Elsevier

Combustion and Flame

Volume 212, February 2020, Pages 67-78
Combustion and Flame

A comprehensive mechanism for liquid-phase decomposition of 1,3,5,7-tetranitro-1,3,5,7-tetrazoctane (HMX): Thermolysis experiments and detailed kinetic modeling

https://doi.org/10.1016/j.combustflame.2019.10.025Get rights and content

Abstract

The nitramines 1,3,5,7-tetranitro-1,3,5,7-tetrazoctane (HMX), and 1,3,5-trinitro-1,3,5-triazinane (RDX) are energetic materials commonly used in solid propellants and explosives. In order to predict ignition and deflagration of propellants containing these ingredients, their thermal decomposition behaviors must be thoroughly understood. In this study, the thermal decomposition of HMX was investigated using synergetic application of experimental and computational methods. Mole fraction profiles of the gaseous decomposition products evolving from the liquid-phase HMX were obtained using Fourier transform infrared (FTIR) spectroscopy for two types of thermolysis experiments – thermogravimetric analysis (TGA) and confined rapid thermolysis (CRT). Four heating rates (5, 10, 15, and 20 K/min) in TGA experiments and four set temperatures (290, 300, 310 and 320°C) in CRT experiments were considered. In the TGA and differential scanning calorimetry (DSC) results, steep mass loss and rapid decomposition were observed after the melting of the HMX at 280°C. CH2O and N2O were identified as the major decomposition products. Smaller quantities of H2O, HCN, NO and NO2, CO and CO2 were also formed. In the complementary computational study, liquid-phase elementary reactions were investigated using quantum mechanics calculations at B3LYP/6-311++G(d,p) level of theory with the conductor-like polarizable continuum model (CPCM). A zero-dimensional model was developed to simulate the TGA and CRT experiments based on conservation of mass and species in the condensed-phase and the gas-phase control volumes. The predicted mass loss and gas-phase mole fraction profiles of the decomposition products are in good agreements with the corresponding experimental results, indicating that the comprehensive mechanism proposed here captures the important reactions occurring during liquid-phase decomposition of HMX.

Introduction

The cyclic nitramines HMX and RDX are important energetic ingredients commonly used in many applications, including, among others, explosives and rocket propellants. A quantitative understanding of chemical kinetics during thermal decomposition is necessary for predictive modeling of ignition, combustion and detonation behavior of explosives and solid propellants containing these ingredients. Thus, RDX and HMX have been extensively investigated with a focus on structural characterization [1], [2], [3], [4], [5], [6], phase transformation [7], [8], [9], [10], [11], [12], thermo-physical properties [13], [14], [15], thermal decomposition behavior [16], [17], [18], [19], [20], [21], [22], [23], [24], [25], detonation characteristics [26], [27], [28], [29], [30], [31], [32], as well as burn-rate and gas-phase flame structure [33], [34], [35], [36], [37], [38], [39]. During the past few decades, theoretical studies have also become increasingly important for predictive modeling of the ignition, combustion and detonation behaviors of these energetic materials [40], [41], [42], [43], [44], [45], [46].

Brill and Karpowicz [8] used FTIR spectroscopy to obtain Arrhenius parameters for the β-δ phase transition in HMX and proposed that the breakdown of intermolecular forces controls the rate of thermal decomposition. A detailed quantum molecular dynamics study was performed by Ye et al. [12] to explain the structural changes associated with β-HMX to δ-HMX phase transition and the initial unimolecular reaction pathways of these two conformers. Second-harmonic imaging microscopy was used to investigate the β-δ phase transition mechanism [9] and develop a kinetic model [10,11]. Products formed during thermal decomposition of HMX have been identified by rapid-scan/FTIR spectroscopy [16] and triple quadruple mass spectrometry [19]. Temporal evolution profiles of gaseous decomposition products of various energetic materials have been probed using simultaneous thermogravimetric modulated beam mass spectrometry (STMBMS) [20,21] and confined rapid thermolysis (CRT)/FTIR spectroscopy [22]. The kinetics of phase change, thermal ignition, and decomposition have been investigated using cook-off studies [26,27]. Using the measured gas formation rates from STMBMS experiments of Behrens, kinetic parameters for global decomposition models were obtained and applied to slow cookoff in accident scenarios [23,24].

The times-to-explosion for HMX-based plastic-bonded explosives calculated by Tarver and Tran [29] using a global kinetic model are in reasonable agreement with the experimental results [30], [31], [32]. Nonetheless, the times-to-explosion are over-predicted using the global model above the liquefaction temperature of HMX, i.e., 553 K, which may be attributed to the inability of the model to account for faster autocatalytic decomposition in the liquid phase. Recently, Hobbs et al. [25] have successfully modeled the cookoff of HMX-based plastic bonded explosive (PBX 9501) using a melt rate accelerator factor at temperatures above melting point. This approach is reasonable from an engineering perspective, however, for a fundamental understanding at the molecular level, the detailed liquid-phase mechanism is required to explain faster rates at higher temperatures.

