Oxidation of a Levitated Droplet of 1-Allyl-3-methylimidazolium … · 2019. 1. 17. · ABSTRACT:...

17
Oxidation of a Levitated Droplet of 1Allyl-3-methylimidazolium Dicyanamide by Nitrogen Dioxide Michael Lucas, Stephen J. Brotton, Shashi Kant Shukla, Jiang Yu, Scott L. Anderson, and Ralf I. Kaiser* ,Department of Chemistry, University of Hawaii at Manoa, Honolulu, Hawaii 96822, United States Department of Chemistry, University of Utah, Salt Lake City, Utah 84112, United States * S Supporting Information ABSTRACT: Understanding the reaction mechanisms of ionic liquids and their oxidizers is necessary to develop the next generation of hypergolic, ionic-liquid- based fuels. We studied reactions between a levitated droplet of 1-allyl-3- methylimidazolium dicyanamide ([AMIM][DCA]), with and without hydrogen- capped boron nanoparticles, and nitrogen dioxide (NO 2 ). The reactions were monitored with Fourier-transform infrared (FTIR) and Raman spectroscopy. The emergence of new structures in the FTIR and Raman spectra is consistent with the formation of functional groups including organic nitrites (RONO), nitroamines (R 1 R 2 NNO 2 ), and carbonitrates (R 1 R 2 C=NO 2 ). Possible reaction mechanisms based on these new functional groups are discussed. The reaction rates were deduced at various temperatures by heating the levitated droplets with a carbon dioxide laser. We thereby determined an overall activation energy of 38.5 ± 2.3 kJ mol 1 for the oxidation of [AMIM][DCA] for the rst time. 1. INTRODUCTION Ionic liquids (ILs) are dened as liquid salts consisting of ion pairs with melting points below 373 K (100 °C). 1 There has been an exponential growth in research on ionic liquids over the last couple of decades because their unique, tunable properties have interest in multiple disciplines including organic, material, and physical chemistry due to their potential roles in green chemistry, 2,3 catalysis, 2,4,5 fuels, 4,611 nuclear waste separation, 12,13 electrochemistry, 14,15 and pharmaceut- icals. 16,17 In particular, energetic ionic liquids (EILs) with hypergolic properties provide a potential alternative for hydrazine-based fuels (N 2 H 4 ) that can be safer, greener, and cheaper. 611 Some EILs have potential as an alternative fuel because these have been shown to be hypergolic with white fuming nitric acid (WFNA), red fuming nitric acid (RFNA), and hydrogen peroxide (H 2 O 2 ). 7,9,10 These EILs consist of ion pairs including nitrogen-containing cations such as imidazo- lium, triazolium, and tetrazolium along with anions dicyana- mide ([DCA] ), nitrocyanamide (NCNNO 2 ), azide (N 3 ), borohydride (BH 4 ), and cyanoborohydride (BH 3 CN ). The current standard for hypergolic fuels involves hydrazine along with its derivatives and the oxidizer dinitrogen tetroxide in equilibrium with nitrogen dioxide (N 2 O 4 2NO 2 ), which have critical disadvantages, however, such as a high vapor pressure, ammability, toxicity, and a relatively low energy density that can lead to high handling and storage costs. 18 Studies of the ignition of EILs have suggested that the anion plays a critical role in the ignition of EILs, whereas the cation aects the physical properties of the EIL such as the density and viscosity. 10,1922 The ignition of [DCA] -based EILs with WFNA has been well studied and a mechanism has been established. 2224 The initial step is proposed to be the protonation of the [DCA] at a nitrile nitrogen. This creates an electrophilic carbon where NO 3 can attack followed by a 1,3-shift of the NO 2 moiety to the terminal nitrogen and the formation of a C=O bond. A similar reaction sequence can occur at the second terminal nitrogen followed by a protonation of the central nitrogen to produce dinitrobiuret (DNB), which decomposes into the observed products carbon dioxide (CO 2 ), nitrous oxide (N 2 O), and isocyanic acid (HNCO). 2224 Although EILs show potential as hypergolic fuels, critical improvements are still necessary such as increasing the energy Received: August 29, 2018 Revised: October 1, 2018 Published: October 18, 2018 Scheme 1. Molecular Structures of [MAT][DCA] (1), [BMIM][DCA] (2), and [AMIM][DCA] (3) The numbering convention for the positions of the atoms in the imidazolium ring is shown. Article pubs.acs.org/JPCA Cite This: J. Phys. Chem. A 2019, 123, 400-416 © 2018 American Chemical Society 400 DOI: 10.1021/acs.jpca.8b08395 J. Phys. Chem. A 2019, 123, 400416 Downloaded via UNIV OF HAWAII on January 17, 2019 at 21:18:00 (UTC). See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

Transcript of Oxidation of a Levitated Droplet of 1-Allyl-3-methylimidazolium … · 2019. 1. 17. · ABSTRACT:...

  • Oxidation of a Levitated Droplet of 1‑Allyl-3-methylimidazoliumDicyanamide by Nitrogen DioxideMichael Lucas,† Stephen J. Brotton,† Shashi Kant Shukla,† Jiang Yu,‡ Scott L. Anderson,‡

    and Ralf I. Kaiser*,†

    †Department of Chemistry, University of Hawaii at Manoa, Honolulu, Hawaii 96822, United States‡Department of Chemistry, University of Utah, Salt Lake City, Utah 84112, United States

    *S Supporting Information

    ABSTRACT: Understanding the reaction mechanisms of ionic liquids and theiroxidizers is necessary to develop the next generation of hypergolic, ionic-liquid-based fuels. We studied reactions between a levitated droplet of 1-allyl-3-methylimidazolium dicyanamide ([AMIM][DCA]), with and without hydrogen-capped boron nanoparticles, and nitrogen dioxide (NO2). The reactions weremonitored with Fourier-transform infrared (FTIR) and Raman spectroscopy. Theemergence of new structures in the FTIR and Raman spectra is consistent with theformation of functional groups including organic nitrites (RONO), nitroamines(R1R2NNO2), and carbonitrates (R

    1R2C=NO2−). Possible reaction mechanisms

    based on these new functional groups are discussed. The reaction rates werededuced at various temperatures by heating the levitated droplets with a carbondioxide laser. We thereby determined an overall activation energy of 38.5 ± 2.3 kJmol−1 for the oxidation of [AMIM][DCA] for the first time.

    1. INTRODUCTIONIonic liquids (ILs) are defined as liquid salts consisting of ionpairs with melting points below 373 K (100 °C).1 There has

    been an exponential growth in research on ionic liquids overthe last couple of decades because their unique, tunableproperties have interest in multiple disciplines includingorganic, material, and physical chemistry due to their potentialroles in green chemistry,2,3 catalysis,2,4,5 fuels,4,6−11 nuclearwaste separation,12,13 electrochemistry,14,15 and pharmaceut-icals.16,17 In particular, energetic ionic liquids (EILs) withhypergolic properties provide a potential alternative forhydrazine-based fuels (N2H4) that can be safer, greener, andcheaper.6−11 Some EILs have potential as an alternative fuelbecause these have been shown to be hypergolic with whitefuming nitric acid (WFNA), red fuming nitric acid (RFNA),and hydrogen peroxide (H2O2).

    7,9,10 These EILs consist of ionpairs including nitrogen-containing cations such as imidazo-

    lium, triazolium, and tetrazolium along with anions dicyana-mide ([DCA]−), nitrocyanamide (NCNNO2), azide (N3

    −),borohydride (BH4

    −), and cyanoborohydride (BH3CN−). The

    current standard for hypergolic fuels involves hydrazine alongwith its derivatives and the oxidizer dinitrogen tetroxide inequilibrium with nitrogen dioxide (N2O4 ⇌ 2NO2), whichhave critical disadvantages, however, such as a high vaporpressure, flammability, toxicity, and a relatively low energydensity that can lead to high handling and storage costs.18

    Studies of the ignition of EILs have suggested that the anionplays a critical role in the ignition of EILs, whereas the cationaffects the physical properties of the EIL such as the densityand viscosity.10,19−22 The ignition of [DCA]−-based EILs withWFNA has been well studied and a mechanism has beenestablished.22−24 The initial step is proposed to be theprotonation of the [DCA]− at a nitrile nitrogen. This createsan electrophilic carbon where NO3

    − can attack followed by a1,3-shift of the NO2 moiety to the terminal nitrogen and theformation of a C=O bond. A similar reaction sequence canoccur at the second terminal nitrogen followed by aprotonation of the central nitrogen to produce dinitrobiuret(DNB), which decomposes into the observed products carbondioxide (CO2), nitrous oxide (N2O), and isocyanic acid(HNCO).22−24

    Although EILs show potential as hypergolic fuels, criticalimprovements are still necessary such as increasing the energy

    Received: August 29, 2018Revised: October 1, 2018Published: October 18, 2018

    Scheme 1. Molecular Structures of [MAT][DCA] (1),[BMIM][DCA] (2), and [AMIM][DCA] (3)

    The numbering convention for the positions of the atoms in theimidazolium ring is shown.

    Article

    pubs.acs.org/JPCACite This: J. Phys. Chem. A 2019, 123, 400−416

    © 2018 American Chemical Society 400 DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    Dow

    nloa

    ded

    via

    UN

    IV O

    F H

    AW

    AII

    on

    Janu

    ary

    17, 2

    019

    at 2

    1:18

    :00

    (UT

    C).

