Eapen Thomas,
Kunduchi Periya Vijayalakshmi* and
Benny Kattikkanal George
Analytical and Spectroscopy Division, Analytical, Spectroscopy and Ceramics Group, Propellants, Polymers, Chemicals and Materials Entity, Vikram Sarabhai Space Centre, Thiruvananthapuram – 695022, India. E-mail: vijisura@gmail.com; kp_vijayalakshmi@vssc.gov.in
First published on 18th August 2015
A large variety of 1-ethyl-3-methylimidazolium ([EMIm]+) based energetic ionic liquids (ILs) have been studied via their ion pair ([EMIm]+[X]−) formation using high accuracy G3MP2 method and density functional theory (DFT) methods M06L, M05-2X, M06-2X and B3LYP. The selected X− includes nitrogen rich derivatives of tetrazolate and triazolate, dinitramine, dicyanamide as well as conventional anions BF4− and PF6−. The nitrogen enrichment in the system produces energetic ionic liquids (EILs) which showed comparable and in some cases superior thermochemical, fluid and specific impulse (Isp) properties than conventional ionic liquids. The binding energy values for [EMIm]+[X]− are in the range 336–400 kJ mol−1 at DFT levels while the atomization procedure used to compute their heat of formation (ΔfH°) at the G3MP2 level produced results in very close agreement with available experimental data (maximum deviation < 5%). The ΔfH° of conventional ILs is negative whereas that of EILs (167–559 kJ mol−1) confirmed their high energy state. The predicted Isp of all EILs are slightly lower compared to hydrazine in monopropellant systems whereas a significant increase in Isp is observed with the addition of hydroxyl ammonium nitrate (HAN). A good linear correlation between Isp and the wt% of (N + O) content of the EIL is also observed. Our results suggest that imidazolium based energetic ionic liquids have attractive thermochemical properties for use as green substitutes to hazardous hydrazine for monopropellant application in spacecraft technology.
Among various ILs in use, imidazolium based ILs are known for their chemical stability and fluid properties. The unique properties of these cations stem from the electronic structure of the aromatic cations. The electronic structure of imidazolium cations is characterized by 3-center-4-electron configuration across the N1–C2–N3 bonds, and a double bond between C4 and C5. The C2 is positively charged owing to the electron scarcity in the CN bond, imparting slight acidity to the C2–H hydrogen atom (Fig. 1). Hence the C2–H⋯anion interactions play an important role in the overall chemical and physical properties of the compound. The liquescency features of imidazolium based ILs is also well explained with these C2–H⋯anion interactions12 and are relevant to storability of these compounds as propellants. Apart from fluidity parameters, heat of formation (ΔfH°) turns out to be a decisive factor for propellant properties. The energy release during combustion of a high energy material is given by the difference between the heat of formation of the compound and that of the combustion products which increases with increasing positive heat of formation of the compound. In spite of the innumerous reports on structure-physical property correlation of ILs,13–15 studies on the thermo chemical properties are very limited. Recently, ab initio methods have been employed for accurate calculation of heat of formation parameters of energetic salts.16,17 Heintz et al.18 studied heat of formation of the ionic liquid 1-butyl-3-methylimidazolium dicyanamide (BMImdc) using atomization reaction and isodesmic reactions and observed a concurrence between theoretical and experimental results. In the present study, the formation and stabilities of several 1,3-dialkyl substituted imidazolium based ILs are predicted using DFT and ab initio calculations. The ILs and a binary mixture of ionic liquid with an oxidizer are explored and the performance is compared with that of hydrazine.
