Next Article in Journal
Fault Characterization of a Proton Exchange Membrane Fuel Cell Stack
Next Article in Special Issue
Thermodynamic Analysis on the Aging of THPP, ZPP and BKNO3 Explosive Charges in PMDs
Previous Article in Journal
Evaluation of Direct Horizontal Irradiance in China Using a Physically-Based Model and Machine Learning Methods
Previous Article in Special Issue
Considering Multiple Factors to Forecast CO2 Emissions: A Hybrid Multivariable Grey Forecasting and Genetic Programming Approach
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

A Combined Study of TEM-EDS/XPS and Molecular Modeling on the Aging of THPP, ZPP, and BKNO3 Explosive Charges in PMDs under Accelerated Aging Conditions

1
Department of Chemical Engineering, Pukyong National University, Busan 48547, Korea
2
Department of Chemical Engineering, Pohang University of Science and Technology (POSTECH), Pohang 37673, Kyoungbuk, Korea
3
Defense R&D Center, Hanwha Corporation, Daejeon 34068, Korea
4
Agency for Defense Development, Daejeon 34183, Korea
*
Author to whom correspondence should be addressed.
These authors contributed equally to this work.
Energies 2019, 12(1), 151; https://doi.org/10.3390/en12010151
Submission received: 19 November 2018 / Revised: 29 December 2018 / Accepted: 29 December 2018 / Published: 2 January 2019

Abstract

:
The aging mechanism of explosive charges in pyrotechnic mechanical devices (PMDs) is pre-oxidations of their fuels (TiH2 for THPP, Zr for ZPP, and B for BKNO3) by external oxygen. The effect of water on the aging of explosive charges was thus investigated by TEM-EDS/XPS and DFT-based molecular modeling under accelerated aging with 71 °C and 100% relative humidity. The formation of oxide shell and its thickness on the surface of fuels by the aging were observed by TEM-EDS. It failed to detect any oxide on the surface of TiH2 (no sign of Ti-O peaks in XPS) regardless of the aging time, while the thickness of oxide shell increases linearly with the time for ZPP and is saturated at a certain point for BKNO3. It suggested that THPP is highly robust to aging compared to the others (the order of THPP >> BKNO3 > ZPP). Then, DFT-based vacuum slab calculations visualized the diffusion of oxygen from the surface of fuels into the interior, confirming that the activation barrier for the oxygen diffusion is much lower for Zr and B than TiH2 (37, 107, and 512 kcal/mol for Zr, B, and TiH2, respectively), in agreement with experimental results.

1. Introduction

Pyrotechnic mechanical devices (PMDs) convert the explosive power (or pressure) of explosive charges into mechanical force to ignite the serial explosion of the bomb or rocket [1]. The three primary explosive charges used in fields are boron potassium nitrate (BKNO3), zirconium potassium perchlorate (ZPP), and titanium hydride potassium perchlorate (THPP) [2,3,4]. However, the aging of those explosive charges has rarely been studied intensively, although it significantly affects the long-term performance (or stability) of PMDs [5,6].
Recently, we have studied how the aging of BKNO3 and ZPP is affected by internal and external factors [7,8]. It turned out that internal factors such as spontaneous pre-reactions have no effect, but pre-oxidations of fuels (B and Zr) by external oxygen source is the key aging mechanism of explosive charges. According to AKTS (advanced kinetics and technology solutions) simulations, any spontaneous pre-reaction such as laminac (a polymer binder used for integrating metal and oxidizer in explosive charges) decomposition would not occur in less than 500 years even at a storage temperature of 120 °C [7,8]. Meanwhile, TEM (transmission electron microscopy)-EDS (energy dispersion spectroscopy) and XPS (X-ray photoelectron spectroscopy) studies confirmed the formation of oxide shells on the surface of fuels and their growth with time under accelerated aging conditions [7,8]. A substantial decrease of relative heat (18% and 40% for BKNO3 and ZPP, respectively) was also detected by DSC (differential scanning calorimetry) measurements for 16-week aged samples under accelerated aging conditions, confirming that the aging of explosive charges is mostly affected by extra oxygen sources, such as moisture [7,8]. In addition, thermodynamic calculations showed the flame temperature decrease (or explosive power decrease) of BKNO3 due to pre-oxidations of B by extra oxygen source, complied with experimental results [9,10]. However, the growth of oxide shells with time was different between B and Zr, suggesting that the extent of aging depends on the materials of explosive charges [7,8].
In this study, we employed the same TEM-EDS and XPS study for THPP finally to compare the extent of aging by time among three explosive charges (BKNO3, ZPP, and THPP) and proposed which one is the most robust to the aging and why. DSC measurements were used complementarily. Furthermore, DFT (density functional theory)-based computational calculations with the vacuum slab model were introduced to support our TEM-EDS/XPS studies by the calculated energetics of oxygen diffusion from the surface of fuels (TiH2 for THPP, Zr for ZPP, and B for BKNO3) of explosive charges to the interior.

