12
Azole energetic materials: Initial mechanisms for the energy release from electronical excited nitropyrazoles Bing Yuan, Zijun Yu, and Elliot R. Bernstein Citation: The Journal of Chemical Physics 140, 034320 (2014); doi: 10.1063/1.4861670 View online: http://dx.doi.org/10.1063/1.4861670 View Table of Contents: http://aip.scitation.org/toc/jcp/140/3 Published by the American Institute of Physics

excited nitropyrazoles Bing Yuan, Zijun Yu, and Elliot R ...thermal stability and impact/shock resistance in order to min-imize the risk of accidental discharge. Environmental and

  • Upload
    others

  • View
    0

  • Download
    0

Embed Size (px)

Citation preview

  • Azole energetic materials: Initial mechanisms for the energy release from electronicalexcited nitropyrazolesBing Yuan, Zijun Yu, and Elliot R. Bernstein

    Citation: The Journal of Chemical Physics 140, 034320 (2014); doi: 10.1063/1.4861670View online: http://dx.doi.org/10.1063/1.4861670View Table of Contents: http://aip.scitation.org/toc/jcp/140/3Published by the American Institute of Physics

    http://oasc12039.247realmedia.com/RealMedia/ads/click_lx.ads/www.aip.org/pt/adcenter/pdfcover_test/L-37/20939943/x01/AIP-PT/JCP_ArticleDL_0117/PTBG_orange_1640x440.jpg/434f71374e315a556e61414141774c75?xhttp://aip.scitation.org/author/Yuan%2C+Binghttp://aip.scitation.org/author/Yu%2C+Zijunhttp://aip.scitation.org/author/Bernstein%2C+Elliot+R/loi/jcphttp://dx.doi.org/10.1063/1.4861670http://aip.scitation.org/toc/jcp/140/3http://aip.scitation.org/publisher/

  • THE JOURNAL OF CHEMICAL PHYSICS 140, 034320 (2014)

    Azole energetic materials: Initial mechanisms for the energy releasefrom electronical excited nitropyrazoles

    Bing Yuan, Zijun Yu, and Elliot R. Bernsteina)Department of Chemistry, Colorado State University, Fort Collins, Colorado 80523-1872, USA

    (Received 4 November 2013; accepted 24 December 2013; published online 21 January 2014)

    Decomposition of energetic material 3,4-dinitropyrazole (DNP) and two model molecules4-nitropyrazole and 1-nitropyrazole is investigated both theoretically and experimentally. The initialdecomposition mechanisms for these three nitropyrazoles are explored with complete active spaceself-consistent field (CASSCF) level. The NO molecule is observed as an initial decomposition prod-uct from all three materials subsequent to UV excitation. Observed NO products are rotationally cold(pyrazole.15Using time-resolved pulsed photoacoustic spectroscopy toexcite some of the resonant modes of compounds at differ-ent temperatures, NO2 is observed as a major ground elec-tronic state product and the same stability orders are found.16

    The properties of pyrazoles substituted with NO2 at differentpyrazole ring positions are quite useful for further studies ofDNP. The more NO2 groups on the pyrazole ring the less sta-ble the molecule becomes. Nonetheless, detailed comparisonsbetween energetic dinitro- and non-energetic (model) mononitro-pyrazole molecules with regard to their decompositionreactions and mechanisms should prove very illuminating and

    0021-9606/2014/140(3)/034320/11/$30.00 © 2014 AIP Publishing LLC140, 034320-1

    http://dx.doi.org/10.1063/1.4861670http://dx.doi.org/10.1063/1.4861670http://dx.doi.org/10.1063/1.4861670mailto: [email protected]://crossmark.crossref.org/dialog/?doi=10.1063/1.4861670&domain=pdf&date_stamp=2014-01-21

  • 034320-2 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    FIG. 1. Chemical structures of 4-nitropyrazole, 1-nitropyrazole, and DNP.

    insightful, as at present the mechanistic differences betweenenergetic and model system behavior is not understood forthese species.

    Many energetic materials are organic molecules withNO2 groups bonded to a carbon, nitrogen, or oxygen atom.From previous studies, the initial step in the decomposi-tion process of these isolated molecules, following electronicexcitation, is X-NO2 (X = C,N,O) to X-ONO isomeriza-tion (a nitro-to-nitrite, unimolecular reaction): the isomer-ized species then decomposes to form an NO product.3, 18

    Electronic excitation of condensed phase energetic specieshas been suggested and recognized as a major factor in therelease of stored energy from energetic molecules over thepast 50 years.1–6, 18–23 In this instance, conical intersectionsbetween different potential energy surfaces play a key rolein the ultrafast (∼100 fs) decomposition mechanisms.5, 6, 24In nonadiabatic unimolecular chemistry, a conical intersec-tion is the crossing (interaction) point between two elec-tronic states or two potential energy surfaces: conical inter-sections are widely employed to explain and understand theexcited state chemistry of organic and inorganic molecules.In our previous study of energetic and model molecules, in-cluding 1,3,5-trinitro-1,3,5-triazacclohexane (RDX), 1,3,5,7-tetranitro-1,3,5,7-tetraazacyclooctane (HXM), Hexanitrohex-aazaisowurtzitane (CL-20), s-tetrazine, 3-amino-6-chloro-1,2,4,5-tetrazine-2,4-dioxide (ACTO), 3,6-diamino-1,2,4,5-tetrazine-1,4-dioxide (DATO), diaminofurazan (DAF), fu-razan, pentaerythritol tetranitrate (PETN), 2-nitroimidazole,4-nitroimidazole, 1-methyl-5-nitroimidazole, and dimethyl-nitramine (DMNA), the existence of conical intersectionsin the X-ONO isomerization reaction coordinate providesa direct, nonadiabatic, decomposition pathway in the ex-cited state, which successfully explains the experimentalobservations.25–27, 29–32, 34–38

    In this work, we focus on understanding the first-stepin the unimolecular decomposition mechanisms for the en-ergetic material 3,4-dinitropyrazole and two model, non-energetic molecules 1-nitropyrazole and 4-nitropyrazole, fol-lowing electronic excitation. Experimentally, energy resolvedspectroscopy is studied to define dynamics of the reactionthrough elucidating the rotational and vibrational tempera-tures of the initial decomposition product NO. Employingquantum chemistry calculations (Gaussian 09, CASSCF), po-tential energy surfaces for a number of excited states and theground electronic state are explored. The detailed decompo-sition mechanisms for these three molecules are determinedand discussed. Through comparisons between energetic andmodel molecules, a more complete comprehension of the

    energetic molecule properties obtains. Such insights can beemployed to develop new and better energy storage materialsfor both general use and specific purposes.

