You are on page 1of 17

molecules

Article
The Role of Graphene Oxide in the Exothermic Mechanism of
Al/CuO Nanocomposites
Jiaxin Su 1,2 , Yan Hu 1,2, * , Bin Zhou 1,2 , Yinghua Ye 1,2 and Ruiqi Shen 1,2

1 School of Chemistry and Chemical Engineering, Nanjing University of Science and Technology,
Nanjing 210094, China
2 Micro-Nano Energetic Devices Key Laboratory, Ministry of Industry and Information Technology,
Nanjing 210094, China
* Correspondence: huyan@njust.edu.cn

Abstract: Metastable intermixed composites (MICs) have received increasing attention in the field of
energy materials in recent years due to their high energy and good combustion performance. The
exploration of ways of improving their potential release of heat is still underway. In this study, Al–
CuO/graphene oxide (GO) nanocomposites were prepared using a combination of the self-assembly
and in-suit synthesis methods. The formulation and experimental conditions were also optimized to
maximize the exothermic heat. The DSC analysis shows that the addition of the GO made a significant
contribution to the exothermic effect of the nanothermite. Compared with the Al–CuO nanothermite,
the exothermic heat of the Al–CuO/GO nanocomposites increase by 306.9–1166.3 J/g and the peak
temperatures dropped by 7.9–26.4 ◦ C with different GO content. The reaction mechanism of the
nanocomposite was investigated using a DSC and thermal reaction kinetics analysis. It was found
that, compared with typical thermite reactions, the addition of the GO changed the reaction pathway
of the nanothermite. The reaction products included CuAlO2 . Moreover, the combustion properties
of nanocomposite were investigated. This work reveals the unique mechanism of GO in thermite
reactions, which may promote the application of carbon materials in nanothermite.
Citation: Su, J.; Hu, Y.; Zhou, B.; Ye,
Y.; Shen, R. The Role of Graphene Keywords: graphene oxide; metastable intermixed composites (MICs); nanothermite; thermokinetics;
Oxide in the Exothermic Mechanism
combustion
of Al/CuO Nanocomposites.
Molecules 2022, 27, 7614. https://
doi.org/10.3390/molecules27217614

Academic Editors: Chongjun Zhao 1. Introduction


and Andrea Dorigato For decades, metastable intermixed composites (MICs) have attracted more and more
Received: 15 September 2022
interest in the area of energetic materials (EMs) [1–5]. EMs are critical to the advancement
Accepted: 3 November 2022
of microscale energy-demanding systems such as propulsion units, actuation parts, power
Published: 6 November 2022
and igniters [6–9]. MICs are widely used in propellants, the combustion synthesis of so-
phisticated materials, and high explosives due to their extraordinary energy density, great
Publisher’s Note: MDPI stays neutral
ignition performance, and high burning rate [10–14]. Generally, MICs consist of a metal fuel
with regard to jurisdictional claims in
(aluminum (Al), boron, magnesium, etc.) and an oxidizer (copper oxide (CuO), bismuth tri-
published maps and institutional affil-
oxide, ferric oxide, etc.), and at least one of them is on the nanometer scale [9,15–18]. Among
iations.
them, Al/CuO has received much attention due to its high energy release [19]. The proper-
ties of Al/CuO are actually influenced by its particle morphology [20,21]. Chen et al. [22]
prepared three different morphologies of Al/CuO nano particles (NPs). By changing the
Copyright: © 2022 by the authors.
nanoparticle morphology, the properties of the Al/CuO nanothermite changed significantly.
Licensee MDPI, Basel, Switzerland. Therefore, the performance of nanothermite agents can be improved by controlling the
This article is an open access article particle morphology.
distributed under the terms and Nanoscale particles are highly reactive materials and are desirable for better com-
conditions of the Creative Commons bustion efficiency. However, their small particle size also poses some challenges, such
Attribution (CC BY) license (https:// as their tendency to agglomerate, which leads to a reduction in specific surface area [23]
creativecommons.org/licenses/by/ and consequently affects the heat release of the nanothermite. In order to solve above
4.0/). problem, the immobilization of nanoparticles on certain substrates has been widely used to

Molecules 2022, 27, 7614. https://doi.org/10.3390/molecules27217614 https://www.mdpi.com/journal/molecules


