Next Article in Journal
Study of Optical Rectification in Polycrystalline Materials Based on Random Quasi-Phase Matching
Next Article in Special Issue
Synthetic Methods towards Energetic Heterocyclic N-Oxides via Several Cyclization Reactions
Previous Article in Journal
On the Definition of Phase Diagram
Previous Article in Special Issue
Effect of 5-Amino-1H-Tetrazole on Combustion Pyrolysis Characteristics and Kinetics of a Combustion Tear Gas Mixture
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

The Oxidation Process and Methods for Improving Reactivity of Al

National Special Superfine Powder Engineering Research Center of China, School of Chemistry and Chemical Engineering, Nanjing University of Science and Technology, Nanjing 210094, China
*
Author to whom correspondence should be addressed.
Crystals 2022, 12(9), 1187; https://doi.org/10.3390/cryst12091187
Submission received: 4 July 2022 / Revised: 8 August 2022 / Accepted: 10 August 2022 / Published: 24 August 2022
(This article belongs to the Special Issue Advanced Energetic Materials: Testing and Modeling)

Abstract

:
Aluminum (Al) has been widely used in micro-electromechanical systems (MEMS), polymer bonded explosives (PBXs) and solid propellants. Its typical core-shell structure (the inside active Al core and the external alumina (Al2O3) shell) determines its oxidation process, which is mainly influenced by oxidant diffusion, Al2O3 crystal transformation and melt-dispersion of the inside active Al. Consequently, the properties of Al can be controlled by changing these factors. Metastable intermixed composites (MICs), flake Al and nano Al can improve the properties of Al by increasing the diffusion efficiency of the oxidant. Fluorine, Titanium carbide (TiC), and alloy can crack the Al2O3 shell to improve the properties of Al. Furthermore, those materials with good thermal conductivity can increase the heat transferred to the internal active Al, which can also improve the reactivity of Al. Now, the integration of different modification methods is employed to further improve the properties of Al. With the ever-increasing demands on the performance of MEMS, PBXs and solid propellants, Al-based composite materials with high stability during storage and transportation, and high reactivity for usage will become a new research focus in the future.

1. Introduction

Aluminum (Al) is an important solid metal fuel, which has been widely used in connection between crystalline silicon [1], micro-electromechanical systems (MEMS) [2,3,4,5], polymer bonded explosives (PBXs) [6,7,8,9,10,11,12], and solid propellants [13,14,15,16,17,18] to provide energy. As a combustion agent, Al displays obvious advantages, such as low oxidant consumption, high combustion calorific value (31,070 J/g), and high measured specific impulse [19,20,21]. The dense and high melting point (about 2000 °C) alumina (Al2O3) shell will be formed when Al is exposed to air, which can improve the safety of aluminum powder in the process of production, storage, transportation and usage [22,23,24]. However, the oxidation process of Al is also closely related to the Al2O3 shell, which causes the inside active Al to become hindered (it cannot come into contact with oxidation components) [25,26,27,28]. As a result, the ignition and combustion reaction activity of Al is limited by the Al2O3 shell [21,29,30,31]. In addition, the combustion product Al2O3 will wrap onto the surface of the active Al again, which further hinders the combustion chemical reaction, resulting in an incomplete combustion of Al, and reduced energy release efficiency [29].
To solve the problem of the Al2O3 shell limiting oxidation activity and combustion efficiency, the oxidation process of Al has been studied extensively. To describe the ignition and combustion process of Al, some mechanisms, such as the oxidant diffusion mechanism [32,33,34], the Al2O3 crystal transformation mechanism [35,36,37,38,39,40,41], and the melt-dispersion mechanism [42,43,44,45,46], have been proposed. Meanwhile, to improve the activity of Al, some new composite materials, such as Al-based metastable intermixed composites (MICs), flake Al-based composite materials, alloying and so on, have been found and used gradually [12,47,48,49,50,51,52,53,54,55,56,57].
A large number of reviews on Al have been published. Xiang Zhou [58] summarized the synthesis, ignition and combustion modeling, and applications of Al-based nanocomposites. Wei He [59] introduced the preparation and characterization of Al-based MICs. In addition, Xiaoxia Ma [60] focused on the preparation and fundamental properties of the Al-based core–shell structured nanoenergetic. However, they discussed the preparation and characterization methods, and hardly any reviews systematically summarized the relationship between the modification method and the different ignition mechanism of Al during the past decades. This review examines systematically the key influencing factors and the mechanisms during the Al oxidation process, and classified modification methods by the influencing factors of its transformation. Furthermore, it provides ideas for selecting modification methods of Al under different application conditions. What is more, the new trend of Al-based composite materials has been indicated.

2. The Key Influencing Factors and the Reaction Mechanism during the Al Oxidation Process

The Al particle is a typical core-shell structure with active Al wrapped in dense Al2O3. Therefore, the oxidation process of the Al particle is the evolution process of the oxidant, the alumina shell, and the inside active Al. Furthermore, the diffusion of the oxidant, the growth and rupture of the Al2O3 shell, and the melting and gasification of the inside active Al are the key influencing factors of the Al oxidation process.

2.1. Effect of Diffusion of the Oxidant on the Al Oxidation Process

The diffusion of the oxidant is a necessary process for Al oxidation. Zachariah [32,33,34] proposed a three-step oxidation process of Al (oxidant diffusion mechanism) on the basis of a well-known idealized “shrinking core” mechanism for spherical particles, which involves a reaction front moving radially inward separating an unreacted core with a completely reacted ash outer layer. As shown in Figure 1, the diffusion of the oxidant is influenced by three factors: the Al2O3 shell (generated on the surface in air), the Al2O3 ash (the product of oxidation reaction) and the chemical reaction of the oxidant with the inside active Al.
Step 1: Diffusion of the oxidant through the Al2O3 shell to ash (Al2O3 generated by reaction, Al2O3 shell resistance).
Step 2: Diffusion of the oxidant through the ash (Al2O3 generated by reaction) to the surface of the inside active Al (ash resistance).
Step 3: Chemical reaction of the oxidant with the inside active Al at the unreacted core surface (chemical reaction resistance).
Since ash (Al2O3 generated by reaction) forms very rapidly, step 1 (Al2O3 shell resistance) can be ignored. This still leaves steps 2 and 3 from which to deduce the rate-limiting step. Then it is proved by experiments that the reaction of Al is controlled by diffusion. Therefore, the diffusion of the oxidant through the Al2O3 ash to the inside active Al is the rate-limiting step.

2.2. Effect of the Al2O3 Shell on the Al Oxidation Process

The Al2O3 shell on the surface of Al has a crucial effect on the oxidation process of Al. Trunov [35] analyzed the processes of simultaneous growth and phase transformations of Al2O3 during oxidation of the Al particle (Al2O3 crystal transformation mechanism). The process of Al oxidation can be divided into four stages, which can be seen in Figure 2. Firstly, amorphous Al2O3 on the Al surface gradually grows thicker and the reaction rate is controlled by the outward diffusion of Al cations [36]. Secondly, amorphous Al2O3 on the Al surface transforms to γ- Al2O3 when the critical thickness is approached, or when the temperature becomes sufficiently high [37,38,39,40]. Since the density of γ-Al2O3 exceeds that of the amorphous Al2O3, the Al2O3 shell on the Al surface ruptures. In addition, the inside active Al can come into contact with oxide, which greatly increases the reaction rate. With the growth and healing of the γ-Al2O3 shell, the reaction rate decreases significantly. Eventually, a regular polycrystalline layer of γ-Al2O3 forms by the end of the second stage. In stage three, the growth of γ-Al2O3 continues. In the meantime, γ-Al2O3 transforms the crystal phase to δ-Al2O3 or θ-Al2O3. Due to the density of δ-Al2O3, and θ-Al2O3 is similar to that of γ-Al2O3, the shell of Al2O3 will not rupture and the reaction will not change significantly. Additionally, the reaction rate is limited by the inward grain boundary diffusion of oxidant anions in stage three [36,41]. When the stable and denser α-Al2O3 forms by increased temperature, stage three ends. Stage four starts when Al2O3 is completely transformed to α-Al2O3. In stage three, the thickness of the γ-Al2O3 layer decreases, and the oxidation rate increases momentarily. Once most of the oxide layer is transformed to coarse and dense α-Al2O3, the contact between the internal active aluminum and the oxidized components is completely blocked, and the reaction rate decreases rapidly.