    Se

    e ht

    tps:

    //pub

    s.ac

    s.or

    g/sh

    arin

    ggui

    delin

    es f

    or o

    ptio

    ns o

    n ho

    w to

    legi

    timat

    ely

    shar

    e pu

    blis

    hed

    artic

    les.

    pubs.acs.org/JPCAhttp://pubs.acs.org/action/showCitFormats?doi=10.1021/acs.jpca.8b08395http://dx.doi.org/10.1021/acs.jpca.8b08395

  • density. One way to enhance the energy density of EILs is toinclude ions that contain multiple nitrogen atoms and hencenitrogen−nitrogen bonds.25,26 Another potential strategy is toadd energetic solids to the EIL. This pathway is alreadycommon in solid rocket propellants. One such additive isboron, which has a high gravimetric and volumetric energydensity of 51 kJ g−1 and 138 kJ cm−3 for combustion in oxygen,respectively, which is higher than traditional hydrocarbon fuels(35−40 kJ cm−3) and combustible metals such as aluminum(84 kJ cm−3) or magnesium (43 kJ cm−3).27 Unfortunately, thecombustion of boron has limitations that prevent full energyrelease. Ulas et al.28 suggested two reasons for this limitation:(1) ignition of boron is delayed due to a boron oxide (B2O3)layer that occurs from exposure to air, and (2) combustion ofboron in the presence of hydrogen-containing gases can lead tothe formation of gaseous metaboric acid (HOBO) instead ofboron oxide (B2O3), thereby creating an “energy trap” thatprevents the release of all the energy available from theoxidation of boron. The formation of HOBO and B2O3 arecompetitive at temperatures between their boiling points of1398 and 2158 K, respectively.28 The combustion of boron is aheterogeneous process limited by the diffusion to and from thesurface.29−33 One purposed solution to minimize the diffusionlimitation is to increase the surface area-to-volume ratio by

    using boron nanoparticles. However, the mass fraction of theoxide layer also increases with the larger surface area. Cappingthe boron nanoparticles with suitable ILs has been shown toprevent the formation of the oxide layer.31,34−36 The Andersongroup had success in creating air-stable, unoxidized particles bymilling solid boron in a hydrogen atmosphere creatinghydrogen-loaded boron nanoparticles that are then cappedwith an IL.36−38

    It is important to understand the underlying reactionmechanisms of EILs with oxidizers to improve the performanceof EIL-based fuels. Some of the potential EIL candidatesinvolve nitrogen-rich cations, such as imidazolium ortriazolium, and dicyanamide anions.24,39−41 We have pre-viously studied the oxidation of levitated 1-methyl-4-amino-1,2,4-triazolium dicyanamide ([MAT][DCA]; 1, Scheme 1)and 1-butyl-3-methylimidazolium dicyanamide ([BMIM]-[DCA]; 2, Scheme 1) droplets by nitrogen dioxide (NO2)with and without hydrogen-capped boron nanoparticles byFourier-transform infrared (FTIR), Raman, and ultraviolet−visible (UV−Vis) spectroscopy.42,43 Electronic structurecalculations for the reaction of [MAT][DCA] with nitrogendioxide were used to aid the investigation of one possible initialreaction intermediate, [O2N-NCNCN]

    −, where the nitrogenatom of the NO2 bonded to a terminal nitrogen on the

    Figure 1. Photographs of droplets of [AMIM][DCA] (a,c) and boron-doped [AMIM][DCA] (b,d) levitated in 100% Ar (a,b) and 0.4% NO2 and99.6% Ar (c,d).

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    401

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • [DCA]−.42 More recently, we studied the oxidation of[BMIM][DCA] droplets,43 analyzed in more depth theformation of the new peaks in the FTIR, Raman, and UV−Vis spectra, and produced a list of functional groups likely toproduce the new peaks, such as organic nitrites (RONO),nitroamines (R1R2NNO2), aromatic nitro-compounds(ArNO2), and carbonitrates (R

    1R2C=NO2−). In addition, the

    reaction rate constants, k, were determined for pure and boron-doped [BMIM][DCA] droplets as well as for a pure[MAT][DCA] droplet. The addition of the boron nano-particles caused the reaction rate of [BMIM][DCA] todecrease by a factor of ∼2, whereas the reaction rate of pure[MAT][DCA] was slower by a factor of ∼20 than for pure[BMIM][DCA].43

    In the current work, we extended our studies on theoxidation of EILs by probing the oxidation reaction of 1-allyl-3-methylimidazolium dicyanamide, [AMIM][DCA] (3;Scheme 1), with nitrogen dioxide. The main differencebetween the three ILs shown in Scheme 1 is that [AMIM]+

    contains an allyl group (−CH2−CH=CH2) and an imidazo-lium ring, whereas the [BMIM]+ contains a butyl group(−C4H9) and an imidazolium ring, and the [MAT]+ carries anamino group (−NH2) and has a triazolium ring (Scheme 1). Inthis study, we first aim to identify the key reactionintermediates and functional groups formed during theoxidation of [AMIM][DCA] with nitrogen dioxide. Second,we determine the temperature dependence of the rate constantfor the oxidation processes by heating a droplet of the ILlevitated in NO2. Finally, these data are exploited to extract theactivation energies, Ea, for the oxidation of a pure [AMIM]-[DCA] droplet by NO2.

    2. EXPERIMENTAL METHODS

    A single droplet of the ionic liquid was levitated in anultrasonic levitator, which is housed within a pressure-compatible process chamber. This levitator apparatus hasbeen discussed in detail elsewhere.43−46 Briefly, ultrasonicsound waves produced by a piezoelectric transducer oscillatingat 58 kHz reflect between the transducer and a concave-shapedreflector.44 The droplet can be levitated in the resultingstanding wave as the sound waves exert acoustic radiationpressure on the droplet.47 The surrounding process chamberallows for levitation in inert gases as well as in highly reactiveand toxic gases. In the present experiments, the chamber wasfilled to a total pressure of 880 Torr with either argon (Ar;Airgas, 99.9999%) or a gas mixture of 1.3% nitrogen dioxide(NO2; Aldrich, ≥99.5%) and 98.7% Ar or 0.4% NO2 and99.6% Ar.[AMIM][DCA] was purchased from IoLiTec with a purity

    of >98%, but the IL had a yellowish color that requiredpurification to be removed. For [AMIM][DCA] to be purified,10 mL of the IL was mixed with ∼60 g of activated charcoaland 100 mL of dichloromethane (DCM) for 24−48 h. Aftermixing, the charcoal solution was filtered with a Buchnerfunnel and washed with 100 mL of DCM. This solution wasfurther filtered with a 0.20 μm syringe filter to ensure theremoval of any particulates. The DCM solvent was removedusing a rotary evaporator. Lastly, to ensure all the solvent wasremoved, the IL was stirred under a vacuum until theformation of bubbles stopped followed by an additional 30min of stirring under vacuum. The purified [AMIM][DCA]was stored in a dry cabinet. The [AMIM][DCA]-capped,hydrogenated boron nanoparticles were produced by high-energy balling-milling boron powder in a hydrogen atmosphere

    Figure 2. Selected thermal images of three different heating experiments used to determine the temperature of the droplets. The scale on the rightis in °C.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    402

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • Figure 3. FTIR spectra of [AMIM][DCA] (left column) and [AMIM][DCA] with hydrogen-capped boron nanoparticles (right column) alongwith the individual peak fits. The [AMIM]+ and [DCA]− peaks are labeled by 2−50 and 2*−9*, respectively. The peak assignments in Table 1 arebased on refs 48−51.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    403

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • followed by capping the hydrogen-terminated nanoparticleswith [AMIM][DCA]. This results in unoxidized nanoparticlesthat are air-stable.38

    A microliter droplet deposition system was used tointroduce the [AMIM][DCA] droplet into the levitator.43

    Briefly, a syringe is interfaced to the chamber and connected toa microneedle that is attached to a wobble stick, which allowsfor full translational motion of the needle and hence permits asingle droplet to be deposited into a pressure minimum(Figure 1). The levitated droplets were oblate spheroids with amean horizontal diameter of 3.2 ± 0.4 mm and a mean verticaldiameter of 0.9 ± 0.2 mm. The levitated droplets were heatedwith a carbon dioxide laser emitting at 10.6 μm (SynradFirestar v40). The output power was increased by varying theduty cycle of the discharge for each of the heating experimentsup to a maximum of 4.4 W. The temperature of the dropletwas measured using an FLIR A6703sc thermal imaging camera(Figure 2). The camera uses a cooled detector composed of640 × 512 pixels, covers the spectral range of 3−5 μm, and canmeasure temperatures between 253 and 1773 K with anaccuracy of ±2% of the reading.The chemical modifications of the levitated droplet were

    monitored with Raman and Fourier-transform infrared (FTIR)spectrometers. The spectrometers have been described indetail previously.44 Briefly, a Q-switched Nd:YAG laseroperating at 532 nm was used to excite the Raman transitions.