![]() | ||
Fig. 1 Chemical structures of imidazolium cation and the selected counter anions. The abbreviations used to name these systems are also given. |
The dispersion included Minnesota DFT method M06L23 as implemented in Gaussian 09 (ref. 24) is used in conjunction with 6-311+G(d,p) basis set for optimizing all the molecular geometries in their gas phase. Since the focus of this study is on the propellant applications of energetic ionic liquids (EILs) and considering that the process of ignition and combustion of EILs takes place in the gas phase, the gas phase results are expected to be reliable. Very recently, a benchmark study by Remya and Suresh showed that this method is the most effective to reproduce CCSD(T) level geometry and binding energy (Eb) of noncovalently bonded molecular dimers of organic molecules.25 Several orientations of the anion–cation pairs were considered to locate the most stable IL structure (ESI†). Normal modes of vibrations were analyzed for all the optimized geometries to ensure that every structure is a minimum on the potential energy surface. Basis set superposition error corrected Eb of the ion pair [EMIm]+[X]− formation in the gas phase (eqn (1)) was calculated at M06L/6-311+G(d,p) level.
[EMIm]+ + [X]− ⇌ [EMIm]+[X]− | (1) |
In the present study, atomization data obtained from the literature along with the G3MP2 level thermodynamic data is used to calculate the heat of formation for all IL systems. The G3MP2 is one of the popular high accuracy methods for calculating enthalpy quantities which is based on complex energy computations involving several pre-defined calculations on the specified molecular system.26
Calculating enthalpies of formation of the molecule at 298 K, ΔfH°(M, 298 K) can be split into a couple of steps which is given below.27
(a) Atomization energy of the molecule, ∑Do(M)
∑Do(M) = ∑nX(εo) − M(εo+ZPE) | (2) |
(b) Heat of formation of the molecule at 0 K, ΔfH°(M, 0 K)
ΔfH°(M, 0 K) = ∑nΔfH°(X, 0 K) − ∑Do(M) | (3) |
The experimental values of ΔfH°(X, 0 K) of the first and second row atomic elements at 0 K are tabulated in Table 1.
(c) Enthalpy correction for the molecule, H°M(298 K) − H°M(0 K)
H°M(298 K) − H°M(0 K) = Hcorr − M(εZPE) | (4) |
The enthalpy corrections, H°(298 K) − H°(0 K) is used to convert the atomic heats of formation at 0 K to those at 298.15 K, which is obtained from Gaussian output as “thermal correction to enthalpy”.
(d) Heat of formation of the molecule at 298 K, ΔfH°(M, 298 K).
ΔfH°(M, 298 K) = ΔfH°(M, 0 K) + [H°M(298 K) − H°M(0 K)] − [∑n(H°(298 K) − H°(0 K))] | (5) |
To estimate the accuracy of the proposed methods, an error function based on the normalized standard deviation is calculated as follows (eqn (6)),
![]() | (6) |
The term ‘specific Impulse’ (Isp) is used in rocket propulsion to define the impulse created per unit weight of propellant. It is equal to the number of pounds of thrust produced per pound of propellant burned per second. Isp is usually expressed in “seconds”. Using the G3MP2 level heat of formation of the molecule, rocket performance analysis is done to determine the capability of the selected ILs (combination of [EMIm]+ and any of the selected anions) to function as chemical monopropellants. Equilibrium combustion analysis is done by using the NASA Chemical Equilibrium with Applications (CEA) computer code.29 Isp is calculated by assuming equilibrium composition during expansion from infinite area combustor. CEA is employed for 100% fuel and then with a shifting equilibrium assumption for binary mixture, fuel and oxidiser. In the binary mixture of IL with HAN, Isp is calculated as a function of percent oxidiser by weight.