    II. EXPERIMENTAL PROCEDURES

    The experimental setup consists of a matrix-assisted laserdesorption (MALD) system, a supersonic jet expansion noz-zle, and a time of flight mass spectrometer. Details of the in-strumental design are described in our previous papers.25, 26

    The nozzle used for the molecular beam generation is con-structed from a Jordan Co. pulsed valve and a laser abla-tion attachment. The laser desorption head is attached to thefront of the pulsed valve with three significant parts: (1) a2 × 60 mm channel for the expansion gas from the nozzle,(2) a conical channel (3 mm at the outside and 1 mm at the in-tersection with gas expansion channel) for the ablation laserbeam perpendicular to the expansion gas channel, and (3) a40 mm diameter hole for the sample drum. The sample drumfits into the 40 mm hole and is simultaneously rotated andtranslated by a motor and gear system in the vacuum in orderto present a fresh sample region to the ablation laser for eachpulse. The nonvolatile samples are desorbed from the drumby 532 nm ablation laser, entrained in the flow of He carriedgas under a pressure of 80 psi through the 2 × 60 mm chan-nel in the laser desorption head and expanded into the vac-uum chamber. With 80 psi He backing pressure for the closedpulsed valve, the chamber pressure remains 2 × 10−7 torr;with the valve open at 10 Hz, the chamber pressure increasesto 4 × 10−7 torr.

    All sample drums for MALD are prepared by wrappinga piece of porous filter paper around a clean Al drum. A so-lution of 0.02 mol/l matrix (R6G) and 0.01 mol/l sample inacetone is uniformly sprayed on the drum surface while it isrotating under a halogen heat lamp in a fume hood to makesure the sample coating is dry. An air atomizing spray nozzle(Spraying System Co.) with siphon pressure of 10 psi is usedto deposit sample plus matrix on the filter paper surface. Thedried drum with well-distributed sample is then placed in thelaser ablation head assembly and put into the vacuum cham-ber for decomposition reaction studies. The DNP is suppliedby Dr. Rao Surapaneni and Mr. Reddy Damavarapu (ARL,Picatinny Arsenal, N.J.). 1-nitropyraozle and 4-nitropyrazoleare bought from Sigma-aldrich Co. at 98% and 97% purity,respectively.

    In addition to ablation laser, one or two others laser arerequired to photoexcite the sample in the beam and then de-tect the photo dissociated fragments. For NO detection, asingle pump/probe laser is used at 226/236/248 nm to bothinitiation of energetic sample and detection of the NO prod-uct following a one color (1+1) resonance-enhanced two pho-ton ionization (R2PI) scheme [A(v′ = 0) ← X(v′′ = 0-2)and I ← A transitions] through TOFMS. The proper UV laserwavelengths for this process are generated by a dye laser,pumped by the second harmonic (532 nm) of a Nd: yttriumaluminum garnet laser’s fundamental output (1.064 μm), inconjunction with a wavelength extension system. The typicalpulse energy of the UV laser is 70–700 μJ/pulse, which givesan intensity of ∼2.0 × 107–2.8 × 108 W/cm2 for a 8 ns pulse

  • 034320-3 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    duration. The spectra are not saturated and only single pho-ton absorption occurs for the sample and the NO product at∼50 μJ/pulse between 248 and 226 nm. The molecular beamis perpendicularly crossed by the UV laser beam, which is fo-cused to a spot size of about 0.2 mm diameter at the ionizationregion of the TOFMS.

    The timing sequence of pulsed nozzle, ablation laser,and excitation/ionization laser are controlled by time delaygenerators (SRS DG535). The experiment is run at a repeti-tion rate of 10 Hz. Ion signals in the TOFMS are detectedby a microchannel plate (MCP) and signals are recordedand processed on a personal computer (PC) using an ADCcard (Analog Devices RTI-800) and a boxcar averager (SRSSR 250).

    III. COMPUTATIONAL METHODS

    All calculations are executed at the CASSCF(12,8)/6-31+G(d) level of theory with the Gaussian 09 program. Nosymmetry restrictions are applied for the calculations. Equi-librium geometry calculations are conducted taking the totalcharge as neutral and the spin multiplicity as 1. To explorethe excited state potential energy surfaces, the active spacecomprises 12 electrons distributed in 8 orbitals, denoted asCASSCF (12,8). Orbitals used for the 4-nitropyrazole activespace are two π -bonding orbitals of the five-member ringπ1 and π2, two NO nonbonding orbital nσ NO1 and nσ NO2,one π -nonbonding orbital nπO, one σ nonbonding orbitalnσ O, one delocalized ONO π -antibonding orbital πONO∗, andone π -antibonding orbital of the five member ring π∗ asshown in Figure 2(a). Orbitals of 1-nitropyrazole are sim-ilar. Orbitals used for the DNP active space are shown inFigure 2(b); they include two π -nonbonding orbital πONO-C4and πONO-C3, three σ nonbonding orbital nσ NO-C3(1),nσ NO-C3(2), and nσ NO-C4, one π -nonbonding orbital of the five-member ring πRING, and two delocalized ONO π -antibondingorbital π∗ONO-C4 and π∗ONO-C3. To avoid convergence failureproblems, a reduced (6,5) active space is also used initially forthe excited state calculations: the CASSCF(12,8) is appliedto get a more accurate energy values. At the larger CASSCFlevel orbitals for NO2 and the five-member ring in DNP/singlenitropyraozles are coupled and interact with each other. Thisinteraction under the larger CASSCF is especially importantfor representation of highly excited states. Excitation ener-gies are calculated by state averaging over the ground and ex-cited states with equal weights for each state. Larger basis setsthan 6-31(d) for these CASSCF calculations do not substan-tially improve the results and understanding of the reactionmechanisms.37

    Critical points (minima and transition state structures) arecharacterized by analytical frequency calculations, and min-imum energy paths are calculated using an intrinsic reactioncoordinate (IRC) algorithm implemented in the Gaussian 09.DNP contains two different NO2 groups which are not sym-metric and to calculate the different reaction paths and crit-ical points (conical intersections and transition states) withcertain C-NO2 bond breaking, CASSCF(6,5) is employed inthe beginning with chosen NO2 orbitals located only on oneNO2 group. Following this, the (12,8) active space is applied

    FIG. 2. (a) Orbitals used in the active space for CASSCF calculations for 4-nitropyrazole. The (12,8) active space comprises nonbonding orbitals, bond-ing, and antibonding π orbitals. Similar orbitals are employed to describe thestates of 1-nitropyrazole. (b) Orbitals used in the active space for CASSCFcalculations for 3,4-dinitropyrazole (DNP).

    to get a better potential energy surface. To find the transi-tion states and intermediate states along the reaction path-ways, a relaxed scan optimization algorithm as implementedis employed in which all geometrical parameters except forthe specified bond distance are optimized and electronic en-ergies are monitored as the specified bond is elongated. In thescan, the structure with peak potential energy is most likelya transition state, and the structure with potential energy in avalley is most likely an intermediate state. To verify this con-clusion and obtain a more accurate potential energy surfacefor the transition/intermediate states, the molecular structureprovided in the scan is used as the initial structure in the fol-lowing optimization calculation with CASSCF(12,8) as theactive space. The accuracies of the calculations along the re-action pathway are difficult to estimate since experimental in-formation about the conical intersections and the transitionstates is not available. Calculations presented in this paper,

  • 034320-4 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    however, are based on the experimental observations includ-ing decomposition products and the internal energy distribu-tions within these products. Therefore, the proposed reactionpathways based on the computational results provide a rea-sonable and, at minimum qualitative, interpretation for theexperimental observations.