    The Raman-shifted photons, backscattered from the droplet,were focused by a lens into a HoloSpec f/1.8 holographicimaging spectrograph equipped with a PI-Max 2 ICCD camera(Princeton Instruments). The spectra were collected simulta-neously over two wavenumber ranges, 170−2450 and 2400−4400 cm−1, with a resolution of 9 cm−1. The FTIR spectrawere collected in the range of 400−4000 cm−1 before and afterthe reaction using an attenuated total reflection (ATR)accessory located within a Nicolet 6700 FTIR spectrometer(Thermo Scientific). A high-speed camera (Phantom Miro3a10) was used to photograph visual changes occurring duringthe reaction. The camera can record up to 1850 frames persecond with a maximum pixel resolution of 1280 × 1280 and iscapable of exposure times as short as 1 μs. A Navitar Zoom6000 modular lens system can be used to zoom in on thedroplet. The field of view (FOV) can be adjusted from 28 mm× 28 mm to 4.4 mm × 4.4 mm at a working distance of 220mm corresponding to a magnification range of 0.46−2.93. Thelowest FOV is equivalent to a full angle of 1.1° and provideshigh-resolution images of the droplet.43

    3. RESULTS3.1. [AMIM][DCA]. 3.1.1. Infrared Spectroscopy. The first

    step to identify any changes in the spectra after the reactionwith nitrogen dioxide is to record spectra of the unreacted[AMIM][DCA] and boron-doped [AMIM][DCA] for refer-

    Table 1. Vibrational Mode Assignments for the Observed Peaks in the FTIR Spectra of [AMIM][DCA] and Boron-Doped[AMIM][DCA] Based on Refs 48−51 and 53

    experimental wavenumber (cm−1)a

    mode [AMIM][DCA] boron-doped [AMIM][DCA] expected wavenumber (cm−1) assignmentb

    2 3153 3152 3146 v(C4−H, C5−H) out-of-phase3 3097 ± 3 3100 ± 4 3090 v(C2H)4 3014 3013 3022 vas(CH2(C = C))9 2956 ± 3 2970 ± 10 2958 vas(CH2)11 2856 ± 2 2858 ± 2 2888 vs(CH3)B−H 2400−2600 2480−2565 v(BH)6* + 7* 2227 2227 2227 DCA− combination9* 2191 2191 2192 vs(C≡N)8* 2122 2123 2133 vas(C≡N)14 1647 1647 1646 v(CC) + β(CH)15 1569 1568 1575 v(CC)19 1471 ± 3 1468 ± 5 1470 βas(CH2)20 1448 1447 1448 βs(CH3)21 1423 1423 1423 β(CH2)23 1385 1387 ± 2 1385 Ras + β(CH)7* 1303 1303 1309 vas(C−N)31 1164 1164 1166 v(CC) + β(CH)32 1110 ± 20 1120 ± 20 1108 β(CH) + Ras33 1090 ± 20 1090 ± 10 1088 β(CH) + Ras36 1022 ± 2 1022 ± 3 1020 Ras+v(CN)38 995 995 997 ω(CH)40 945 946 947 v(CC)6* 902 902 904 vs(C−N)43 847 847 846 ω(CH)44 754 754 758 v(CN) + β(CCN) + τ(CH)5* 669 667 666 βs(CNC)49 621 622 622 τ(Ring)50 565 564 568 τ(CH) + τ(Ring) + v(CN)2* 520 519 517 βas(NCN)

    aThe uncertainties are equal to, or less than, 1 cm−1 unless otherwise stated. bKey: v, stretching; β, bending; R, ring; ω, wagging; τ, torsion; s,symmetric; as, antisymmetric

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    404

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • ence (Figure 3). The spectra were decomposed into theircomponents peaks by a fitting procedure, and the resultingpeak wavenumbers are compiled in Table 1. It should be notedthat there are regions in the spectra that contain manyoverlapping peaks leading to degeneracy in the fittingprocedure, and therefore the fitted peak wavenumbers areonly compiled for clearly defined peaks and prominentshoulders. The fitted peaks were also observed to besuperimposed on broad backgrounds produced by thespectrometer, and these backgrounds were modeled byoptimizing a polynomial of fifth to eighth order simultaneouslywith the peak fits. The vibrational mode assignments weremade by comparing the fitted peak wavenumbers with themeasured and theoretical values of published [AMIM][DCA]spectra.48−51 Xuan et al. studied the cation−anion interactionsof [AMIM]+-based ionic liquids using density functionaltheory (DFT) with a B3LYP functional.48 The calculations

    found three stable conformers for the isolated [AMIM]+ cationand eight stable conformers for the [AMIM][DCA] ion pairs.In addition, Xuan et al. measured the FTIR and Raman spectraof [AMIM][DCA] and [AMIM]Cl and compared the resultswith their calculated vibrational frequencies. The assignmentswere based on the potential energy distributions of the normalmodes calculated. Unfortunately, the theoretical vibrationalfrequencies were insufficiently accurate to distinguish betweenthe different conformers. The vibrational modes presented forthe FTIR (Figure 3, Table 1) and Raman (Figure 4, Table 2)spectra can be separated into two parts: the vibrational modesassociated with the [AMIM]+ cation, which are labeled 1-55,based on the observed IR and Raman spectra in refs 48 and 49,and the vibrational modes associated with the [DCA]− anion,which are labeled 2*−9*, based on the observed IR andRaman spectra in refs 50 and 51.

    Figure 4. Raman spectra of [AMIM][DCA] (top and center row) and [AMIM][DCA] with hydrogen-capped boron nanoparticles (bottom row)in argon. The [AMIM]+ and [DCA]− peaks are labeled 1−55 and 2*−9*, respectively, and the assignments are compiled in Table 2.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    405

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • The FTIR spectra of pure and boron-doped [AMIM][DCA]are shown in Figure 3. The assignments in Table 1 are nearlyequivalent with or without boron nanoparticles. The vibra-tional modes can be divided into two parts: [AMIM]+ and[DCA]− modes. We first discuss the modes of the [AMIM]+

    cation. The peaks in the high-wavenumber region from 2800 to3200 cm−1 are assigned to the C−H stretching modes of the[AMIM]+ cation.48,49 More specifically, the C−H stretchingmodes of the imidazolium ring occur in the range of 3050−3200 cm−1 with the peaks at 3097 and 3153 cm−1 produced bythe C(2)-H and C(4,5)-H stretching modes, respectively (3,Scheme 1). In contrast, the modes in the range of 2800−3050cm−1 are assigned to C−H stretching modes of the allyl andmethyl side chains; that is, the peaks at 2856 cm−1, 2956 cm−1,and 3014 cm−1 are produced by the stretches of the CH3, CH2,and the terminal allylic CH2, respectively. The vibrationalmodes below 1700 cm−1 are associated with various bending,ring, wagging, torsion, and combination modes of the[AMIM]+ cation.48,49 The strongest of these peaks occur at1164, 754, and 621 cm−1 and are associated with stretching,combination, and ring torsion modes of the imidazolium ring,respectively. The vibrational modes associated with allyl group

    are observed at 1647, 1569, and 945 cm−1. The largestvibrational modes for the [DCA]− anion are observed at 2122,2191, and 2227 cm−1, which are assigned to the asymmetricC≡N stretching mode, symmetric C≡N stretching mode, anda vs(C−N) and vas(C−N) combination band, respectively.50,51This combination band is in a Fermi resonance with thevas(C−N) fundamental mode, which causes a shift in thefrequency and an increase in the intensity of the combinationband.50,52 The vs(C−N) and vas(C−N) stretching modes areobserved separately at 1303 and 902 cm−1. In addition to thestretching modes, the asymmetric NCN bending mode,βas(NCN), and symmetric bending mode, βs(CNC), wereobserved at 520 and 669 cm−1, respectively. The onlydiscernable difference between the pure and boron-doped[AMIM][DCA] spectra is the small but broad absorptionfeature in the 2400−2600 cm−1 region of the boron-dopedspectrum, which is assigned to a B−H stretching mode.53

    3.1.2. Raman Spectroscopy. Figure 4 presents the Ramanspectra of an [AMIM][DCA] droplet and a boron-doped[AMIM][DCA] droplet. The optimized wavenumbers of thefitted peaks are compiled and assigned in Table 2. Thevibrational assignments are based on refs 48−51 and 53. Theaddition of the boron nanoparticles caused the sharp peaks ofthe Raman spectrum to vanish, and so this section will focus onthe pure [AMIM][DCA]. In the fitting procedure, the broadbackgrounds in the Raman spectra were treated in the samemanner as mentioned in section 3.1.1. The selection rulesdiffer for IR and Raman spectroscopy,54 and therefore themodes excited and their relative intensities usually differbetween the two methods. The differences are apparent in themodes of the [DCA]− anion. In the Raman spectrum, mode 9*(vs(C≡N)) is much larger than mode 8* (νas(C≡N)) andmode 5* (βs(CNC)) is significantly more prominent thanmode 2* (βas(CNC)), whereas in the FTIR spectra the relativeamplitudes of these peaks is reversed with 9* < 8* and 5* <2*. Furthermore, the 6* + 7* combination band is observedonly in the FTIR spectrum. For the Raman spectrum, the alkylgroup C−H stretching region between 2800 and 3050 cm−1 islarger than the ring C−H stretching region between 3050 cm−1and 3200 cm−1, whereas in the FTIR spectrum the relativeprominence of the ring and alkyl groups is reversed. In thelower wavenumber region, the peaks at 1569, 1164, 754, and621 cm−1 produced by modes 15 (ν(CC)), 31 (ν(CC) +β(CH)), 44 (ν(CN) + β(CCN) + τ(CH)), and 49 (τ(Ring)),respectively, have higher intensities in the FTIR spectrum. Incontrast, the peaks at 1641, 1410, 1015, and 406 cm−1

    produced by modes 14 (ν(CC) + β(CH)), 22 (β(CH) +ν(CN)), 35 (Rs + β(CH)), and 55 (β(CNC,CCC)),respectively, are significantly larger in the Raman spectrum.