2. Materials and Methods

2.1. TEM-EDS and XPS Study under Accelerated Aging Conditions

THPP samples (30.0 wt% of TiH2, 65.0 wt% of KClO4, and 5.0 wt% of VitonB) were provided from Hanwha Corp. (Daejeon, Korea) and aged under accelerated aging conditions at 71 °C in 100% RH (relative humidity) for up to 16 weeks; for reference, BKNO3 consisted of 23.7 wt% of boron, 70.7 wt% of KNO3, and 5.6 wt% Laminac and ZPP 52.0 wt% consisted of Zirconium, 43.0 wt% of KClO4, and 5 wt% of VitonB. Then, the samples for TEM-EDS studies were prepared by immersing TEM meshes into ethanol containing suspended aged TiH2 particles. After being dried in air, the TEM meshes were used for TEM imaging, elemental mapping, and EDS analysis by a Titan G2 ChemiSTEM Cs probe (FEI Company, Eindhoven, The Netherlands) with an acceleration voltage of 200 kV. XPS spectra were obtained using a MultiLab 2000 with monochromated Al Kα X-rays and a hemispherical analyzer with a pass energy of 30 eV. The THPP samples were prepared on conductive copper tape. Background subtraction was performed using a Shirley background and MonoXPS.

2.2. DSC Measurements

DSC measurements were performed using a DSC 4000 (PerkinElmer, Waltham, MA, USA). The heats of reaction were measured as the samples were heated to 550 °C at a scanning rate of 2 °C/min. A slow heating rate could generate thermodynamic information with a sufficient relaxation time.

2.3. DFT-Based Computational Calculation

The vacuum slab model was employed to calculate the energetics of oxygen diffusion from the surface of fuels (TiH2 for THPP, Zr for ZPP, and B for BKNO3) into the interior [11,12]. The calculated activation (or kinetic) barrier was then compared to evaluate the extent of surface oxidation with time among three explosive charges. Firstly, supercells are constructed by 2 × 2 × 2 expansion from the primitive cells of fuels in literatures; simple rhombohedral B (space group R-3m), hexagonal closed packed Zr (P63/mmc), and cubic TiH2 (Fm-3m) structures [13,14,15]. Secondly, (100) surfaces were cleaved out from the supercells and vacuum slabs were built with a sufficiently large vacuum thickness of 5 Å to avoid any artificial effect from the upper and lower slabs [11]. Thirdly, an oxygen atom (or radical) is placed on the cleaved surface and the optimized geometry of the corresponding local minimum was located. The energy of this geometry was used as a baseline for the energetics of oxygen diffusion. In every calculation, the fractional coordinates of all the layers were fixed because the coordinate relaxation is considered to have limited effect for the comparison of the energetics and the computation time was also a factor. Finally, the energetics was constructed by moving the position of oxygen atom into the interior and calculating the corresponding energy in every step. Although extra oxygen sources, such as water, have to be decomposed into oxygen atoms (or radicals) on the surface prior to the subsequent oxygen diffusion, our concern was on the kinetics of the growth of oxide shells and thus the decomposition step was not counted here.
DFT calculations for the energetics were carried out using the CASTEP program suite in Materials Studio (Accelrys Inc., San Diego, CA, USA) [16]. The DFT exchange-correlation potential used in the calculations was the generalized gradient approximation (GGA) along with the PW91 (Perdew-Wang) functional [17]. Core and valence electron interactions were considered using on-the-fly generation (OTFG) ultrasoft pseudopotentials and a plane-wave basis set with a cutoff energy of 381 eV [18]. The tolerance for energy convergence was set to as low as 5.0 × 105 eV/atom to speed up the calculations.