In addition to cook-off models, various theoretical models, as reviewed by Beckstead et al. [41] have been formulated to study ignition [46] and deflagration [42], [43], [44] of HMX. These models analyze the combustion wave structure in three regions (solid phase, melt layer and gas phase) by solving the equations for the conservation for mass, momentum, energy, and species. In the solid phase, β-HMX to δ-HMX phase transition is modeled. Other solid-phase decomposition reactions are assumed not to occur in these combustion models. The melt layer is either ignored or modeled by a three-step global mechanism [47]. A detailed reaction mechanism, originally developed by Melius [48,49], refined by Yetter et al. [50], and further updated by Chakraborty et al. [51] is used in the gas phase.

During thermal decomposition of HMX, various gaseous species, including CH2O, N2O, HCN, NO, NO2, CO, CO2, and H2O, evolve from the condensed phase. Based on the analysis of these decomposition products, various reaction pathways were proposed. However, due to fast reactions and short-lived intermediates, experimental identification of the reaction products and the elementary reactions is quite challenging. Recently, quantum mechanics calculations have been performed in many theoretical studies to probe solid-phase and gas-phase decomposition of HMX. Truong and co-workers [52], [53], [54] performed direct dynamics calculations for the gas-phase decomposition of α-HMX at B3LYP/cc-pVDZ level of theory. Based on the analysis of rate constants calculated using transition state theory, they found that the Nsingle bondNO2 bond fission is the preferred decomposition pathway. Molt et al. [5,6] used high-level ab initio methods to compute the geometries and energetic ordering of various conformers of RDX and HMX in the gas phase. In a subsequent study [55], they were the first to locate a well-defined transition state for Nsingle bondNO2 bond homolysis during gas-phase decomposition of RDX. With a barrier of 41.9 kcal/mol, HONO elimination was found to be the dominant initial reaction. The equivalent barrier for Nsingle bondNO2 homolysis was 53.9 kcal/mol. However, a recent study [56] on rate constants using variable reaction coordinate transition state theory have concluded Nsingle bondNO2 bond dissociation as the dominant pathway in the gas phase.

Decomposition of RDX and HMX from their excited electronic states and related anionic species was investigated by Bernstein and co-workers [57], [58], [59]. NO2 and HONO were ruled out as sources for NO formation, which indicates that the decomposition mechanism is different in the excited states as compared to the ground states. Zhang et al. [60] found that H+and OH accelerates the decomposition of HMX in the gas phase, however, in the aqueous solution, this is true for OH only. Kuklja and co-workers [61], [62], [63], [64] investigated gas-phase and solid-phase decomposition of β- and δ-HMX, and observed that Nsingle bondNO2 bond fission is the dominant pathway in the gas phase. In the condensed phase, Nsingle bondNO2 bond fission is also the dominant pathway for β-HMX, however, the exothermic HONO elimination channel is found to be more favorable in δ-HMX.

Most of these theoretical studies have investigated the gas-phase or solid-phase decomposition of HMX and RDX. A large number of reactive molecular dynamics studies [12,[65], [66], [67], [68], [69], [70] have also been performed with a focus on detonation characteristics. However, very little attention has been given to liquid-phase decomposition of HMX, even though the melt layer has been observed at the surface of burning propellants [71]. The liquid-phase decomposition in various combustion models is still represented by a global mechanism [42], [43], [44], which is inadequate for rigorous combustion modeling. The global mechanisms do not accurately capture the evolution of various decomposition products from the condensed phase into the gas phase above the surface of the burning propellants. This may also lead to inaccuracies in the gas-phase combustion mechanisms. Hence, liquid-phase chemical kinetics needs further investigation.

In this study, we demonstrate synergetic application of quantum mechanics calculations, chemical kinetic modeling and thermolysis experiments to develop and validate a comprehensive mechanism for liquid-phase decomposition of energetic materials consisting of elementary reactions. HMX is chosen as the representative energetic material but the methodology is applicable to other energetic materials as well. The first objective is to quantify species evolution rates during thermal decomposition of HMX using FTIR spectroscopy for low heating rates using TGA setup and for high heating rates using CRT setup. In our previous work [72,73], quantum mechanics calculations were performed to develop initiation mechanism for the liquid-phase decomposition of RDX and HMX. Thus, the second objective is to further expand and validate the HMX initiation mechanism consisting of elementary reactions. The detailed chemical kinetic mechanism in the liquid phase developed and validated here can then be used in three-phase deflagration models along with appropriate solid-phase and gas-phase mechanisms to predict the ignition and combustion behavior of HMX.