    3.2. [AMIM][DCA] Oxidation. 3.2.1. Infrared Spectros-copy. A droplet of [AMIM][DCA] or boron-doped [AMIM]-[DCA] was levitated in a 1.3% nitrogen dioxide and 98.7%argon gas mixture. Figures 5 and 6 show the ATR-FTIRspectra and optimized peak fits of the pure and boron-dopeddroplets after the reaction, respectively. The unreacted spectraare also shown in Figures 5 and 6 and allow the newly formedpeaks to be identified. The wavenumbers of the new peaks arecompiled in Table 3. After oxidation, both the pure and boron-doped spectra exhibit nearly the same new structures, whichindicates that the nitrogen dioxide reacts with the IL to formorganic nitro-compounds but does not react significantly withthe boron nanoparticles. The analysis of the [AMIM][DCA]spectrum to be discussed below therefore also applies to the

    Table 2. Vibrational Mode Assignments for the ObservedPeaks in the Raman Spectrum of an [AMIM][DCA] DropletBased on Refs 48−51 and 53

    modeexperimental

    wavenumber (cm−1)a

    expectedwavenumber

    (cm−1) assignmentb

    1 3148 3160 v(C4−H, C5−H)in-phase

    3 3085 3090 v(C2H)4 3016 3024 vas(CH2(C = C))6 2987 2987 vs(CH2(C = C))10 2957 2957 vs(CH2)11 2841 ± 3 2888 vs(CH3)9* 2182 2192 vs(C≡N)8* 2123 2133 vas(C≡N)14 1641 1646 v(CC) + β(CH)16 1560 1565 v(CC)20 1439 ± 3 1448 βs(CH3)22 1410 1412 β(CH) + v(CN)23 1378 1385 Ras + β(CH)26 1327 1329 Rs7* 1286 1309 vas(C−N)29 1210 1229 v(CN) + β(CH)31 1162 1166 v(CC) + β(CH)32 1098 ± 4 1108 β(CH) + Ras33 1081 1088 β(CH) + Ras35 1015 1023 Rs + β(CH)40 947 ± 2 947 v(CC)6* 904 ± 2 904 vs(C−N)43 870 ± 10 846 ω(CH)44 759 758 v(CN) + β(CCN) +

    τ(CH)5* 664 666 βs(CNC)49 622 622 τ(Ring)50 571 ± 6 568 τ(CH) + τ(Ring) +

    v(CN)2* 504 ± 6 517 βas(NCN)55 406 ± 2 393 β(CNC,CCC)

    aThe uncertainties are equal to, or less than, 1 cm−1 unless otherwisestated. bKey: v, stretching; β, bending; R, ring; ω, wagging; τ, torsion;s, symmetric; as, antisymmetric

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    406

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • boron-doped spectrum. Because of the complexity of theoxidized [AMIM][DCA] spectrum, it is unlikely that specificmolecules can be identified from the FTIR spectra, but thespectra can help determine which functional groups formed inthe oxidation products. The possible functional groups ofcompounds that could form from the oxidation of [AMIM]-[DCA] by nitrogen dioxide with absorption ranges includingthe newly formed peaks are compiled in Table 3. The list ofpossible functional groups includes organic nitrites (RONO),nitroamines (R1R2NNO2), aromatic nitro-compounds(ArNO2), aliphatic nitro-compounds (RNO2), and carboni-trates (R1R2C=NO2

    −). One of the most noticeable regions ofnew peaks occurs in the 1500−1800 cm−1 range (Figures 5 and6). The peaks at 1789, 1749, and 1716 cm−1 are in the rangedominated by the stretching modes of the carbonyl functionalgroups, C=O, which may have been produced by higher-orderreactions.53 Alternatively, these new peaks might be accountedfor by the vibrational modes of NO2 or N2O4 dissolved in theIL as suggested by UV−Vis spectroscopy.55 The N=O andNO2 stretching modes of organic nitrites, nitroamines, andaromatic nitro-compounds could explain the new peaksobserved at 1678, 1613, and 1524 cm−1, respectively.53 Themost intense new peaks occur at 1356 and 1269 cm−1 andcould be due to the NO2 stretching modes of aliphatic andaromatic nitro-compounds, carbonitrates, and nitroamines,whereas NO2 and N−O stretches of the carbonitrates andorganic nitrites can account for the peaks at 1040 and 827cm−1, respectively. The peak at 1040 cm−1 could also be due to

    CN stretching of an aromatic nitro-compound. Lastly, the newpeak at 711 cm−1 might be caused by NO2 wagging of thenitroamines or the NO2 deformation mode of the carboni-trates, and the new peak at 528 cm−1 can be produced by NO2rocking of an aliphatic nitro-compound.53

    In addition to the appearance of new peaks in Figures 5 and6, there is a significant reduction of the vas(C≡N), vs(C≡N),and vs(C−N) + vas(C−N) modes of [DCA]− at 2122, 2191,and 2227 cm−1, respectively. This suggests that the [DCA]−

    anion has been degraded from oxidation by nitrogen dioxide.In contrast, the structures assigned to the [AMIM]+ cationshow little change after the reaction, as evident in the C−Hstretching region from 2800 to 3200 cm−1. This suggests thatthe [DCA]− anion reacts to a greater extent than the [AMIM]+

    cation.3.2.2. Raman Spectroscopy. The Raman spectra of an

    [AMIM][DCA] droplet or a boron-doped [AMIM][DCA]droplet levitated in 1.3% nitrogen dioxide and 98.7% argon areshown in Figure 7. The unreacted spectra (blue lines) are alsoshown in Figure 7. As for the FTIR spectra in Figures 5 and 6,the pure and boron-doped Raman spectra after oxidationexhibit similar new structures at comparable wavenumbers(Table 4), indicating that the nitrogen dioxide reacted with theIL and not the boron nanoparticles. Because both spectra mustbe produced by the same reaction products, the Raman activityof the ionic liquid and reaction products on the surface of theboron-doped droplets must be higher than any contributionfrom the boron nanoparticles. The reaction of nitrogen dioxide

    Figure 5. FTIR spectra of [AMIM][DCA] after reaction with nitrogen dioxide. The spectra of unreacted [AMIM][DCA] (blue lines) are shownfor comparison. The wavenumbers of the newly formed peaks are labeled, and their vibrational assignments are shown in Table 3. The unreactedspectrum in the 2600−1900 cm−1 region is scaled by a factor of 0.1.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    407

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • with the IL in the high Raman-shift region (3300−2400 cm−1)caused the unreacted peaks to reduce in amplitude and bereplaced by a structureless continuum. Similarly, the NO2caused a reduction of the peak amplitudes in the lower Raman-shift region (2400−400 cm−1), but new features also formed(Figure 7). The resulting new features are labeled a−s inFigure 7, and the wavenumbers of the individual peak fits arecompiled in Table 4. Similarly to the FTIR spectra, thepossible functional groups that could explain the new peakswere determined by comparing the known spectral regions ofthe organic nitro-compounds with our experimentallydetermined wavenumbers of the new peaks. The resultingpossible functional groups include organic nitrites (RONO),nitroamines (R1R2NNO2), carbonitrates (R

    1R2C=NO2−),

    nitroalkanes (R1R2CHNO2), dinitroalkanes (R1R2CH(NO2)2),

    organic nitroso compounds (RNO), aliphatic (RNO2) andaromatic (ArNO2) nitro compounds, and o-aminonitro-aromatic compounds (NH2ArNO2). The measured wave-numbers and assignments for Figure 7 are compiled in Table 4.These functional groups of the organic nitro compounds alonecannot explain all the new peaks in the Raman spectra and,more specifically, do not account for the peaks at wavenumbersabove 1700 cm−1 or below 450 cm−1. Many oxygen- andnitrogen-containing molecules are Raman active outside the1700 to 450 cm−1 region. To explain the new peaks in theseregions, moieties were therefore chosen that formed part of theorganic nitro compounds produced by the reaction of the ILand NO2. Lastly, unreacted [AMIM][DCA] may also

    contribute to the spectra after oxidation in regions where thepure [AMIM][DCA] produces larger peaks.