    IV. EXPERIMENTAL RESULTS AND DISCUSSION

    NO product is observed as a major decompositionproduct from electronically excited DNP at 226/236/248 nmby TOFMS. The 226/236/248 nm excitation wavelength alsocorresponds to the resonance (0-0)/(0-1)/(0-2) vibronic bandsof the A2�+ ← X2� electronic transition of the NO product.Therefore, by scanning these laser excitation wavelengths, a(1+1) R2PI rotationally resolved spectrum of the NO prod-uct from DNP obtains. The line width of the NO mass peak is10 ns (the laser pulse width), consistent with that expectedfor decomposition from the lowest few excited states. If DNPwere to absorb two or more photons sequentially, the NOmass signal would broaden as we have reported earlier.27

    Thus, the NO product is associated with single photon absorp-tion for these three samples. In this experiment, laser beam in-tensity is varied without change in the NO TOFMS line width,so that multiphoton dissociation of DNP is unlikely. Excitedelectronic states of DNP, which might be generated in the ab-lation process, are effectively relaxed and cooled in the highlycollisional expansion process, through the supersonic nozzle.

    Figure 3 shows the spectra of three different A(v′ = 0) ← X(v′′ = 0,1,2) rovibronic transitions of the NOmolecule generated from DNP excited to its first excitedelectronic state. All spectra have similar rotational patternsbut a varying vibrational intensity for each vibronic band.The most intense figure in each spectrum of NO correspondsto the (Q11+P12) band and the lower intensity features

    FIG. 3. One color (1+1) R2PI spectra of the vibronic transitions A2�+(v′= 0) ← X2�(v′ ′ = 0,1,2) of the NO product from the excited electronic statedecomposition of DNP. Rotational simulations with a Boltzmann populationdistribution show that the three observed vibrational levels of the ground elec-tronic state have a similar rotational temperature of ∼25 K with ±10 K uncer-tainty. The black line is the experimental data and the red one is the theoreticalsimulation for NO (A-X) transition at rotational temperature 25 K.

    for each vibronic transition correspond to other rotationaltransitions.28 Spectral simulations based on Boltzmannpopulation distributions for the three vibronic transitionsof the NO product from DNP produce similar rotationaltemperatures (25 ± 10 K) for the three vibrational levels(v′′ = 0,1,2) in the ground electronic state of NO. Thevibrational temperature of the NO product can be obtainedby simulating the relative intensities among the observedvibronic bonds, using a Boltzmann population distributionanalysis and the Frank-Condon factors. The vibrational tem-perature of NO product from DNP is calculated to be (3850± 50) K and details of the vibrational temperature calculationwill be described below. In summary, the NO product fromelectronically excited DNP has a low rotational temperatureand a hot vibrational temperature, which is similar to ourprevious results for all NO2 containing energetic molecules.

    Even though the MALD technique is known to desorbeasily fragile molecules in the gas phase without fragmen-tation, NO product is proven to be created in the excitationregion in our previous study in three ways: comparison of NOvelocity distributions from the nozzle, determination of arrivaltime for the NO signal intensity as a function of nozzle/pumplaser timing, and NO rotational and vibrational temperaturedetermination as a function of source of NO.25

    NO product can be produced from several possible ways,including N-O bond breaking after nitro-nitrite isomeriza-tion, NO2 photolysis, and HONO elimination. In our previ-ous study of NO2 photolysis, the rotational temperature ofNO product from the expansion cooled NO2 gas is about130 K for the A(v′ = 0) ← X(v′′ = 0) vibronic transitionand the temperature goes up to 400 K for vibronic transi-tion A(v′ = 0) ← X(v′′ = 1). As the rotational temperatureof the NO product from DNP is much lower than that forNO2 photolysis, NO2 is not the precursor for NO product, andNO2 loss is not the major reaction channel for the decompo-sition of electronically excited DNP. The observed rotationaland vibrational temperatures assigned for NO in these exper-iments show clearly that the NO must come from decompo-sition of nitropyrazoles and not NO2.25, 29–31 HONO is alsoone of the possible intermediates in the formation of NO fromDNP or single nitro-pyrazoles. HONO elimination for previ-ously studied X-NO2 system is less than 5% of the total NOelimination following the nitro-nitrite isomerization; more-over, the rotational temperature of NO product from HONOphotolysis is extremely high.25, 29 Based on calculations for 4-nitropyrazole, the energy required for HONO elimination onthe ground state surface is 107 kcal/mol, 21 kcal/mol higherthan for the formation of NO through the nitro-nitrite isomer-ization, which makes HONO elimination less likely in nitro-pyrazoles: this result is similar to those for the nitro-imidazolesysems.32 Thus, neither the NO2 loss channel nor theHONO channel is present as a competitor to the nitro-nitriteisomerization/NO elimination channel.

    In efforts to improve our understanding of the ex-cited state decomposition mechanisms of DNP, two modelmolecules 1-nitropyrazole and 4-nitropyrazole are explored.Figure 4 and Figure 1S in the supplementary material39 il-lustrate the A ← X (0-0,1,2) transition spectra with rotationalspectra simulations for the initial decomposition product NO

  • 034320-5 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    FIG. 4. One color (1 + 1) R2PI spectra of the vibronic transitions A2�+(v′= 0) ← X2�(v′ ′ = 0,1,2) of the NO product from the excited electronic statedecomposition of 4-nitropyrazole. Rotational simulations with a Boltzmannpopulation distribution show that the rotation temperatures of the three ob-served vibrational levels on the ground electronic state (0-0), (0-1), and (0-2)are 25 K, 30 K, and 50 K, respectively, with ±10 K uncertainty. The blackline is the experimental data and the red one is theoretical simulation for NO(A-X) transition at rotational temperature 25 K, 30 K, and 50 K from (0-0) to(0-2).

    for the two model molecules. The simulated rotational tem-peratures for these two model molecules are similar and bothincrease from (25 ± 10) K to (50 ± 10) K as the vibronicbands change from the (0-0) to (0-2) transition. The vibra-tional temperatures of NO from the two model materialsare both calculated to be (2300 ± 300) K, 1350 K lowerthan NO from the energetic molecule DNP. The two mononitro-pyrazoles may follow similar decomposition routes, butthey should be different from that followed by the energeticmaterial DNP.