2.3. Effect of the Inside Active Al on the Al Oxidation Process

Levitas [42,43,44,45,46] proposed a melt-dispersion mechanism when Al is heated rapidly. As shown in Figure 3, the volume change (6% volume expansion) due to the melting of the inside active Al can make the pressure of the Al2O3 shell reach about 11 GPa. That can cause spallation of the Al2O3 shell and the inside active Al can come into contact with the oxidant. Furthermore, the pressure inside the molten active Al will disperse the active Al into small droplets. Then the oxidation process of the active Al would occur rapidly.
From the above discussion, it can be seen that the diffusion of the oxidant, the crystallization and the growth of the Al2O3 shell, and the melting and dispersion of the internal active Al, all have a crucial impact on the oxidation process of Al. The leading factors will change with the particle size and the heating rate of Al. Therefore, the oxidation process of Al under different conditions is not the same. At a slow heating rate, the diffusion of the oxidant plays a leading role, and the oxidation process of nano Al, which has a large specific surface area, conforms to the three-step oxidation process (proposed by Zachariah [32,33]). At the slow heating rate, the transformation and growth of the Al2O3 shell have the greatest influence on the oxidation process of micron Al, so the oxidation process of micron Al conforms to the Al2O3 shell transformation and growth mechanism (proposed by Trunov [35]). At a fast heating rate, the melting and dispersion of the internal active Al have the greatest influence on the oxidation process of nano Al, so the oxidation process of nano Al conforms to the melting diffusion mechanism (proposed by Levitas [42,43,44,45,46]).
The three crucial factors that determine the oxidation process of Al are: the diffusion of the oxidant, the transformation and growth of the Al2O3 shell, and the melting and dispersion of the internal active Al. As the key factors affecting the oxidation process of Al have been investigated, the oxidation process of Al under different conditions can be accurately described. Then the method of changing the Al oxidation process can be found, and the properties of Al can be improved.

2.4. Effect of the Gas Phase Reaction on the Al Oxidation Process

The gasification temperature of the internal active Al is lower than that of the Al2O3 shell, which leads to the gas phase reaction involving gas phase Al during Al oxidation [61]. As the models of the combustion process of micron Al, which were introduced both by Law [62] and Glassman [61], and later expanded upon by Brooks and Beckstead [63], the gas phase flame can drive the surface reactions of the Al particle. However, it is difficult to accurately characterize the process and products of gas phase reactions. Lynch [64,65] found that volatile suboxides existed during Al combustion, and Tappan [66] found the reaction between the gas phase Al and the Al2O3 shell. Although the gas phase reaction has been found, the reaction process under the different reaction conditions has not been accurately confirmed, and its effect on the oxidation of Al continues to be studied.

3. Methods for Improving Al Properties

To improve the properties of Al, the following methods have been investigated: metastable intermixed composites (MICs, reducing the diffusion of the oxidant), nano Al and flake Al (increasing the specific surface area of Al) for improving diffusion efficiency, fluorine modification [67], Titanium carbide (TiC) modification and alloying for cracking the Al2O3 shell and the modification of good thermal conductivity materials (Ag, graphene) for improving heat transfer efficiency.

3.1. Improving Diffusion Efficiency

3.1.1. MICs

Metastable intermixed composites (MICs) are usually composed of a metal fuel and an oxidizer. The Los Alamos National Laboratory in the US was the first to study the combustion performance of MICs. Aumann [68] prepared the Al/MoO3 nano Al, which bulk energy density can reach 16 kJ/cm3, and the combustion rate is more than 1000 times that of the traditional thermit. The oxidizer of the Al-based MICs is usually metal oxide (bismuth oxide (Bi2O3), copper oxide (CuO), cuprous oxide (Cu2O), molybdenum oxide (MoO3), ferric oxide (Fe2O3), nickel oxide (NiO), tungsten trioxide (WO3)), halate (potassium perchlorate (KClO4), potassium periodate (KIO4), iron iodate (Fe(IO3)3), copper iodate (Cu(IO3)2), bismuth iodate (Bi(IO3)3)), and persulfide (potassium persulfate(K2S2O8)) etc. Al-based MICs effectively reduce the diffusion distance of oxidation, which has significant advantages in volume energy density, ignition, and burning rate [59]. Although, MICs have not been able to achieve efficient, safe and low cost batch preparation, and the reaction process of MICs is difficult to accurately control. Therefore, MICs are difficult to apply in solid propellants and PBXs these days [69].
Ludovic Glavier [70] prepared Al-based MICs by Al, Bi2O3, CuO and MoO3. Among them, as shown in Table 1, Al/Bi2O3 have the shorter the ignition delay time (5 μs), the highest burning rate (420 m/s), and the fastest pressurization rate (5762 kPa/μs). Egor A. Lebedev [71] prepared the MICs layer composed of Al, CuO and Cu2O by electrophoretic deposition, and the maximum heat release of MICs layer is 1954 J/g. Aifeng Jiang [72] prepared Fe2O3/nano Al by ball milling. The initial combustion temperature of Fe2O3/nano is about 600 °C. Ning Wang [73] prepared a Al@NiO core-shell structure composite micro-unit; the ignition temperature can be advanced to 531.5 °C, the heat release is 1410.2 J/g. Chunpei Yu [74] prepared 3D ordered macroporous Al/NiO MICs. Its heat releases up to 2462.27 J/g. Wei He [75] prepared Al/Energetic metal organic frameworks (EMOF) MICs, which can activate Al by eliminating Al2O3 shell and produce metal oxide by decomposing of EMOF. The heat release of Al/EMOF is 3464 J/g; the burning rate is 2.8 m/s.
W. Lee Perry [76] prepared WO3·H2O/Al MICs, which had an energy release of approximately 1.8 MJ/kg at a rate of approximately 215 GW/m2. They found that the enhanced behavior of the hydrated MICs formulation resulted from the reaction of Al with the interstitially bound H2O, which had additional energy release and generated hydrogen gas.
Ahmed Fahd [77] compared and analyzed the thermal behavior of different nanothermite tertiary compositions based on nano Al, graphene oxide (GO), various salt and metallic oxidizers (as shown in Figure 4). The addition of GO enhances the reactivity of nanothermites with both salt and metallic oxidizers by reducing the reaction onset temperature, activation energy and increasing the heat release. For nanothermites with oxidizing salts, the heterogeneous solid–gas reaction mechanism plays a more important role than the condensed phase reactions. In general, nanothermites based on oxidizing salts are more reactive than those with metallic ones, as indicated in both theoretical and experimental data. Among them, the GO/Al/KClO4 nanothermite exhibits the highest heat release (9614 J/g), while the GO/Al/K2S2O8 nanothermite shows the lowest onset temperature and activation energy (380 °C and 105 kJ/mol−1).