3. Results and Discussion

3.1. Oxide Shell Formation

The formation of oxide shells on the surface of TiH2 particles stored in accelerated aging conditions was monitored by TEM-EDS characterizations. However, it failed to observe any oxide layer on the surface regardless of the duration of accelerated aging conditions, completely different from B and Zr in our previous studies. Figure 1a shows the TEM-EDS result for the 8-week sample, where atomic contents were characterized across the surface of a TiH2 particle from the vacuum to the interior of the particle. There is no oxygen peak surge on the surface, unlike in B and Zr particles [7,8]. It is all the same for the pristine and 16-week aged samples, suggesting that THPP is highly robust to the aging by extra oxygen source. XPS spectra showed no trace of Ti-oxide more clearly (see Figure 1b). Only Ti-hydride peaks are present. The decrease of relative peak intensity of Ti-hydride peaks depending on the duration of accelerating aging proposes a possibility of physical surface deformation such as local metal coagulation.
Quantitative heat analysis by using DSC measurements was carried out as well. As shown in Figure 2, there are endothermic peaks at 300 °C linked with the crystal structure change of KClO4 from orthorhombic to face-centered cubic and exothermic peaks related to combustion reactions from 350 to 550 °C [19]. To consider the error in weighing highly energetic materials, relative heat released was calculated by the ratio of exothermic heat to endothermic heats; 51.9, 49.9, and 49.0 for the pristine, 8-week, and 16-week aged samples, respectively. A 6% decrease was measured for the 16-week sample, compared to 18% and 40% decreases for BKNO3 and ZPP, respectively [7,8]. In accordance, the thickness of oxide shells with the duration of aging shows different behavior among explosive charges as shown in Figure 3. While the thickness of oxide shell increases linearly with the duration for Zr and is saturated at a certain point for B, there is no change for TiH2. The combination of these two results (DSC results and the thickness change of oxide shells), suggests that the aging effect is negligible in THPP compared to BKNO3 and ZPP, in the order of ZPP (high) > BKNO3 >> THPP (negligible).

3.2. Computational Interpretation of Oxygen Diffusion

Figure 3 shows the change of oxide shell thickness with the duration of accelerated aging for the three explosive charges. It is apparent that ZPP experiences performance degradation when exposed to water for a long time because oxygen keeps penetrating into the interior of Zr, allowing growth of oxide shells by pre-oxidation. Pre-oxidation means the pre-consumption of explosive power, leading to the decreased power of aged explosive charges [9,10]. On the contrary, THPP is superbly robust to aging. Because pre-oxidations are in the regime of kinetics, the comparison of barriers for the oxygen diffusion into the interior (in other words, activation energies for the oxygen diffusion) among the three explosive charges could give quantitative support for experimental results.
Figure 4 shows constructed vacuum slab systems. The baseline is the energy when the oxygen atom (or radical) is in the most stable surface position. At first glance, the B structure is too dense for oxygen penetration (see Figure 4c) compared to Zr and TiH2, although the chemical affinity of oxygen to B, Zr, or Ti will be counted later by calculations. Meanwhile, the Ti matrix (see Figure 4b) seems to be as spacious as that of Zr (see Figure 4a), but the presence of interstitial hydrogen atoms probably inhibits the diffusion of oxygen atoms.
Figure 5 presents the calculated energetics for the oxygen penetration into the interior of Zr, TiH2, and B. Especially, Figure 5a visualizes how the energy calculation was done according to the position of penetrating oxygen as an example. Along with the baseline position (the 1st point in Figure 5a), we detected a local minimum in the vacancy just underneath the surface (the 6th point in Figure 5a). Thus, in the journey from the baseline to the local minimum inside, the energy related to the position of penetrating oxygen increases due to the neighboring Zr atoms and decreases at a certain point down to that of the internal local minimum. The energy increases again for further penetration. The anomalous energy increase (the 7th point in Figure 5a) probably resulted from the small depth of constructed vacuum slab. Meanwhile, it is noted that the local minimum inside is quite stable for Zr, compared to TiH2 and B (the 6th points in Figure 5a–c), favored by the presence of spacious vacancies in Zr matrix and the chemical affinity of oxygen to Zr. From the energetics in Figure 5, the activation energies were estimated as ~37, ~512, ~107 kcal/mol for Zr, TiH2, and B, respectively. It means that oxygen penetration into the interior is almost impossible for TiH2. The order of estimated activation energies (TiH2 >> B > Zr) are consistent with the findings in Figure 3. In this calculation, the absolute accuracy of calculated values is limited due to the usage of rather simplified calculation model and scheme. But it has its rationale to compare the values calculated by the same methodology because any inaccuracy will be cancelled out in comparison.