    The rotational and vibrational temperatures for the NOproduct from the three materials in the experiment are sum-marized in Table I. The signal intensities for the NO prod-uct of 4-nitropyrazole, 1-nitropyrazole, and DNP at differentvibrational transitions A(v′ = 0) ← X(v′′ = 0,1,2) and tran-sition ratios I(0-2)/(0-0) and I(0-1)/(0-0) are summarized inTable II and the Franck-Condon factors for the NO (A-X)transition in the temperature range from 1900 K to 4000 Kare listed in Table III. The Franck-Condon factors are cal-culated using the program, Lifbase: Data Base and Spec-tra Simulation (version 2.1.1). For the rotational temperaturerange 25–50 K, these simulated Franck-Condon factors are

    TABLE I. Rotational and vibrational temperatures for NO product fromDNP, 1-nitropyrazole, and 4-nitropyrazole. The uncertainty of experimentalrotational temperature is ±10 K.

    NO rotational temperature (K) Vibrational

    0-0 0-1 0-2 temperature (K)

    3,4-dinitropyrazole(DNP, energetic)

    25 25 25 3850 ± 50

    4-nitropyrazole 25 30 50 2300 ± 3001-nitropyrazole 25 35 50 2300 ± 300

    unchanged. For NO from 4-nitropyrazole, the experimentalI(0-1)/(0-0) ratio equals 0.36, yielding a vibrational temper-ature of about 2600 K from Table III. The I(0-2)/(0-0) ratiofor NO from 4-nitropyrazole is 0.068 which gives a ∼2000 Kvibrational temperature in Table III. The two values are dif-ferent due to experimental uncertainties; for example, laserintensity is not stable, sample spaying is not exactly uniform,etc. The vibrational temperature of 4-nitropyrazole is deter-mined to be (2300 ± 300) K. Similarly, the vibrational tem-perature of NO from 1-nitropyraozle and 3,4-dinitropyraozleare (2300 ± 300) K and (3850 ± 50) K, respectively.

    The experimental results prove that NO is a primary de-composition product of DNP and the NO product from DNPhas a low rotational temperature and a high vibrational tem-perature. The reaction decomposition mechanisms of ener-getic material DNP and model molecules should be differentas the vibrational temperature of the NO product from DNPis 1350 K higher than that of NO from the model molecules.The NO product from DNP is most likely formed in a lowerelectronic state than that of the model molecules: the detailsof the reaction mechanism and reaction potential energy sur-faces will be discussed based on the theoretical calculationspresented in Sec. V.

    V. THEORETICAL RESULTS AND DISCUSSION

    The experimental results yield that the NO molecule isan initial nanosecond UV decomposition product for bothelectronically excited energetic and model systems. In orderto understand the experimental data more completely, the-oretical calculations of molecular geometries and energiesfor the Franck Condon structure, conical intersections, tran-sition states, intermediate states, and nitro-nitrite isomeriza-tions along the ground and excited state potential energy sur-faces, are performed for the energetic system DNP and model

    TABLE II. Experimental maximum signal intensities of the vibrational transitions A(v′ = 0) ← X(v′ ′ = 0,1,2)of NO product for materials 4-nitropyrazole, 1-nitropyrazole, and DNP when the laser intensity is 0.65 mJ.

    Signal intensity/V (laser intensity 0.65 mJ) Ratios

    Materials v = (0-0) v = (0-1) v = (0-2) (0-1)/(0-0) (0-2)/(0-0)

    4-nitropyrazole 2.5 0.8 0.17 0.35 0.0681-nitropyrazole 1.1 0.39 0.08 0.35 0.0733,4-dinitropyrazole 4.0 1.97 1.0 0.49 0.25

  • 034320-6 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    TABLE III. The Frank-Condon factors for transitions (0-2)/(0-0) and (0-1)/(0-0) calculated using Lifbase: Data base and spectra simulation (version2.1.1). I(. . . ) is intensity for the transition ratio indicated.

    Vibrationaltemperature/1000 K 1.9 2.0 2.1 2.5 2.6 2.7 3.8 3.9 4.0

    I(0-1)/(0-0) 0.24 0.26 0.28 0.34 0.35 0.37 0.49 0.50 0.51I(0-2)/(0-0) 0.06 0.07 0.08 0.12 0.13 0.14 0.24 0.25 0.26

    systems 4-nitropyrazole and 1-nitropyrazole. The theoreticalreaction paths with potential energies and molecular geome-tries are shown in Figures 5–8 and 2S–5S (supplementarymaterial).39

    Two possible decomposition mechanisms are explored inthe theoretical study. The first mechanism includes an excitedstate decomposition through transition states and nitro-nitriteisomerization on the first excited electronic state S1 poten-tial energy surface. The second mechanism is a ground statedecomposition process which goes through a conical intersec-tion (S1/S0)CI between the S1 and S0 states followed by an S0nitro-nitrite isomerization. These two channels are chosen forcomparison because the decomposition of energetic materialDNP generates NO product with higher vibrational temper-ature than the model systems. The actual reaction pathwaywill depend on different factors, for example, the rate of in-ternal vibrational energy redistribution, the heights of reac-tion barriers, and the rate of nonadiabatic transition through

    FIG. 5. A schematic one-dimensional projection of the multi-dimensionalenergy surfaces for 4-nitropyrazole dissociation computed at theCASSCF(12,8)/6-31+G(d) level of theory. The red arrows representthe more favorable reaction path for NO dissociation, which occurs onthe excited electronic state S1. The black arrows represent the other, lessfavorable, reaction path, along which NO is formed on the ground electronicstate S0. FC geometry S0,FC is the optimized minimum energy structure of4-nitropyrazole on S0. S1,min and S2,min are the structure on the first andsecond excited electronic states with minimum potential energy, respectively.(S1/S0)CI is the conical intersection between S0 and S1 states, and (S2/S1)CIis the conical intersection between S2 and S1 states. S1,TS1 to S1,TS4 are theexcited transition states on the S1 surface, S0,TS is transition state on the S0surface, and S1,IM is the intermediate state on S1 state. S0,nitrite and S1,nitriteare the nitro-nitrite isomerization intermediates on the S0 and S1 states,respectively, and S0,NO and S1,NO are NO dissociated products on these twostates.

    FIG. 6. The structures of all the critical points and conical intersections men-tioned in Figure 5 along the 4-nitropyrazole dissociation reaction path. Ar-rows in the structures of the transition states show the reaction coordinateassociated with the imaginary frequency (I.F.).

    the different conical intersections which can only be assessedby dynamical calculations.

