [Google Scholar] [CrossRef]
  46. Khakhalev, A.V.; Anisimov, A.A.; Sheremetev, A.B. 3,4-Bis(bromomethyl)furazan and its N-oxide. Chem. Heterocycl. Compd. 2022, 58, 263–266. [Google Scholar] [CrossRef]
  47. Ivanova, O.A.; Averina, E.B.; Kuznetsova, T.S.; Zefirov, N.S. Synthesis of new 3,4-disubstituted furazans. Chem. Heterocycl. Compd. 2001, 36, 1091–1096. [Google Scholar] [CrossRef]
  48. Sheremetev, A.B.; Mel’nikova, S.F.; Kokareva, E.S.; Nekrutenko, R.E.; Strizhenko, K.V.; Suponitsky, K.Y.; Pham, T.D.; Pivkina, A.N.; Sinditskii, V.P. Nitroxy- and azidomethyl azofurazans as advanced energetic materials. Def. Technol. 2022, 18, 1369–1381. [Google Scholar] [CrossRef]
  49. Min, B.S.; Kim, S.J.; Kim, W.H.; Moon, H.S.; Hwang, G.W.; Jeon, H.B.; Cho, H.R.; Jung, J.Y. An Energetic Prepolymer for Solid Propellant Binder and Manufacturing Method Thereof. Korean Patent KR2386460, 15 April 2022. [Google Scholar]
  50. Gribov, P.S.; Suponitsky, K.Y.; Sheremetev, A.B. Efficient Synthesis of N-(Chloromethyl)nitramines via TiCl4-Catalyzed Chlorodeacetoxylation. New J. Chem. 2022, 46, 17548–17553. [Google Scholar] [CrossRef]
  51. Rowland, R.S.; Taylor, R. Intermolecular nonbonded contact distances in organic crystal structures: Comparison with distances expected from van der Waals radii. J. Phys. Chem. 1996, 100, 7384–7391. [Google Scholar] [CrossRef]
  52. Bader, R.F.W. Atoms in Molecules. In A Quantum Theory; Clarendon Press: Oxford, UK, 1990. [Google Scholar]
  53. Keith, T.A. AIMAll (Version 15.05.18); TK Gristmill Software: Overland Park, KS, USA, 2015. [Google Scholar]
  54. Gribov, P.S.; Kon’kova, T.S.; Suponitsky, K.Y.; Sheremetev, A.B. Dipropargyl ethers possessing nitramine units. Mendeleev Commun. 2023, 34, 466–468. [Google Scholar]
  55. Karpov, S.V.; Petrov, A.O.; Malkov, G.V.; Badamshina, E.R. The Gaussian G4 enthalpy of formation of propargylamine and propargyloxy derivatives of triazido-s-triazine. Mendeleev Commun. 2022, 32, 338–340. [Google Scholar] [CrossRef]
  56. Wamhoff, H. 1,2,3-Triazoles and their Benzo Derivatives. Compr. Heterocycl. Chem. 1984, 5, 669–732. [Google Scholar] [CrossRef]
  57. Hein, J.E.; Fokin, V.V. Copper-catalyzed azide–alkyne cycloaddition (CuAAC) and beyond: New reactivity of copper(I) acetylides. Chem. Soc. Rev. 2010, 39, 1302–1315. [Google Scholar] [CrossRef]
  58. Binke, N.; Rong, L.; Xianqi, C.; Yuan, W.; Rongzu, H.; Qingsen, Y. Study on the Melting Process of Nitrocellulose by Thermal Analysis Method. J. Therm. Anal. Calorim. 1999, 58, 249–256. [Google Scholar] [CrossRef]