![]() | ||
Fig. 2 The optimized structures (left) of the representative set of EILs predicted for propellant applications and MESP plots (right). MESP-derived charge on anion is also depicted. Distances in Å. |
IL | Eb (kJ mol−1) | ΔE* (eV) | MK charge |
---|---|---|---|
[EMIm]+[BF4]− | 370.0 | 5.38 | −0.86 |
[EMIm]+[PF6]− | 351.2 | 5.54 | −0.85 |
[EMIm]+[dc]− | 361.9 | 2.83 | −0.80 |
[EMIm]+[trz]− | 398.7 | 3.31 | −0.67 |
[EMIm]+[dtrz]− | 335.5 | 3.14 | −0.83 |
[EMIm]+[dn]− | 355.9 | 3.30 | −0.85 |
[EMIm]+[tz]− | 388.0 | 3.68 | −0.69 |
[EMIm]+[mtz]− | 389.7 | 3.62 | −0.68 |
[EMIm]+[CNtz]− | 351.5 | 3.83 | −0.77 |
[EMIm]+[NH2tz]− | 399.8 | 2.94 | −0.72 |
[EMIm]+[NO2tz]− | 353.6 | 3.12 | −0.78 |
[EMIm]+[NO2Otz]− | 347.2 | 2.39 | −0.82 |
The cation–anion charge separation is significant in all the cases which is well evident in the molecular electrostatic potential (MESP) plots given in Fig. 2, where the anion has the MESP-derived Merz–Singh–Kollman (MK) charge in the range of −0.68 to −0.86 (Table 2) and cation has the counter positive values. The MK charge can be considered as a measure of the amount of charge transfer from the anion to the cation. The charge transfer (29–34%) is clearly more dominant in four energetic anions, viz. [trz]−, [tz]−, [mtz]− and [NH2tz]− which show Eb in the range 388–400 kJ mol−1 while the rest of the energetic anion show Eb in the range 336–370 kJ mol−1 with charge transfer 20 ± 5%.
The stability of compounds are also examined based on the energy gap (ΔE*) between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO). Conventional ILs (first two entries in Table 2) possess ΔE* of ∼5.5 eV while energetic ILs (EILs) are characterized with a lower band gap energy (2–4 eV). In general, a small ΔE* helps the electrons to cross the band gap easily, leading to poor compound stability30,31 as this process destroys charge separation in the ion-pair. The derivatives of tetrazolate anions show good stability in terms of the predicted band gap. Among different derivatives of tetrazolate based EILs, ΔE* predicts [EMIm]+[CNtz]− as the most stable as it possesses the highest ΔE* value 3.83 eV. The very high Eb (398.7 kJ mol−1) obtained for [EMIm]+[trz]− was decreased to 335.5 kJ mol−1 by dinitro substitution in the triazolate anion ([EMIm]+[dtrz]−). The electronic features have successfully been employed to tailor thermodynamic properties such as melting point of ILs. Generally, highly delocalised charge in the anion and shielded positive charge in the cation assist the formation of low melting ILs. In the given set of ion pairs, the best delocalization is observed for highly symmetric anions like [dc]− and [dtrz]− as revealed from the MESP plot given in Fig. 2. There are many strategies reported to fluidize imidazolium based ionic liquids. Introducing weakly polar anions generally reduce the interaction energy between cations and anions resulting in reduced melting points and decreased viscosities.32 Another strategy is to introduce asymmetry into the imidazolium cation by using different substituents at the N1 and N3 positions.33 Strong, directional and localized hydrogen bonds are also reported to enhance the fluid properties of imidazolium based ILs.34
We have considered all these factors while designing the compounds and sought possible correlation between computed Eb and melting points of known compounds. Generally higher Eb correlates to higher melting point which suggest that the melting points of energetic ILs could be higher than the traditional ILs (first two entries in Table 2) and among them the [EMIm]+[NH2tz]− is expected to possess the highest melting point. Organic salts of [dc]− anions, in general are known to possess relatively lower melting points35 and reported melting point of [EMIm]+[dc]− is −27 °C. Among the named EILs, [EMIm]+[CNtz]−, [EMIm]+[dn]−, [EMIm]+[dtrz]−, [EMIm]+[NO2tz]− and [EMIm]+[NO2Otz]− possess lower Eb than the [EMIm]+[dc]− and indicate their existence in liquid state.