3.1.2. Flake Al and Nano Al

Flake Al and nano Al have larger specific surface areas than that of sphere Al, which can improve the diffusion efficiency of oxidation [12,49,50,51,78,79,80,81,82]. However, the large-scale preparation process of nano Al and the effective Al content are still the key problems restricting its application.
Al has good ductility when it is subjected to external force, which would deform firstly. So flake Al is usually prepared by ball milling. As shown in Figure 5, in the process of ball milling, the sphere Al is extruded and sheared by ball milling beads. Under the action of force, sphere Al is changed to cake-like, firstly. In the second stage, the caked Al continues to be subjected to force and becomes flake Al; in the third stage, the flake Al would be broken and becomes smaller flake Al. So different thicknesses of flake Al can be prepared by controlling the parameters of ball milling.
Lei Xiao [51] prepared flake Al powder by ball milling, which could increase the detonation heat of Al-containing explosives by 6.48%. William Wilson [49] found that the flake Al became small particles more easily by shock waves during combustion. Qingming Liu [78] found that the flake Al dust–air mixture could be ignited, and self-sustained detonation by an electric spark of 40 J. DeQi Wang [81] applied the flake Al powder to a solid propellant, which increased the burning rate of the propellant by 5.5%. A.L. Kuhl [50,82] found the flake-Al could increase the impulse of the TNT composite charges by the fast combustion of the flake-Al.
The main methods of preparing nano Al include the electric explosion of wires method [83] and the plasma-arc recondensation [84]. In order to protect the nano Al from oxidation during storage, it is usually necessary to form a coating film on the surface of the nano Al. Deluca L T [21,85] prepared a nano-coated Al with stearic acid (L-ALEX), palmitic acid (P-ALEX), and trihydroperfluoro-undecyl alcohol (F-ALEX), et al. (as shown in Figure 6).
Nano Al is widely used in MICs and solid propellants, etc [86]. Deyun Tang [87] used tannic acid (TA) to coat on nano Al as an interfacial layer to bind with (Fe(IO3)3), copper iodate (Cu(IO3)2), bismuth iodate (Bi(IO3)3), respectively. For the energy release, Fe(IO3)3-based MICs can be increased to 24.1 kJ/cm3 (14.5% higher), whereas the Cu(IO3)2-based MICs to 22.8 kJ/cm3 (19.4%), and Bi(IO3)3-based MICs to 20.2 kJ/cm3 (3.1%). Deluca L [85] found that 100–200 nm Al can clearly increase the burning rate of the propellant.
Nano Al has high reactivity, which also produces some problems such as a high safety risk, a low content of active aluminum and sintering during combustion. Uncoated nano Al is classified by the International Air Transport Association as a highly flammable solid [88]. Weismiller [52] researched the effect of the particle size of Al and the oxidant on the properties of Al/CuO and Al/MoO3. As shown in Table 2, the properties of Al/CuO and Al/MoO3 are greatly improved due to the rapid decomposition of nano oxidants. However, due to the high Al2O3 content of nano Al, the nano Al does not show theoretical property advantages.
Zachariah [89] found that the combustion product of nanothermites has two distinct populations of particles, as shown in Figure 7. The large particles include aluminum, oxidant, and reduced metal while the nano-sized particle is composed of reduced metal/metal oxide. As such large particles cannot be formed from the vapor phase condensation during the available transit time to the substrate, they must be formed in the condensed state as molten material.

3.2. Cracking the Al2O3 Shell

3.2.1. Fluorine Modification

Fluoropolymers and fluorides can crack the Al2O3 shell of Al by the reaction between fluorine and Al2O3 shell. The curve of DSC will appear as a small exothermic peak before the main exothermic peak, caused by oxidation of Al. This phenomenon is known as preignition reaction (PIR), which is found by Osborne [90], and is verified by Zachariah by a quadrupole mass spectrometer and TG-DSC-MS coupling techniques [91]. The properties of the fluorine-modified Al are greatly improved due to the crack of the Al2O3 shell by PIR. Fluorine-modified Al can be used in MEMS and PBXs, but compatibility of fluoride with a propellant system is a key problem to realize its application in the propellant system.
Siva K. Valluri [92] prepared composite micro-units containing Al and NiF2 by reaction inhibition ball milling, which shortened the ignition delay time and improved the combustion efficiency of Al powder. Sergey Matveev [48] prepared Al/BiF3, which could produce 3200 K temperatures. Aifeng Jiang [72] prepared FeF3/Al@vinyltrimethoxysilane by ball milling, which had a large specific surface area (26.33 m2/g) and could be well-preserved from the air atmosphere and water. Additionally, the maximum heat release of FeF3/nano Al@vinyltrimethoxysilane can go up to 12,852 J/g.
Jena McCollum [93] composited Al with perfluoropolyethylene to advance the ignition temperature of Al to about 330 °C. Dong Won Kim [94] coated polyvinylidene fluoride (PVDF) onto the surface of Al (Al@PVDF), and the heat release of Al@PVDF at 900 °C was 11,040 J/g. Jun Wang from the Institute of Chemical Materials [95] prepared Carbon nanotubes/polytetrafluoroethylene (PTFE)/Al nanocomposites, which had a lower initial reaction temperature (reduced by 80 °C) and a shorter ignition delay time (reduced by 0.21 ms). Xiang Ke [96] coated PVDF onto the surface of Al to prepare reactive film materials, which appear PIR at 430 °C.

3.2.2. TiC Modification

The reaction (as shown in Equation (1)) of Al2O3 and Titanium carbide (TiC) has been found during the preparation process of ceramics, which also can generate gas [97,98,99]. If this reaction can take place before the oxidation reaction of Al, the Al2O3 shell can be removed effectively, and the generated gas can also break the Al into small particles under high pressure. DeQi Wang [56] prepared the thick flake Al/TiC, which can crack the Al2O3 shell in-situ before 633 °C. As shown in Figure 8, the reaction of TiC with the Al2O3 shell on the Al surface to produce gas has been experimentally confirmed. The heat release of it can reach 21,419 J/g, and this powdered material has good application prospects in solid propellants and polymer bonded explosives (PBXs). This provides a new research idea for solving the limitation of the Al2O3 shell on the Al oxidation process. However, the application of it in PBXs and the propellant system needs further exploration.
Al2O3(s) + TiC(s) → Al2O(g) + TiO(g) + CO(g)

3.2.3. Alloying

The oxide film on the surface of the Al-alloy particle may not be pure Al2O3 shell, which may have higher transmittance to oxidation. Therefore, the Al-alloy can be oxidized at lower temperatures. The oxidation product is not a dense structure like the Al2O3 shell. Accordingly, alloying is one of the effective ways to destroy the structure of the Al2O3 shell. Some metals (such as Li, Mg and Zr et al.) can be oxidized at a lower temperature than Al and can provide activation energy for the oxidation of Al, so that oxidation activity of Al-alloy is higher [53,54,55]. The preparation method of the high-density alloy is the key factor restricting the application of this modification method.
Hao Fu [55] prepared the Al- europium (Eu) alloy, which could be oxidized at 1065 °C and the heat release of the Al-Eu alloy is almost 5 times that of pure Al. Aobo Hu [53,54] prepared the Al-Zr alloy and the Al-W alloy. New alloy phases, ZrAl3, formed in the Al-Zr alloy, which changed the oxidation process (as shown in Figure 9a). Therefore, the Al-Zr alloy can complete combustion under high pressure. Furthermore, the Al-W alloy is almost completely oxidized in air at 1500 °C. In addition, as shown in Figure 9b,c, the gas product, WO3, increases the contact area between the active Al and the oxidant. As a result, the properties of the Al-W alloy have been improved greatly. Fahad Noor [100] prepared the Al-Cu alloy, which had the lower ignition temperature (565 °C).

3.3. Accelerating the Melting of the inside Active Al

The composite made of high thermal conductivity material and Al can increase the rate of Al absorbing heat, which can accelerate the melting of active Al inside and break through the limitation of Al2O3 shell. Consequently, the inside active Al can make contact with the oxidant and the reactivity of Al will be improved. However, as with MICs, how to achieve an efficient, safe and low cost batch preparation is the key problem when it is used in PBXs and propellants.
Jinpeng Shen [101] investigated the effects of nano-Ag on the combustion wave behavior of Al/CuO. The experimental observations confirm that the presence of nano-Ag particles improves the heat transfer efficiency, and the first exothermic peak temperature decreases from 607.8 °C to 567.6 °C.
Ahmed Fahd [77] compared and analyzed the thermal behavior of different nanothermite tertiary compositions based on Al, graphene oxide (GO), and various salt and metallic oxidizers. The addition of GO enhances the reactivity of nanothermites with both salt and metallic oxidizers by reducing the reaction onset temperature, activation energy and increasing the heat release.