    Calculations at the CASSCF(12,8)/6-31+G(d) theoreti-cal level are employed to determine the vertical excitation en-ergies for 4-nitropyrazole, 1-nitropyrazole, and DNP from theground electronic state S0 to the first excited state S1 and thesecond excited state S2. The vertical excitation energies of the

    FIG. 7. A schematic one-dimensional projection of the multi-dimensionalenergy surfaces for 3,4-dinitropyrazole with the NO2 bonded to C4 on thepyrazole ring dissociated as the first step computed at the CASSCF(12,8)/6-31+G(d) level of theory. The dashed arrows represent the reaction path ofNO dissociation occurring on the excited electronic state S1: this pathway isnot available based on energy considerations. The red arrows represent thereaction path on which NO is formed on the ground electronic state S0: thispathway is energetically allowed. FC geometry S0,FC is the optimized min-imum energy structure of 3,4-dinitropyrazole on the ground electronic stateS0. S1,min and S2,min are the structure on the first and second excited electronicstates with minimum potential energy, respectively. (S1/S0)CI is the conicalintersection between S0 and S1 states and (S2/S1)CI is the conical intersectionbetween S2 and S1 states. S1,TS1 and S1,TS2 are the excited transition stateson the S1 surface and S0,TS is transition state on the S0 surface. S0,nitrite andS1,nitrite are the nitro-nitrite isomerization in S0 and S1 states, respectively,while S0,NO NO dissociated products in ground electronic state.

  • 034320-7 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    FIG. 8. The structures of all critical points and conical intersections men-tioned in Figure 7 in which 3,4-dinitropyrazole with one NO2 connected toC4 on the pyrazole ring dissociates as the first step. Arrows on the struc-tures of the transition states show the reaction coordinates of the imaginaryfrequency (I.F.).

    three materials are very close to the energy of single photonexcitations in the range from 226 nm to 248 nm (5.0 eV to5.5 eV) which have been used to excite these molecules. Sincethere is no experimental data for vertical excitation energiesof nitropyrazole molecules published so far, the experimen-tal UV-Vis absorption spectra of nitropyrazoles are taken todetermine the accuracy of the CASSCF calculations. Verti-cal energies for the three materials and the related UV-Vismaximum absorption wavelength14, 33, 34 are summarized inTable IV. As shown in Table IV, the vertical excitations calcu-lated through CASSCF(12,8)/6-31+G(d) for the two lowestlying excited electronic states (S1 and S2) of 4-nitropyrazoleare 4.16 eV and 4.78 eV, respectively. The UV-VIS absorp-tion of 4-nitropyrazole has two maximum at 3.91 ± 0.67 eVand 4.62 ± 0.62 eV. This comparison reveals that theCASSCF(12,8) method gives a reasonable treatment of therelevant excited states as it falls into the commonly accept-able uncertainty range of ±0.5 eV. Similarly, CASSCF(12,8)calculation for 1-nitropyrazole and DNP are also reasonableas the theoretical value is close to the experimental data. TheCASSCF calculations show that the two lowest-lying excitedstates (S1 and S2) for two model materials are (n,π∗) transi-tions from the σ nonbonding orbital nσ O(S1) or nσ NO (S2)of NO2 to delocalized ONO π -antibonding orbital π∗ONO. Inthe third electronic state S3, the transition becomes (π ,π∗)

    from π -bonding orbitals of the five-member ring π1 or π2 toπ∗ONO (orbitals shown in Figure 2(a)). The three lowest-lyingexcited states (S1 to S3) for DNP are (n,π∗) transitions fromthe σ nonbonding orbital nσ NO of NO2 to delocalized ONOπ -antibonding orbital π∗ONO of NO2. The transition to the S1state is from nσ NO-C3(1) of NO2 on C3 to π∗NON-C3 of NO2on C3, to the S2 state is from nσ NO-C4 on C4 to π∗NON-C4 onC4, and to the S3state is from the other σ nonbonding orbitalnσ NO-C3(2) on C3 to π∗NON-C3 on C3 (orbitals shown in Fig-ure 2(b)). From CIS calculations, the four lowest transition ofDNP are all (n,π∗) and the fifth one becomes (π ,π∗) as DNPcontains one more NO2 group than model systems. Thesetransitions are similar to those found for the nitro-imidazolesystems in previous study.31, 32 The potential energy surfacesfor the S3 state of these three molecules are above the laserexcitation maximum. Thus, the S3 state is not considered inthe potential energy surface calculations. From our previousstudies, enlarging the basis set does not significantly improvethe relative excitation energies of critical points, and there-fore, a 6-31+G(d) basis set is used for all calculations tomaintain optimum conditions for computational cost, accu-racy, and comparisons.

    A. Calculations for 4-nitropyrazole

    A schematic one-dimensional projection of the multidi-mensional singlet potential energy surfaces (S0, S1 and S2)of 4-nitropyrazole with locations and potential energies (thepresented energies are not corrected for zero point energy)for different critical points and conical intersections along theminimum energy reaction path is plotted in Figure 5. Thereaction coordinates depicted in Figure 5 include C-N bondlengths, ONC bond angles, and ONCC dihedral angles of theactive site of 4-nitropyrazole. Arrows in Figure 5 indicate dif-ferent possible channels for 4-nitropyrazole decomposition.The structures of each critical point and conical intersectionare summarized in Figure 6 and arrows near the transitionstates and conical intersections show the imaginary frequen-cies (I.F.). In Figures 5 and 6, FC geometry S0,FC is the op-timized minimum energy, planar structure of 4-nitropyrazoleon the S0 state. (S1/S0)CI is the conical intersection betweenthe S0 and S1 states and (S2/S1)CI is the conical intersectionbetween the S2 and S1 states. S1,TS1 to S1,TS4 are the excitedtransition states on the S1 surface and S1,IM is the intermedi-ate state. S0,nitrite and S1,nitrite are the nitro-nitrite isomerizationstates on the S0 and S1 states, respectively, while S0,NO and

    TABLE IV. Comparison of previous experimental data and present CASSCF calculations of vertical excitationenergy S1 and S2 for 4-nitropyrazole, 1-nitropyrazole, and 3,4-dinitropyraozle (DNP).