Compound | ΔfH° (kJ mol−1) | |||||
---|---|---|---|---|---|---|
DFT | Ab initio | Experimental | ||||
B3LYP | M06L | M05-2X | M06-2X | G3MP2 | ||
1,3-H-Imidazolium cation | 758.8 | 749.5 | 716.5 | 752.8 | 725.8 | 719.8 (ref. 36) |
1,4-H-1,2,4-Triazolium cation | 867.0 | 869.3 | 908.1 | 877.7 | 847.3 | 836.0 (ref. 37) |
Tetrazolate anion | 192.4 | 202.4 | 222.4 | 228.3 | 186.1 | 200.2 (ref. 37) |
1,2,4-Triazolate anion | 120.6 | 121.2 | 142.3 | 135.1 | 100.8 | 102.8 (ref. 37) |
[BMIm]+[dc]− | 499.7 | 404.0 | 592.1 | 445.7 | 360.9 | 363.4 (ref. 18) |
[EMIm]+[dc]− | 510.3 | 422.2 | 576.8 | 466.4 | 404.1 | 391.7 (ref. 38) |
Hydrazine | 91.9 | 151.6 | 158.5 | 124.2 | 92.1 | 90.0 (ref. 39), 95.0 (ref. 40) |
Monomethyl hydrazine | 104.0 | 158.1 | 185.3 | 128.2 | 82.9 | 92.0 (ref. 39) |
Pyridine | 193.4 | 141.1 | 219.8 | 177.3 | 133.5 | 140.0 (ref. 41) |
Pyridazine | 322.7 | 285.9 | 356.9 | 325.1 | 277.4 | 278.3 (ref. 41) |
Pyrazine | 245.2 | 210.5 | 276.0 | 243.0 | 201.8 | 196.0 (ref. 41) |
1,2,3-Triazine | 435.9 | 411.5 | 455.6 | 477.4 | 403.3 | 416.0 (ref. 41) |
1,2,4-Triazine | 368.1 | 348.9 | 404.6 | 382.6 | 337.1 | 334.0 (ref. 41) |
1,3,5-Triazine | 256.8 | 239.8 | 282.9 | 263.4 | 226.4 | 226.0 (ref. 41) |
For EILs to be used for propellant applications, they require high energy density which is often associated with large positive heat of formation. This in turn results in high combustion chamber temperature and hence a higher Isp. Conventional ILs, viz. [EMIm]+[BF4]− and [EMIm]+[PF6]− have negative heat of formation suggesting a lower energy content. The heat of formation increases from −1692.2 to +558.6 kJ mol−1 when the anionic part is substituted by energetic groups (Table 4). The computed ΔfH° of all the nitrogen rich energetic ILs possess much higher heat of formation than hydrazine and suggests their use for propellant applications. The highest value of ΔfH°(g) is obtained for CNtz (558.6 kJ mol−1) derivative. The lower values of ΔfH°(g) for [EMIm]+[trz]−, [EMIm]+[Mtz]− and [EMIm]+[dc]− can be attributed to their higher carbon content.
Compound | Molecular formula | ΔfH° (kJ mol−1) | (N + O) (wt%) | Isp (s) | % HAN with IL for Isp = 263 s | Isp (s) maximum (% HAN) |
---|---|---|---|---|---|---|
a A condensed species H3BO3 is formed. | ||||||
Hydrazine | N2H4 | 90.0 | 87.5 | 263 | — | — |
[EMIm]+[BF4]− | C6N2H11BF4 | −1692.2 | 14.2 | —a | 68 | 295 (80%) |
[EMIm]+[PF6]− | C6N2H11PF6 | −1909.2 | 10.8 | 105 | 63 | 287 (80%) |
[EMIm]+[tz]− | C7N6H12 | 441.9 | 46.7 | 221 | 54 | 317 (80%) |
[EMIm]+[dc]− | C8N5H11 | 401.3 | 39.5 | 208 | 58 | 314 (80%) |
[EMIm]+[mtz]− | C8N6H14 | 404.2 | 43.3 | 213 | 57 | 315 (80%) |
[EMIm]+[trz]− | C8N5H13 | 346.9 | 39.1 | 207 | 59 | 314 (80%) |
[EMIm]+[dtrz]− | C8N7H11O4 | 305.9 | 60.2 | 241 | 34 | 315 (70%) |
[EMIm]+[dn]− | C6N5H11O4 | 166.9 | 61.7 | 252 | 23 | 319 (70%) |
[EMIm]+[CNtz]− | C8N7H11 | 558.6 | 47.8 | 221 | 54 | 316 (80%) |
[EMIm]+[NH2tz]− | C7N7H13 | 444.8 | 50.3 | 219 | 53 | 317 (80%) |
[EMIm]+[NO2tz]− | C7N7H11O2 | 418.8 | 57.7 | 239 | 40 | 315 (80%) |
[EMIm]+[NO2Otz]− | C7N7H11O3 | 387.1 | 60.6 | 246 | 33 | 316 (70%) |
HAN | N2H4O4 | −198.2 | 95.8 | 238 | — | — |