Molecules 2022, 27, 7614 2 of 17

avoid agglomeration [24]. Graphene oxide (GO) nanoflakes have a large surface area and
abundant oxygen-containing groups, which are increasingly chosen as suitable substrates
for anchoring NPs to prevent their aggregation [25–28]. GO is usually prepared by chemical
oxidation and the flaking of graphite powders, which has been widely researched in the
area of energetic materials, biology, and chemistry [29,30]. It has surface activity and can
reduce interfacial energy because its basal plane has abundant oxygen-containing groups
(hydroxyl, carboxylic, epoxide, etc.) [31–33]. The existence of these functional groups
makes it possible for GO to be functionalized with other materials [34–36]. In addition,
GO is thermally unstable, which means it can release a significant amount of heat with
only a little heat excitation. When a heavy GO chunk is heated on a heating source, it has
the potential to explode in a matter of seconds [28,37]. Rajagopalan et al. [38] prepared
Al-Bi2 O3 /graphene sheet composites using a self-assembly method, and the energy release
of the nanothermite was enhanced from 739 to 1421 J/g. However, the large particle size of
the oxide particles and the high loss during the self-assembly process led to the incomplete
release of potential heat from the MICs.
In this work, Al–CuO/GO nanocomposites were prepared using a combination of
in-suit synthesis and self-assembly routes using Cu(CH3 COO)2 ·H2 O as the copper source,
and the experimental conditions and formulations were optimized for investigation. The
nanocomposites and their reaction products were characterized by various means, such as X-
ray diffraction (XRD), energy dispersive spectrum (EDS), transmission electron microscopy
(TEM), etc. Moreover, the thermal behavior of the nanocomposites was studied using
a differential scanning calorimeter (DSC), and their reaction kinetics were investigated.
Finally, the combustion properties of the nanocomposite were studied.

2. Results and Discussion


2.1. Formula Optimization
2.1.1. Effect of Al/CuO Equivalence Ratios
By adjusting the quality of the added raw materials, a series of Al–CuO/GO nanocom-
posites with different Al:CuO equivalent ratios were prepared. The GO content of the
nanocomposites was 5 wt%. The XRD results are shown in Figure 1. In Figure 1a, the
characteristic diffraction peaks of the GO (001) crystal plane appear at about 2θ = 10◦ . The
characteristic diffraction peaks of the CuO (−111) and (111) crystal planes corresponding
to the monoclinic system appeared at 35.2◦ and 38.5◦ , respectively. In addition, the corre-
sponding characteristic diffraction peaks of the Al (111), (200), (200), and (311) crystal planes
appeared at 38.5◦ , 44.7◦ , 65.1◦ , and 78.2◦ , respectively. Nevertheless, the disappearance of
the reflection peak (001) of GO in the composite may demonstrate that the ruled lamellar
pattern of the GO had been disrupted, and that exfoliated GO sheets had formed due to the
loading of Al particles and the growth of CuO nanocrystals. No significant Al peaks were
detected in the nanocomposites due to the small amount of Al in the system, as shown in
Figure 1b. As the Al content grew, the characteristic Al peaks could be identified.
Furthermore, we observed of the morphology and dimensions of the nanostructures in
the composites using SEM. Figure 2 shows the micrographs of the GO, different proportions
of Al to CuO nanocomposites, including a partially enlarged image. As displayed in
Figure 2a, the initial GO sheets were large and smooth with a thickness of about 2.2 nm.
After the reaction, the whole GO was divided into a small piece of GO, and the layer became
rough. As the Al content rose, the agglomeration of Al particles on the GO layer increased.
The dispersion of the loaded particles on the GO was not affected by the change in CuO
and Al equivalence ratios. A large amount of Al can cause particle agglomeration. Figure 2f
is the local enlarged image of the composite of CuO:Al = 1.5:1. Here it can be seen more
clearly that the complete GO layer was divided into small pieces with a thickness of about
0.4 nm. The SEM results show that Al and CuO nanoparticles were successfully loaded
onto the surface of the GO lamellae. Figure 3 shows energy-dispersive X-ray spectroscopy
(EDS) images of the nanocomposite (CuO:Al = 1.5:1). The EDS images indicate that the Al
and CuO were uniformly spread in the composite particles.
Molecules 2022, 27, 7614 3 of 18
Molecules 2022, 27, 7614 3 of 17

■ Al
● CuO

■ ♦ GO
● ■
Intensity(a.u.)
● ■ (e) ■

(d)

(c)

(b)

(a)

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80

2θ(°)
Molecules 2022, 27, 7614 Figure 1. XRD patterns of Al–CuO/GO nanocomposites: (a) GO, (b) CuO:Al = 2.5:1, (c) CuO:Al 4 of 18
=
Figure 1. XRD patterns of Al–CuO/GO nanocomposites:
2:1, (d) CuO:Al = 1.5:1, and (e) CuO:Al = 1:1. (a) GO, (b) CuO:Al = 2.5:1, (c) CuO:Al = 2:1,
(d) CuO:Al = 1.5:1, and (e) CuO:Al = 1:1.