    3.3. Reaction Rates. Our experimental setup allows us toheat and monitor the oxidation of [AMIM][DCA] or boron-doped [AMIM][DCA] droplets by Raman spectroscopy insitu. We can therefore extract the rates for the reactionbetween the IL and nitrogen dioxide at various temperatures. Asmall time delay is caused by the setup of the levitationapparatus for the following reasons. For a droplet to belevitated, the process chamber must already be filled with a gas,such as the nitrogen dioxide and argon gas mixture used in thepresent experiments. Consequently, the reaction begins assoon as the IL is exposed on the needle tip. The minimumtime delay is ∼15−20 s from the first appearance of the IL onthe needle tip to completing the collection of the spectrum anddepends on the number of accumulations in a spectral scan. Tominimize this time delay, the tip of the needle was placed nearthe pressure minimum where the droplet will be levitated, anda drop or two of the liquid was expended to avoid exposing thedroplet before moving the needle into position. The new peaksshown in Figure 7 fully formed within this time delay, makingit infeasible to determine their formation rate with the presentexperimental setup. Because we cannot prevent the initial timedelay with the current experimental apparatus, we decreasedthe nitrogen dioxide concentration to slow the reaction ratesufficiently to allow monitoring. The concentration of nitrogendioxide was therefore reduced to 0.4%, and the resultingRaman spectra are presented in Figure 8. The four peaksstudied are labeled (a−d). Peaks (a), (c), and (d) are three

    Figure 6. FTIR spectra of boron-doped [AMIM][DCA] after reaction with nitrogen dioxide. The spectra of unreacted boron-doped[AMIM][DCA] (blue lines) are shown for comparison. The wavenumbers of the newly formed peaks are labeled, and their vibrational assignmentsare shown in Table 3. The unreacted spectrum in the 2600−1900 cm−1 region was scaled by a factor of 1/15.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    408

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • major peaks in the unreacted [AMIM][DCA] spectrumobserved at 2128, 1015, and 664 cm−1 (Table 2), respectively,whereas peak (b) at 1115 cm−1 is newly formed and appears tothe left of peak 31. The possible assignment of peak (b)includes the CN stretching mode of an aromatic nitrocompound or the symmetric NO2 stretching mode ofcarbonitrates (Table 4). Multiple Raman spectra wererecorded at different time intervals from when the dropletwas first levitated to the completion of the reaction. Todetermine the reaction rate of the oxidation, the heights ofpeaks (a−d) were first measured as a function of time as shownin Figures S1−S4. The resulting decay or growth curves foreach peak were studied at temperatures between 306 and 491K.To determine the rate constant, the decay profiles of peaks

    (a), (c), and (d) were fit with a first-order rate equation, y =Ae−kt (Figures S1, S3, and S4). The growth curve in Figure S2

    exhibits a sharp increase followed by a gradual decline.Therefore, the relevant formula56

    =−

    −− −yk b

    k ke e( )k t k t1

    2 1

    1 2

    (1)

    was fit to data for peak (b) to determine the rate constants k1and k2 at each temperature, where b is an adjustable scalefactor. Although eq 1 gives a rate constant, k2, for the decayportion of the curve, we are only concerned with the growthrate, so the decay rate constant is not discussed further. Theresulting decay, k, and growth rate, k1, constants are compiledin Table 5. The rate constants for the decay of peaks (a), (c),and (d) range from (0.92−1.1) × 10−3 s−1 at 306 K to 0.18−0.22 s−1 at 491 K. Peaks (a) and (d) are associated with the[DCA]− anion, whereas peak (c) is associated with the[AMIM]+ cation. The comparable rate constants for thesethree peaks suggests that the anion and cation are similarlydegraded by the nitrogen dioxide. The rate constants for the

    Table 3. Assignments of the New Vibrational Modes in the FTIR Spectra of Pure or Boron-Doped [AMIM][DCA] Producedby Reaction with 1.3% Nitrogen Dioxide and 98.7% Argon with the Functional Groups, Regions of Maximum Absorption, andVibrational Modes from the Cited References

    aThe uncertainties are equal to, or less than, 1 cm−1 unless otherwise stated.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    409

    http://pubs.acs.org/doi/suppl/10.1021/acs.jpca.8b08395/suppl_file/jp8b08395_si_001.pdfhttp://pubs.acs.org/doi/suppl/10.1021/acs.jpca.8b08395/suppl_file/jp8b08395_si_001.pdfhttp://pubs.acs.org/doi/suppl/10.1021/acs.jpca.8b08395/suppl_file/jp8b08395_si_001.pdfhttp://dx.doi.org/10.1021/acs.jpca.8b08395

  • growth of peak (b) are nearly the same at all temperatures andlie in the range of 0.27−0.31 s−1. This suggests that the growthis occurring at a rate faster than our experimental setup canmeasure and, hence, we can only report the lower limits of therate constant, k1.The temperature dependence of the rate constant is given by

    the Arrhenius equation, k = Ae−Ea/RT, where Ea is the activationenergy, A is the pre-exponential factor, and R is the ideal gasconstant. As shown in Figure 9, an Arrhenius plot of the rateconstants as a function of temperature for peaks (a), (c), and(d) was used to determine three values of the activationenergy, Ea, for the oxidation of [AMIM][DCA] by nitrogendioxide. Table 6 presents the values of Ea thereby determined.Comparing the Ea in Table 6 with the known activation

    energies for reactions involving nitrogen dioxide may providesome details into the reaction mechanisms for the oxidation of[AMIM][DCA]. Table 7 presents values of Ea for selectedreactions involving NO2 obtained from the NIST ChemicalKinetics Database.57 These include NO2 reacting with doublebonds in the molecules isobutene (i-C4H8), 1-butene (1-C4H8), 2-butene (2-C4H8), 1-pentene (l-C5H10), propene(CH3CHCH2), isoprene (CH2C(CH3)CHCH2), ethylene(C2H4), and 1,3-cyclohexadiene (c-1,3-C6H8); triple bonds indimethylacetylene (CH3CCCH3), methylacetylene(CH3CCH), and acetylene (C2H2); and the N-containingmolecules hydrogen cyanide (HCN) and cyanonitrene(NCN). The activation energy for the oxidation of [AMIM]-[DCA] when averaged over the three peaks studied at 38.5 ±2.3 kJ mol−1 lies between the values for propene (33.0 kJ

    mol−1) and NCN (45.4 kJ mol−1), suggesting that NO2 reactswith the [DCA]− anion or the allyl side chain of the [AMIM]+

    cation. Chambreau et al. proposed that the initial reaction of 1-propagyl-3-methylimidazolium dicyanamide with nitric acid(HNO3) involves the nitric acid reacting with the anion in thepreignition phase.22 The initial step was the protonation of the[DCA]− followed by the NO3

    − attacking one of the C atoms ofthe anion. The NO2 oxidizer in our experiments clearly cannotact as the proton source. For protonation to occur in ourreaction, proton transfer must therefore occur from the cationto the anion. It has been reported previously that [DCA]− canreceive a proton from imidazolium-based cations.58 After theproton transfer, the NO2 may react with the protonated[DCA]− similarly to the known reaction of HNO3 with[DCA]−, leading to the decay of the [DCA]− vibrational peaksbecause of the formation of possible new functional groupssuch as organic nitrites, nitroamines, and carbonites. Addi-tionally, the loss of a proton from the cation may allow areaction of the NO2 with the cation, which may lead to theformation of functional groups such as organic nitrites,aromatic and aliphatic nitro compounds, carbonitrates, andnitroamines. The addition of the nitrogen atom of the nitrogendioxide oxidizer to a nitrogen atom of [DCA]− may be thesimplest explanation for the formation of nitroamines.Furthermore, the addition of the nitrogen atom of the NO2to a carbon atom of the anion could also account for theformation of carbonitrates.The rate constants for the oxidation of [AMIM][DCA] are

    compared below to the corresponding rate constants

    Figure 7. Raman spectra of [AMIM][DCA] (top row) and boron-doped [AMIM][DCA] (bottom row) after reaction with nitrogen dioxide. Theunreacted spectra (blue lines) are shown for comparison. The wavenumbers of the newly formed peaks are labeled (a−s) and their vibrationalassignments are compiled in Table 4. The unreacted spectra for the [AMIM][DCA] sample were scaled by a factor of 2/7 in the 3300−2400 cm−1region and 1/2 in the 2400−400 cm−1 region, respectively.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    410

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • Table 4. Possible Assignments of New Peaks in the Raman Spectra of an [AMIM][DCA] or a Boron-Doped [AMIM][DCA]Droplet Levitated in 1.3% Nitrogen Dioxide and 98.7% Argon with the Functional Groups, Regions of Maximum Scattering,and Vibrational Modes Obtained from the References

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    411

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • previously reported for [BMIM][DCA] and [MAT][DCA].43

    The experiments discussed in ref 43 for [BMIM][DCA] and[MAT][DCA] were performed at room temperature and at aconcentration of 0.2% NO2, so the rate constants can only becompared with the present values at 306 K. The fits to thedecay profiles in section 3.3 suggest first-order reactionprocesses, and therefore the reaction rates will scale inproportion to the concentration of the nitrogen dioxide.When comparing the reaction rates for the different ionicliquids below, we will consequently first scale the reaction ratespresented in ref 43 for [BMIM][DCA] and [MAT][DCA] bya factor of two. The rate constants of [BMIM][DCA] rangingfrom (3.0−5.0) × 10−3 s−1 are higher than the rate constantsfor [AMIM][DCA] from (0.98−1.1) × 10−3 s−1. In contrast,the rate constants for [MAT][DCA] from (1.5−2.1) × 10−4s−1 are one order of magnitude lower than for [AMIM][DCA].Thus, the oxidation rate by nitrogen dioxide increases in orderof [MAT][DCA], [AMIM][DCA], and [BMIM][DCA]. If thereaction of nitrogen dioxide with the [DCA]−-based ILsrequires the initial protonation of [DCA]−, then the differencesin the rate constant may be because [BMIM]+ has the most Hatoms available for proton transfer followed by [AMIM]+ andfinally [MAT]+. Another possible reason for the differentreaction rates may be the result of differences in the interactionstrength between the anion and cation, because a stronger

    cation−anion interaction will make it harder for the nitrogendioxide to interact with the ions.