  59. Kissinger, H.E. Reaction kinetics in differential thermal analysis. Anal Chem. 1957, 29, 1702–1706. [Google Scholar] [CrossRef]
  60. Kubasov, A.A. Chemical Kinetics and Catalysis; Part 1; Moscow State University Publishing House: Moscow, Russia, 2005. (In Russian) [Google Scholar]
  61. Prokudin, V.G.; Nazin, G.M. Gas-phase cyclodecomposition of furazane and furazane N-oxide. Bull. Acad. Sci. USSR Div. Chem. Sci. 1987, 36, 199–201. [Google Scholar] [CrossRef]
  62. Kumar, A.S.; Ghule, V.D.; Subrahmanyam, S.; Sahoo, A.K. Synthesis of Thermally Stable Energetic 1,2,3-Triazole Derivatives. Chem. A Eur. J. 2013, 19, 509–518. [Google Scholar] [CrossRef] [PubMed]
  63. Yao, W.; Xue, Y.; Qian, L.; Yang, H.; Cheng, G. Combination of 1,2,3-triazole and 1,2,4-triazole frameworks for new high-energy and low-sensitivity compounds. Energetic Mater. Front. 2021, 2, 131–138. [Google Scholar] [CrossRef]
  64. Pourmortazavi, S.M.; Hosseini, S.G.; Rahimi-Nasrabadi, M.; Hajimirsadeghi, S.S.; Momenian, H. Effect of nitrate content on thermal decomposition of nitrocellulose. J. Hazard. Mater. 2009, 162, 1141–1144. [Google Scholar] [CrossRef] [PubMed]
  65. Zinoviev, V.M.; Kutsenko, G.V.; Ermilov, A.S.; Boldavnin, I.I. High-Energy Plasticizers of Solid Rocket and Gun Propellants; Perm STU Publishing House: Perm, Russia, 2010. (In Russian) [Google Scholar]
  66. Vinogradov, D.B.; Bulatov, P.V.; Petrov, E.Y.; Gribov, P.S.; Kondakova, N.N.; Il’icheva, N.N.; Stepanova, E.R.; Denesyuk, A.P.; Sizov, V.A.; Sinditskii, V.P.; et al. Promising oxygen- and nitrogen-rich azidonitramino ether plasticizers for energetic materials. Molecules 2022, 27, 7749. [Google Scholar] [CrossRef] [PubMed]
  67. Sutton, J.P.; Biblarz, O. Rocket Propulsion Elements, 9th ed.; Wiley: New York, NY, USA, 2017. [Google Scholar]
  68. Agrawal, J.P. High Energy Materials: Propellants, Explosives and Pyrotechnics; Wiley-VCH: Weinheim, Germany, 2010. [Google Scholar]
  69. Sinditskii, V.P.; Egorshev, V.Y.; Serushkin, V.V.; Levshenkov, A.I.; Berezin, M.V.; Filatov, S.A.; Smirnov, S.P. Evaluation of Decomposition Kinetics of Energetic Materials in the Combustion Wave. Thermochim. Acta 2009, 496, 1–12. [Google Scholar] [CrossRef]
  70. Denisyuk, A.P.; Shepelev, Y.G.; Yudaev, S.V.; Kalashnikov, I.V. Combustion of systems containing linear nitramines. Combust Explos Shock Waves 2005, 41, 206–214. [Google Scholar] [CrossRef]
  71. Sarner, S.F. Propellant Chemistry; Reinhold: New York, NY, USA, 1966. [Google Scholar]
  72. Kostochko, A.V.; Kazban, B.M. Gunpowder, Solid Composite Propellants and Their Properties; INTRA-M: Moscow, Russia, 2015. (In Russian) [Google Scholar]