Furthermore, we observed of the morphology and dimensions of the nanostructures


in the composites using SEM. Figure 2 shows the micrographs of the GO, different pro-
portions of Al to CuO nanocomposites, including a partially enlarged image. As displayed
in Figure 2a, the initial GO sheets were large and smooth with a thickness of about 2.2 nm.
After the reaction, the whole GO was divided into a small piece of GO, and the layer be-
came rough. As the Al content rose, the agglomeration of Al particles on the GO layer
increased. The dispersion of the loaded particles on the GO was not affected by the change
in CuO and Al equivalence ratios. A large amount of Al can cause particle agglomeration.
Figure 2f is the local enlarged image of the composite of CuO:Al = 1.5:1. Here it can be
seen more clearly that the complete GO layer was divided into small pieces with a thick-
ness of about 0.4 nm. The SEM results show that Al and CuO nanoparticles were success-
fully loaded onto the surface of the GO lamellae. Figure 3 shows energy-dispersive X-ray
spectroscopy (EDS) images of the nanocomposite (CuO:Al = 1.5:1). The EDS images indi-
cate that the Al and CuO were uniformly spread in the composite particles.
In order to further observe the morphology of the composites, they were character-
ized by transmission electron microscopy (TEM). The composites without GO were pre-
pared by this method as the control (Figure 4b). The Al particles were observed to be reg-
ular spheres with clear edges and a particle size of about 70–100 nm. The CuO NPs were
clastic, with a particle size of 5–10 nm, and they were wrapped on the surface of the Al
particles. Figure 4c is a TEM image of the Al–CuO/GO with a CuO:Al ratio of 1.5:1. The
Figure
result shows2. SEM
thatimages of Al–CuO/GO
after adding GO to the nanocomposites: (a) GO, (b)
reaction, compared CuO:Al
with = 2.5:1, (c)
the smooth CuO:Alof= 2:1,
surface
Figure 2. SEM images of Al–CuO/GO nanocomposites: (a) GO, (b) CuO:Al = 2.5:1, (c) CuO:Al = 2:1,
(d) CuO:Al = 1.5:1, (e) CuO:Al = 1:1, and (f) a partially enlarged image.
the initial GO =
(d) CuO:Al with
1.5:1,a (e)
width of about
CuO:Al 5 μm
= 1:1, and (f) ainpartially
Figure enlarged
4a, the GO layer was covered with
image.
particles, and the particles loaded on the GO layer were uniform. The lamellae were de-
stroyed andIn order
dividedto further observe
into smaller the morphology
lamellae with a width of the composites,
of about 1.5 μm.they
From were
thecharacterized
partially
enlarged image (Figure 4d), it can be seen more clearly that the CuO particles were
by transmission electron microscopy (TEM). The composites without GO and the prepared
Al
by this
particles method
were as the uniformly
distributed control (Figure 4b).GO
over the The Al particles
sheets, and that were observed
the CuO to be regular
was wrapped
withspheres with clear
Al particles, showingedges and
the a particle
same shape size
as the of composites
about 70–100 nm. The
without GO CuO NPs 4b).
(Figure wereItclastic,
is
with a particle size of 5–10 nm, and they were wrapped on the surface
notable that, except for the CuO particles on the surface of the GO sheets, some of the CuO of the Al particles.
Figure
particles 4c is
were a TEMinto
inserted image of the
the GO Al–CuO/GO
sheets. with aCu
This is because CuO:Al ratio
2+ ions are notofonly
1.5:1. The result
adsorbed
on the surface of GO, but also intercalated GO sheets. The CuO crystals then nucleatethe
shows that after adding GO to the reaction, compared with the smooth surface of andinitial
GO with a width of about 5 µm
grow, which causes the GO lamella to peel off. in Figure 4a, the GO layer was covered with particles,
and the particles loaded on the GO layer were uniform. The lamellae were destroyed and
divided into smaller lamellae with a width of about 1.5 µm. From the partially enlarged
image (Figure 4d), it can be seen more clearly that the CuO particles and the Al particles

Figure 3. EDS images of Al–CuO/GO with a CuO:Al ratio of 1.5:1.

The results of the SEM and TEM show that the Al and CuO particles loaded uni-
Molecules 2022, 27, 7614 4 of 17

were distributed uniformly over the GO sheets, and that the CuO was wrapped with Al
particles, showing the same shape as the composites without GO (Figure 4b). It is notable
that, except for the CuO particles on the surface of the GO sheets, some of the CuO particles
were inserted into the GO sheets. This is because Cu2+ ions are not only adsorbed on the
surface2.of
Figure GO,
SEM but also
images intercalated
of Al–CuO/GO GO sheets. The
nanocomposites: (a)CuO crystals
GO, (b) then
CuO:Al nucleate
= 2.5:1, and grow,
(c) CuO:Al = 2:1,
which
(d) causes
CuO:Al the (e)
= 1.5:1, GOCuO:Al
lamella to peel
= 1:1, off.a partially enlarged image.
and (f)

Molecules 2022, 27, 7614 5 of 18


Figure
Figure 3.
3. EDS
EDS images
images of
of Al–CuO/GO
Al–CuO/GOwith
withaaCuO:Al
CuO:Alratio
ratioof
of1.5:1.
1.5:1.