4. New Trend of Improving Reactivity of Aluminum Powder

With the deepening of research on the Al oxidation process, various factors restricting Al oxidation have been found, gradually. The methods of improving the reactivity of Al have been integrated, which can simultaneously change multiple conditions in the oxidation process of aluminum powder to improve the properties of Al from multiple perspectives.
Jena McCollum [102] investigated the reactivity of Al/MoO3@perfluoropolyethers (PFPE) and Al/CuO@PFPE. As shown in Table 3, fluorine−Al-based surface reaction can improve the reactivity of Al/MoO3. However, the reactivity of CuO reduces when the PFPE concentration is increased. Lei Xiao [47] successfully assembled Al/CuO/PVDF/RDX, and the combustion properties of microspheres is mainly affected by the content of RDX. Aifeng Jiang [72] prepared Fe2O3/nano Al and FeF3/nano Al@vinyltrimethoxysilane by ball milling. The initial combustion temperature of Fe2O3/nano is about 600 °C. FeF3/nano Al@vinyltrimethoxysilane had a large specific surface area (26.33 m2/g) and could be well-preserved from air atmosphere and water. Furthermore, the maximum heat release of FeF3/nano Al@vinyltrimethoxysilane can go up to 12,852 J/g.
In the meantime, the safety of Al during storage, transport and usage also should be addressed when improving the properties of Al. The design and preparation of Al composites with high stability during storage and transportation, and high reactivity during usage, will become a research focus in the future.
DeQi Wang [56] designed flake Al/TiC, with TiC embedded on the surface of flake Al, which retains the Al2O3 shell to keep Al stable during storage and transportation, and cracks the Al2O3 shell to improve the reaction activity and combustion efficiency by the reaction of Al2O3 and TiC before 633 °C. In addition, the heat release of flake Al/TiC is 21,419 J/g at 985.6 °C.

5. Summary and Prospect

With increasing research, the oxidation process of Al has been gradually revealed, which is closely related to the reaction conditions. The oxidation process of nano Al (large specific surface area) is mainly restricted by oxidant diffusion under a slow heating rate, and conforms to a three-step oxidation process, proposed by Zachariah. The Al2O3 shell has the greatest influence on the micron Al oxidation process and Al2O3 shell transformation, and the growth mechanism can well explain this process, which was proposed by Trunov under the slow heating rate. The melting and dispersion of the internal active Al has a great impact on the oxidation process of nano Al under a fast heating rate, which can be described by the melting diffusion mechanism proposed by Levitas.
Therefore, to improve the properties of Al, the key influencing factors of the Al oxidation process need to be changed. MICs, flake Al and nano Al can improve the diffusion efficiency of the oxidant during Al oxidation, which can influence the properties of Al under a slow heating rate. Since the Al2O3 shell can be cracked by fluorine, TiC, and alloy, the properties of Al under a fast heating rate can be changed by them. Those materials with good thermal conductivity can increase the heat transferred to the internal active Al, so the properties of Al under a fast heating rate can be improved by good thermal conductivity materials modification.
In order to improve the properties of Al more comprehensively, an integration of different modification methods has been employed, such as fluoride coated nano Al-based MICs, flake Al/TiC, various material-covered Al and so on. Furthermore, the safety of Al during storage, transport and usage also needs to be addressed when improving the properties of Al. The design and preparation of Al composites with high stability during storage and transportation, and high reactivity during usage, will become a research focus in the future.