Section snippets

Slow thermolysis: TGA-FTIR setup

The TGA technique in conjunction with FTIR spectroscopy is used to study thermal decomposition of HMX for low heating rates as shown in Fig. 1. HMX samples of 1 mg mass were placed in aluminum crucibles and decomposed in a Netzsch STA 449 F5 Jupiter TGA/DSC apparatus coupled to a Bruker Vertex 80 FTIR spectrometer. Non-isothermal decomposition of HMX from 30 to 320°C was investigated for four heating rates: 5, 10, 15, and 20 K/min. The decomposition products were transferred to the TGA-FTIR

Quantum mechanics calculations

In the complementary computational study, elementary reactions in the liquid phase were investigated using quantum mechanics calculations performed in Gaussian 09 [78], Density functional theory (DFT) calculations are used for geometry optimization with B3LYP functional [79] and 6–311++G(d,p) basis set. For liquid-phase calculations, implicit conductor-like solvation model CPCM [80] is used with water as a solvent. A detailed comparsion of various combinations of solvation models and solvents

TGA and DSC results

The TGA curves for the four heating rates are given in Fig. 4(a) and the corresponding DSC curves are given in Fig. 4(b). These results are the average of three repeatable experiments with similar sample masses.

The β-HMX to δ-HMX phase transformation is observed at around 190°C as indicated by a slight endothermic peak in the DSC data (given in Supplementary information as Fig. S9). As the temperature is further increased above 260°C, the mass loss starts very slowly. On further heating, the

Conclusions

Thermal decomposition of HMX was studied using TGA-FTIR setup for different low heating rates and using CRT-FTIR setup for high heating rates to isothermal conditions at different set temperatures. A phase transformation from β-HMX to δ-HMX was observed around 190°C. DSC data shows the endothermic melting at approximately 280°C. This was followed by rapid mass loss and exothermic decomposition. For heating rates of 10, 15 and 20 K/min, sublimation of HMX was very small and it decomposed

Declaration of Competing Interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Acknowledgment

This material is based upon work supported by, or in part by, the U.S. Army Research Laboratory and the U. S. Army Research Office under grant number W911NF-15-1-0202.

References (98)

  • K. Prasad et al.

    An eigenvalue method for computing the burning rates of hmx propellants

    Combust. Flame

    (1998)
  • J.E. Davidson et al.

    A three-phase model of hmx combustion

    Symp. Combust.

    (1996)
  • M.L. Gross et al.

    Fast cook-off modeling of hmx

    Combust. Flame

    (2015)
  • K.V. Meredith et al.

    Laser-induced ignition modeling of hmx

    Combust. Flame

    (2015)
  • X. Chen et al.

    Predictive kinetics for the thermal decomposition of rdx

    Proc. Combust. Inst.

    (2019)
  • M. Greenfield et al.

    Ultrafast photodissociation dynamics of hmx and rdx from their excited electronic states via femtosecond laser pump-probe techniques

    Chem. Phys. Lett.

    (2006)
  • L. Patidar et al.

    Quantum mechanics investigation of initial reaction pathways and early ring-opening reactions in thermal decomposition of liquid-phase rdx

    Combust. Flame

    (2017)
  • L. Patidar et al.

    Identification of initial decomposition reactions in liquid-phase hmx using quantum mechanics calculations

    Combust. Flame

    (2018)
  • A. Chowdhury et al.

    Confined rapid thermolysis/FTIR/ToF studies of imidazolium-based ionic liquids

    Thermochim. Acta

    (2006)
  • L.S. Rothman et al.

    The hitran 2008 molecular spectroscopic database

    J. Quant. Spectrosc. Radiat. Transf.

    (2009)
  • M. Khichar et al.

    Improvement and validation of a detailed reaction mechanism for thermal decomposition of rdx in liquid phase

    Combust. Flame

    (2018)
  • Y. Oyumi et al.

    Thermal decomposition of energetic materials 3. A high-rate, in situ, ftir study of the thermolysis of rdx and hmx with pressure and heating rate as variables

    Combust. Flame

    (1985)
  • J.S. You et al.

    Thermal decomposition kinetics of gap ETPE/RDX-based solid propellant

    Thermochim. Acta

    (2012)
  • C.F. Melius et al.

    Initial reaction steps in the condensed-phase decomposition of propellants

    Proc. Combust. Inst.

    (2002)
  • H.H. Cady et al.

    The crystal structure of α-HMX and a refinement of the structure of β-HMX

    Acta Crystallogr.

    (1963)
  • C.S. Choi et al.

