Next Article in Journal
Linear Optimisation of a Settlement Towards the Energy-Plus House Standard
Next Article in Special Issue
Liquid-Phase Hydrogenation of Maleic Acid over Pd/Al2O3 Catalysts Prepared via Deposition–Precipitation Method
Previous Article in Journal
Igniting Soaring Droplets of Promising Fuel Slurries
Previous Article in Special Issue
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
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Thermodynamic Analysis on the Aging of THPP, ZPP and BKNO3 Explosive Charges in PMDs

1
Department of Chemical Engineering, Pukyong National University, Busan 48547, Korea
2
Defence R&D center, Hanwha Corporation, Daejeon 34068, Korea
3
Agency for Defense Development, Daejeon 34183, Korea
*
Author to whom correspondence should be addressed.
Energies 2019, 12(2), 209; https://doi.org/10.3390/en12020209
Submission received: 19 November 2018 / Revised: 19 December 2018 / Accepted: 9 January 2019 / Published: 10 January 2019

Abstract

:
The aging of explosive charges in pyrotechnic mechanical devices (PMDs) significantly affects the reliability of their performance. Aging decreases the explosive power of explosive charges via pre-oxidation by external sources of oxygen. In this study, the flame temperature was calculated based on the amount of added oxygen source, and the effect of aging was quantified in terms of the performance of explosive charges. Three primary explosive charges (THPP, ZPP, and BKNO3) were selected, and the explosion was assumed to have occurred in a closed adiabatic system. The flame temperature decreased with increasing oxygen addition, suggesting that the aging decreased the pressure exerted by the PMDs. The explosive power decreased in the following order: ZPP >> THPP > BKNO3 (4100 >> 3260 > 3050 K), and it was decreased by aging as follows: ZPP >> BKNO3 > THPP (802 >> 219 > 95 K per mole of added water). As a result, ZPP is appropriate for the generation of the huge thrust of PMDs despite its increased susceptibility to aging, while THPP is superior for reliable PMD performance.

1. Introduction

Pyrotechnic mechanical devices (PMDs) convert the energy of explosive charges into mechanical energy in military and aerospace applications [1]. The precision and reliability of PMDs is very important, and therefore, aged PMDs are periodically scrapped due to unreliable long-term performance of the aged explosive charges in PMDs. The study investigating the aging of explosive charges is thus critical in efforts designed to extend the shelf-life of PMDs and further reduce the military defense budget [2,3]. The three widely used explosive charges in PMDs include the following: BKNO3 (boron potassium nitrate), [4] ZPP (zirconium potassium perchlorate), [5] and THPP (titanium hydride potassium perchlorate) [6]. They occur as mixtures of metallic fuels, oxidizers, and polymer binders [7,8,9]. External shocks, such as electric arcs, generate a large amount of heat instantaneously via rapid oxidation of metal and generation of increased pressure for the propulsion of PMDs [10,11,12]. Meanwhile, external factors such as temperature and humidity affect the properties of explosive charges [13]. Exposure to sources of oxygen such as H2O and O2, in particular, leads to pre-oxidation between the metal and oxidizer and decreases the explosive power of PMDs in the long run. Thus, storage under high humidity to prevent undesirable explosions during the fabrication of PMDs contributes to aging. Although the explosive charges are dried after the fabrication of PMDs, they would have already been exposed to a considerable amount of H2O. In addition, the explosive charges are exposed to air due to imperfect sealing of the chamber of PMDs.
We have already proposed an approach based on the thermodynamic analysis of the flame temperature to investigate the effect of oxygen sources (H2O or O2) on the aging of BKNO3. The rationale of the approach is presented in Figure 1 [14]. Explosive charges are aged by oxygen sources in an open system during the fabrication of PMDs and storage. As the explosion is triggered, the system can be regarded as closed in an instant time lapse following the event. The closed system consists of several sub-processes: an adiabatic process in which the huge energy generated by the oxidation reactions between the metal and the oxidizer is utilized in heating reaction products and the phase transition of condensed matter into gases, the expansion (or increased kinetic energy) of gaseous species for the propulsion of mechanical gadgets in PMDs, and heat loss through the metal casing. The kinetic barriers of primary oxidations and subsequent reactions may be ignored because the huge energy generated by the explosion may overcome the barriers easily. Thermodynamically, the concept of “adiabatic flame temperature” is a good indicator of energy released by an explosion. We also reported that the pressure exerted by PMDs is correlated to the flame temperature of the system [15,16]. The decrease in flame temperature suggests that the resulting gaseous products are less energetic and thereby weakens the propulsion of PMDs.
Meanwhile, we conducted our experimental study on the aging of BKNO3 and ZPP by exposing them to water vapor (H2O) in an accelerated aging condition with high RH (relative humidity) [17,18]. DSC (differential scanning calorimetry)-based thermal analysis confirmed the reaction (or explosive) energy decrease with the duration of exposure to a highly humid atmosphere, and TEM (transmission electron microscopy)-EDS (energy dispersive spectroscopy) specified the gradual growth of an oxide layer on the surface of B and Zr—metallic fuels of BKNO3 and ZPP, respectively. Then, the relative robustness of BKNO3, ZPP, and THPP to aging was reported based on our experimental works and a complementary computational kinetic study [19].
In this study, we adopted the same approach [14] used in our previous study for two other explosive charges (ZPP and THPP). The flame temperature was calculated with respect to the amount of added oxygen source (H2O or O2) to illustrate the sensitivity of explosive charges to thermodynamic aging. Finally, the aging effect of the three primary explosive charges (THPP, ZPP, and BKNO3) was compared.