  73. APEX2 and SAINT; Bruker AXS Inc.: Madison, WI, USA, 2014.
  74. Sheldrick, G.M. Crystal structure refinement with SHELXL. Acta Cryst. C 2015, 71, 3–8. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  75. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Montgomery, J.A.; Kudin, K.N., Jr.; Burant, J.C.; Millam, J.M.; et al. Gaussian 03, Revision E.01; Gaussian, Inc.: Wallingford, UK, 2004. [Google Scholar]
  76. Suponitsky, K.Y.; Smol’yakov, A.F.; Ananyev, I.V.; Khakhalev, A.V.; Gidaspov, A.A.; Sheremetev, A.B. 3,4-Dinitrofurazan: Structural Nonequivalence of ortho-Nitro Groups as a Key Feature of the Crystal Structure and Density. ChemistrySelect 2020, 5, 14543–14548. [Google Scholar] [CrossRef]
  77. Suponitsky, K.Y.; Fedyanin, I.V.; Karnoukhova, V.A.; Zalomlenkov, V.A.; Gidaspov, A.A.; Bakharev, V.V.; Sheremetev, A.B. Energetic co-crystal of a primary metal-free explosive with BTF. Ideal pair for co-crystallization. Molecules 2021, 26, 7452. [Google Scholar] [CrossRef]
  78. Suponitsky, K.Y.; Masunov, A.E.; Antipin, M.Y. Conformational dependence of the first molecular hyperpolarizability in the computational design of nonlinear optical materials for optical switching. Mendeleev. Commun. 2008, 18, 265–267. [Google Scholar] [CrossRef]
Scheme 1. Furazan-based energetic polymer 1–3, which are available in the literature to date.
Scheme 1. Furazan-based energetic polymer 1–3, which are available in the literature to date.
Ijms 24 09645 sch001
Scheme 2. Synthesis of diazido-monomer 7.
Scheme 2. Synthesis of diazido-monomer 7.
Ijms 24 09645 sch002
Scheme 3. Synthesis of dipropargylic-monomer 9 and analog 10 with the desired alkyne and nitramine functionalities.
Scheme 3. Synthesis of dipropargylic-monomer 9 and analog 10 with the desired alkyne and nitramine functionalities.
Ijms 24 09645 sch003
Figure 1. On the left: a general view of compound 9 showing the atomic numbering scheme. Atoms are represented as thermals ellipsoids with a probability of 50%. Close C2 … N2 (C2A … N2A) contacts are shown by dotted lines. On the right: a crystal packing fragment of compound 9. The CCDC number is 2251638.
Figure 1. On the left: a general view of compound 9 showing the atomic numbering scheme. Atoms are represented as thermals ellipsoids with a probability of 50%. Close C2 … N2 (C2A … N2A) contacts are shown by dotted lines. On the right: a crystal packing fragment of compound 9. The CCDC number is 2251638.
Ijms 24 09645 g001
Scheme 4. A model azide–alkyne cycloaddition reaction.
Scheme 4. A model azide–alkyne cycloaddition reaction.
Ijms 24 09645 sch004
Figure 2. 1H NMR spectrum (DMSO-d6) of three regioisomeric di-1,2,3-triazoles 11a, 11b, and 11c.
Figure 2. 1H NMR spectrum (DMSO-d6) of three regioisomeric di-1,2,3-triazoles 11a, 11b, and 11c.
Ijms 24 09645 g002
Scheme 5. Cycloaddition polymerization of diazide 7 and diacetylene 9.
Scheme 5. Cycloaddition polymerization of diazide 7 and diacetylene 9.
Ijms 24 09645 sch005
Figure 3. Comparison of the FTIR spectra of 7 (green line), 9 (red line), and 11 (blue line) with polymer 12 (purple line).
Figure 3. Comparison of the FTIR spectra of 7 (green line), 9 (red line), and 11 (blue line) with polymer 12 (purple line).
Ijms 24 09645 g003
Figure 4. 1H NMR spectrum (DMSO-d6) of polymer 12: three regioisomeric 1,2,3-triazole units are visible.
Figure 4. 1H NMR spectrum (DMSO-d6) of polymer 12: three regioisomeric 1,2,3-triazole units are visible.
Ijms 24 09645 g004
Figure 5. DSC plot of comonomer 9 (a) and polymer 12 (b) with scanning at 10 °C min−1.
Figure 5. DSC plot of comonomer 9 (a) and polymer 12 (b) with scanning at 10 °C min−1.
Ijms 24 09645 g005
Figure 6. DSC plots of polymer 12 at different heating rates.
Figure 6. DSC plots of polymer 12 at different heating rates.
Ijms 24 09645 g006
Figure 7. Comparison of the decomposition rate constants of polymer 12, monomer 9, NC, and DMF.
Figure 7. Comparison of the decomposition rate constants of polymer 12, monomer 9, NC, and DMF.