    A study of the crystal structure of β-cyclotetramethylene tetranitramine by neutron diffraction

    Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem.

    (1970)
  • R.E. Cobbledick et al.

    The crystal structure of the δ-form of 1,3,5,7-tetranitro-1,3,5,7-tetraazacyclooctane (δ-HMX)

    Acta Crystallogr. Sect. B Struct. Crystallogr. Cryst. Chem.

    (1974)
  • W. Zhu et al.

    First-principles study of the four polymorphs of crystalline octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine

    J. Phys. Chem. B

    (2007)
  • R.W. Molt et al.

    RDX geometries, excited states, and revised energy ordering of conformers via MP2 and CCSD(T) methodologies: insights into decomposition mechanism

    J. Phys. Chem. A

    (2011)
  • R.W. Molt et al.

    The great diversity of hmx conformers: probing the potential energy surface using CCSD(T)

    J. Phys. Chem. A

    (2013)
  • R.J. Karpowicz et al.

    The β-δ transformation of hmx – its thermal analysis and relationship to propellants

    AIAA J.

    (1982)
  • T.B. Brill et al.

    Solid phase transition kinetics. The role of intermolecular forces in the condensed-phase decomposition of octahydro-1, 3, 5, 7-tetranitro-1, 3, 5, 7-tetrazocine

    J. Phys. Chem.

    (1982)
  • B.F. Henson et al.

    Dynamic measurement of the hmx β-δ phase transition by second harmonic generation

    Phys. Rev. Lett.

    (1999)
  • B.F. Henson et al.

    The β–δ phase transition in the energetic nitramine octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine: thermodynamics

    J. Chem. Phys.

    (2002)
  • L. Smilowitz et al.

    The β–δ phase transition in the energetic nitramine-octahydro-1,3,5,7-tetranitro-1,3,5,7-tetrazocine: kinetics

    J. Chem. Phys.

    (2002)
  • C.-.C. Ye et al.

    Initial decomposition of hmx energetic material from quantum molecular dynamics and the molecular structure transition of β-HMX to δ-HMX

    J. Phys. Chem. C

    (2019)
  • P.C. Myint et al.

    Thermodynamics of hmx polymorphs and hmx/rdx mixtures

    Ind. Eng. Chem. Res.

    (2016)
  • S.K. Bhattacharia et al.

    Melting behavior and heat of fusion of compounds that undergo simultaneous melting and decomposition: an investigation with hmx

    J. Chem. Eng. Data

    (2017)
  • Y.Y. Maksimov

    Boiling point and enthalpy of evaporation of liquid hexogen and octogen

    Russ. J. Phys. Chem.

    (1992)
  • M.A. Schroeder

    Critical analysis of nitramine decomposition data: Product distributions from hmx and rdx decomposition

    (1985)
  • R. Behrens

    Identification of octahydro208;1, 3, 5, 7‐tetranitro‐1, 3, 5, 7‐;tetrazocine (HMX) pyrolysis products by simultaneous thermogravimetric modulated beam mass spectrometry and time‐of‐flight velocity‐spectra measurements

    Int. J. Chem. Kinet.

    (1990)
  • S. Bulusu et al.

    A review of the thermal decomposition pathways in RDX, hmx and other closely related cyclic nitramines

    Def. Sci. J.

    (1996)
  • M.L. Hobbs

    A global hmx decomposition model

  • R. Behrens Jr et al.

    A zero-dimensional model of experimental thermal decomposition of hmx

  • P.M. Dickson et al.

    Measurement of phase change and thermal decomposition kinetics during cookoff of pbx 9501

  • M.J. Kaneshige et al.

    Cook-off experiments for model validation at sandia national laboratories

  • P.M. Dickson et al.

    Thermal cook–off response of confined pbx 9501

    Proc. R. Soc. Lond. A Math. Phys. Eng. Sci.

    (2004)
  • E. Catalano et al.

    Thermal decomposition and reaction of confined explosives

  • T.D. Tran et al.

    Thermal decomposition of trinitrotoluene (TNT) with a new one-dimensional time to explosion (ODTX) apparatus

  • Cited by (18)

    • A reduced mechanism with optimal rate-kinetics parameters for liquid-phase decomposition of bis(triaminoguanidinium) 5,5'-azotetrazolate (TAGzT): Quantum chemical calculations, thermolysis experiments and kinetic modeling

      2021, Thermochimica Acta
      Citation Excerpt :

      In both sets of experiments, the evolved product gases freely escape the heated sample and mix with inert nitrogen gas. Hence, the experimental technique is suitable to study the liquid-phase decomposition [29,30]. Experiments with a closed crucible or a crucible with pinhole lid were not performed to minimize gas-phase reactions.

    View all citing articles on Scopus
    View full text