4. Conclusion

We investigated how susceptible three explosive charges (BKNO3, ZPP, and THPP) are to aging with an extra oxygen source (i.e., water). The presence of an oxide shell on the surface of THPP and its thickness change with the duration of accelerated aging were monitored by TEM-EDS, but no sign of oxide formation was detected. XPS also confirmed the presence of Ti-hydride, but without any Ti-oxide. Compared to BKNO3 and ZPP in our previous studies, THPP is highly robust to aging (the order of THPP >> BKNO3 > ZPP). A minor decrease of relative heat released from DSC measurements for THPP indicated the same conclusion as well.
Meanwhile, the energetics of oxygen diffusion from the surface of explosive charges into the interior were obtained by DFT-based molecular modeling using the vacuum slab model. The calculated activation barrier for oxygen diffusion is much lower for ZPP and BKNO3 compared to THPP (the order of ZPP < BKNO3 << THPP), suggesting that ZPP is the most susceptible to the pre-oxidation (or the aging) by an extra oxygen source.

Author Contributions

Conceptualization, Y.S.W.; Methodology, J.L., K.M.K. and S.I.C.; Software, K.M.K. and S.I.C.; Validation, Y.H.K.; Formal analysis, J.L.; Investigation, J.L., K.M.K. and S.I.C.; Resources, Y.S.W.; Data curation, J.L., K.M.K. and S.I.C.; Writing—original draft preparation, J.L. and Y.S.W.; Writing—review and editing, Y.S.W.; Visualization, J.L., K.M.K. and S.I.C.; Supervision, J.G.P.; Project administration, G.H.A.; Funding acquisition, B.T.R.

Funding

This research received no external funding.

Acknowledgments

This work was supported by the Agency for Defense Development under the Precise Energy Release for the Pyrotechnic Mechanical Device program and the BB21+ Project in 2018. The authors sincerely thanks Professor Taiho Park in the Department of Chemical Engineering in POSTECH for the analytical supports in this manuscript.