Author Contributions

Conceptualization, D.W.; formal analysis, D.W.; investigation, G.X. and W.D.; writing—original draft preparation, T.T.; writing—review and editing, S.L.; supervision, F.L.; project administration, J.L. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Sui, H.; Huda, N.; Shen, Z.; Wen, J.Z. Al–NiO energetic composites as heat source for joining silicon wafer. J. Mater. Processing Technol. 2020, 279, 116572. [Google Scholar] [CrossRef]
  2. Nagarjuna, Y.; Lin, J.C.; Wang, S.C.; Hsiao, W.T.; Hsiao, Y.J. AZO-Based ZnO Nanosheet MEMS Sensor with Different Al Concentrations for Enhanced H2S Gas Sensing. Nanomaterials 2021, 11, 3377. [Google Scholar] [CrossRef]
  3. Kabra, S.; Gharde, S.; Gore, P.M.; Jain, S.; Khire, V.H.; Kandasubramanian, B. Recent trends in nanothermites: Fabrication, characteristics and applications. Nano Express 2020, 1, 032001. [Google Scholar] [CrossRef]
  4. Guo, X.; Sun, Q.; Liang, T.; Giwa, A.S. Controllable Electrically Guided Nano-Al/MoO3 Energetic-Film Formation on a Semiconductor Bridge with High Reactivity and Combustion Performance. Nanomaterials 2020, 10, 955. [Google Scholar] [CrossRef]
  5. Ke, X.; Zhou, X.; Gao, H.; Hao, G.; Xiao, L.; Chen, T.; Liu, J.; Jiang, W. Surface functionalized core/shell structured CuO/Al nanothermite with long-term storage stability and steady combustion performance. Mater. Des. 2018, 140, 179–187. [Google Scholar] [CrossRef]
  6. Cheng, W.; Mu, J.; Li, K.; Xie, Z.; Zhang, P.; An, C.; Ye, B.; Wang, J. Evolution of HTPB/RDX/Al/DOA mixed explosives with 90% solid loading in resonance acoustic mixing process. J. Energetic Mater. 2021, 1–20. [Google Scholar] [CrossRef]
  7. Warren, A.D.; Lawrence, G.W.; Jouet, R.J.; Elert, M.; Furnish, M.D.; Chau, R.; Holmes, N.; Nguyen, J. Investigation of Formulations Containing Perfluorocoated Oxide-Free Nano-Aluminum; American Institute of Physics: New York, NY, USA, 2007; pp. 1018–1021. [Google Scholar]
  8. Hardin, D.B.; Zhou, M.; Horie, Y. Ignition Behavior of an Aluminum-Bonded Explosive (ABX). In AIP Conference Proceedings; AIP Publishing LLC: Melville, NY, USA, 2017; p. 040013. [Google Scholar]
  9. Guo, X.L.; Cao, W.; Duan, Y.L.; Han, Y.; Ran, J.L.; Lu, X.J. Experimental study and numerical simulation of the corner turning of TATB based and CL-20 based polymer bonded explosives. Combust. Explos. Shock. Waves 2016, 52, 719–726. [Google Scholar] [CrossRef]
  10. Elia, T.; Baudin, G.; Genetier, M.; Lefrançois, A.; Osmont, A.; Catoire, L. Shock to Detonation Transition of Plastic Bonded Aluminized Explosives. Propellants Explos. Pyrotech. 2020, 45, 554–567. [Google Scholar] [CrossRef]
  11. Yeager, J.D.; Bowden, P.R.; Guildenbecher, D.R.; Olles, J.D. Characterization of hypervelocity metal fragments for explosive initiation. J. Appl. Phys. 2017, 122, 035901. [Google Scholar] [CrossRef]
  12. Rumchik, C.G.; Jordan, J.L.; Elert, M.; Furnish, M.D.; Chau, R.; Holmes, N.; Nguyen, J. Effect of Aluminum Particle Size on the High Strain Rate Properties of Pressed Aluminized Explosives. Shock. Compress. Condens. Matter 2007, 955, 795–798. [Google Scholar]
  13. Ao, W.; Wen, Z.; Liu, L.; Wang, Y.; Zhang, Y.; Liu, P.; Qin, Z.; Li, L.K.B. Combustion and agglomeration characteristics of aluminized propellants containing Al/CuO/PVDF metastable intermolecular composites: A highly adjustable functional catalyst. Combust. Flame 2022, 241, 112110. [Google Scholar] [CrossRef]
  14. Ji, J.; Zhu, W. Thermal decomposition of core–shell structured HMX@Al nanoparticle simulated by reactive molecular dynamics. Comput. Mater. Sci. 2022, 209, 111405. [Google Scholar] [CrossRef]
  15. Chalghoum, F.; Trache, D.; Benziane, M.; Chelouche, S. Effect of complex metal hydride on the thermal decomposition behavior of AP/HTPB-based aluminized solid rocket propellant. J. Therm. Anal. Calorim. 2022, 1–28. [Google Scholar] [CrossRef]
  16. Tang, W.; Yang, R.; Zeng, T.; Li, J.; Hu, J.; Zhou, X.; Jiang, E.; Zhang, Y. Positive effects of organic fluoride on reduction of slag accumulation in static testing of solid rocket motors of different diameters. Acta Astronaut. 2022, 194, 277–285. [Google Scholar] [CrossRef]
  17. Wang, D.; Guo, X.; Wang, Z.; Wang, S.; Wu, C.; Zhao, W.; Fang, H. On the Promotion of TKX-50 Thermal Activity with Aluminum. Propellants Explos. Pyrotech. 2022, 47, e202200012. [Google Scholar] [CrossRef]
  18. Tian, H.; Wang, Z.; Guo, Z.; Yu, R.; Cai, G.; Zhang, Y. Effect of metal and metalloid solid-fuel additives on performance and nozzle ablation in a hydroxy-terminated polybutadiene based hybrid rocket motor. Aerosp. Sci. Technol. 2022, 123, 107493. [Google Scholar] [CrossRef]
  19. Maggi, F.; Gariani, G.; Galfetti, L.; DeLuca, L.T. Theoretical analysis of hydrides in solid and hybrid rocket propulsion. Int. J. Hydrog. Energy 2012, 37, 1760–1769. [Google Scholar] [CrossRef]
  20. Yavor, Y.; Rosenband, V.; Gany, A. Reduced agglomeration in solid propellants containing porous aluminum. Proc. Inst. Mech. Eng. Part G J. Aerosp. Eng. 2014, 228, 1857–1862. [Google Scholar]
  21. Deluca, L.T.; Shimada, T.; Sinditskii, V.P.; Calabro, M. Chemical Rocket Propulsion—A Comprehensive Survey of Energetic Materials; Springer International Publishing: Cham, Switzerland, 2017. [Google Scholar]
  22. Joshi, N.; Mathur, N.; Mane, T.; Sundaram, D. Size effect on melting temperatures of alumina nanocrystals: Molecular dynamics simulations and thermodynamic modeling. Comput. Mater. Sci. 2018, 145, 140–153. [Google Scholar] [CrossRef]
  23. Nazarenko, O.B.; Amelkovich, Y.A.; Sechin, A.I. Characterization of aluminum nanopowders after long-term storage. Appl. Surf. Sci. 2014, 321, 475–480. [Google Scholar] [CrossRef]
  24. Jin, X.; Li, S.; Yang, Y.; Yang, Y.; Huang, X. Effect of Heating Rate on Ignition Characteristics of Newly Prepared and Aged Aluminum Nanoparticles. Propellants Explos. Pyrotech. 2020, 45, 1428–1435. [Google Scholar] [CrossRef]
  25. Dreizin, E.L.; Allen, D.J.; Glumac, N.G. Depression of melting point for protective aluminum oxide films. Chem. Phys. Lett. 2015, 618, 63–65. [Google Scholar] [CrossRef]
  26. Tang, Y.; Kong, C.; Zong, Y.; Li, S.; Zhuo, J.; Yao, Q. Combustion of aluminum nanoparticle agglomerates: From mild oxidation to microexplosion. Proc. Combust. Inst. 2017, 36, 2325–2332. [Google Scholar] [CrossRef]
  27. Sundaram, D.S.; Puri, P.; Yang, V. A general theory of ignition and combustion of nano- and micron-sized aluminum particles. Combust. Flame 2016, 169, 94–109. [Google Scholar] [CrossRef]
  28. Wu, B.; Wu, F.; Zhu, Y.; He, A.; Wang, P.; Wu, H. Fast reaction of aluminum nanoparticles promoted by oxide shell. J. Appl. Phys. 2019, 126, 144305. [Google Scholar] [CrossRef]
  29. Braconnier, A.; Chauveau, C.; Halter, F.; Gallier, S. Experimental investigation of the aluminum combustion in different O2 oxidizing mixtures: Effect of the diluent gases. Exp. Therm. Fluid Sci. 2020, 117, 110110. [Google Scholar] [CrossRef]