    CASSCF calculation (eV) UV-vis absorption in experiment (eV)

    S1 S2 λmax,1 λmax,2

    3,4-dinitropyrazole 4.20 4.78 3.97a . . .4-nitropyrazole 4.16 4.78 3.91 ± 0.67b 4.61 ± 0.62b1-nitropyrazole 4.17 4.97 . . . 4.64c

    aUV-vis absorption in 0.05 N NaOH.34bUV-vis absorption in EtOH.33cUV-vis absorption in 1M H2SO4.14

  • 034320-8 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    S1,NO are NO dissociated products in these two states. Theoptimized x, y, z coordinates in angstroms of all the station-ary points for the 4-nitrpyrazole reaction are shown in TableIS(a) in the supplementary material.39 Figure 5 shows that fol-lowing excitation to the second excited electronic state S2, 4-nitropyrazole undergoes a rapid internal conversion from theS2 state to the S1 state through the (S2/S1)CI conical inter-section: molecules undergoing this process would store suffi-cient vibrational energy in the S1 state, transferred from theS2 electronic energy, to dissociate. In this nonadiabatic radi-ationless internal conversion, no significant change occurs inthe nuclear configuration of 4-nitropyrazole (only some rota-tion of the NO2 group) and no barrier exists. The adiabaticenergy gap between the S1 and S2 surfaces near (S1/S2)CI iscomputed to be 88 cm−1 which means the S1 and S2 sur-faces are strongly nonadiabatically coupled with one anotherand the small energy gap increases the probability of nona-diabatic transition from upper to lower electronic states. Asmolecules move from S2 to S1 through (S2/S1)CI, the mostfavorable path for 4-nitropyrazole decomposition shown inFigure 5 is the one colored in red. Along this nitro-nitriteisomerization path, molecules surmount the energy barriersfor two concerted transition states S1,TS2 and S1,TS3 at ener-gies 5.00 eV and 4.94 eV. Then, 4-nitropyrazole undergoesa nitro-nitrite isomerization in the S1 state (S1,nitrite) and fi-nally NO departs the pyrazole ring moiety after passing overa small barrier of 0.26 eV with respect to S1,nitrite. S1,TS2 andS1,TS3 are tight transition states, the negative frequency or un-stable normal mode of which is characterized by out of planebending of the ONCC moiety (a nitro-nitrite isomerization).The IRC algorithm is associated with the bending mode ofthe C-NO2 moiety, which ensures that these two transitionstates lead to the nitro-nitrite isomerization. As NO movesaway from the pyrazole ring, no obvious torque is generatedon the NO product. Therefore, the NO product should have acold rotational temperature, as is consistent with the experi-mental results. The other possible conversion path shown inFigure 5 is that instead of decomposing on the S1 state, 4-nitropyrazole surmounts the energy barrier of S1,TS1(5.21 eV),undergoes another nonadiabatic internal conversion to the S0state through (S1/S0)CI (the energy difference between S1 andS0 state is 7 cm−1) and decomposes on the S0 state via S0,nitriteisomerization. This pathway is energy acceptable under226 nm and 236 nm excitation, as the highest potential energypeak S1,TS1 has lower energy than the laser source; nonethe-less, this potential peak is 0.21 eV higher than the one (S1,TS2)on the S1 decomposition path, making the S0 decompositionpathway less favorable. Under 248 nm excitation, the S0 de-composition pathway for 4-nitropyrazole becomes forbiddendue to energy constraints for the S1,TS1 barrier. In sum, un-der 226 nm and 236 nm excitation, the decomposition of4-nitropyrazole can occur either on the S1 or S0 state sur-faces: the S1 state decomposition is more favorable, how-ever, due to a lower energy barrier. At 248 nm excitation, thedecomposition can only take place on the S1 state surface.

    From the experimental results, NO product from 4-nitropyrazole has a cold rotational temperature (

  • 034320-9 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    Figure 2S, 1-nitropyrazole is excited to its second excitedelectronic state S2 (226–248 nm) and then undergoes a rapidinternal conversion from S2 to S1 through the (S2/S1)CI con-ical intersection. In this nonadiabatic, radiationless internalconversion, no significant change occurs in the nuclear con-figuration and no barrier exists. The adiabatic energy gap be-tween the S1 and S2 surfaces near (S1/S2)CI is computed tobe 153 cm−1. After molecules transition from the S2 to theS1 state through (S2/S1)CI, two decomposition pathways arepossible, as shown in Figure 2S. On one path, molecules sur-mount the energy barriers for two concerted nitro-nitrite iso-merization transition states S1,TS2 and S1,TS3 with 3.93 eV and3.78 eV energy on the S1 state surface, respectively. Thenthe 1-nitropyrazole molecule undergoes a nitro-nitrite isomer-ization (S1,nitrite), and finally NO leaves after passing a smallbarrier of 0.30 eV with respect to S1,nitrite on the S1 state po-tential energy surface. S1,TS2 and S1,TS3 are loose transitionstates, the imaginary frequency is characterized by the out ofplane bending of the ONCC moiety, which corresponds toa true nitro-nitrite isomerization. As NO moves away fromthe pyrazole ring, no obvious torque is found in the calcu-lation acting on the NO product: the NO product should,thereby, have a cold rotational temperature. The other pos-sible reaction path involves as follows: 1-nitropyrazole sur-mounts the energy barrier S1,TS1 (3.93 eV), undergoes anothernonadiabatic internal conversion through (S1/S0)CI conical in-tersection (the energy difference between S1 and S0 state is35 cm−1), and decomposition the S0 state via the S0,nitrite iso-merization with a small barrier of 0.42 eV. Both pathwaysare roughly equally probable, based on their energy valuesand conical intersection adiabatic separations. If the moleculegoes through (S1/S0)CI, it has another two possibilities on theS0 surface: it can return to S0,FC (no reaction) or move to thenitro-nitrite isomerization S0,nitrite, and finally form the NOproduct. Both of these two pathways are energetically ac-ceptable and the potential energy of the molecule changessmoothly along the S0 potential surface without any barriers.In a simple (−�E/�r) reaction force calculation, in which Eis the potential energy and r is the distance along the reactioncoordinate the absolute force for the S0,nitrite reaction coordi-nate is higher by a factor of three for nitro-nitrite isomeriza-tion reaction coordinate. This force difference suggests thatthe S0,nitrite pathway is more favorable as it presents a steeperdescent potential energy surface. In this case, the probabilityof an NO product forming on the S0 state is much greater thanan NO2 product. Thus 1-nitropyrazole is expected, based onthese calculations, to generate NO as an initial decompositionproduct. In sum, under all three excitation conditions, the de-composition of 1-nitropyrazole can occur either on the S1 orS0 state potential energy surface (or perhaps both), as bothreaction paths are energetically accessible and most probablefor NO generation.

    In the NO dissociation process, the molecular energydecreases quickly from S1,TS4 in Figure 2S until the NO groupis 3.5 A away from the pyrazole ring at which point the en-ergy reaches a minimum: the total electronic energy decreasein this process is 2.11 eV. During the molecular decomposi-tion on the S1 state, most of the energy stored in the moleculeis transferred to the vibrational and translational energy of the

    NO product. This energy transfer to the nuclear coordinatesenables the NO product to have a relatively high vibrationaltemperature. The NO bond length at the nitro-nitrite isomer-ization (1.31 A), compared to the equilibrium bond length ofthe NO product (1.14 A), ensures significant excitation of thevibration of NO as the final O-N bond breaks.