2. Materials and Methods

Thermodynamic calculations were carried out using the CALPHAD (CALculation of PHAse Diagrams) method using a commercial software known as ‘Thermo-Calc’ [20]. In this method, the elements are defined for a given system initially, and the thermodynamic properties of all the available species comprising defined elements are collected from a database as forms of Gibbs energies. In this study, the SSUB5 substance database was used. The SSUB5 is one of the largest SGTE (Scientific Group Thermodata Europe) databases, containing 5580 pure condensed compounds and gaseous species. The parameters of the mathematical models were evaluated by optimizing the model fit to the collected information. The calculation algorithm for equilibrium analysis was summarized in our previous study by a flowchart [14]. The adiabatic flame temperature was defined as the final temperature of the system raised by the combustion of explosive charges in a closed adiabatic state (Thermo-Calc. Example #22).
To investigate the aging effect of explosive charges by oxygen sources (H2O and O2), the flame temperature was calculated by increasing the amount of oxygen source delivered into pure explosive charges. The calculation was based on a stoichiometric ratio of 1 mole of metal component to the amount of oxidizer. The amount of oxygen element added to the system was then increased from 0.05 to 0.20 mole by an increment of 0.05. In other words, H2O was added by the same amount (0.05 to 0.20 mole by 0.05), while the levels of O2 were halved because it has two moles of oxygen.

3. Results and Discussion

3.1. Aging of THPP Explosive Charge by Oxygen Sources

The explosion reaction of THPP is represented as follows.
TiH2 + 0.75 KClO4 → 0.75 KCl + TiO2 + H2O
The calculation (also denoted as ‘pure THPP’) is thus based on 1 mol of Ti, 2 moles of H, 0.75 mole of K and Cl, and 3 moles of O. The decrease in flame temperature is shown in Figure 2 based on the oxygen amount added to the system. The decrease was more sensitive following exposure to O2 than H2O.
We then investigated the extent of metal pre-oxidation in explosive charges following the addition of H2O and O2. Thus, equilibrium calculations were introduced to alter the composition, as summarized in Table 1. Equilibrium calculations propose the most stable phases under given conditions. For example, Ti exists mostly as TiO2 (s) at low temperatures (such as 298 K), regardless of the amount of added oxygen source. It does not imply spontaneous TiO2 (s) formation at low temperatures. However, the oxidation of metal is extremely favorable even at low temperatures if kinetic barriers are overcome by the energy supply (ignition) or aging. To compare the relative extent of Ti pre-oxidation with H2O and O2 addition, calculations were conducted at temperatures as high as the flame temperatures under the given conditions. The amount of TiO (s) was decreased by the addition of oxygen sources (H2O or O2) due to the oxidation of TiO (s) to TiO2 (s or l). The amount of TiO (s) was, however, generally higher with H2O addition. A reducing environment was created by the presence of three moles of hydrogen: one mole derived from TiH and two moles from H2O. Thus, the pre-oxidation of THPP explosive charge by the extra oxygen source was enhanced by O2 addition. The decrease in flame temperature by pre-oxidation (or the loss of explosive energy) was accordingly more significant with O2 addition.