Conflicts of Interest

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

References

  1. Bement, L.J.; Multhaup, H.A. Determining Functional Reliability of Pyrotechnic Mechanical Devices. AIAA J. 1999, 37, 357–363. [Google Scholar] [CrossRef] [Green Version]
  2. Lee, J.; Lin, L.; Lin, C.; Ch’en, P.; Huang, C.; Chang, S. A Study of Zirconium/Potassium Perchlorate Primer Mixtures. Thermochim. Acta 1990, 173, 211–218. [Google Scholar] [CrossRef]
  3. Ulas, A.; Risha, G.A.; Kuo, K.K. An Investigation of the Performance of a Boron/Potassium-nitrate Based Pyrotechnic Igniter. Propellants Explos. Pyrotech. 2006, 31, 311–317. [Google Scholar] [CrossRef]
  4. Yan, N.; Bao, B.; Zheng, F.; Li, C. Ignition Characteristics of Micro-energy Semiconductor Bridges with Different Ignition Compositions. Propellants Explos. Pyrotech. 2016, 41, 223–227. [Google Scholar] [CrossRef]
  5. Olmos, R.P.; Rios, A.; Martín, M.P.; Lapa, R.A.S.; Lima, J.C.F.C. Construction and Evaluation of Ion Selective Electrodes for Perchlorate with a Summing Operational Amplifier: Application to Pyrotechnics Mixtures Analysis. Analyst 1999, 124, 97–100. [Google Scholar] [CrossRef]
  6. Lai, K.S. Boron Potassium Nitrate (BKNO3) Aging Study. In Proceedings of the 34th AIAA/ASME/SAE/ASEE Joint Propulsion Conference and Exhibit, Cleveland, OH, USA, 13–15 July 1998. [Google Scholar]
  7. Lee, J.; Kim, T.; Ryu, S.U.; Choi, K.; Ahn, G.H.; Paik, J.G.; Ryu, B.; Park, T.; Won, Y.S. Study on the Aging Mechanism of Boron Potassium Nitrate (BKNO3) for Sustainable Efficiency in Pyrotechnic Mechanical Devices. Sci. Rep. 2018, 8, 11745–11754. [Google Scholar] [CrossRef] [PubMed]
  8. Lee, J.; Choi, K.; Ryu, S.U.; Ahn, G.H.; Paik, J.G.; Ryu, B.; Won, Y.S. Aging Mechanism of Zirconium Potassium Perchlorate Propellant in Pyrotechnic Mechanical Devices. Nanosci. Nanotechnol. Lett. 2018, 10, 735–740. [Google Scholar] [CrossRef]
  9. Eom, K.H.; An, H.Y.; Kim, K.M.; Ahn, G.H.; Paik, J.G.; Ryu, B.T.; Im, D.J.; Won, Y.S. Equilibrium Analysis on the Aging of a BKNO3 Igniter. J. Nanosci. Nanotechnol. 2017, 17, 7685–7688. [Google Scholar] [CrossRef]
  10. Kong, T.K.; Won, Y.S.; Ryu, B.; Ahn, G.H.; Im, D.J. Mathematical Modeling of ZrKClO4 Nano Particle Energy Release. J. Nanosci. Nanotechnol. 2017, 17, 8372–8377. [Google Scholar] [CrossRef]
  11. Park, H.; Kim, K.M.; Kim, H.; Kim, D.; Won, Y.S.; Kim, S. Electrodeposition-fabricated PtCu-alloy Cathode Catalysts for High-temperature Proton Exchange Membrane Fuel Cells. Korean J. Chem. Eng. 2018, 35, 1547–1555. [Google Scholar] [CrossRef]
  12. Savizi, I.S.P.; Janik, M.J. Acetate and Phosphate Anion Adsorption Linear Sweep Voltammograms Simulated Using Density Functional Theory. Electrochim. Acta 2011, 56, 3996–4006. [Google Scholar] [CrossRef]
  13. Decker, B.F.; Kasper, J.S. The Crystal Structure of a Simple Rhombohedral Form of Boron Locality: Synthetic. Acta Crystallogr. 1959, 12, 503–506. [Google Scholar] [CrossRef]
  14. Wyckoff, R.W.G. Crystal Structures; John Wiley: New York, NY, USA, 1963. [Google Scholar]
  15. Villars, P. Springer & Material Phases Data System (MPDS); Switzerland & National Institute for Materials Science (NIMS): Tsukuba, Japan, 2016. [Google Scholar]
  16. Segall, M.D.; Lindan, P.J.D.; Probert, M.J.; Pickard, C.J.; Hasnip, P.J.; Clark, S.J.; Payne, M.C. First-principles Simulation: Ideas, Illustrations and the CASTEP Code. J. Phys. Condens. Matter 2002, 14, 2717–2744. [Google Scholar] [CrossRef]
  17. Perdew, J.P.; Chevary, J.A.; Vosko, S.H.; Jackson, K.A.; Pederson, M.R.; Singh, D.J.; Carlos, F. Atoms, Molecules, Solids, and Surfaces: Applications of the Generalized Gradient Approximation for Exchange and Correlation. Phys. Rev. B 1992, 46, 6671–6687. [Google Scholar] [CrossRef]
  18. Vanderbilt, D. Soft Self-Consistent Pseudopotentials in a Generalized Eigenvalue Formalism. Phys. Rev. B 1990, 41, 7982. [Google Scholar] [CrossRef]
  19. Denise, B.; Debeau, M.; Depondt, P.; Heger, G. Orientational Disorder in the High Temperature Phase of KCIO4. J. Phys. France 1988, 49, 1203–1210. [Google Scholar] [CrossRef]
Figure 1. TEM-EDS (a) and XPS (b) characterizations of THPP under accelerated aging conditions for 8 weeks. Unlike BKNO3 and ZPP, the formation of an oxide layer on the surface was not detected, regardless of the duration of accelerated aging.
Figure 1. TEM-EDS (a) and XPS (b) characterizations of THPP under accelerated aging conditions for 8 weeks. Unlike BKNO3 and ZPP, the formation of an oxide layer on the surface was not detected, regardless of the duration of accelerated aging.
Energies 12 00151 g001
Figure 2. DSC profiles of THPP under accelerated aging conditions. The relative heat released rarely changed with the duration of accelerated aging.
Figure 2. DSC profiles of THPP under accelerated aging conditions. The relative heat released rarely changed with the duration of accelerated aging.
Energies 12 00151 g002
Figure 3. The thickness change of oxide shell vs. aging time. The aging effect has the order of ZPP >> BKNO3 > THPP [7,8].
Figure 3. The thickness change of oxide shell vs. aging time. The aging effect has the order of ZPP >> BKNO3 > THPP [7,8].
Energies 12 00151 g003
Figure 4. Constructed vacuum slab systems (side view) and their dimensions for Zr (a), TiH2 (b), and B (c).
Figure 4. Constructed vacuum slab systems (side view) and their dimensions for Zr (a), TiH2 (b), and B (c).
Energies 12 00151 g004
Figure 5. Calculated energetics for oxygen penetration into the inside of Zr (a), TiH2 (b), and B (c).
Figure 5. Calculated energetics for oxygen penetration into the inside of Zr (a), TiH2 (b), and B (c).
Energies 12 00151 g005