  30. Liang, L.; Guo, X.; Liao, X.; Chang, Z. Improve the interfacial adhesion, corrosion resistance and combustion properties of aluminum powder by modification of nickel and dopamine. Appl. Surf. Sci. 2020, 508, 144790. [Google Scholar] [CrossRef]
  31. Liang, D.; Xiao, R.; Liu, J.; Wang, Y. Ignition and heterogeneous combustion of aluminum boride and boron–aluminum blend. Aerosp. Sci. Technol. 2019, 84, 1081–1091. [Google Scholar] [CrossRef]
  32. Rai, A.; Park, K.; Zhou, L.; Zachariah, M.R. Understanding the mechanism of aluminium nanoparticle oxidation. Combust. Theory Model. 2006, 10, 843–859. [Google Scholar] [CrossRef]
  33. Park, K.; Lee, D.; Rai, A.; Mukherjee, D.; Zachariah, M.R. Size-Resolved Kinetic Measurements of Aluminum Nanoparticle Oxidation with Single Particle Mass Spectrometry. J. Phys. Chem. B 2005, 109, 7290–7299. [Google Scholar] [CrossRef]
  34. Chowdhury, S.; Sullivan, K.; Piekiel, N.; Zhou, L.; Zachariah, M.R. Diffusive vs Explosive Reaction at the Nanoscale. J. Phys. Chem. C 2010, 114, 9191–9195. [Google Scholar] [CrossRef]
  35. Trunov, M.A.; Schoenitz, M.; Dreizin, E.L. Effect of polymorphic phase transformations in alumina layer on ignition of aluminium particles. Combust. Theory Model. 2006, 10, 603–623. [Google Scholar] [CrossRef]
  36. Jeurgens, L.P.H.; Sloof, W.G.; Tichelaar, F.D.; Mittemeijer, E.J. Growth kinetics and mechanisms of aluminum-oxide films formed by thermal oxidation of aluminum. J. Appl. Phys. 2002, 92, 1649–1656. [Google Scholar] [CrossRef]
  37. Jeurgens, L.P.H.; Sloof, W.G.; Tichelaar, F.D.; Mittemeijer, E.J. Thermodynamic stability of amorphous oxide films on metals: Application to aluminum oxide films on aluminum substrates. Phys. Rev. B 2000, 62, 4707–4719. [Google Scholar] [CrossRef] [Green Version]
  38. Jeurgens, L.P.H.; Sloof, W.G.; Tichelaar, F.D.; Mittemeijer, E.J. Structure and morphology of aluminium-oxide films formed by thermal oxidation of aluminium. Thin Solid Film. 2002, 418, 89–101. [Google Scholar] [CrossRef]
  39. Levin, I.; Brandon, D. Metastable Alumina Polymorphs: Crystal Structures and Transition Sequences. J. Am. Ceram. Soc. 1998, 81, 1995–2012. [Google Scholar] [CrossRef]
  40. Dwivedi, R.K.; Gowda, G. Thermal stability of aluminium oxides prepared from gel. J. Mater. Sci. Lett. 1985, 4, 331–334. [Google Scholar] [CrossRef]
  41. Ruano, O.A.; Wadsworth, J.; Sherby, O.D. Deformation of fine-grained alumina by grain boundary sliding accommodated by slip. Acta Mater. 2003, 51, 3617–3634. [Google Scholar] [CrossRef]
  42. Levitas, V.I.; Asay, B.W.; Son, S.F.; Pantoya, M. Melt dispersion mechanism for fast reaction of nanothermites. Appl. Phys. Lett. 2006, 89, 071909. [Google Scholar] [CrossRef]
  43. Levitas, V.I.; Pantoya, M.L.; Chauhan, G.; Rivero, I. Effect of the Alumina Shell on the Melting Temperature Depression for Aluminum Nanoparticles. J. Phys. Chem. C 2009, 113, 14088–14096. [Google Scholar] [CrossRef] [Green Version]
  44. Levitas, V.I.; Pantoya, M.L.; Dean, S. Melt dispersion mechanism for fast reaction of aluminum nano- and micron-scale particles: Flame propagation and SEM studies. Combust. Flame 2014, 161, 1668–1677. [Google Scholar] [CrossRef]
  45. Levitas, V.I.; Pantoya, M.L.; Dikici, B. Melt dispersion versus diffusive oxidation mechanism for aluminum nanoparticles: Critical experiments and controlling parameters. Appl. Phys. Lett. 2008, 92, 011921. [Google Scholar] [CrossRef] [Green Version]
  46. Levitas, V. Burn time of aluminum nanoparticles: Strong effect of the heating rate and melt-dispersion mechanism. Combust. Flame 2009, 156, 543–546. [Google Scholar] [CrossRef]
  47. Xiao, L.; Zhao, L.; Ke, X.; Zhang, T.; Hao, G.; Hu, Y.; Zhang, G.; Guo, H.; Jiang, W. Energetic metastable Al/CuO/PVDF/RDX microspheres with enhanced combustion performance. Chem. Eng. Sci. 2021, 231, 116302. [Google Scholar] [CrossRef]
  48. Matveev, S.; Dlott, D.D.; Valluri, S.K.; Mursalat, M.; Dreizin, E.L. Fast energy release from reactive materials under shock compression. Appl. Phys. Lett. 2021, 118, 101902. [Google Scholar] [CrossRef]
  49. Yoshinaka, A.; Zhang, F.; Wilson, W.; Elert, M.; Furnish, M.D.; Chau, R.; Holmes, N.; Nguyen, J. Effect of Shock Compression on Aluminum Particles in Condensed Media; American Institute of Physics: New York, NY, USA, 2008; pp. 1057–1060. [Google Scholar]
  50. Kuhl, A.L.; Neuwald, P.; Reichenbach, H. Effectiveness of combustion of shock-dispersed fuels in calorimeters of various volumes. Combust. Explos. Shock. Waves 2006, 42, 731–734. [Google Scholar] [CrossRef]
  51. Xiao, L.; Liu, J.; Hao, G.; Ke, X.; Chen, T.; Gao, H.; Rong, Y.-b.; Jin, C.-s.; Li, J.-l.; Jiang, W. Preparation and study of ultrafine flake-aluminum with high reactivity. Def. Technol. 2017, 13, 234–238. [Google Scholar] [CrossRef]
  52. Weismiller, M.R.; Malchi, J.Y.; Lee, J.G.; Yetter, R.A.; Foley, T.J. Effects of fuel and oxidizer particle dimensions on the propagation of aluminum containing thermites. Proc. Combust. Inst. 2011, 33, 1989–1996. [Google Scholar] [CrossRef]
  53. Hu, A.; Zou, H.; Shi, W.; Pang, A.m.; Cai, S. Preparation, Microstructure and Thermal Property of ZrAl3/Al Composite Fuels. Propellants Explos. Pyrotech. 2019, 44, 1454–1465. [Google Scholar] [CrossRef]
  54. Hu, A.; Cai, S. Research on the novel Al–W alloy powder with high volumetric combustion enthalpy. J. Mater. Res. Technol. 2021, 13, 311–320. [Google Scholar] [CrossRef]
  55. Fu, H.; Zou, H.; Cai, S.-z. The role of microstructure refinement in improving the thermal behavior of gas atomized Al-Eu alloy powder. Adv. Powder Technol. 2016, 27, 1898–1904. [Google Scholar] [CrossRef]
  56. Wang, D.; Cao, X.; Liu, J.; Zhang, Z.; Jin, X.; Gao, J.; Yu, H.; Sun, S.; Li, F. TF-Al/TiC highly reactive composite particle for application potential in solid propellants. Chem. Eng. J. 2021, 425, 130674. [Google Scholar] [CrossRef]
  57. Yao, E.; Zhao, N.; Qin, Z.; Ma, H.; Li, H.; Xu, S.; An, T.; Yi, J.; Zhao, F. Thermal Decomposition Behavior and Thermal Safety of Nitrocellulose with Different Shape CuO and Al/CuO Nanothermites. Nanomaterials 2020, 10, 725. [Google Scholar] [CrossRef] [Green Version]
  58. Zhou, X.; Torabi, M.; Lu, J.; Shen, R.; Zhang, K. Nanostructured energetic composites: Synthesis, ignition/combustion modeling, and applications. ACS Appl. Mater. Interfaces 2014, 6, 3058–3074. [Google Scholar] [CrossRef]
  59. He, W.; Liu, P.-J.; He, G.-Q.; Gozin, M.; Yan, Q.-L. Highly Reactive Metastable Intermixed Composites (MICs): Preparation and Characterization. Adv. Mater. 2018, 30, e1706293. [Google Scholar] [CrossRef]
  60. Ma, X.; Li, Y.; Hussain, I.; Shen, R.; Yang, G.; Zhang, K. Core-Shell Structured Nanoenergetic Materials: Preparation and Fundamental Properties. Adv. Mater. 2020, 32, e2001291. [Google Scholar] [CrossRef]
  61. Irvin, G.; Yetter, R.A. Combustion, 4th ed.; Academic Press: San Diego, CA, USA, 2008. [Google Scholar]
  62. Law, C.K. A Simplified Theoretical Model for the Vapor-Phase Combustion of Metal Particles. Combust. Sci. Technol. 1973, 7, 197–212. [Google Scholar] [CrossRef]