3.2. Aging of ZPP Explosive Charge by Oxygen Sources

The explosive reaction of ZPP is represented as follows:
Zr + 0.5 KClO4 → ZrO2 + 0.5 KClO
The calculation denoted by ‘pure ZPP’ is based on 1 mol of Zr, 0.05 mole of K and Cl, and 2 moles of O. The flame temperature decrease is shown in Figure 3 with respect to the oxygen amount added to the system. Interestingly enough, the decreasing effect was observed only with H2O addition and remained negligible with O2.
Using the same approach, the compositional changes of ZrOx species were determined by equilibrium calculations as summarized in Table 2. Again, to determine the relative oxidative ability of H2O and O2, calculations were conducted at the flame temperatures under the given conditions. In fact, we found that the oxygen permeated the Zr metal surface with time in accelerated aging experiments under high humidity (not shown here), suggesting that kinetic barriers were overcome by Zr oxidation under aging [15]. However, with increasing exposure to oxygen sources, the amount of ZrO (s) gradually decreased via conversion to ZrO2 (s or l). The amount of ZrO (s) was, however, higher with O2 addition, compared with THPP. It was found that the additional oxygen sources were also consumed to generate radicals such as KO and ClO via O2 addition (see upper part in Table 2). The H and OH radicals were generated by the decomposition of added H2O (see lower part in Table 2). Thus, pre-oxidation of ZPP by the additional oxygen source was promoted by H2O addition. Accordingly, the decrease in flame temperature was steeper with H2O addition. The Zr pre-oxidations with O2 addition probably competed with the oxidation of K and Cl elements.

3.3. Comparison of the Aging Effect Among THPP, ZPP, and BKNO3

As mentioned in the introduction, the most probable source for the aging of explosive charges is water vapor (H2O) during the PMD fabrication under high humidity. Thus, the decrease in flame temperature was compared with H2O addition among the three primary explosive charges—THPP, ZPP, and BKNO3, as shown in Figure 4. The flame temperature ranges were 4100 ± 80 K, 3260 ± 12 K, and 3050 ± 30 K for ZPP, THPP, and BKNO3, respectively, based on the amount of H2O added to 0.2 mole. The explosive power thus decreased in the order of ZPP >> THPP > BKNO3, because higher energy was released at elevated flame temperatures. Meanwhile, the flame temperature decreased (or explosive power decreased) in the following order: 801.90 K, 219.28 K, and 95.28 K per mole of added H2O for ZPP, BKNO3, and THPP, respectively. Although energy was utilized to heat the derivatives of added oxygen sources, resulting in a decrease in the flame temperature, substantial differences in the slope indicate that additional oxygen sources certainly contributed to the thermodynamic changes in the system. In summary, the decrease in the explosive power of ZPP was very sensitive to the aging effect of water vapor. By contrast, THPP and BKNO3 were robust to the aging effects, despite their lower explosive powers.

4. Conclusions

The decrease in flame temperature following the addition of oxygen sources into the system was effectively used to evaluate the aging of explosive charges in PMDs in this and in our previous studies. The aging is mainly attributed to exposure to water vapor during the PMD fabrication under high humidity. With increasing H2O addition to THPP and ZPP explosive charges, the flame temperature decreases (or the explosive power decreased) in general as a result of pre-oxidation of explosive charges. The comparison of THPP, ZPP, and BKNO3 demonstrated that the explosive power followed the order of ZPP >> THPP > BKNO3; however, the explosive power decrease by aging occurred in the order of ZPP >> BKNO3 > THPP. Thus, THPP may represent an appropriate choice for the reliable retention of PMD performance with time, compared with ZPP, which is very sensitive to aging. In other words, it is recommended to increase the ratio of THPP in explosive charge mixtures to extend the shelf-life of PMDs and further reduce the military defense budget.

Author Contributions

Conceptualization, Y.S.W.; methodology, K.M.K., S.I.C., and K.H.E.; software, K.M.K., S.I.C., and K.H.E.; validation, Y.H.K.; formal analysis, K.M.K., S.I.C. and K.H.E.; investigation, K.M.K., S.I.C., and K.H.E.; resources, Y.S.W.; data curation, K.M.K., S.I.C. and K.H.E.; writing—original draft preparation, Y.S.W.; writing—review and editing, Y.S.W.; visualization, K.M.K., S.I.C., and K.H.E.; 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.