    1-nitropyrazole and 4-nitropyrazole have different NO2group positions on the pyrazole ring. Though the reactionmechanism is slightly different, both of them can decomposeeither on the S1 electronic state or S0 electronic state produc-ing NO with similar cold rotational temperature and warm vi-brational temperature. Compared to other model molecules,the behavior of nitropyrazole model materials is closer tothe behavior of nitroimidazole based materials: unlike othermodel systems with hot rotational temperature for NO prod-uct, both species have relatively cold rotational temperatures(

  • 034320-10 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    energy to surmount the transition state S1,TS2 threshold toproduce NO on the S1 potential energy surface. In this case,only the nonadiabatic relaxation pathway through (S1/S0)CIis open for DNP. The S1 potential energy surface of DNP canlead to a nonadiabatically coupled conical intersection fromthe S1 surface to the S0 surface through (S1/S0)CI followingthe minimum energy path. The molecule isomerizes to thenitrite form on its S0 surface and undergoes NO elimination,surmounting a small energy barrier (∼0.70–0.88 eV) tothe nitro-nitrite isomerization transition state. This conicalintersection is energetically accessible and the molecule canrapidly come back to the ground state transferring electronicexcitation energy to its vibrational degrees of freedom. Thestructure near (S1/S0)CI is a loose nitro-nitrite transition stategeometry for which an NO2 moiety at the related active siteof DNP interacts with the rest of the molecule from a longdistance (∼3.0 A) for the different NO2 groups. This NO2group has an out of plane NO2 bending mode with respectto the C-NO2 plane: near (S1/S0)CI the energy differencesbetween S1 and S0 for the two NO2 group are 15 cm−1 (C3)and 4 cm−1 (C4), respectively. No torque is applied to NOduring the decomposition along this reaction coordinate, thusthe NO product should have a cold rotational temperature,consistent with experimental results. In sum, DNP canonly decompose on its S0 ground state under all excitationconditions, generating cold rotational and hot vibrationaltemperatures for the NO initial product. Initial decompositionfor either of the two NO2 groups at C3 or C4 positionsof the pyrazole ring is energetically and structurally quitesimilar.

    The most significant difference between the ener-getic system DNP and model systems 4-nitropyrazole and1-nitropyrazole, is that the model systems have two competi-tive decomposition pathways: model molecules can form NOproduct in either S0 or S1 state, while the energetic systemcan only decompose in the ground state S0, in which more vi-brational energy is stored. The difference is manifested in theexperimental results, as the NO product from DNP is 1350 Kvibrationally hotter than that from the model species. Fromour previous study on energetic/model systems, the NO prod-uct from energetic materials usually has cold rotational tem-perature and hot vibrational temperature while from modelmaterial the NO product is usually rotationally hot and vibra-tionally cold.25–27, 29–32, 34–38 DNP behavior is consistent withthat of previous studies of energetics, while these model sys-tems display somewhat different behavior from that of othermodel systems. Nonetheless, both mono-nitropyrazoles and-nitroimidazoles display similar internal energy distributionsfor the initial NO product. Both nitropyrazoles and nitroimi-dazoles have aromatic rings that have a significant affect thereaction pathways; moreover, their first few electronic excitedstates are of a similar orbital excitation description involv-ing nπ∗ transition on local NO2 groups. The rotational excita-tion of the NO product depends on the structure and reactioncoordinate of the final transition state for the decomposingparent molecule. The vibrational excitation of NO dependson the stored vibrational energy in the particular electronicstate in which the decomposition occurs (e.g., S0 or S1 inthis instance).The general result is that for all NO2 containing

    energetic materials decomposition following electronic exci-tation, NO is the initial product from S0 with a higher degreeof vibrational excitation than found for comparable modelspecies, which typically decompose from an excited elec-tronic state.

    The vibrational temperature of the initial dissociationproduct is important in recognizing energetic molecules be-cause initial decomposition products with high vibrationalexcitation are able to propagate a chain reaction, followingthe initial stimulus, which leads to detonation. The modelmolecules do not have an efficient single pathway to releasetheir entire chemically stored energy during the excited elec-tronic state decomposition process. Multi-nitropyrazoles de-scend to the ground state through a series of conical intersec-tions and dissociate on the ground state PES after nitro-nitriteisomerization to produce an NO initial decomposition prod-uct. This Sn → . . . → So pathway through a series of conicalintersections leaves the molecule on a new part of the S0 po-tential energy surface, not necessary near the FC equilibriumpoint.31, 32 Thereby, all the excitation energy is available tobreak internal bonds to generate reactive fragments and radi-cals for further reactions as maybe required to be classified asan energetic molecule.

    VI. CONCLUSIONS

    The decomposition of the energetic material DNP andtwo model materials 4-nitropyraozle and 1-nitropyrazole fol-lowing electronic excitation has been explored via nanosec-ond, energy resolved spectroscopy. These materials create NOas an initial decomposition product at the nanosecond ex-citation energies (5.0–5.5 eV) with vibrationally hot (modelsystem 2300 ± 300 K, DNP 3850 ± 50 K) and rotation-ally cold (

  • 034320-11 Yuan, Yu, and Bernstein J. Chem. Phys. 140, 034320 (2014)

    ACKNOWLEDGMENTS

    This study is supported by a grant from the U.S. ArmyResearch Office (ARO, FA9550-10-1-0454) and in part by theU.S. National Science Foundation (NSF) through the XSEDEsupercomputer resources provided by NCSA under Grant No.TG-CHE110083. We also want to thank Dr. Rao Surapaneniand Mr. Reddy Damavarapu (ARL, Picatinny Arsenal, N.J.)for supplying the DNP sample used in this study.

    1B. M. Rice, S. Sahu, and F. J. Owens, J. Mol. Struct.: THEOCHEM 583,69 (2002).

    2B. Tan, X. Long, R. Peng, H. Li, B. Jin, S. Chu, and H. Dong, J. Hazard.Mater. 183, 908 (2010).

    3M. R. Manaa, L. E. Fried, C. F. Melius, M. Elstner, and Th. Frauenheim, J.Phys. Chem. A 106, 9024 (2002).

    4A. B. Kunz, M. M. Kuklja, T. R. Botcher, and T. P. Russell, Thermochim.Acta 384, 279 (2002).

    5M. M. Kuklja, J. Appl. Phys. 89, 4156 (2001).6E. J. Reed, J. D. Joannopoulos, and L. E. Fried, Phys. Rev. B 62, 16500(2000).

    7K. E. Gutowski, R. D. Rogers, and D. A. Dixon, J. Phys. Chem. A 110,11890 (2006).

    8K. E. Gutowski, R. D. Rogers, and D. A. Dixon, J. Phys. Chem. B 111,4788 (2007).

    9S. Du, Y. Wang, L. Chen, W. Shi, F. Ren, Y. Li, J. Wang, and D. Cao, J.Mol. Model. 18, 2105 (2012).

    10S. Blanco, J. C. Lopez, J. L. Alonso, O. Mo, M. Yanez, N. Jagerovic, andJ. Elguero, J. Mol. Struct. 344, 241 (1995).