  63. Brooks, K.P.; Beckstead, M.W. Dynamics of aluminum combustion. J. Propuls. Power 1995, 11, 769–780. [Google Scholar] [CrossRef]
  64. Lynch, P.; Krier, H.; Glumac, N. A correlation for burn time of aluminum particles in the transition regime. Proc. Combust. Inst. 2009, 32, 1887–1893. [Google Scholar] [CrossRef]
  65. Lynch, P.; Fiore, G.; Krier, H.; Glumac, N. Gas-Phase Reaction in Nanoaluminum Combustion. Combust. Sci. Technol. 2010, 182, 842–857. [Google Scholar] [CrossRef]
  66. Tappan, B.C.; Dirmyer, M.R.; Risha, G.A. Evidence of a kinetic isotope effect in nanoaluminum and water combustion. Angew. Chem. Int. Ed. Engl. 2014, 53, 9218–9221. [Google Scholar] [CrossRef]
  67. Ji, Y.; Sun, Y.; Zhu, B.; Liu, J.; Wu, Y. Calcium fluoride promoting the combustion of aluminum powder. Energy 2022, 250, 123772. [Google Scholar] [CrossRef]
  68. Aumann, C.E.; Skofronick, G.L.; Martin, J.A. Oxidation behavior of aluminum nanopowders. J. Vac. Sci. Technol. B Microelectron. Nanometer Struct. 1995, 13. [Google Scholar] [CrossRef]
  69. Zaky, M.; Elbeih, A.; Elshenawy, T. Review of Nano-thermites; a Pathway to Enhanced Energetic Materials. Cent. Eur. J. Energetic Mater. 2021, 18, 63–85. [Google Scholar] [CrossRef]
  70. Glavier, L.; Taton, G.; Ducéré, J.-M.; Baijot, V.; Pinon, S.; Calais, T.; Estève, A.; Djafari Rouhani, M.; Rossi, C. Nanoenergetics as pressure generator for nontoxic impact primers: Comparison of Al/Bi2O3, Al/CuO, Al/MoO3 nanothermites and Al/PTFE. Combust. Flame 2015, 162, 1813–1820. [Google Scholar] [CrossRef]
  71. Lebedev, E.A.; Sorokina, L.I.; Trifonov, A.Y.; Ryazanov, R.M.; Pereverzeva, S.Y.; Gavrilov, S.A.; Gromov, D.G. Influence of Composition on Energetic Properties of Copper Oxide—Aluminum Powder Nanothermite Materials Formed by Electrophoretic Deposition. Propellants Explos. Pyrotech. 2021, 47, e202100292. [Google Scholar] [CrossRef]
  72. Jiang, A.; Xia, D.; Li, M.; Qiang, L.; Fan, R.; Lin, K.; Yang, Y. Ball Milling Produced FeF3-Containing Nanothermites: Investigations of Its Thermal and Inflaming Properties. ChemistrySelect 2019, 4, 12662–12667. [Google Scholar] [CrossRef]
  73. Wang, N.; Hu, Y.; Ke, X.; Xiao, L.; Zhou, X.; Peng, S.; Hao, G.; Jiang, W. Enhanced-absorption template method for preparation of double-shell NiO hollow nanospheres with controllable particle size for nanothermite application. Chem. Eng. J. 2020, 379, 122330. [Google Scholar] [CrossRef]
  74. Yu, C.; Zhang, W.; Shen, R.; Xu, X.; Cheng, J.; Ye, J.; Qin, Z.; Chao, Y. 3D ordered macroporous NiO/Al nanothermite film with significantly improved higher heat output, lower ignition temperature and less gas production. Mater. Des. 2016, 110, 304–310. [Google Scholar] [CrossRef] [Green Version]
  75. He, W.; Li, Z.-H.; Chen, S.; Yang, G.; Yang, Z.; Liu, P.-J.; Yan, Q.-L. Energetic metastable n-Al@PVDF/EMOF composite nanofibers with improved combustion performances. Chem. Eng. J. 2020, 383, 123146. [Google Scholar] [CrossRef]
  76. Lee Perry, W.; Tappan, B.C.; Reardon, B.L.; Sanders, V.E.; Son, S.F. Energy release characteristics of the nanoscale aluminum-tungsten oxide hydrate metastable intermolecular composite. J. Appl. Phys. 2007, 101, 064313. [Google Scholar] [CrossRef] [Green Version]
  77. Fahd, A.; Dubois, C.; Chaouki, J.; Wen, J.Z. Combustion behaviour and reaction kinetics of GO/Al/oxidizing salts ternary nanothermites. J. Therm. Anal. Calorim. 2022, 1–13. [Google Scholar] [CrossRef]
  78. Liu, Q.; Li, X.; Bai, C. Deflagration to detonation transition in aluminum dust–air mixture under weak ignition condition. Combust. Flame 2009, 156, 914–921. [Google Scholar] [CrossRef]
  79. Antipina, S.A.; Zmanovskii, S.V.; Gromov, A.A.; Teipel, U. Air and water oxidation of aluminum flake particles. Powder Technol. 2017, 307, 184–189. [Google Scholar] [CrossRef]
  80. Houim, R.W.; Boyd, E. Combustion of aluminum flakes in the post-flame zone of a hencken burner. Int. J. Energetic Mater. Chem. Propuls. 2008, 7, 55–71. [Google Scholar] [CrossRef]
  81. Wang, D.Q.; Yu, H.M.; Liu, J.; Li, F.S.; Jin, X.X.; Zheng, S.J.; Zheng, T.T.; Li, Y.; Zhang, Z.J.; Li, D.; et al. Preparation and Properties of a Flake Aluminum Powder in an Ammonium-Perchlorate-Based Composite Modified Double-Base Propellant. Combust. Explos. Shock. Waves 2020, 56, 691–696. [Google Scholar] [CrossRef]
  82. Kuhl, A.L.; Reichenbach, H. Combustion effects in confined explosions. Proc. Combust. Inst. 2009, 32, 2291–2298. [Google Scholar] [CrossRef] [Green Version]
  83. Kotov, Y.A. Electric Explosion of Wires as a Method for Preparation of Nanopowders. J. Nanoparticle Res. 2003, 5, 539–550. [Google Scholar] [CrossRef]
  84. Lerner, M.; Vorozhtsov, A.; Guseinov, S.; Storozhenko, P. Metal Nanopowders Production. Met. Nanopowders 2014, 79–106. [Google Scholar] [CrossRef]
  85. DeLuca, L.; Galfetti, L. Burning of Metalized Composite Solid Rocket Propellants: From Micrometric to Nanometric Aluminum Size. In Proceedings of the Asian Joint Conference on Propulsion and Power, Gyeongju, Korea, 6–8 March 2008; pp. 6–8. [Google Scholar]
  86. Trubert, J.-F.; Hommel, J.; Lambert, D.; Fabignon, Y.; Orlandi, O. New HTPB/AP/Al propellant combustion process in the presence of aluminum nano-particles. Int. J. Energetic Mater. Chem. Propuls. 2008, 7, 99–122. [Google Scholar] [CrossRef]
  87. Tang, D.-Y.; Lyu, J.; He, W.; Chen, J.; Yang, G.; Liu, P.-J.; Yan, Q.-L. Metastable intermixed Core-shell Al@M(IO3)x nanocomposites with improved combustion efficiency by using tannic acid as a functional interfacial layer. Chem. Eng. J. 2020, 384, 123369. [Google Scholar] [CrossRef]
  88. Lerner, M.; Vorozhtsov, A.; Eisenreich, N. Safety Aspects of Metal Nanopowders. Met. Nanopowders 2014, 153–162. [Google Scholar] [CrossRef]
  89. Jacob, R.J.; Jian, G.; Guerieri, P.M.; Zachariah, M.R. Energy release pathways in nanothermites follow through the condensed state. Combust. Flame 2015, 162, 258–264. [Google Scholar] [CrossRef]
  90. Osborne, D.T.; Pantoya, M.L. Effect of Al Particle Size on the Thermal Degradation of Al/Teflon Mixtures. Combust. Sci. Technol. 2007, 179, 1467–1480. [Google Scholar] [CrossRef]
  91. DeLisio, J.B.; Hu, X.; Wu, T.; Egan, G.C.; Young, G.; Zachariah, M.R. Probing the Reaction Mechanism of Aluminum/Poly(vinylidene fluoride) Composites. J. Phys. Chem. B 2016, 120, 5534–5542. [Google Scholar] [CrossRef]
  92. Valluri, S.K.; Bushiri, D.; Schoenitz, M.; Dreizin, E. Fuel-rich aluminum–nickel fluoride reactive composites. Combust. Flame 2019, 210, 439–453. [Google Scholar] [CrossRef]
  93. McCollum, J.; Pantoya, M.L.; Iacono, S.T. Catalyzing aluminum particle reactivity with a fluorine oligomer surface coating for energy generating applications. J. Fluor. Chem. 2015, 180, 265–271. [Google Scholar] [CrossRef]
  94. Kim, D.W.; Kim, K.T.; Min, T.S.; Kim, K.J.; Kim, S.H. Improved Energetic-Behaviors of Spontaneously Surface-Mediated Al Particles. Sci. Rep. 2017, 7, 4659. [Google Scholar] [CrossRef] [Green Version]