    4. CONCLUSIONWe have studied the oxidation by nitrogen dioxide of levitateddroplets of [AMIM][DCA] and [AMIM][DCA] doped withhydrogen-capped boron nanoparticles and thereby extendedour previous work on the oxidation of the [DCA]−-containingionic liquids [MAT][DCA] and [BMIM][DCA].42,43 Theoxidation reaction was studied using FTIR and Ramanspectroscopy. The first step was to collect the FTIR andRaman spectra of a pure or boron-doped [AMIM][DCA]droplet before the reaction. The vibrational modes were

    Table 4. continued

    aThe uncertainties are equal to, or less than, 1 cm−1 unless otherwise stated.

    Figure 8. Raman spectrum of an [AMIM][DCA] droplet levitated in0.4% nitrogen dioxide and 99.6% argon (red line). The Ramanspectrum of an [AMIM][DCA] droplet before the reaction with NO2is shown (black line). The peaks labeled (a), (c), and (d) are presentin the unreacted spectrum, whereas peak (b) is a newly formed peakproduced by the reaction with nitrogen dioxide.

    Table 5. Rate Constants, k, for Peaks (a−d) in the RamanSpectra of an [AMIM][DCA] Droplet Levitated in 0.4%Nitrogen Dioxide and 99.6% Argon at DifferentTemperatures

    peakwavenumber

    (cm−1) assignmenta temp (K) rate constant (s−1)

    a 2182 vs(C≡N) 306 ± 2 (9.8 ± 0.8) × 10−4

    308 ± 1 (9 ± 1) × 10−4

    338 ± 1 (1.6 ± 0.1) × 10−3

    368 ± 5 (1.9 ± 0.4) × 10−3

    470 ± 10 0.12 ± 0.01491 ± 4 0.18 ± 0.02

    b 1115 see text 306 ± 2 0.27 ± 0.06308 ± 1 0.3 ± 0.2338 ± 1 0.3 ± 0.2368 ± 5 0.3 ± 0.1470 ± 10 0.31 ± 0.01491 ± 4 0.28 ± 0.09

    c 1015 Rs + β(CH) 306 ± 2 (9 ± 2) × 10−4

    308 ± 1 (9 ± 1) × 10−4

    338 ± 1 (1.8 ± 0.1) × 10−3

    368 ± 5 (1.9 ± 0.1) × 10−3

    470 ± 10 0.18 ± 0.01491 ± 4 0.18 ± 0.04

    d 664 βs(CNC) 306 ± 2 (1.1 ± 0.1) × 10−3

    308 ± 1 (1.1 ± 0.1) × 10−3

    338 ± 1 (1.7 ± 0.1) × 10−3

    368 ± 5 (1.7 ± 0.1) × 10−3

    470 ± 10 0.17 ± 0.01491 ± 4 0.2 ± 0.1

    aKey: v, stretching; β, bending; R, ring; ω, wagging; τ, torsion; s,symmetric; as, antisymmetric.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    412

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • assigned based on published spectra of [AMIM][DCA].48−51

    The FTIR spectra of [AMIM][DCA] with and without boronnanoparticles were found to be very similar except for a smallabsorption feature in the region of 2400−2600 cm−1, whichwas assigned to a B−H stretching mode. Unlike the FTIRspectra, the addition of the boron nanoparticles caused thesharp peaks in the Raman spectrum of the pure IL to bereplaced by a broad, featureless continuum.

    For the reaction with nitrogen dioxide, a pure or boron-doped [AMIM][DCA] droplet was levitated in 1.3% nitrogendioxide and 98.7% argon. By comparing the FTIR spectrabefore and after the reaction, new peaks could clearly beidentified. Because of the complexity of the oxidation reactionand the numerous vibrational modes of the product molecule,FTIR and Raman spectroscopy alone cannot identify themolecules formed but can be used to identify possibleemerging functional groups. To explain the new structures,we suggested that the nitrogen or oxygen atom of the NO2bonds to a carbon or nitrogen atom of the IL. The resultingpossible functional groups formed with absorption rangesincluding the new peaks were compiled, namely, organicnitrites, nitroamines, aromatic and aliphatic nitro compounds,and carbonitrates. These functional groups can account formost of the new features, although not for the new peaks in the1700−1800 cm−1 region. These peaks could be produced byN2O4 dissolved in the IL droplet or the carbonyl functionalgroup. Boron-doped [AMIM][DCA] after the reactionproduces a spectrum with nearly the same new structures asemerged in the oxidized [AMIM][DCA] spectrum, andtherefore the new structures for the boron-doped sample canbe explained by the same functional groups as formed for thepure [AMIM][DCA]. This also indicates that the nitrogendioxide reacts with the [AMIM][DCA] rather than the boronnanoparticles. Lastly, comparing the change in the intensitiesof peaks before and after the reaction indicates that the[DCA]− anion was degraded to a greater degree by NO2 thanthe [BMIM]+ cation.The reaction of nitrogen dioxide with [AMIM][DCA]

    caused the sharp peaks in the Raman spectrum to reduce andbe replaced by multiple broader new features. Similarly to theFTIR spectrum, the new features can be explained by theformation of new nitrogen- and oxygen-containing functionalgroups. The boron-doped [AMIM][DCA] spectrum beforethe reaction is a broad, featureless continuum; however, afterthe reaction, the boron-doped spectrum contains many broadnew peaks similar to the oxidized spectrum without boronnanoparticles. This confirms that nitrogen dioxide reacts withthe IL and not the boron nanoparticles.

    Figure 9. Fits to the Arrhenius plots for the rate constants k of peaks(a) (black), (c) (red), and (d) (blue) versus temperature todetermine the activation energies Ea.

    Table 6. Activation Energies, Ea, Calculated from ArrheniusPlots for Peaks (a), (c), and (d)

    peak wavenumber (cm−1) Ea (kJ mol−1)

    a 2182 38.7 ± 2.7c 1115 38.3 ± 5.7d 664 37.8 ± 6.3weighted mean 38.5 ± 2.3

    Table 7. Activation Energies, Ea, for Selected ReactionsInvolving NO2 Obtained from the NIST Chemical KineticsDatabase57

    reaction Ea (kJ mol−1)

    NO2 + CH3CCCCH3 50.1NO2 + i-C4H8 16.7NO2 + l-C4H8 30.5NO2 + 2-C4H8 47.7

    46.948.6

    NO2 + l-C5H10 30.1NO2 + CH3CHCH2 23.5

    33.1NO2 + CH2C(CH3)CHCH2 8.8NO2 + CH3CCH 53.5NO2 + C2H2 60.3NO2 + C2H4 75.3

    52.333.958.5

    NO2 + HCN 269.0NO2 + NCN 45.4NO2 + c-1,3-C6H8 25.5

    Figure 10. Summary of the possible functional groups formed whenNO2 reacts with the [DCA]

    − anion (left) or the [AMIM]+ cation(right).

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    413

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • The reaction rates for the oxidation of an [AMIM][DCA]droplet levitated in a 0.4% nitrogen dioxide and 99.6% argongas mixture were monitored using Raman spectroscopy. Thechange in the height of the new peak and those present in theunreacted spectrum were determined as a function of time.From the resulting decay curves for the degradation of[AMIM][DCA] and the growth curve for the newly formedpeak, the rate constants, k, were determined. The reaction rateconstants for the oxidation reactions were determined atvarious temperatures by heating the droplet with a carbondioxide laser. For the decay of selected [AMIM][DCA] peaks,the rates k ranged from (0.92−1.1) × 10−3 s−1 at 306 K to0.18−0.22 s−1 at 491 K. An Arrhenius plot for the rateconstants as a function of temperature was used to determinethe mean activation energy, Ea, from the three peaks of 38.5 ±2.3 kJ mol−1 for the oxidation of [AMIM][DCA]. For theselected peaks studied, the rate constants of [BMIM][DCA]and [MAT][DCA] were previously determined to be (1.5−2.5) × 10−3 s−1 and (7.3−10.4) × 10−5 s−1, respectively, atroom temperature for an NO2 concentration of 0.2%. Afteraccounting for the different NO2 concentrations between thetwo experiments,43 the oxidation rate of [AMIM][DCA] islower than for [BMIM][DCA], whereas [AMIM][DCA]oxidized one order of magnitude faster than [MAT][DCA].The similarity of the pure and boron-doped [AMIM][DCA]

    spectra after oxidation with NO2 in both the FTIR and Ramandata suggests that the nitrogen dioxide reacts with the IL andnot significantly with the boron nanoparticles. Therefore, the[AMIM][DCA] capping of the nanoparticles significantlyreduced the oxidation of the boron. This work providesfundamental insight into the oxidation of [AMIM][DCA] bysuggesting which new functional groups form. The formationof these possible new functional groups may be explained by areaction mechanism that initially requires a proton to betransferred from the [AMIM]+ cation to the [DCA]− anion;the NO2 can then react with the protonated anion or thedeprotonated cation. In addition, a direct reaction of NO2 withthe [DCA]− anion might also be possible. A summary of thepossible functional groups that could form via the reaction ofNO2 with each ion is shown in Figure 10. Further theoreticalstudies are required to fully understand the complete reactionmechanisms.

    ■ ASSOCIATED CONTENT*S Supporting InformationThe Supporting Information is available free of charge on theACS Publications website at DOI: 10.1021/acs.jpca.8b08395.