Ijms 24 09645 g007
Figure 8. Polymer 12 at (a) room temperature and (b) after thermostating at 60 °C.
Figure 8. Polymer 12 at (a) room temperature and (b) after thermostating at 60 °C.
Ijms 24 09645 g008
Figure 9. DSC plot for a mixture of NC/NG (50:50).
Figure 9. DSC plot for a mixture of NC/NG (50:50).
Ijms 24 09645 g009
Figure 10. DSC profiles of mixture of polymer 12/NG. a—60:40; b—50:50; c—40:60; d—30:70.
Figure 10. DSC profiles of mixture of polymer 12/NG. a—60:40; b—50:50; c—40:60; d—30:70.
Ijms 24 09645 g010
Figure 11. DSC plot for a mixture of 12/DEGDN (80:20).
Figure 11. DSC plot for a mixture of 12/DEGDN (80:20).
Ijms 24 09645 g011
Figure 12. Plasticizers used for modification of polymer 12.
Figure 12. Plasticizers used for modification of polymer 12.
Ijms 24 09645 g012
Figure 13. DSC profiles of plasticized polymer 12 (50:50). a—with Z3; b—with Z8; c—with Z1.
Figure 13. DSC profiles of plasticized polymer 12 (50:50). a—with Z3; b—with Z8; c—with Z1.
Ijms 24 09645 g013
Figure 14. Appearance of 12/NG compositions with different NG content.
Figure 14. Appearance of 12/NG compositions with different NG content.
Ijms 24 09645 g014
Figure 15. Dependence of viscosity (η) on shear stress (τ) for 12/NG compositions: 60:40 (line a); 50:50 (line b), 40:60 (line c); 30:70 (line d).
Figure 15. Dependence of viscosity (η) on shear stress (τ) for 12/NG compositions: 60:40 (line a); 50:50 (line b), 40:60 (line c); 30:70 (line d).
Ijms 24 09645 g015
Figure 16. Plot of the burning rate vs. the pressure dependence of NC (line a), high-molecular polymer 12 (line b), low-molecular polymer 12 (line c), and comonomer 9 (line d).
Figure 16. Plot of the burning rate vs. the pressure dependence of NC (line a), high-molecular polymer 12 (line b), low-molecular polymer 12 (line c), and comonomer 9 (line d).
Ijms 24 09645 g016
Figure 17. Comparison of the burning rates of NC (line a), high-molecular polymer 12 (line b), with composition 12/NG (50:50) (line e), and NC/NG (50:50) (line f).
Figure 17. Comparison of the burning rates of NC (line a), high-molecular polymer 12 (line b), with composition 12/NG (50:50) (line e), and NC/NG (50:50) (line f).
Ijms 24 09645 g017
Table 1. Cycloaddition polymerisation 1 of diazide 7 with diacetylene 9 at 60 °C.
Table 1. Cycloaddition polymerisation 1 of diazide 7 with diacetylene 9 at 60 °C.
EntryReaction Time
(h)
Mn2
(g mol−1)
MW3
(g mol−1)
ĐM4
12121017201.42
210181033381.84
318245045391.85
442253251222.02
566276667472.44
6 56614,20042,4002.99
1 Reaction conditions: diazide 7 (50 mmol), diacetylene 9 (50 mmol). 2 Mn is number-average molecular weight. 3 MW is weight-average molecular weight. 4 Dispersity (ĐM) and Mn determined by gel permeation chromatography (N-methylpyrrolidone) analysis against polystyrene (PS) standards. 5 Reaction at 50 → 80 °C.
Table 2. Decomposition results (DSC-TGA measurements).
Table 2. Decomposition results (DSC-TGA measurements).
SampleDSCTGA
Tonset, 1 °CTpeak, 2 °CHd, 3 J g−1Tonset, 4 °CΔm, 5 %
7207215- 610398
9237270−227321893
12243280−216824860
NC (12.5% N)197204−155016698
1 Decomposition temperature (onset) measured at a heating rate of 10 °C min−1. 2 Decomposition temperature (peak). 3 Heat of decomposition. 4 Onset temperature of mass loss. 5 Mass loss during decomposition. 6 Due to the evaporation of the compound, it was not possible to obtain values for the enthalpy of decomposition.