  95. Wang, J.; Zeng, C.; Zhan, C.; Zhang, L. Tuning the reactivity and combustion characteristics of PTFE/Al through carbon nanotubes and grapheme. Thermochim. Acta 2019, 676, 276–281. [Google Scholar] [CrossRef]
  96. Ke, X.; Guo, S.; Zhang, G.; Zhou, X.; Xiao, L.; Hao, G.; Wang, N.; Jiang, W. Safe preparation, energetic performance and reaction mechanism of corrosion-resistant Al/PVDF nanocomposite films. J. Mater. Chem. A 2018, 6, 17713–17723. [Google Scholar] [CrossRef]
  97. Meir, S.; Kalabukhov, S.; Hayun, S. Low temperature spark plasma sintering of Al2O3–TiC composites. Ceram. Int. 2014, 40, 12187–12192. [Google Scholar] [CrossRef]
  98. Cai, K.F.; McLachlan, D.S.; Axen, N.; Manyatsa, R. Preparation, microstructures and properties of Al2O3–TiC composites. Ceram. Int. 2002, 28, 217–222. [Google Scholar] [CrossRef]
  99. Cheng, Y.; Sun, S.; Hu, H. Preparation of Al2O3/TiC micro-composite ceramic tool materials by microwave sintering and their microstructure and properties. Ceram. Int. 2014, 40, 16761–16766. [Google Scholar] [CrossRef]
  100. Noor, F.; Vorozhtsov, A.; Lerner, M.; Bandarra Filho, E.P.; Wen, D. Thermal-Chemical Characteristics of Al–Cu Alloy Nanoparticles. J. Phys. Chem. C 2015, 119, 14001–14009. [Google Scholar] [CrossRef] [Green Version]
  101. Shen, J.; Qiao, Z.; Zhang, K.; Wang, J.; Li, R.; Xu, H.; Yang, G.; Nie, F. Effects of nano-Ag on the combustion process of Al–CuO metastable intermolecular composite. Appl. Therm. Eng. 2014, 62, 732–737. [Google Scholar] [CrossRef]
  102. McCollum, J.; Pantoya, M.L.; Iacono, S.T. Activating Aluminum Reactivity with Fluoropolymer Coatings for Improved Energetic Composite Combustion. ACS Appl. Mater. Interfaces 2015, 7, 18742–18749. [Google Scholar] [CrossRef]
Figure 1. The three-step oxidation process of Al proposed by Zachariah.
Figure 1. The three-step oxidation process of Al proposed by Zachariah.
Crystals 12 01187 g001
Figure 2. The change of the Al2O3 shell during the oxidation process of the Al particle, reprinted/adapted with permission from Ref. [35]. 2006, Dreizin, E. L.
Figure 2. The change of the Al2O3 shell during the oxidation process of the Al particle, reprinted/adapted with permission from Ref. [35]. 2006, Dreizin, E. L.
Crystals 12 01187 g002
Figure 3. The process of melt and dispersion of the active Al during the oxidation process of the Al particle, reprinted/adapted with permission from Ref. [46]. 2008, Levitas Valery I.
Figure 3. The process of melt and dispersion of the active Al during the oxidation process of the Al particle, reprinted/adapted with permission from Ref. [46]. 2008, Levitas Valery I.
Crystals 12 01187 g003
Figure 4. Relation between the oxidant release temperature of different oxidizers and the onset temperature of their nanothermites, reprinted/adapted with permission from Ref. [77]. 2022, Dubois, Charles.
Figure 4. Relation between the oxidant release temperature of different oxidizers and the onset temperature of their nanothermites, reprinted/adapted with permission from Ref. [77]. 2022, Dubois, Charles.
Crystals 12 01187 g004
Figure 5. The deformation process of Al during ball milling.
Figure 5. The deformation process of Al during ball milling.
Crystals 12 01187 g005
Figure 6. The SEM image of P-ALEX (a) and F-ALEX (b), reprinted/adapted with permission from Ref. [21]. 2017, Deluca, L.T.
Figure 6. The SEM image of P-ALEX (a) and F-ALEX (b), reprinted/adapted with permission from Ref. [21]. 2017, Deluca, L.T.
Crystals 12 01187 g006
Figure 7. The combustion product of nanothermites, reprinted/adapted with permission from Ref. [89]. 2015, Zachariah, Michael R.
Figure 7. The combustion product of nanothermites, reprinted/adapted with permission from Ref. [89]. 2015, Zachariah, Michael R.
Crystals 12 01187 g007
Figure 8. The pores morphology of TF-Al/5%TiC surface at different temperature: 633 °C (a), 650 °C (b), 670 °C (c) and 690 °C (d), reprinted/adapted with permission from Ref. [56]. 2021, Jie Liu.
Figure 8. The pores morphology of TF-Al/5%TiC surface at different temperature: 633 °C (a), 650 °C (b), 670 °C (c) and 690 °C (d), reprinted/adapted with permission from Ref. [56]. 2021, Jie Liu.
Crystals 12 01187 g008
Figure 9. The oxidation process of Al-Zr alloy (a), Al-W alloy (b) and Al-W alloy after combustion (c), reprinted/adapted with permission from Refs. [53,54]. 2019 and 2021, Ai-min Pang and Shuizhou Cai.
Figure 9. The oxidation process of Al-Zr alloy (a), Al-W alloy (b) and Al-W alloy after combustion (c), reprinted/adapted with permission from Refs. [53,54]. 2019 and 2021, Ai-min Pang and Shuizhou Cai.
Crystals 12 01187 g009
Table 1. The thermal and combustion properties of Al-based MICs, reprinted/adapted with permission from Ref. [70]. 2015, Rossi Carole.
Table 1. The thermal and combustion properties of Al-based MICs, reprinted/adapted with permission from Ref. [70]. 2015, Rossi Carole.
SampleHeat Release
J/g
Delay Time
μs
Burning Rate
m/s
Pressurization Rate
kPa/μs
Al/Bi2O3154154205762
Al/CuO105715340172
Al/MoO3188311010035
Table 2. Properties for Al/CuO and Al/MoO3, reprinted/adapted with permission from Ref. [52]. 2011, Weismiller, M. R.
Table 2. Properties for Al/CuO and Al/MoO3, reprinted/adapted with permission from Ref. [52]. 2011, Weismiller, M. R.
SampleLinear Burning Rate
m/s
Mass Burning Rate
kg/s
Pressurization Rate
MPa/μs
Nano Al/nano CuO9803.80.67
Micron Al/nano CuO6604.81.82
Nano Al/micron CuO2001.30.28
Micron Al/micron CuO1802.00.11
Nano Al/nano MoO36802.00.68
Micron Al/nano MoO33601.50.44
Nano Al/micron MoO31500.450.20
Micron Al/micron MoO3470.520.17
Table 3. The data DSC of Al/MoO3@PFPE and Al/CuO@PFPE, reprinted/adapted with permission from Ref. [102]. 2015, Michelle L. Pantoya.
Table 3. The data DSC of Al/MoO3@PFPE and Al/CuO@PFPE, reprinted/adapted with permission from Ref. [102]. 2015, Michelle L. Pantoya.
SamplePIR Onset Temperature
°C
PIR Heat Release
J/g
Thermite Reaction Onset Temperature
°C
Thermite Heat Release
J/g
Al@PFPE31519.80561133
Al/MoO3--5082078
Al/MoO3@5%PFPE29821.795341370
Al/MoO3@10%PFPE30135.175411672
Al/MoO3@20%PFPE305103.15661889
Al/CuO--517763
Al/CuO@PFPE29819.565691658
Al/CuO@10%PFPE29929.705811305
Al/CuO@20%PFPE30351.37583843
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Wang, D.; Xu, G.; Tan, T.; Liu, S.; Dong, W.; Li, F.; Liu, J. The Oxidation Process and Methods for Improving Reactivity of Al. Crystals 2022, 12, 1187. https://doi.org/10.3390/cryst12091187

AMA Style

Wang D, Xu G, Tan T, Liu S, Dong W, Li F, Liu J. The Oxidation Process and Methods for Improving Reactivity of Al. Crystals. 2022; 12(9):1187. https://doi.org/10.3390/cryst12091187

Chicago/Turabian Style

Wang, Deqi, Guozhen Xu, Tianyu Tan, Shishuo Liu, Wei Dong, Fengsheng Li, and Jie Liu. 2022. "The Oxidation Process and Methods for Improving Reactivity of Al" Crystals 12, no. 9: 1187. https://doi.org/10.3390/cryst12091187

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