Share and Cite

MDPI and ACS Style

Kim, K.M.; Lee, J.; Choi, S.I.; Ahn, G.H.; Paik, J.G.; Ryu, B.T.; Kim, Y.H.; Won, Y.S. A Combined Study of TEM-EDS/XPS and Molecular Modeling on the Aging of THPP, ZPP, and BKNO3 Explosive Charges in PMDs under Accelerated Aging Conditions. Energies 2019, 12, 151. https://doi.org/10.3390/en12010151

AMA Style

Kim KM, Lee J, Choi SI, Ahn GH, Paik JG, Ryu BT, Kim YH, Won YS. A Combined Study of TEM-EDS/XPS and Molecular Modeling on the Aging of THPP, ZPP, and BKNO3 Explosive Charges in PMDs under Accelerated Aging Conditions. Energies. 2019; 12(1):151. https://doi.org/10.3390/en12010151

Chicago/Turabian Style

Kim, Kyung Min, Junwoo Lee, Sung Il Choi, Gil Hwan Ahn, Jong Gyu Paik, Byung Tae Ryu, Yong Ha Kim, and Yong Sun Won. 2019. "A Combined Study of TEM-EDS/XPS and Molecular Modeling on the Aging of THPP, ZPP, and BKNO3 Explosive Charges in PMDs under Accelerated Aging Conditions" Energies 12, no. 1: 151. https://doi.org/10.3390/en12010151

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop