Zhang-Et-Al. Chemical-Modification-Of-Hydroxyl-Terminated-Polybutadiene-And-Its-Application-In-Composite-Propellants (HTPB) (Propellant) (Recipe)

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 11

pubs.acs.

org/IECR Article

Chemical Modification of Hydroxyl-Terminated Polybutadiene and


Its Application in Composite Propellants
Pingan Zhang, Wenjiayi Tan, Xilong Zhang,* Juan Chen, Jianmin Yuan, and Jianru Deng*
Cite This: Ind. Eng. Chem. Res. 2021, 60, 3819−3829 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: In this article, hydroxyl-terminated polybutadiene


(HTPB) was chemically transformed into multiamine, hydroxy
polybutadiene (AEHTPB) by the adoption of appropriate
synthesis strategies. The structure of AEHTPB was confirmed by
Downloaded via CLAREMONT COLG on March 4, 2024 at 17:08:49 (UTC).

Fourier transform infrared spectroscopy (FT-IR) and nuclear


magnetic resonance (NMR). The modification degree was altered
by adjusting reaction conditions to find out the effect of structural
change on various physical properties, including the viscosity and
glass transition temperature (Tg) of AEHTPB. Adding AEHTPB to
the HTPB binder system could improve the cross-link density and
strength of the polyurethane elastomer. Meanwhile, the experiment
results indicated that AEHTPB could inhibit the low-temperature
decomposition of ammonium perchlorate (AP) by forming a dense
coating on the surface of AP. Then, the propellant level studies were conducted with the use of AEHTPB/HTPB mixture as a
binder. As a result, the burning rate of propellant containing AEHTPB (5 wt %) was found to be reduced by 10.8% compared to
HTPB-based propellant, and the mechanical properties of the propellant were superior to those of HTPB-based propellant with an
improvement of 46.2% in tensile strength and 35.1% enhancement in elongation at break. Finally, tensile testing and observation of
fracture surface morphology further demonstrated that AEHTPB had a significant bonding effect on AP-based propellant.

1. INTRODUCTION substances have poor compatibility with HTPB propellants,


HTPB propellant is a type of composite solid propellant with and a small amount of addition will cause adverse effects on
hydroxyl-terminated polybutadiene (HTPB) as a polymeric the propellants.
binder. Because of its technical maturity, low cost, and good On the other hand, with the continuous requirements of
combination property, it is widely applied to various strategic energy in propellants, the content of solid fillers is increasing
and tactical solid missiles.1−5 At present, the propellant is higher and higher (up to 90%), which brings an enormous
developing toward high thrust and strong antioverload challenge to the mechanical properties of propellant grains.
capability, which requires the solid propellant used to have However, as a nonpolar binder, HTPB does not has strong
both high energy density and good mechanical properties.6,7 interfacial bonding with polar fillers, and thus undergoes
Meanwhile, in practice, some new booster rockets and cruise dewetting (debonding) upon deformation. Dewetting will
missile engines also need the propellant to have a lower adversely affect the mechanical properties, combustion
burning rate to ensure long-term sustained and stable thrust.8,9 performance, and storage stability of the propellant.21,22 To
For HTPB propellant, the reduction of the burning rate is improve the performance of the propellant, it is usually
mainly achieved by adding the burning-rate depressant to necessary to add a variety of functional additives to the
inhibit the thermal decomposition of ammonium perchlorate propellant formulation, for example, bonding agents,23,24
(AP).10,11 The most commonly used burning-rate depressants burning-rate modifiers,25,26 process additives,27 and plasti-
in HTPB propellants are inorganic salts, such as ammonium cizers.28 However, this makes the propellant composition more
oxalate ((NH4)2C2O4),12,13 calcium carbonate (CaCO3),14,15 complicated and it is more difficult to control quality. Also, the
and strontium carbonate (SrCO3).16,17 However, these
burning-rate depressants do not contribute to the energy of
the propellant, and it is necessary to use these additives in Received: December 16, 2020
rather substantial quantities if the desired burning suppression Revised: February 22, 2021
is to be obtained, which will reduce the energy density and Accepted: February 22, 2021
mechanical properties of the propellant.18,19 Some of the Published: March 4, 2021
literature reported that quaternary ammonium salts have a
significant effect of reducing the burning rate,20 but these

© 2021 American Chemical Society https://dx.doi.org/10.1021/acs.iecr.0c06172


3819 Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

effect of these additives will be reduced due to the antagonism Scheme 2. Synthesis Scheme for AEHTPB
between components and poor thermodynamic compatibility
with the binder matrix.
In this paper, we report a novel approach of functionaliza-
tion of HTPB to multiamine, hydroxy polybutadiene
(AEHTPB) by chemical modification. The hydroxyl group in
the AEHTPB molecule can increase the cross-link density of
the binder matrix through the curing reaction, and the amine
group not only has an interface bonding effect on AP but also bubbles in the mixture. Finally, the mixture was cured at 60 °C
can inhibit its thermal decomposition. Above all, the AEHTPB for 7 days. The cured elastomers were cut into appropriate size
has natural miscibility with HTPB. This work has described the specimens for mechanical performance testing.
synthesis and characterization of AEHTPB, and we have also 2.4. Preparation of Composite Propellants. The
made efforts to research the influence of AEHTPB on the prepared AEHTPB was further applied to the HTPB
thermal decomposition of AP as well as the effect on the propellant, and the detailed propellant composition is provided
mechanical properties of the binder matrix and propellant. in Table 1. Figure 1 shows the preparation process of the

2. EXPERIMENTAL SECTION Table 1. HTPB Propellant Compositiona


2.1. Raw Materials. The HTPB used has a molecular constituent percentage (wt %)
weight (Mn) of ∼4000 g/mol (composed by 20 mol % 1,2- binder system (HTPB/AEHTPB/IPDI) 12
addition structure and 80 mol % 1,4-addition structure RDX 10
microstructures) and a hydroxyl value of 28.5 mg of KOH/g. AP 63
It was purchased from Zibo Qilong Chemical Industry Co. Ltd. Al 15
(China) and used as received without any further purification. a
AEHTPB accounts for 5% of the binder system mass. Curing ratio
Isophorone diisocyanate (IPDI), hydrogen peroxide (H2O2), NCO/OH = 1.0.
diethylamine, and acetic acid were procured from Sinopharm
Chemical Reagent Co. Ltd. (China). Ammonium perchlorate
(AP, 100 mesh) and hexogen (RDX, 100 mesh) were kindly
provided by Jianghe State Chemical Plant (China). All the
solvents were dried and distilled before use. All other chemicals
were at the highest purity available and used without further
purification.
2.2. Synthesis. 2.2.1. Synthesis of Epoxidized HTPB. The
epoxidation of HTPB (EHTPB) was carried out by peroxide
acetic acid in situ (Scheme 1).29,30 Briefly, the procedure for

Scheme 1. Synthesis Scheme for EHTPB

Figure 1. Manufacturing process of HTPB propellant.


the synthesis of EHTPB is as follows: HTPB (30 g) was
dissolved in a mixture of dichloromethane (100 mL) and acetic propellant sample. The AEHTPB was first mixed with HTPB
acid (24 g) solution and then introduced into the reactor, at 50 °C in the mixer. Then, the mixture was moved to a
which was submerged in a well-controlled-temperature water vertical kneader; Al powder, RDX, and AP were added
bath. Sodium tungstate dihydrate (0.3 g) was completely gradually with stirring. Finally, the required amounts of curing
dissolved in a 30 wt % hydrogen peroxide aqueous solution (60 agents were added and mixed thoroughly, followed by
g). The solution was put into the reactor while stirring to start degassing in a vacuum oven at 40 °C and curing at 60 °C
the reaction. The reaction was carried out at 60 °C for 3 h. for 7 days. The cured propellant samples were tested for
After that, the stirring was stopped, and the two phases were mechanical properties.
separated. The organic solvents were washed with distilled 2.5. Characterization Techniques. 2.5.1. Structural
water (3 × 100 mL) and allowed to evaporate. Then, the Characterization. Fourier transform infrared (FTIR) spectra
residue was dried under a vacuum at 60 °C for 12 h. The were recorded on a Nicolet IS10 FTIR spectrometer in the
product is a colorless viscous liquid with a yield of 98%. range 500−4000 cm−1, and the polymers were coated on a KBr
2.2.2. Synthesis of Multiamine, Hydroxy Polybutadiene. disk to form the film. 1H NMR and 13C NMR spectra were
Scheme 2 shows the synthesis scheme for multiamine, hydroxy recorded on a Bruker Avance-III 400 (400 MHz) spectrometer
polybutadiene (AEHTPB). in CDCl3 solvent.
2.3. Preparation of Elastomers. The elastomers based on 2.5.2. Thermal Study. The thermal behaviors of AEHTPB
theHTPB/AEHTPB binder system were prepared with IPDI at and the AEHTPB/HTPB mixture were analyzed with a
a stoichiometry ratio (NCO/OH) of 1.0 without any catalyst. differential scanning calorimeter (DSC) technique. DSC was
First, HTPB and AEHTPB were stirred evenly at 50 °C and performed with a TA Instruments Q500 in a nitrogen
then mixed with a calculated amount of IPDI. Furthermore, atmosphere. About 5 mg of the sample was heat-treated
the above mixture was poured into a polytetrafluoroethylene from −120 °C to room temperature at a heating rate of 10 °C/
mold and held under vacuum for 30 min at 40 °C to remove min.
3820 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

The effect of AEHTPB on the decomposition behavior of are reduced, but the position of the 1,2-vinyl double bonds
AP was evaluated by thermogravimetry−differential scanning remained constant, revealing that the epoxidation reaction
calorimetry (TG−DSC) analysis. The TG−DSC measure- mainly occurs on the 1,4-(cis and trans) double bonds. The
ments were carried out with a Netzsch STA449F5 under adsorption peak at 1240 cm−1 confirms the presence of the
nitrogen atmosphere (50 mL/min) at 10 °C/min heating rate, epoxide group (C−O−C, stretching vibration) in EHTPB.32
from room temperature to 600 °C. In the case of AEHTPB (Figure 2c), an obvious peak centered
2.5.3. Determination of Cross-Link Density. The cross-link at 1070 cm−1 corresponds to the secondary alcohol (C−O,
density of the elastomer was measured by the swelling method. stretching vibration) and disappearance of epoxide group
The cured elastomer samples were cut into pieces of absorption at 1240 cm−1. Also, the broad signal at 3400−3600
approximately 7 × 7 × 3 mm3 size and soaked in toluene for cm−1 and that related to the hydroxyl group (hydrogen bond)
72 h. The sample was weighed after the solvent was gently of AEHTPB are enhanced, indicating a complete conversion of
wiped off. From the swell ratio, the cross-link density was EHTPB to AEHTPB.
calculated by using the Flory−Rehner equation. Figure 3 displays 13C NMR spectra of (a) HTPB, (b)
2.5.4. Mechanical Property. The static mechanical proper- HTPB, and (c) AEHTPB. In the 1H NMR spectrum of HTPB,
ties of the cured polymers and propellants were evaluated by the peaks at 5.4−5.5 (a) and 4.9 ppm (b) are assigned to 1,4-
using a universal testing machine (INSTRON Model 5582). (cis and trans) protons and 1,2-vinyl protons, respectively. For
Dumbbell specimens were cut following ASTM D412 and EHTPB (Figure 3b), the absorption at 2.7 ppm (c)
GJB770B-2005 standard methods, and the tests were corresponds to the trans epoxy protons, whereas the one at
performed at a temperature of 25 °C at a cross-head speed 2.9 ppm (d) corresponds to the cis epoxy protons. According
of 100 mm/min. Three specimens of each elastomer and to the research from Gemmer and Golub,33 the vinyl epoxy
propellants at least were tested, and the average values were group proton has absorption at 2.5 ppm (e). However, we did
reported. not find a peak at 2.5 ppm, which further verified the
2.5.5. Burning Rate. The burning rates were measured by conclusion from the FTIR results that the epoxidation reaction
the strand burner method (GJB 770B-2005 test method of occurred mainly on the 1,4-(cis and trans) double bonds of
propellant) at 20 °C. Each batch of sample was tested under HTPB. From Figure 3c, it is clear that the peaks corresponding
four different pressure points between 2 and 10 MPa with to epoxy protons have disappeared. Moreover, the multiple
three repetitions, and the average values were reported. Then, peaks at 1.0−1.1 (f) and 2.4−2.7 ppm (g) are attributed to the
the pressure exponents were calculated by Vieille’s law (r = methyl protons and methylene protons of diethylamine,
bpn). respectively. All the above results suggest that AEHTPB is
2.5.6. Microscopic Study. The fracture morphologies of successfully prepared. The corresponding assignments of
propellant specimens were shown by a scanning electronic chemical shifts (δ, ppm) from 1H NMR and 13C NMR of
microscope (Hitachi S-4800, Japan) operated at 15 kV. HTPB, EHTPB, and AEHTPB are presented in the Supporting
Information.
3. RESULTS AND DISCUSSION By controlling the epoxidation reaction of HTPB, we
3.1. Polymer Characterization. The FTIR spectra of prepared a series of AEHTPBs with modification degrees of
HTPB, EHTPB, and AEHTPB are enumerated in Figure 2. In 10, 15, 20, and 25% and named them AEHTPB10% ,
the FTIR spectrum of HTPB (Figure 2a), the characteristic AEHTPB15%, AEHTPB20%, and AEHTPB25%, respectively
peak at 725 cm−1 is related to 1,4-cis double bonds, that at 965 (the numbers represent the percentage of double bonds
cm−1 is related to 1,4-trans double bonds, and that at 910 cm−1 modified in HTPB). The physicochemical properties, such as
is related to 1,2-vinyl double bonds.31 For EHTPB (Figure density, viscosity, glass transition temperature (Tg), hydroxyl
2b), the absorption peaks of 1,4-cis and 1,4-trans double bonds value, and amine value of the product, were measured. The
results are displayed in Table 2.
Viscosity and glass transition temperature are important
performance indicators of propellant binder, which determine
the processing performance and low-temperature mechanical
properties. As can be seen from Table 2, the modification
degree is critical in determining the properties of the product.
The viscosity and the glass transition temperature of AEHTPB
are both increased with the increasing modification degree.
The reasons are mainly from two aspects: first, the
modification process destroyed the double bond in HTPB
and introduced bulky groups (hydroxyl and tertiary amine
groups) onto the backbone, which reduced the flexibility of the
molecule; second, these polar groups interact with each other,
increasing the interaction between molecules.36 When the
modification degree on the HTPB exceeds 20%, the viscosity
of AEHTPB increases sharply, which will affect its processing
and usage performance. However, the potential to change the
properties is an important point since different users may have
different requirements, such as more or fewer functional
groups, higher or lower viscosity, and higher or lower Tg, all of
Figure 2. FTIR spectra of (a) HTPB, (b) EHTPB, and (c) AEHTPB. which are controllable by variation in the modification
3821 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 3. 13C NMR spectra of (a) HTPB, (b) EHTPB, and (c) AEHTPB.

Table 2. Physicochemical Properties of AEHTPB with Different Degrees of Modification


viscosity (Pa·s)
sample a 3
density (kg/m ) Mn 30 °C 50 °C Tg (°C) hydroxyl valueb (mmol/g) amine valuec (mmol/g)
HTPB 0.887 4255 9.6 5.0 −91.5 0.470 −
AEHTPB10% 0.920 4986 46.3 14.2 −72.7 1.648 1.178
AEHTPB15% 0.925 5260 88.4 23.5 −55.5 2.252 1.785
AEHTPB20% 0.934 5685 275.0 62.4 −45.9 2.721 2.238
AEHTPB25% 0.946 5952 568.1 168.3 −31.7 3.031 2.551
a
Density is measured by specific gravity method at room temperature. bHydroxyl value is determined by acetic anhydride/N-methylimidazole
method.34 cAmine value is determined by nonaqueous titration.35

degree.37 The follow-up research uses AEHTPB15% as the


analysis object.
3.2. Compatibility Analysis. As a propellant binder
system, AEHTPB and HTPB should have good miscibility to
ensure that AEHTPB can evenly disperse in the HTPB matrix.
The miscibility of blends can be characterized using
thermoanalytical techniques, such as differential scanning
calorimetry (DSC).38,39 As a rule, miscible mixtures show a
single Tg between the two components, if the two component
Tg values are more than 20 °C apart. In contrast, the
immiscible mixtures often possess several glass transition
temperatures corresponding to the components of the blend.
In other words, changes in Tg are governed by the miscibility
of two mixing components.40
Physical mixtures of AEHTPB and HTPB were prepared at
three mass ratios (1:9, 2:8, 3:7) by geometric mixing. Figure 4
depicts the DSC thermograms of AEHTPB/HTPB mixtures at Figure 4. DSC thermograms of (a) HTPB, (b) AEHTPB/HTPB
various ratios. As we can see, the glass transition temperatures (1:9), (c) AEHTPB/HTPB (2:8), (d) AEHTPB/HTPB (3:7), and
of HTPB and AEHTPB are detected to be −91.2 and −55.7 (e) AEHTPB.
°C, respectively. In terms of AEHTPB/HTPB mixtures, there
is only one obvious step in each curve, which means the
3822 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 5. Stress−strain curves of elastomers with different mass ratios of AEHTPB: (a) 0−5 and (b) 5−25 wt % AEHTPB content.

AEHTPB/HTPB mixture exhibits a single Tg, and the Tg is calculated by using the data obtained from the swelling
close to that of HTPB. Therefore, the AEHTPB/HTPB experiment based on the Flory−Rehner equation:44,45
systems are completely compatible and have good low-
temperature performance. From the perspective of molecular Me = −Vsρp (Vp1/3 − Vp/2)/[ln(1 − Vp) + Vp + χVp2]
structure, although a large number of tertiary amine groups (1)
and hydroxyl groups are introduced into the AEHTPB where Vs is the molar volume of the solvent, Vp is the volume
molecule, the main chain structure is still similar to that of fraction of the polymer in the swollen specimen, and χ is the is
HTPB, and the hydroxyl groups in the AEHTPB and HTPB the Flory−Huggins interaction parameter between polymer
molecules can interact through hydrogen bonds, which further and solvent. Vp is computed from
improves the compatibility between AEHTPB and HTPB.
Vp = 1/[1 + (m1/m0 − 1)ρp /ρs ] (2)
3.3. Mechanical Properties. The stress−strain curves of
elastomers with different mass ratios of AEHTPB are exhibited where m0 is the dry weight of the polymer and m1 is the wet
in Figure 5. We can observe that the tensile strength of the weight of the polymer. ρs and ρp are the densities of the solvent
elastomer increases with the AEHTPB content, while the and the polymer, respectively. A value of 0.35 is taken for χ
elongation at break gradually decreases. The elongation at (for HTPB−toluene interaction46). The cross-link density and
break of HTPB elastomer is as high as 1250%, but the tensile corresponding mechanical properties of each elastomer are
strength is only 0.32 MPa. When the addition of AEHTPB is listed in Table 3.
5%, the elongation at break is 997% and the tensile strength is
increased to 0.54 MPa, demonstrating that AEHTPB can Table 3. Cross-Link Network Structure Parameters and
effectively enhance the strength of the binder matrix. It is Mechanical Properties of Different Elastomers
reported that the average functionality of HTPB molecules is
AEHTPB content (wt %) ρp (g/cm3) Me σ (MPa) ε (%)
2.1−2.3,41 which means the cross-link density of the elastomer
prepared from HTPB and IPDI is relatively low. However, 0 0.904 41278 0.32 1225
AEHTPB is a high functionality polymer, which can act as a 1 0.905 39217 0.33 1183
2 0.906 37532 0.34 1155
cross-linking agent during the curing process to increase the
3 0.908 34558 0.37 1103
matrix cross-link density. Meanwhile, the hard segment
4 0.910 27028 0.45 1030
content is increased with increasing AEHTPB. As a result,
5 0.913 22053 0.54 997
the physical cross-link points in the system also increase
10 0.916 17662 0.66 921
accordingly.42 15 0.919 10164 0.92 702
Also, it is worth noting that, when the content of AEHTPB 20 0.922 6317 1.07 458
is low, the impact on the elastic mechanical properties is also 25 0.926 4448 1.06 158
relatively small (Figure 5a). For example, when the addition of
AEHTPB is 3%, the tensile strength of the elastomer is 0.37
Table 3 shows that the cross-link density gradually increased
MPa and the elongation at break is 1103%, which has changed with the increase of AEHTPB, and the changing trend is
little compared with the HTPB elastomer. This is different similar to the mechanical properties. The cross-link network
from the traditional cross-linking agent (such as glycerol); the structure formed by HTPB, AEHTPB, and IPDI is depicted in
addition of the small molecule cross-linking agent is usually Figure 6. Since AEHTPB is a polymer with a large number of
very low (less than 1%).43 hydroxyl groups, it can create highly cross-linked areas in the
3.4. Analysis of Cross-Link Network Structure. The binder network during the curing process. These regions are
macromechanical properties of polyurethane elastomers are composed of very short chains and would not be elastic in
closely related to their microstructures. Here, the swelling comparison with the regular HTPB chains. From the
method was used to determine the cross-link density of the macroscopic view, these short chains in aggregates could be
AEHTPB/HTPB elastomers. The cross-link density of the regarded as large multibranched cross-link points (the red area
polymer is usually characterized by the average molecular in Figure 6). When the amount of AEHTPB added is small,
weight (Me) of adjacent cross-link points, and Me can be these macro-cross-link points are also less, which has little
3823 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

NH4ClO4 (s) ⇆ NH3(s) + HClO4 (s) (3)

In the third stage (the high-temperature decomposition


(HTD)), the exothermic peak appears at a relatively higher
temperature of 440.6 °C, and this corresponds to the complete
decomposition of AP.48 In this process, the adsorbed NH3 and
HClO4 are gradually desorbed into gaseous molecules, and the
redox reactions can take place under a certain temperature:

NH4ClO4 (s) → NH3(g) + HClO4 (g) → O2 (g) + N2(g)


+ Cl 2(g) + H 2O(g) (4)
Figure 6. Cross-link network structure of HTPB/AEHTPB/IPDI
system, where Nc* is the number of carbon atoms between the cross- Parts b and c in Figure 7 are the TG−DSC curves of HTPB
link points. and AEHTPB, respectively. The results show that the thermal
decomposition of HTPB has mainly gone through two stages.
effect on the cross-link density and mechanical properties of For the first stage, the exothermic peak is about 371.5 °C and
the elastomer. Only when the addition of AEHTPB reaches a is accompanied by 21.3% mass loss. The reactions are primarily
certain level will the system’s cross-link density increase depolymerization, cyclization, and cross-linking, accompanied
significantly. by partial decomposition of the cyclized product.49 The second
To evaluate the hypothesis discussed above, we prepared stage is an endothermic process, with the endothermic peak at
AEHTPB15%/HTPB and AEHTPB20%/HTPB elastomers 485 °C and with rapid mass loss. This stage primarily involves
under the same conditions and tested their mechanical the decomposition of the cyclized product with increasing
properties. The addition amounts of AEHTPB20% (hydroxyl temperature. Similarly, the thermal decomposition of AEHTPB
value = 2.721 mmol/g) and AEHTPB15% (hydroxyl value = also involved two stages. The first exothermic peak around
2.252 mmol/g) were 10 and 13.2%, respectively, to ensure that 364.9 °C corresponds to 39.9% weight loss. We could find that
the total number of hydroxyl groups in their AEHTPB/HTPB the initial decomposition temperature of AEHTPB is lower
mixture and the amount of curing agent added were the same than that of HTPB, and the weight loss is higher. The main
(R(NCO/OH) = 1.0). Table 4 tabulates the test results of reason is that the thermal degradation activation energy of the
C−N bonds and C−O bonds in the AEHTPB molecule is low
Table 4. Cross-Link Density and Mechanical Data Observed and the molecule is easier to decompose.
for AEHTPB15%/HTPB and AEHTPB20%/HTPB Elastomers Figure 7d,e illustrates the TG−DSC curves of HTPB/AP
AEHTPB content
and AEHTPB/HTPB/AP. In Figure 7d, the crystallographic
sample (wt %) Me σ (MPa) ε (%) transition temperature and the LTD exothermic peak temper-
AEHTPB15/HTPB 13.2 11611 0.87 714 ature in the HTPB/AP mixture system are 248.6 and 316.4 °C,
AEHTPB20/HTPB 10 15780 0.69 885 respectively, which are the same as those of pure AP.
Surprisingly, the HTD exothermic peak of AP shifted from
440.6 to 407.5 °C. The reason is as follows: when the
cross-link density and mechanical properties. As we can see, temperature is lower than the first exothermic peak temper-
the cross-link density and tensile strength of the AEHTPB15%/ ature of HTPB (371.5 °C), the HTPB is relatively stable, so
HTPB elastomer are higher than those of AEHTPB20%/HTPB. AP would not be exposed. As the temperature continues to
Therefore, we can conclude that the content of AEHTPB is the rise, HTPB begins to rapidly decompose and exposes AP.
main factor affecting the mechanical properties of AEHTPB/ More importantly, at this time the temperature has exceeded
HTPB elastomer, not the degree of modification. the first exothermic peak temperature of AP, causing
3.5. Effect of AEHTPB on AP Thermal Decomposition. accelerated decomposition of AP and instant release of a
The TG−DSC thermal analysis method was applied to large amount of energy, and it advances the high-temperature
investigate the effects of HTPB and AEHTPB on the thermal decomposition of AP. For the AEHTPB/HTPB/AP mixture
decomposition performance of AP. Figure 7 displays the system, the same endothermic peak at 248 °C is also found.
thermal decomposition curves of pure AP, HTPB, AEHTPB, However, the exothermic peak of LTD rises to 337.5 °C, which
HTPB/AP (by mixing in a mass ratio of 2:8), and AEHTPB/ is nearly 20 °C higher than that of pure AP, and the
HTPB/AP (by mixing in a mass ratio of 1:19:80), respectively. exothermic peak area is reduced. Moreover, the exothermic
In addition, the thermal decomposition performances of RDX, peak of HTD has no change compared with the HTPB/AP
HTPB/RDX, and AEHTPB/HTPB/RDX are provided in the mixture system. From the above results, it can be seen that the
Supporting Information. addition of AEHTPB effectively inhibited the low-temperature-
It can be seen from Figure 7a that the thermal decomposition process of AP. Accordingly, in combination
decomposition of pure AP presents three stages: the first with all our observations, we put forward the mechanism of
stage occurs at 246.8 °C, which represents the crystal structural inhibiting the decomposition of AP: First, the AEHTPB
transition from orthorhombic to cubic without mass loss.18 In molecule contains tertiary amine groups, which can react with
the second stage, the exothermic peak at 317.7 °C is attributed the surface of AP particles to form ammonium salt ion bonds
to low-temperature decomposition and accompanied by 19.9% and release ammonia.50,51 The reaction formula is as follows:
mass loss.47 According to the mechanism of the thermal
decomposition of AP, the low-temperature-decomposition (CH3CH 2)2 NR + NH4ClO4
(LTD) step is a solid−gas multiphase reaction, including
decomposition and sublimation: → (CH3CH 2)2 RN+HClO4 − + NH3↑ (5)

3824 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 7. TG−DSC curves of (a) pure AP, (b) pure HTPB, (c) pure AEHTPB, (d) HTPB/AP mixture, and (e) AEHTPB/HTPB/AP mixture in
nitrogen atmosphere.

In the meantime, the AEHTPB molecules are firmly the burning rates of propellants with different contents of
adsorbed on the AP surface by the bonding effect of ionic AEHTPB. The results show that AEHTPB can effectively
bonds, forming a dense coating layer, and this transition layer reduce the burning rate of HTPB propellant. In the presence of
temporarily prevents the buildup of a perchloric acid shell on 5 wt % AEHTPB, the propellant burning rate could decrease
the AP surface during LTD (that is, inhibits the desorption of from 8.54 to 7.62 mm/s, dropping by 10.8%. However, as the
NH3) (Figure 8). By contrast, the interfacial interaction content of AEHTPB increases, the burning rate of the
between HTPB and AP is weak, and the relatively loose propellant no longer changes significantly. According to the
adsorption layer formed by the HTPB cannot prevent the mechanism of AEHTPB, we can speculate that, when the
desorption of ammonia. content of AEHTPB is 5 wt %, the AP particles in the
To further explore the effect of AEHTPB on reducing the propellant have been completely coated. Therefore, continuing
burning rate of the propellant, we applied AEHTPB to the to increase the amount of addition will not have a better
HTPB propellant and tested the burning rate. Table 5 presents reduction effect.
3825 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

MPa, ε = 51.9%), indicating that AEHTPB has the interface


bonding effect and can improve the mechanical strength of the
propellant.
Furthermore, to understand the bonding effect of AEHTPB
on different fillers, we prepared propellant samples containing
a single solid filler and studied them through scanning electron
microscopy (SEM) and mechanical testing. The mechanical
properties and SEM images are shown in Table 6 and Figure

Table 6. Mechanical Properties of Single Filler Samples with


and without AEHTPB

Figure 8. Schematic diagram of the mechanisms of AEHTPB tensile strength elongation


fillera binderb (MPa) (%)
inhibiting AP thermal decomposition.
AP (50 wt %) HTPB 0.57 54.5
Table 5. Burning Rates and Pressure Exponents of (D50 = 121.37 μm) AEHTPB/HTPB 0.96 102.6
Propellants with Different Contents of AEHTPB RDX (50 wt %) HTPB 0.69 46.4
(D50 = 25.11 μm) AEHTPB/HTPB 0.78 54.9
burning rate (mm/s) Al (50 wt %) HTPB 1.40 255.2
AEHTPB content (D50 = 7.12 μm) AEHTPB/HTPB 1.61 280.6
(wt %) 3.5 MPa 5.5 MPa 7.5 MPa 9.5 MPa n a
Weight percent relative to overall sample. b
The mass ratio of
0 5.53 6.71 7.69 8.54 0.436 AEHTPB to HTPB is 1:19.
5 5.03 6.06 6.90 7.62 0.415
10 4.99 6.02 6.84 7.55 0.413
15 4.97 6.02 6.82 7.53 0.413
10. The tensile testing results indicate that, even without
3.6. Interfacial Bonding Effect of AEHTPB. As AEHTPB, the sample of HTPB/Al can obtain good
mentioned above, AEHTPB molecules can be adsorbed on mechanical properties, but the testing results of HTPB/AP
the AP surface through ionic bonds and form a dense coating and HTPB/RDX systems are not satisfactory. After the
layer. During the curing process of the propellant grains, the addition of AEHTPB, the mechanical properties of the
hydroxyl groups in the AEHTPB molecules can react with the AEHTPB/HTPB/Al system are further improved. Also, both
curing agent and connect to the binder matrix network. Thus, the tensile strength and the elongation at break of the
AEHTPB should has an interface bonding function in the AEHTPB/HTPB/AP system are significantly increased,
propellant.52 To examine the practical effect, some propellant especially the elongation at break, which increased from 54.5
level studies were conducted. The propellant used a mixture of to 102.6%. However, the improvement of the AEHTPB/
AEHTPB/HTPB as a binder (AEHTPB content is 5 wt %), HTPB/RDX system is not evident. On the basis of the above
with AP, RDX as the oxidizer, and aluminum powder as analysis, we can draw the following conclusions: (1) The
metallic fuel. For comparison, the performance of the adhesion of AP and binder matrix as well as RDX and binder
propellant processed with HTPB−IPDI polyurethane as a matrix are the main factors that affect the mechanical
binder was also evaluated. properties of propellant. (2) AEHTPB has a significant
Figure 9 displays the maximum tensile strength (σ) and the bonding effect on AP particles, but the bonding effect on
elongation at break (ε) of propellant. As shown, the RDX particles is relatively smaller.
mechanical properties of propellants based on the AEHTPB/ Figure 10 displays the images of filler particles and fracture
HTPB binder system are superior to those in HTPB surface of the propellant samples with and without AEHTPB.
propellant. The tensile strength of the AEHTPB/HTPB As depicted in Figure 10e, a large number of AP particles are
propellant is 0.95 MPa and elongation at break is 70.1%, exposed at the fracture surface and not covered by the binder.
which are higher than those for HTPB propellant (σ = 0.65 After the addition of AEHTPPB (Figure 10i), the fracture
surface of the propellant becomes smooth and the exposed AP
particles are invisible, indicating that the interface between AP
and binder matrix is firmly bonded and the fracture process is
mainly internal tearing of the binder matrix. Since both HTPB
and AEHTPB molecules lack functional groups that can
interact with RDX,53,54 samples containing RDX have poor
adhesion between the filler and the adhesive matrix (Figure
10f,g). Typically, metal fillers show strong interaction with the
binder matrix because of its larger specific surface area and
ability to form coordination bonds with electronegative atoms
in the binder matrix (such as N and O atoms), which is much
stronger than van der Waals interaction.55,56 As expected, the
Al powders in Figure 10g−l are all completely covered by the
bonding matrix, and no interface peeling occurs. The SEM
Figure 9. Maximum tensile strength (σ) and maximum elongation (ε) observation results are consistent with the mechanical
of propellant. properties.
3826 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Figure 10. SEM images of (a) AP, (b) RDX, and (c, d) Al powder. Fracture surfaces of propellant samples of (e) HTPB/AP, (f) HTPB/RDX, (g,
h) HTPB/Al, (i) AEHTPB/HTPB/AP, (j) AEHTPB/HTPB/RDX, and (k, l) AEHTPB/HTPB/Al.

4. CONCLUSIONS
Starting from commercial HTPB, multiamine, hydroxy

*
ASSOCIATED CONTENT
sı Supporting Information

polybutadiene (AEHTPB) was synthesized through chemical The Supporting Information is available free of charge at
modification. FTIR and NMR data have confirmed the https://pubs.acs.org/doi/10.1021/acs.iecr.0c06172.
modification was successful, and we also determined 13
C NMR spectra of HTPB, EHTPB, and AEHTPB; δ
physicochemical properties, such as viscosity, hydroxyl value, assignment from 1H NMR of HTPB, EHTPB, and
amine value, and Tg. The results showed that 15% modification AEHTPB; effect of reaction time on double bond
conversion rate and calculation equation of double bond
on the HTPB backbone gives a balance of a different conversion; TG−DSC curves of RDX, HTPB/RDX
performance, and the product exhibits good compatibility mixture, and AEHTPB/HTPB/RDX mixture in nitrogen
with HTPB. Then, we prepared the elastomer by using a blend atmosphere; burning rates and pressure exponents of
of AEHTPB and HTPB, and the results demonstrate that the propellants with different contents of AEHTPB (PDF)


introduction of AEHTPB can effectively enhance the
mechanical strength of the elastomer. The TG−DSC thermal AUTHOR INFORMATION
analysis study proves that AEHTPB could inhibit the low- Corresponding Authors
temperature decomposition of AP. Furthermore, studies of the Xilong Zhang − Hu Bei Sanjiang Aerospace Jianghe Chemical
propellant level were conducted, and the propellant based on Technology Co., Ltd., Yichang 444200, People’s Republic of
the AEHTPB/HTPB mixture binder provides mechanical China; Email: zhangxilong@hnu.edu.cn
properties superior (35.1−46.2%) to those of the HTPB Jianru Deng − College of Chemistry and Chemical
propellant, and the burning rate was 10.8% lower than that of Engineering, Hunan University, Changsha 410082, People’s
Republic of China; orcid.org/0000-0002-1386-3062;
the HTPB propellant. Finally, tensile testing of single Email: dengjianru@hnu.edu.cn
component propellant and observation of fracture surface
morphology further indicate that AEHTPB has a significant Authors
bonding effect on AP particles, but it is inefficient for RDX due Pingan Zhang − College of Chemistry and Chemical
to the lack of functional groups that interact with RDX. The Engineering, Hunan University, Changsha 410082, People’s
Republic of China
results have shown a high application potential of the proposed Wenjiayi Tan − College of Chemical Engineering and
modified products in AP-based propellants, and it also points Materials Science, Tianjin University of Science and
out the direction for subsequent research work. Technology, Tianjin 300222, People’s Republic of China
3827 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

Juan Chen − College of Chemistry and Chemical Engineering, (17) Chuiko, S. V.; Sokolovskii, F. S. Effect of small additives on the
Hunan University, Changsha 410082, People’s Republic of combustion of composite solid propellants. Russ. J. Phys. Chem. B
China 2009, 3, 926−935.
Jianmin Yuan − College of Material Science and Engineering, (18) Xiu, L. R.; Rong, J. Y. Study on Burning Rate Inhibitors of
Hunan University, Changsha 410082, People’s Republic of HTPB/AP/Al Propellants. Chin. J. Explos. Propellants 2006, 29, 41−
China; orcid.org/0000-0002-8250-8363 17.
(19) Kulkarni, V. V.; Kulkarni, A. R.; Phawade, P. A.; Agrawal, J. P. A
Complete contact information is available at: Study on the Effect of Additives on Temperature Sensitivityin
https://pubs.acs.org/10.1021/acs.iecr.0c06172 Composite Propellants. Propellants, Explos., Pyrotech. 2001, 26, 125−
129.
Notes (20) Oberth, A. E.; Bruenner, R. S. (Aerojet-General Corp.). Solid
The authors declare no competing financial interest. propellants containing burning rate depressants. US 3625782 A, 1971.


(21) Azoug, A.; Nevièr e, R.; Pradeilles-Duval, R. M.;
Constantinescu, A. Influence of fillers and bonding agents on the
REFERENCES viscoelasticity of highly filled elastomers. J. Appl. Polym. Sci. 2014,
(1) Muthiah, R.; Varghese, T. L.; Rao, S. S.; Ninan, K. N.; 131, 40664.
Krishnamurthy, V. N. Realization of an Eco-Friendly Solid Propellant (22) Lei, M.; Wang, J.; Cheng, J.; Xiao, J.; Wen, L.; Lu, H.; Hou, X.
Based on HTPB-HMX-AP System for Launch Vehicle Applications. A constitutive model of the solid propellants considering the interface
Propellants, Explos., Pyrotech. 1998, 23, 90−93. strength and dewetting. Compos. Sci. Technol. 2020, 185, 107893.
(2) da Silva, G.; Rufino, S. C.; Iha, K. An Overview of the (23) Yadav, A.; Pant, C. S.; Das, S. Research Advances in Bonding
Technological Progress in Propellants Using Hydroxyl-terminated Agents for Composite Propellants. Propellants, Explos., Pyrotech. 2020,
Polybutadiene as Binder During 2002−2012. J. Aerosp. Technol. 45, 695−704.
Manag. 2013, 5, 267−278. (24) Liu, M. E.; Zhang, X. L.; Deng, J. R. Synthesis and application
(3) DeLuca, L. T.; Galfetti, L.; Maggi, F.; Colombo, G.; Merotto, L.; of modified borate ester bonding agent for HTPB propellant. Chin. J.
Boiocchi, M.; Paravan, C.; Reina, A.; Tadini, P.; Fanton, L. Energy Mater. 2016, 24, 550−554.
Characterization of HTPB-based solid fuel formulations: Perform- (25) Shen, S.; Chen, S.; Wu, B. The thermal decomposition of
ance, mechanical properties, and pollution. Acta Astronaut. 2013, 92, ammonium perchlorate (AP) containing a burning-rate modifier.
150−162.
Thermochim. Acta 1993, 223, 135−143.
(4) Joshi, K.; Chaudhuri, S. Hot Spot Interaction with Hydroxyl-
(26) Zhang, Y.; Wang, N.; Huang, Y.; Wu, W.; Huang, C.; Meng, C.
Terminated Polybutadiene Binder in Energetic Composites. J. Phys.
Fabrication and catalytic activity of ultra-long V2O5 nanowires on the
Chem. C 2018, 122, 14434−14446.
(5) Kshirsagar, D. R.; Jain, S.; Jawalkar, S. N.; Naik, N. H.; Pawar, S.; thermal decomposition of ammonium perchlorate. Ceram. Int. 2014,
Maurya, M. Evaluation of Nano-Co3O4 in HTPB-based Composite 40, 11393−11398.
Propellant Formulations. Propellants, Explos., Pyrotech. 2016, 41, 304− (27) Li, H. Processing Aids Used in Manufacturing Solid
311. Propellants. Chin. J. Energy Mater. 1999, 7, 76−78.
(6) Chaturvedi, S.; Dave, P. N. Solid propellants: AP/HTPB (28) Bandgar, B. M.; Krishnamurthy, V. N.; Mukundan, T.; Sharma,
composite propellants. Arabian J. Chem. 2019, 12, 2061−2068. K. C. Mathematical modeling of rheological properties of hydroxyl-
(7) Park, S.; Choi, S.; Kim, K.; Kim, W.; Park, J. Effects of terminated polybutadiene binder and dioctyl adipate plasticizer. J.
Ammonium Perchlorate Particle Size, Ratio, and Total Contents on Appl. Polym. Sci. 2002, 85, 1002−1007.
the Properties of a Composite Solid Propellant. Propellants, Explos., (29) Zhou, Q.; Jie, S.; Li, B. G. Facile synthesis of novel HTPBs and
Pyrotech. 2020, 45, 1376−1381. EHTPBs with high cis-1,4 content and extremely low glass transition
(8) Smith, P.; Bankaitis, H. HTPB propellants for large booster temperature. Polymer 2015, 67, 208−215.
applications. AIAA. J. 1971, DOI: 10.2514/6.1971-708. (30) Wang, Q.; Zhang, X.; Wang, L.; Mi, Z. Kinetics of Epoxidation
(9) Yang, L.; Zhou, R. Effects of Composite Deceleration Agents on of Hydroxyl-Terminated Polybutadiene with Hydrogen Peroxide
the Combustion Characteristics of the Propellant with Low Burning under Phase Transfer Catalysis. J. Mol. Catal. A: Chem. 2009, 309,
Rate. Chin. J. Explos. Propellants 2013, 36, 70−73. 89−94.
(10) Glaskova, A. P. Three possible ways to inhibit the ammonium (31) Reshmi, S.; Arunan, E.; Nair, C. P. R. Azide and Alkyne
perchlorate combustion process. AIAA J. 1975, 13, 438−442. Terminated Polybutadiene Binders: Synthesis, Cross-linking, and
(11) Boldyrev, V. V. Thermal decomposition of ammonium Propellant Studies. Ind. Eng. Chem. Res. 2014, 53, 16612−16620.
perchlorate. Thermochim. Acta 2006, 443, 1−36. (32) Zhang, J.; Zhou, Q.; Jiang, X. H.; Du, A. K.; Zhao, T.; van
(12) Sun, Y. L.; Li, S. F.; Ding, D. H. Effect of ammonium oxalate/ Kasteren, J.; Wang, Y. Oxidation of natural rubber using a sodium
strontium carbonate on the burning rate characteristics of composite tungstate/acetic acid/hydrogen peroxide catalytic system. Polym.
propellants. J. Therm. Anal. Calorim. 2006, 86, 497−503. Degrad. Stab. 2010, 95, 1077−1082.
(13) Ghorpade, V.; Dey, A.; Jawale, L.; Kotbagi, A. M.; Kumar, A.; (33) Gemmer, R. V.; Golub, M. A. 13C-NMR Spectroscopic study
Gupta, M. Study of Burn Rate Suppressants in AP-Based Composite
of epoxidized 1,4-polyisoprene and 1,4-polybutadiene. J. Polym. Sci.,
Propellants. Propellants, Explos., Pyrotech. 2010, 35, 53−56.
Polym. Chem. Ed. 1978, 16, 2985−2990.
(14) Li, X. D.; Yang, R. J.; Yang, Y. Preparation of ultrafine calcium
carbonate with different shapes and their applications in the HTPB (34) Dee, L. A.; Biggers, B. L.; Fiske, M. E. N-Methylimidazole as a
propellant. Chin. J. Energy Mater. 2009, 17, 64−68. catalyst for acetylation of hydroxyl terminated polymers. Anal. Chem.
(15) Dey, A.; Ghorpade, V. G.; Kumar, A.; Gupta, M. Biuret: a 1980, 52, 572−573.
potential burning rate suppressant in ammonium chlorate (VII) based (35) Standard Test Method for Polyurethane Raw Materials:
composite propellants. Cent. Eur. J. Energy Mater. 2014, 11, 3−13. Determination of Basicity in Polyols, Expressed as Percent Nitrogen;
(16) Trache, D.; Maggi, F.; Palmucci, I.; Deluca, L. T.; Khimeche, ASTM D6979-03; ASTM: 2003.
K.; Fassina, M.; Dossi, S.; Colombo, G. Effect of amide-based (36) Rao, B. N.; Yadav, P. J. P.; Malkappa, K.; Jana, T.; Sastry, P.
compounds on the combustion characteristics of composite solid Triazine functionalized hydroxyl terminated polybutadiene polyur-
rocket propellants. Arabian Journal of Chemistry, 12(8). Effect of ethane: Influence of triazine structure. Polymer 2015, 77, 323−333.
amide-based compounds on the combustion characteristics of (37) Colclough, M. E.; Paul, N. C. Nitrated Hydroxy-Terminated
composite solid rocket propellants. Arabian J. Chem. 2019, 12, Polybutadiene: Synthesis and Properties. ACS Symp. Ser. 1996, 623,
3639−3651. 97−103.

3828 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829
Industrial & Engineering Chemistry Research pubs.acs.org/IECR Article

(38) Zhao, X.; Niu, K.; Xu, Y.; Peng, Z.; Jia, L.; Hui, D.; Zhang, L.
Morphology and performance of NR/NBR/ENR ternary rubber
composites. Composites, Part B 2016, 107, 106−112.
(39) Ng, C. W A.; Bellinger, M. A.; Macknight, W. J. Compatibility
Enhancement of Blends of Nylon 4 with Lithium-Neutralized
Sulfonated Polystyrene. Macromolecules 1994, 27, 6942−6947.
(40) Li, T. Y.; Chien, J. T. Evaluation of Rice Starch-Hydrocolloid
Compatibility at Low-Moisture Content by Glass Transitions. J. Food
Sci. 2001, 66, 698−704.
(41) Sadeghi, G. M. M.; Morshedian, J.; Barikani, M. Determination
of OH-number and functionality of polybutadiene-ol by FTIR and
NMR spectroscopy. Polym. Test. 2003, 22, 165−168.
(42) Zhang, P.; Yuan, J.; Pang, A.; Tang, G.; Deng, J. A novel UV-
curing liner for NEPE propellant: Insight from molecular simulations.
Composites, Part B 2020, 195, 108087.
(43) Tsai, Y. M.; Yu, T. L.; Tseng, Y. H. Physical properties of
crosslinked polyurethane. Polym. Int. 1998, 47, 445−450.
(44) Flory, P. J. Principles of Polymer Chemistry; Cornell University
Press: Ithaca, NY, USA, 1953.
(45) Zhu, G. C.; Wang, G. Y.; Hu, C. P. Effect of Crosslink Density
on the Structures and Properties of Aliphatic Polyurethane Elastomer.
Acta Polym. Sin. 2011, 274−280.
(46) Jain, S. R.; Sekkar, V.; Krishnamurthy, V. N. Mechanical and
swelling properties of HTPB-based copolyurethane networks. J. Appl.
Polym. Sci. 1993, 48, 1515−1523.
(47) Zhang, Y.; Liu, X.; Nie, J.; Yu, L.; Zhong, Y.; Huang, C.
Improve the catalytic activity of α-Fe2O3 particles in decomposition
of ammonium perchlorate by coating amorphous carbon on their
surface. J. Solid State Chem. 2011, 184, 387−390.
(48) Chen, L.; Li, L.; Li, G. Synthesis of CuO nanorods and their
catalytic activity in the thermal decomposition of ammonium
Perchlorate. J. Alloys Compd. 2008, 464, 532−536.
(49) Du, T. Thermal decomposition studies of solid propellant
binder HTPB. Thermochim. Acta 1989, 138, 189−197.
(50) Hasegawa, K.; Takizuka, M.; Fukuda, T. Bonding Agents for AP
and Nitramine/HTPB Composite Propellants. AIAA. J. 1983,
DOI: 10.2514/6.1983-1199.
(51) Hori, K.; Iwama, A.; Fukuda, T. FTIR spectroscopic study on
the Interaction between Ammonium Perchlorate and Bonding Agents.
Propellants, Explos., Pyrotech. 1990, 15, 99−102.
(52) Kim, C. S.; Youn, C. H.; Noble, P. N.; Gao, A. Development of
Neutral Polymeric Bonding Agents for Propellants with Polar
Composites Filled with Organic Nitramine Crystals. Propellants,
Explos., Pyrotech. 1992, 17, 38−42.
(53) Landsem, E.; Jensen, T. L.; Hansen, F. K.; Unneberg, E.;
Kristensen, T. E. Neutral Polymeric Bonding Agents (NPBA) and
Their Use in Smokeless Composite Rocket Propellants Based on
HMX-GAP-BuNENA. Propellants, Explos., Pyrotech. 2012, 37, 581−
591.
(54) Zhang, P.; Pang, A.; Tang, G.; Deng, J. Molecular dynamics
simulation study on the mechanism of NPBA enhancing interface
strength of NEPE propellant. Appl. Surf. Sci. 2019, 493, 131−138.
(55) Guo, L.; Song, W.; Hu, M.; Xie, C.; Chen, X. Preparation and
reactivity of aluminum nanopowders coated by hydroxyl-terminated
polybutadiene (HTPB). Appl. Surf. Sci. 2008, 254, 2413−2417.
(56) Zhang, X.; Zheng, J.; Fang, H.; Zhang, Y.; Bai, S.; He, G.
Al2O3/graphene reinforced bio-inspired interlocking polyurethane
composites with superior mechanical and thermal properties for solid
propulsion fuel. Compos. Sci. Technol. 2018, 167, 42−52.

3829 https://dx.doi.org/10.1021/acs.iecr.0c06172
Ind. Eng. Chem. Res. 2021, 60, 3819−3829

You might also like