    Figure S1, decay profiles versus time for peak (a) at 2182cm−1; Figure S2, growth profiles versus time for peak (b)at 1115 cm−1; Figure S3, decay profiles versus time forpeak (c) at 1015 cm−1; Figure S4, decay profiles versustime for peak (d) at 664 cm−1 (PDF)

    ■ AUTHOR INFORMATIONCorresponding Author*E-mail: [email protected] L. Anderson: 0000-0001-9985-8178Ralf I. Kaiser: 0000-0002-7233-7206NotesThe authors declare no competing financial interest.

    ■ ACKNOWLEDGMENTSThis work was supported by the Office of Naval Researchunder Grant N00014-16-1-2078 (Hawaii) and by the Air ForceResearch Laboratory in Space Propulsion (RQRS) andPropellants (RQRP) Branches at Edwards Air Force Base.Particle production was supported by the Office of NavalResearch under Grant N00014-15-1-2681 (Utah).

    ■ REFERENCES(1) Lei, Z.; Chen, B.; Koo, Y.-M.; MacFarlane, D. R. Introduction:Ionic Liquids. Chem. Rev. 2017, 117, 6633−6635.(2) Dai, C.; Zhang, J.; Huang, C.; Lei, Z. Ionic Liquids in SelectiveOxidation: Catalysts and Solvents. Chem. Rev. 2017, 117, 6929−6983.(3) Earle, M. J.; Seddon, K. R. Ionic liquids. Green Solvents for theFuture. Pure Appl. Chem. 2000, 72, 1391−1398.(4) Zhang, Z.; Song, J.; Han, B. Catalytic Transformation ofLignocellulose into Chemicals and Fuel Products in Ionic Liquids.Chem. Rev. 2017, 117, 6834−6880.(5) Parvulescu, V. I.; Hardacre, C. Catalysis in Ionic Liquids. Chem.Rev. 2007, 107, 2615−2665.(6) Watanabe, M.; Thomas, M. L.; Zhang, S.; Ueno, K.; Yasuda, T.;Dokko, K. Application of Ionic Liquids to Energy Storage andConversion Materials and Devices. Chem. Rev. 2017, 117, 7190−7239.(7) Zhang, Y.; Gao, H.; Joo, Y.-H.; Shreeve, J. M. Ionic Liquids asHypergolic Fuels. Angew. Chem., Int. Ed. 2011, 50, 9554−9562.(8) Sebastiao, E.; Cook, C.; Hu, A.; Murugesu, M. RecentDevelopments in the Field of Energetic Ionic Liquids. J. Mater.Chem. A 2014, 2, 8153−8173.(9) Zhang, Q.; Shreeve, J. M. Energetic Ionic Liquids as Explosivesand Propellant Fuels: A New Journey of Ionic Liquid Chemistry.Chem. Rev. 2014, 114, 10527−10574.(10) Zhang, Q.; Shreeve, J. M. Ionic Liquid Propellants: FutureFuels for Space Propulsion. Chem. - Eur. J. 2013, 19, 15446−15451.(11) Zhang, Y. Q.; Gao, H. X.; Joo, Y. H.; Shreeve, J. M. IonicLiquids as Hypergolic Fuels. Angew. Chem., Int. Ed. 2011, 50, 9554−9562.(12) Ventura, S. P. M.; e Silva, F. A.; Quental, M. V.; Mondal, D.;Freire, M. G.; Coutinho, J. A. P. Ionic-Liquid-Mediated Extractionand Separation Processes for Bioactive Compounds: Past, Present,and Future Trends. Chem. Rev. 2017, 117, 6984−7052.(13) Cocalia, V. A.; Jensen, M. P.; Holbrey, J. D.; Spear, S. K.;Stepinski, D. C.; Rogers, R. D. Identical Extraction Behavior andCoordination of Trivalent or Hexavalent f-Element Cations UsingIonic Liquid and Molecular Solvents. Dalton Trans 2005, 1966−1971.(14) Lin, M.-C.; Gong, M.; Lu, B.; Wu, Y.; Wang, D.-Y.; Guan, M.;Angell, M.; Chen, C.; Yang, J.; Hwang, B.-J.; et al. An UltrafastRechargeable Aluminum-Ion Battery. Nature 2015, 520, 324−328.(15) MacFarlane, D. R.; Tachikawa, N.; Forsyth, M.; Pringle, J. M.;Howlett, P. C.; Elliott, G. D.; Davis, J. H.; Watanabe, M.; Simon, P.;Angell, C. A. Energy Applications of Ionic Liquids. Energy Environ. Sci.2014, 7, 232−250.(16) Egorova, K. S.; Gordeev, E. G.; Ananikov, V. P. BiologicalActivity of Ionic Liquids and Their Application in Pharmaceutics andMedicine. Chem. Rev. 2017, 117, 7132−7189.(17) Zakrewsky, M.; Lovejoy, K. S.; Kern, T. L.; Miller, T. E.; Le, V.;Nagy, A.; Goumas, A. M.; Iyer, R. S.; Del Sesto, R. E.; Koppisch, A.T.; et al. Ionic Liquids as a Class of Materials for TransdermalDelivery and Pathogen Neutralization. Proc. Natl. Acad. Sci. U. S. A.2014, 111, 13313−13318.(18) Clark, J. D. Ignition! An Informal History of Liquid RocketPropellants; Rutgers University Press: New Brunswick, NJ, 1972.(19) Sun, C.; Tang, S.; Zhang, X. Role of Cation Structures forEnergetic Performance of Hypergolic Ionic Liquids. Energy Fuels2017, 31, 10055−10059.(20) Vogl, T.; Goodrich, P.; Jacquemin, J.; Passerini, S.; Balducci, A.The Influence of Cation Structure on the Chemical-Physical

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    414

    http://pubs.acs.orghttp://pubs.acs.org/doi/abs/10.1021/acs.jpca.8b08395http://pubs.acs.org/doi/suppl/10.1021/acs.jpca.8b08395/suppl_file/jp8b08395_si_001.pdfmailto:[email protected]://orcid.org/0000-0001-9985-8178http://orcid.org/0000-0002-7233-7206http://dx.doi.org/10.1021/acs.jpca.8b08395

  • Properties of Protic Ionic Liquids. J. Phys. Chem. C 2016, 120, 8525−8533.(21) Carlin, C. M.; Gordon, M. S. Ab Initio Investigation of CationProton Affinity and Proton Transfer Energy for Energetic IonicLiquids. J. Phys. Chem. A 2016, 120, 6059−6063.(22) Chambreau, S. D.; Schneider, S.; Rosander, M.; Hawkins, T.;Gallegos, C. J.; Pastewait, M. F.; Vaghjiani, G. L. Fourier TransformInfrared Studies in Hypergolic Ignition of Ionic Liquids. J. Phys. Chem.A 2008, 112, 7816−7824.(23) Catoire, L.; Chambreau, S. D.; Vaghjiani, G. L. ChemicalKinetics Interpretation of Hypergolicity of Dicyanamide Ionic Liquid-Based Systems. Combust. Flame 2012, 159, 1759−1768.(24) Chambreau, S. D.; Koh, C. J.; Popolan-Vaida, D. M.; Gallegos,C. J.; Hooper, J. B.; Bedrov, D.; Vaghjiani, G. L.; Leone, S. R. Flow-Tube Investigations of Hypergolic Reactions of a Dicyanamide IonicLiquid Via Tunable Vacuum Ultraviolet Aerosol Mass Spectrometry.J. Phys. Chem. A 2016, 120, 8011−8023.(25) Politzer, P.; Murray, J. S.; Seminario, J. M.; Lane, P.; EdwardGrice, M.; Concha, M. C. Computational Characterization ofEnergetic Materials. J. Mol. Struct.: THEOCHEM 2001, 573, 1−10.(26) Rice, B. M.; Pai, S. V.; Hare, J. Predicting Heats of Formationof Energetic Materials using Quantum Mechanical Calculations.Combust. Flame 1999, 118, 445−458.(27) Cox, J. D.; Wagman, D. D.; Medvedev, V. A. CODATA KeyValues for Thermodynamics; Hemisphere Publishing Corp.: New York,NY, 1989.(28) Ulas, A.; Kuo, K. K.; Gotzmer, C. Ignition and Combustion ofBoron Particles in Fluorine-Containing Environments. Combust.Flame 2001, 127, 1935−1957.(29) Antaki, P.; Williams, F. A. Observations on the Combustion ofBoron Slurry Droplets in Air. Combust. Flame 1987, 67, 1−8.(30) King, M. K. Ignition and Combustion of Boron Particles andClouds. J. Spacecr. Rockets 1982, 19, 294−306.(31) Van Devener, B.; Perez, J. P. L.; Jankovich, J.; Anderson, S. L.Oxide-Free, Catalyst-Coated, Fuel-Soluble, Air-Stable Boron Nano-powder as Combined Combustion Catalyst and High Energy DensityFuel. Energy Fuels 2009, 23, 6111−6120.(32) Young, G.; Sullivan, K.; Zachariah, M. R.; Yu, K. CombustionCharacteristics of Boron Nanoparticles. Combust. Flame 2009, 156,322−333.(33) Zhou, W.; Yetter, R. A.; Dryer, F. L.; Rabitz, H.; Brown, R. C.;Kolb, C. E. Multi-Phase Model for Ignition and Combustion of BoronParticles. Combust. Flame 1999, 117, 227−243.(34) Perez, J. P. L.; McMahon, B. W.; Schneider, S.; Boatz, J. A.;Hawkins, T. W.; McCrary, P. D.; Beasley, P. A.; Kelley, S. P.; Rogers,R. D.; Anderson, S. L. Exploring the Structure of Nitrogen-Rich IonicLiquids and Their Binding to the Surface of Oxide-Free BoronNanoparticles. J. Phys. Chem. C 2013, 117, 5693−5707.(35) Perez, J. P. L.; McMahon, B. W.; Anderson, S. L.Functionalization and Passivation of Boron Nanoparticles with aHypergolic Ionic Liquid. J. Propul. Power 2013, 29, 489−495.(36) Perez, J. P. L.; Yu, J.; Sheppard, A. J.; Chambreau, S. D.;Vaghjiani, G. L.; Anderson, S. L. Binding of Alkenes and Ionic Liquidsto B-H-Functionalized Boron Nanoparticles: Creation of Particleswith Controlled Dispersibility and Minimal Surface Oxidation. ACSAppl. Mater. Interfaces 2015, 7, 9991−10003.(37) Perez, J. P. L.; McMahon, B. W.; Schneider, S.; Boatz, J. A.;Hawkins, T. W.; McCrary, P. D.; Beasley, P. A.; Kelley, S. P.; Rogers,R. D.; Anderson, S. L. Exploring the Structure of Nitrogen-Rich IonicLiquids and Their Binding to the Surface of Oxide-Free BoronNanoparticles. J. Phys. Chem. C 2013, 117, 5693−5707.(38) Perez, J. P. L.; McMahon, B. W.; Yu, J.; Schneider, S.; Boatz, J.A.; Hawkins, T. W.; McCrary, P. D.; Flores, L. A.; Rogers, R. D.;Anderson, S. L. Boron Nanoparticles with High Hydrogen Loading:Mechanism for B-H Binding and Potential for Improved Combust-ibility and Specific Impulse. ACS Appl. Mater. Interfaces 2014, 6,8513−8525.