    11I. L. Dalinger, D. V. Khakimov, T. K. Shkineva, I. A. Vatsadze, G. P.Popova, T. S. Pivina, and S. A. Shevelev, Chem. Heterocycl. Compd. 48,1646 (2013).

    12R. K. Toghiani, H. Toghiani, S. W. Maloney, and V. M. Boddu, Fluid PhaseEquilib. 264, 86 (2008).

    13E. N. Rao, P. Ravi, Surya P. Tewari, and S. V. Rao, J. Mol. Struct. 1043,121 (2013).

    14D. Dumanovic, J. Ciric, A. Muk, and V. Nikolic, Talanta 22, 819 (1975).15P. Ravi, A. A. Vargeese, and S. P. Tewari, Thermochim. Acta 550, 83

    (2012).16F. Yehya and A. K. Chaudhary, Appl. Phys. B 110, 15 (2013).

    17J. M. Abboud, R. Notario, M. Yanez, O. Mo, R. Flammang, N. Jagerovic,I. Alkorta, and J. Elguero, J. Phys. Org. Chem. 12, 787 (1999).

    18M. R. Manaa and L. E. Fried, J. Phys. Chem. A 103, 9349 (1999).19B. Tan, R. Peng, X. Long, H. Li, B. Jin, and S. Chu, J. Mol. Model. 18, 583

    (2012).20B. Tan, X. Long, J. Li, F. Nie, and J. Huang, J. Mol. Model. 18, 5127 (2012).21A. B. Kunz and D. R. Beck, Phys. Rev. B 36, 7580 (1987).22C. M. Tarver, AIP. Conf. Proc. 1426, 227 (2012).23S. H. Lee, K. C. Tang, I. C. Chen, M. Schmitt, J. P. Shaffer, T. Schultz, J.

    G. Underwood, M. Z. Zgierski, and A. Stolow, J. Phys Chem. A 106, 8979(2002).

    24A. Bhattacharya, Y. Q. Guo, and E. R. Bernstein, Acc. Chem. Res. 43, 1476(2010).

    25Y. Q. Guo, A. Bhattacharya, and E. R. Bernstein, J. Chem. Phys. 122,244310 (2005).

    26Y. Q. Guo, M. Greemfield, A. Bhattacharya, and E. R. Bernstein, J. Chem.Phys. 127, 154301 (2007).

    27A. Bhattacharya, Y. Q. Guo, and E. R. Bernstein, J. Chem. Phys. 131,194304 (2009).

    28C. J. S. M. Simpson, P. T. Griffiths, H. L. Wallaart, and M. Towrie, Chem.Phys. Lett. 263, 19 (1996).

    29A. Bhattacharya, Y. Q. Guo, and E. R. Bernstein, J. Phys. Chem. A 113,811 (2009).

    30Z. Yu and E. R. Bernstein, J. Chem. Phys. 135, 154305 (2011).31Z. Yu and E. R. Bernstein, J. Chem. Phys. 137, 114303 (2012).32Z. Yu and E. R. Bernstein, J. Phys. Chem. 117, 1756 (2013).33A. I. Vokin, T. N. Komarova, L. I. Larina, and V. A. Lopyrev, Russ. Chem.

    Bull. 46, 297 (1997).34L. Larina and L. Lopyrev, Nitroazoles: Synthesis, Structure and Applica-

    tions (Springer, New York, 2009).35Y. Q. Guo, A. Bhattacharya, and E. R. Bernstein, J. Chem. Phys. 134,

    024318 (2011).36Y. Q. Guo, A. Bhattacharya, and E. R. Bernstein, J. Chem. Phys. 128,

    034303 (2008).37A. Bhattacharya and E. R. Bernstein, J. Phys. Chem. A 115, 4135 (2011).38Y. Q. Guo, A. Bhattacharya, and E. R. Bernstein, J. Phys. Chem. A 115,

    9349 (2011).39See supplementary material at http://dx.doi.org/10.1063/1.4861670 for the

    optimized x, y, z coordinates of the stationary points in all three reac-tions, experimental result for 1-nitropyrazole, theoretical reaction mech-anisms and structures of stationary points for 1-nitropyrazole and 3,4-nitropyrazole(C3) reactions.

    http://dx.doi.org/10.1016/S0166-1280(01)00782-5http://dx.doi.org/10.1016/j.jhazmat.2010.07.115http://dx.doi.org/10.1016/j.jhazmat.2010.07.115http://dx.doi.org/10.1021/jp025668+http://dx.doi.org/10.1021/jp025668+http://dx.doi.org/10.1016/S0040-6031(01)00804-8http://dx.doi.org/10.1016/S0040-6031(01)00804-8http://dx.doi.org/10.1063/1.1350631http://dx.doi.org/10.1103/PhysRevB.62.16500http://dx.doi.org/10.1021/jp0643698http://dx.doi.org/10.1021/jp066420dhttp://dx.doi.org/10.1007/s00894-011-1229-7http://dx.doi.org/10.1007/s00894-011-1229-7http://dx.doi.org/10.1016/0022-2860(94)08434-Jhttp://dx.doi.org/10.1007/s10593-013-1187-3http://dx.doi.org/10.1016/j.fluid.2007.10.018http://dx.doi.org/10.1016/j.fluid.2007.10.018http://dx.doi.org/10.1016/j.molstruc.2013.03.038http://dx.doi.org/10.1016/0039-9140(75)80176-7http://dx.doi.org/10.1016/j.tca.2012.10.003http://dx.doi.org/10.1007/s00340-012-5240-1http://dx.doi.org/10.1002/(SICI)1099-1395(199910)12:103.0.CO;2-Ehttp://dx.doi.org/10.1021/jp992460ohttp://dx.doi.org/10.1007/s00894-011-1096-2http://dx.doi.org/10.1007/s00894-012-1516-yhttp://dx.doi.org/10.1103/PhysRevB.36.7580http://dx.doi.org/10.1063/1.3686260http://dx.doi.org/10.1021%2Fjp021096hhttp://dx.doi.org/10.1021/ar100067fhttp://dx.doi.org/10.1063/1.1929741http://dx.doi.org/10.1063/1.2787587http://dx.doi.org/10.1063/1.2787587http://dx.doi.org/10.1063/1.3262688http://dx.doi.org/10.1016/S0009-2614(97)81412-1http://dx.doi.org/10.1016/S0009-2614(97)81412-1http://dx.doi.org/10.1021/jp807247thttp://dx.doi.org/10.1063/1.3652893http://dx.doi.org/10.1063/1.4752654http://dx.doi.org/10.1021/jp312527uhttp://dx.doi.org/10.1007/BF02494366http://dx.doi.org/10.1007/BF02494366http://dx.doi.org/10.1063/1.3523649http://dx.doi.org/10.1063/1.2822283http://dx.doi.org/10.1021/jp109152phttp://dx.doi.org/10.1021/jp109150uhttp://dx.doi.org/10.1063/1.4861670