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. Paik, J.G.; Ryu, B.T.; Kwon, M. A Study on Acceleration Aging Characteristics of B-KNO3 Igniter. Korean Chem. Eng. Res. 2014, 52, 166–174. [Google Scholar] [CrossRef]
  2. Le, A.Q.; Sun, L.Z.; Miller, T.C. Detectability of Delaminations in Solid Rocket Motors with Embedded Stress Sensors. J. Propul. Power 2013, 29, 299–304. [Google Scholar] [CrossRef]
  3. Hosseini, S.G.; Pourmortazavi, S.M.; Hajimirsadeghi, S.S. Thermal Decomposition of Pyrotechnic Mixtures Containing Sucrose with either Potassium Chlorate or Potassium Perchlorate. Combust. Flame 2005, 141, 322–326. [Google Scholar] [CrossRef]
  4. 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]
  5. Sorensen, D.N.; Quebral, A.P.; Baroody, E.E.; Sanborn, W.B. Investigation of the Thermal Degradation of the Aged Pyrotechnic Titanium Hydride/Potassium Perchlorate. J. Therm. Anal. Calorim. 2006, 85, 151–156. [Google Scholar] [CrossRef]
  6. Lee, J.S.; Lin, L.K.; Lin, C.H.; Chen, P.J.; Huang, C.W.; Chang, S.S. A Study of Zirconium/Potassium Perchlorate Primer Mixture. Thermochim. Acta 1990, 173, 211–218. [Google Scholar] [CrossRef]
  7. Moore, J.E.; Son, S.F.; Sivathanu, Y.R.; Lim, J. Experimental Characteristics of Particle Dynamics within Solid Rocket Motors Environments; Technical Paper (PA #09154); Air Force Research Laboratory: Dayton, OH, USA, 2009. [Google Scholar]
  8. Danali, S.M.; Palaiah, R.S.; Raha, K.C. Developments in Pyrotechnics (Review Paper). Defence Sci. J. 2010, 60, 152–158. [Google Scholar] [CrossRef]
  9. Conkling, J.A.; Mocella, C. Chemistry of Pyrotechnics: Basic Principles and Theory; CRC Press: New York, NY, USA, 2010. [Google Scholar]
  10. Lee, J.R.; Chia, C.C.; Kong, C.W. Review of Pyroshock Wave Measurement and Simulation for Space Systems. Measurement 2012, 45, 631–642. [Google Scholar] [CrossRef]
  11. Babar, Z.; Malik, A.Q. Synthesis of Micro Porous Barium Nitrate with Improved Ignition Reliability as a Reliable Pyrotechnic Oxidant. J. Saudi Chem. Soc. 2014, 18, 707–711. [Google Scholar] [CrossRef]
  12. Kuwahara, T.; Kubota, N. Role of Boron in Burning Rate Augmentation of AP composite propellants. Propellants Explos. Pyrotech. 1989, 14, 43–46. [Google Scholar] [CrossRef]
  13. De Klerk, W.P.C.; Colpa, W.; van Ekeren, P.J. Ageing Studies of Magnesium–Sodium Nitrate Pyrotechnic Compositions. J. Therm. Anal. Calorim. 2006, 85, 203–207. [Google Scholar] [CrossRef]
  14. Eom, K.H.; An, H.Y.; Kim, K.M.; Ahn, G.H.; Paik, J.G.; Ryu, B.T.; Won, Y.S. Equilibrium Analysis on the Aging of a BKNO3 Igniter. J. Nanosci. Nanotechnol. 2017, 17, 7685–7688. [Google Scholar] [CrossRef]
  15. Yang, L.C. Reaction Rate Analysis for Selected Solid-to-Solid-Reaction Pyrotechnic Compositions. In Proceedings of the 48th AIAA/ASME/SAE/ASEE Joint Propulsion Conference & Exhibit, Atlanta, GA, USA, 30 July–1 August 2012. [Google Scholar]
  16. Kong, T.K.; Won, Y.S.; Ryu, B.T.; Ahn, G.H.; Im, D.J. Mathematical Modeling of ZrKClO4 Nano Particle Energy Release. J. Nanosci. Nanotechnol. 2017, 17, 8372–8377. [Google Scholar] [CrossRef]
  17. Lee, J.; Choi, K.; Ryu, S.U.; Ahn, G.H.; Ryu, B.T.; Won, Y.S. Aging Mechanism of Zirconium Potassium Perchlorate Propellant in Pyrotechnic Mechanical Devices. Nanosci. Nanotechnol. Lett. 2018, 10, 735–740. [Google Scholar] [CrossRef]
  18. Lee, J.; Kim, T.; Ryu, S.U.; Choi, K.; Ahn, G.H.; Paik, J.G.; Ryu, B.; Park, T.H.; 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]
  19. 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 2018, 12, 151. [Google Scholar] [CrossRef]
  20. Thermo-Calc Software SSUB5 SGTE Substances Database Version 5.1. Available online: https://www.thermocalc.com/academia/researchers/how-to-cite-thermo-calc-products/ (accessed on 20 March 2016).
Figure 1. Schematic representation of the thermodynamic analysis of the aging of explosive charges.
Figure 1. Schematic representation of the thermodynamic analysis of the aging of explosive charges.
Energies 12 00209 g001