    (39) Schneider, S.; Hawkins, T.; Rosander, M.; Vaghjiani, G.;Chambreau, S.; Drake, G. Ionic Liquids as Hypergolic Fuels. EnergyFuels 2008, 22, 2871−2872.(40) McCrary, P. D.; Beasley, P. A.; Cojocaru, O. A.; Schneider, S.;Hawkins, T. W.; Perez, J. P. L.; McMahon, B. W.; Pfeil, M.; Boatz, J.A.; Anderson, S. L.; et al. Hypergolic Ionic Liquids to Mill, Suspend,and Ignite Boron Nanoparticles. Chem. Commun. 2012, 48, 4311−4313.(41) Schneider, S.; Hawkins, T.; Ahmed, Y.; Rosander, M.;Hudgens, L.; Mills, J. Green Bipropellants: Hydrogen-Rich IonicLiquids that Are Hypergolic with Hydrogen Peroxide. Angew. Chem.,Int. Ed. 2011, 50, 5886−5888.(42) Brotton, S. J.; Lucas, M.; Chambreau, S. D.; Vaghjiani, G. L.;Yu, J.; Anderson, S. L.; Kaiser, R. I. Spectroscopic Investigation of thePrimary Reaction Intermediates in the Oxidation of LevitatedDroplets of Energetic Ionic Liquids. J. Phys. Chem. Lett. 2017, 8,6053−6059.(43) Brotton, S.; Lucas, M.; Jensen, T.; Anderson, S. L.; Kaiser, R. I.A Spectroscopic Study on the Intermediates and Reaction Rates in theOxidation of Levitated Droplets of Energetic Ionic Liquids byNitrogen Dioxide. J. Phys. Chem. A 2018, 122, 7351−7337.(44) Brotton, S. J.; Kaiser, R. I. Novel High-Temperature andPressure-Compatible Ultrasonic Levitator Apparatus Coupled toRaman and Fourier Transform Infrared Spectrometers. Rev. Sci.Instrum. 2013, 84, 055114.(45) Brotton, S. J.; Kaiser, R. I. In Situ Raman Spectroscopic Studyof Gypsum (CaSO4 • 2H2O) and Epsomite (MgSO4 • 7H2O)Dehydration Utilizing an Ultrasonic Levitator. J. Phys. Chem. Lett.2013, 4, 669−673.(46) Bennett, C. J.; Brotton, S. J.; Jones, B. M.; Misra, A. K.; Sharma,S. K.; Kaiser, R. I. High-Sensitivity Raman Spectrometer To StudyPristine and Irradiated Interstellar Ice Analogs. Anal. Chem. 2013, 85,5659−5665.(47) Xie, W. J.; Wei, B. Parametric Study of Single-Axis AcousticLevitation. Appl. Phys. Lett. 2001, 79, 881−883.(48) Xuan, X.-P.; Guo, M.; Pei, Y.-C.; Zheng, Y. Theoretical Studyon Cation-Anion Interaction and Vibrational Spectra of 1-Allyl-3-Methylimidazolium-Based Ionic Liquids. Spectrochim. Acta, Part A2011, 78, 1492−1499.(49) Katsyuba, S. A.; Zvereva, E. E.; Vidis, A.; Dyson, P. J.Application of Density Functional Theory and Vibrational Spectros-copy Toward the Rational Design of Ionic Liquids. J. Phys. Chem. A2007, 111, 352−370.(50) Paschoal, V. H.; Faria, L. F. O.; Ribeiro, M. C. C. VibrationalSpectroscopy of Ionic Liquids. Chem. Rev. 2017, 117, 7053−7112.(51) Kiefer, J.; Noack, K.; Penna, T. C.; Ribeiro, M. C. C.; Weber,H.; Kirchner, B. Vibrational Signatures of Anionic Cyano Groups inImidazolium Ionic Liquids. Vib. Spectrosc. 2017, 91, 141−146.(52) Estillore, A. D.; Visger-Kiefer, L. M.; Ghani, T. A.; Suits, A. G.Dynamics of H and D Abstraction in the Reaction of Cl Atom withButane-1,1,1,4,4,4-d6. Phys. Chem. Chem. Phys. 2011, 13, 8433−8440.(53) Socrates, G. Infrared and Raman Characteristic GroupFrequencies: Tables and Charts, 3rd ed.; Wiley: Chichester; NewYork, 2001; Chapter xv, p 347.(54) Colthup, N. B.; Daly, L. H.; Wiberley, S. E. Introduction toInfrared and Raman Spectroscopy, 3rd ed.; Academic Press: Boston,1990.(55) Bibart, C.; Ewing, G. Vibrational spectrum, torsional potential,and bonding of gaseous N2O4. J. Chem. Phys. 1974, 61, 1284−1292.(56) Steinfeld, J. I.; Francisco, J. S.; Hase, W. L. Chemical Kineticsand Dynamics, 2nd ed.; Prentice Hall: Englewood Cliffs, New Jersey,1999.(57) Manion, J. A.; Huie, R. E.; Levin, R. D.; Burgess, D. R., Jr.;Orkin, V. L.; Tsang, W.; McGivern, W. S.; Hudgens, J. W.; Knyazev,V. D.; Atkinson, D. B., et al., NIST Chemical Kinetics Database;National Institute of Standards and Technology: Gaithersburg, MD,2015.(58) Chambreau, S. D.; Schenk, A. C.; Sheppard, A. J.; Yandek, G.R.; Vaghjiani, G. L.; Maciejewski, J.; Koh, C. J.; Golan, A.; Leone, S.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    415

    http://dx.doi.org/10.1021/acs.jpca.8b08395

  • R. Thermal Decomposition Mechanisms of Alkylimidazolium IonicLiquids with Cyano-Functionalized Anions. J. Phys. Chem. A 2014,118, 11119−11132.(59) Maier, G.; Reisenauer, H. P.; Eckwert, J.; Naumann, M.;DeMarco, M. Isomers of the elemental composition CN2O. Angew.Chem., Int. Ed. Engl. 1997, 36, 1707−1709.(60) Dyke, J. M.; Jonathan, N.; Lewis, A. E.; Mills, J. D.; Morris, A.Vacuum ultraviolet photoelectron spectroscopy of transient species.Part 16. The NCO (X2Π) radical. Mol. Phys. 1983, 50, 77−89.(61) Baker, J.; Butcher, V.; Dyke, J. M.; Morris, A. Vacuumultraviolet photoelectron spectroscopic study of the NH2O and HNOmolecules. J. Chem. Soc., Faraday Trans. 1990, 86, 3843−3851.(62) Noble, M.; Nadler, I.; Reisler, H.; Wittig, C. The rotationallyresolved A 1Á́←X 1Á spectrum of expansion cooled NCNO:Vibrational fundamentals, rotational constants, and perturbations. J.Chem. Phys. 1984, 81, 4333−4340.(63) Thompson, W. E.; Jacox, M. E. The infrared spectrum of NN···CO+ trapped in solid neon. J. Chem. Phys. 2011, 135, 224307.(64) Nonella, M.; Muller, R. P.; Huber, J. R. Infrared spectra, normalcoordinate analysis, and photodecomposition of matrix-isolatedNH2NO2,

    15NH2NO2, ND2NO2, and15ND2NO2. J. Mol. Spectrosc.

    1985, 112, 142−152.

    The Journal of Physical Chemistry A Article

    DOI: 10.1021/acs.jpca.8b08395J. Phys. Chem. A 2019, 123, 400−416

    416

    http://dx.doi.org/10.1021/acs.jpca.8b08395