Table 3. Kinetic data of decomposition of polymer 12 under non-isothermal conditions.
Table 3. Kinetic data of decomposition of polymer 12 under non-isothermal conditions.
Heating Rate, °C min−1Tpeak, °CEnthalpy of Decomposition,
∆Hd, J g−1
k·103, c−1
5271−21765.96
10280−216811.5
15286−222616.9
20290−219922.2
Table 4. Glass transition temperature (Tg) of plasticized polymer 12.
Table 4. Glass transition temperature (Tg) of plasticized polymer 12.
PolymerPlasticizerC, 1 % wt.Tg, °CCp, 2 J g−1·K
12Z150−510.615
Z350−390.607
Z850−430.766
NG40−120,618
50−270.647
60−400.713
70−520.775
DEGDN2020.519
NCNG50−520.591
1 Concentration of plasticizer in a mixture with polymer. 2 Change in isobaric heat capacity.
Table 5. Viscosity of 12/NG compositions, at τ = 3 kPa and 20 °C.
Table 5. Viscosity of 12/NG compositions, at τ = 3 kPa and 20 °C.
Percentage NG,
% wt
Viscosity,
Pa·c
40 *2192
50269
60142
7043
*At 30 °C.
Table 6. Burning rate data.
Table 6. Burning rate data.
MaterialBurning Rate rb = Apn.rb 2,6 mm c−1rb 10, 7 mm c−1
∆p, 3 MPan 4A 5
92–150.481.241.73.7
1210.4–30.642.03.311.5
1220.1–30.541.52.29.4
12/NG (60:40)2–120.603.966.015.8
NC1–120.732.823.310.7
NC/NG (60:40)1–120.71A 34.614.5
1Mw = 42,400 g mol−1. 2 Mw = 6747 g mol−1. 3 Pressure range. 4 Pressure exponent in the burning rate law. 5 Empirical coefficient. 6 Burning rate at 2 MPa. 7 Burning rate at 10 MPa.
Table 7. Calculated thermodynamic properties of plasticized polymer 12 and NC.
Table 7. Calculated thermodynamic properties of plasticized polymer 12 and NC.
CompositionF, 1 kJ kg−1Tc, 2 K
NC/NG (70:30)11553730
12/NG (70:30)9332380
12/NG (60:40)10382584
12/NG (50:50)11472975
1 Powder force. 2 Combustion temperature under isobaric conditions at 300 MPa.
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

Gribov, P.S.; Kondakova, N.N.; Il’icheva, N.N.; Stepanova, E.R.; Denisyuk, A.P.; Sizov, V.A.; Dotsenko, V.D.; Vinogradov, D.B.; Bulatov, P.V.; Sinditskii, V.P.; et al. Energetic Polymer Possessing Furazan, 1,2,3-Triazole, and Nitramine Subunits. Int. J. Mol. Sci. 2023, 24, 9645. https://doi.org/10.3390/ijms24119645

AMA Style

Gribov PS, Kondakova NN, Il’icheva NN, Stepanova ER, Denisyuk AP, Sizov VA, Dotsenko VD, Vinogradov DB, Bulatov PV, Sinditskii VP, et al. Energetic Polymer Possessing Furazan, 1,2,3-Triazole, and Nitramine Subunits. International Journal of Molecular Sciences. 2023; 24(11):9645. https://doi.org/10.3390/ijms24119645

Chicago/Turabian Style

Gribov, Pavel S., Natalia N. Kondakova, Natalia N. Il’icheva, Evgenia R. Stepanova, Anatoly P. Denisyuk, Vladimir A. Sizov, Varvara D. Dotsenko, Dmitry B. Vinogradov, Pavel V. Bulatov, Valery P. Sinditskii, and et al. 2023. "Energetic Polymer Possessing Furazan, 1,2,3-Triazole, and Nitramine Subunits" International Journal of Molecular Sciences 24, no. 11: 9645. https://doi.org/10.3390/ijms24119645

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