Oxygen balance is another significant index of energetic materials which is a measure of oxygen deficiency or excess of oxygen in a compound required to convert all the carbon into carbon monoxide and all the hydrogen into water. Generally oxygen rich compounds have higher Isp as all elements get converted into gaseous products resulting in increased thrust. In the present set of EILs, the absence oxygen content leads to relatively lower Isp for [EMIm]+[dc]−, [EMIm]+[NH2]−, [EMIm]+[CNtz]−, [EMIm]+[tz]−, [EMIm]+[mtz]− and [EMIm]+[trz]− in spite of their higher values of ΔfH°(g) (>400 kJ mol−1). The highest Isp of 252 s is obtained for [EMIm]+[dn]− with two N–NO2 groups in the anion. In fact a linear correlation of Isp with total (N + O) wt% is observed for all the ILs studied in the present work (Fig. 3). The low oxygen balance of imidazolium based EILs can be overcome by using suitable oxidizer such as HAN and solvent properties of ILs can beneficially exploited to form binary-monopropellant mixtures.42
![]() | ||
Fig. 3 Correlation graph between wt% of (N + O) and predicted Isp of [EMIm]+ ion pairs with 11 different anions given in Table 4. |
ILs has been tested for hypergolicity with HAN oxidizer and it showed no visible signs of reactivity at room temperature.43,44 Hence a monopropellant mixture of IL with varying percentage of HAN is analyzed using CEA, which would be thermally stable at room temperature and ignited thermally or catalytically. The Isp gradually increases with increase in the oxidizer concentration. This change is rapid in the 20–50% concentration and reaches the peak region around 60–80% concentration (ESI†).
These results clearly suggest that Isp tuning can be achieved by mixing ILs with suitable amount of HAN and these ILs may find use as fuel component in a binary propellant system. The ILs with oxygen rich anions, viz. [dn]−, [dtrz]− and [NO2Otz]− attain the Isp of currently used propellant hydrazine (263 s) at 23–34% of HAN concentration (Table 4). An enhancement of 56 units in Isp is obtained for [EMIm]+[dn]− for 70% HAN mixtures which provide the limit of Isp that can be achieved using the derivatives of imidazolium based ionic liquids (Table 4).
[EMIm]+ | 1-Ethyl-3-methylimidazolium cation |
[BMIm]+ | 1-Butyl-3-methylimidazolium cation |
Isp | Specific impulse |
DFT | Density functional theory |
EIL | Energetic ionic liquid |
[BF4]− | Tetrafluoroborate |
[PF6]− | Hexafluorophosphate |
[tz]− | 5-H-Tetrazolate |
[dc]− | Dicyanamide |
[mtz]− | 5-Methyl tetrazolate |
[trz]− | 1,2,4-Triazolate |
[dtrz]− | 3,5-Dinitro-1,2,4-triazolate |
[dn]− | Dinitramine |
[CNtz]− | 5-Cyano tetrazolate |
[NH2tz]− | 5-Amino tetrazolate |
[NO2tz]− | 5-Nitrotetrazolate |
[NO2Otz]− | 5-Nitrotetrazolate-2N-oxide |
ΔfH°(M, 0 K) | Heat of formation of the molecule at 0 K |
CCSD(T) | Coupled cluster method |
Δq | Normalized standard deviation |
Eb | Binding energy |
ΔE* | Energy gap between the HOMO and LUMO |
HOMO | Highest occupied molecular orbital |
LUMO | Lowest unoccupied molecular orbital |
MK charge | Merz Singh Kollman charge |
HAN | Hydroxyl ammonium nitrate |
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c5ra09578f |
This journal is © The Royal Society of Chemistry 2015 |