Figure 2. Flame temperature decrease of THPP correlated with the levels of oxygen sources (H2O or O2).
Figure 2. Flame temperature decrease of THPP correlated with the levels of oxygen sources (H2O or O2).
Energies 12 00209 g002
Figure 3. Flame temperature decrease of ZPP correlated with respect to the amount of oxygen sources (H2O or O2) added.
Figure 3. Flame temperature decrease of ZPP correlated with respect to the amount of oxygen sources (H2O or O2) added.
Energies 12 00209 g003
Figure 4. Comparison of the flame temperature decrease with H2O addition among THPP, ZPP, and BKNO3.
Figure 4. Comparison of the flame temperature decrease with H2O addition among THPP, ZPP, and BKNO3.
Energies 12 00209 g004
Table 1. Compositional changes in TiOx species determined by equilibrium calculations at the flame temperatures calculated under given conditions for H2O and O2 addition. The initial composition for equilibrium calculation is 1 mol of Ti, 2 moles of H, 0.75 mole of K and Cl, and 3 moles of O plus added moles of O, and the temperature and pressure for equilibrium calculation are the calculated flame temperatures shown in Figure 1 and Figure 2, respectively.
Table 1. Compositional changes in TiOx species determined by equilibrium calculations at the flame temperatures calculated under given conditions for H2O and O2 addition. The initial composition for equilibrium calculation is 1 mol of Ti, 2 moles of H, 0.75 mole of K and Cl, and 3 moles of O plus added moles of O, and the temperature and pressure for equilibrium calculation are the calculated flame temperatures shown in Figure 1 and Figure 2, respectively.
Added O Amount (mole)Pure THPP0.050.100.150.20
H2O additionTiO2 (s)0.1721 mole0.17160.17120.17090.1707
TiO (s)0.01400.01370.01330.01300.0127
TiO2 (l)0.82140.81340.81510.81580.8163
O2 additionTiO2 (s)0.17210.17290.17340.17370.1738
TiO (s)0.01400.01340.01280.01220.0116
TiO2 (l)0.82140.81340.81350.81380.8143
Table 2. Compositional changes of ZrOx species based on equilibrium calculations at the flame temperature at specific H2O and O2 addition. The initial composition for equilibrium calculation is 1 mol of Zr, 0.05 mole of K and Cl, and 2 moles of O plus added moles of O, and the temperature and pressure for equilibrium calculation are the calculated flame temperatures shown in Figure 1 and Figure 3, respectively.
Table 2. Compositional changes of ZrOx species based on equilibrium calculations at the flame temperature at specific H2O and O2 addition. The initial composition for equilibrium calculation is 1 mol of Zr, 0.05 mole of K and Cl, and 2 moles of O plus added moles of O, and the temperature and pressure for equilibrium calculation are the calculated flame temperatures shown in Figure 1 and Figure 3, respectively.
Added O Amount (Mole)Pure ZPP0.050.100.150.20
H2O additionZrO2 (s)0.1080 mole0.10000.09080.08120.0720
ZrO (s)0.06960.05250.03930.02920.0217
ZrO2 (l)0.82220.84370.86980.88950.9036
KO (rad)0.00780.00990.01170.01330.0146
ClO (rad)0.00040.00050.00050.00060.0007
O2 additionZrO2 (s)0.10800.11370.11710.11890.1198
ZrO (s)0.06960.05970.05260.04710.0457
ZrO2 (l)0.82220.82640.83020.83390.8374
H (rad) 0.07630.14460.20500.2580
OH (rad) 0.01230.02670.04270.0600

Share and Cite

MDPI and ACS Style

Kim, K.M.; Choi, S.I.; Eom, K.H.; Ahn, G.H.; Paik, J.G.; Ryu, B.T.; Kim, Y.H.; Won, Y.S. Thermodynamic Analysis on the Aging of THPP, ZPP and BKNO3 Explosive Charges in PMDs. Energies 2019, 12, 209. https://doi.org/10.3390/en12020209

AMA Style

Kim KM, Choi SI, Eom KH, Ahn GH, Paik JG, Ryu BT, Kim YH, Won YS. Thermodynamic Analysis on the Aging of THPP, ZPP and BKNO3 Explosive Charges in PMDs. Energies. 2019; 12(2):209. https://doi.org/10.3390/en12020209

Chicago/Turabian Style

Kim, Kyung Min, Sung Il Choi, Ki Heon Eom, Gil Hwan Ahn, Jong Gyu Paik, Byung Tae Ryu, Yong Ha Kim, and Yong Sun Won. 2019. "Thermodynamic Analysis on the Aging of THPP, ZPP and BKNO3 Explosive Charges in PMDs" Energies 12, no. 2: 209. https://doi.org/10.3390/en12020209

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