The results of the SEM and TEM show that the Al and CuO particles loaded uni-
formly in the composites. This indicates that it is feasible to prepare Al–CuO/GO MICs
using this method in the water–isopropanol system. Changing the equivalence ratio of
Al–CuO had no noticeable effect on the particle loading. As the aluminum particle content
increased, agglomeration appeared, which indicates that too much Al content is detri-
mental. However, too little Al will affect the heat release of the system. Therefore, the
optimal ratio needs to be further analyzed using a DSC.
Figure 5 presents the DSC results of Al–CuO/GO MICs with different equivalence
ratios of Al and CuO, and the parameters are shown in Table 1. The results show that there
was one heat-absorbing peak and three exothermic peaks at 25–1000 °C. As the Al content
increased, the heat release first increased and then decreased. When CuO:Al = 1.5:1, the
maximum heat release was obtained. The reason is that too much Al content causes the par-
ticles to agglomerate, and the reaction does not complete. Too little Al can prevent the CuO
reaction from completing. This is also consistent with the SEM results (Figure 2) analyzed
previously.

Figure4.4.TEM
Figure TEMimages
imagesof
of(a)
(a)pure
pureGO,
GO,(b)
(b)pure
pureAl–CuO,
Al–CuO,(c)
(c)Al–CuO/GO
Al–CuO/GOwith
withCuO:Al
CuO:Al== 1.5:1,
1.5:1,and
and
(d) a partially enlarged image.
(d) a partially enlarged image.

The results of the SEM and TEM show that the Al and CuO particles loaded uniformly
in the composites. This indicates that it is feasible to prepare Al–CuO/GO MICs using this
method in the water–isopropanol system. Changing the equivalence ratio of Al–CuO had
no noticeable effect on the particle loading. As the aluminum particle content increased,
Molecules 2022, 27, 7614 5 of 17

agglomeration appeared, which indicates that too much Al content is detrimental. However,
too little Al will affect the heat release of the system. Therefore, the optimal ratio needs to
be further analyzed using a DSC.
Figure 5 presents the DSC results of Al–CuO/GO MICs with different equivalence
ratios of Al and CuO, and the parameters are shown in Table 1. The results show that there
was one heat-absorbing peak and three exothermic peaks at 25–1000 ◦ C. As the Al content
increased, the heat release first increased and then decreased. When CuO:Al = 1.5:1, the
maximum heat release was obtained. The reason is that too much Al content causes the
particles to agglomerate, and the reaction does not complete. Too little Al can prevent
the CuO
Figure reaction
4. TEM from
images completing.
of (a) This
pure GO, (b) is also
pure consistent
Al–CuO, with thewith
(c) Al–CuO/GO SEMCuO:Al
results=(Figure 2)
1.5:1, and
analyzed previously.
(d) a partially enlarged image.

Figure
Figure 5.
5. DSC
DSC results
results of
of Al–CuO/GO
Al–CuO/GOMICs
MICswith
withdifferent
differentequivalence
equivalenceratios
ratiosof
ofAl
Aland
andCuO.
CuO.

Table 1. DSC
Table 1. DSC parameters of the
parameters of the Al–CuO/GO
Al–CuO/GO MICs
MICsat
ataaheating
heatingrate
rate of
of 10
10 °C/min.
◦ C/min.

Sample
Sample
Tpk1 (°C)
Tpk1 (◦ C)
Tpk2 (°C)
Tpk2 (◦ C)
Tpk3 (°C)
Tpk3 (◦ C)
Tpk4 (°C)
Tpk4 (◦ C)
ΔH (J/g)
∆H (J/g)
CuO:Al = 1 193.1 279.5 568.7 739.5 711.5
CuO:Al = 1 193.1 279.5 568.7 739.5 711.5
CuO:Al = 1.5
CuO:Al = 1.5
190.8
190.8
279.4
279.4
554.8
554.8
737.8
737.8
1045
1045
CuO:Al
CuO:Al ==22 193.7
193.7 275.8
275.8 555.2
555.2 739.8
739.8 886
886
CuO:Al
CuO:Al == 2.5
2.5 191.6
191.6 278.2
278.2 561.5
561.5 830.5
830.5 811.4
811.4

2.1.2. Effect of the GO Content


In this section, all composites have an equivalent ratio of CuO to Al of 1.5:1. The
samples were named “Al–CuO/GO0.5wt% ”, “Al–CuO/GO1wt% ”, “Al–CuO/GO3wt% ”, and
“Al–CuO/GO5wt% ”, according to the mass fractions of GO. Figure 6 shows the SEM images
of the nanocomposites with varying GO content. Unlike the equivalence ratio of Al/CuO,
changing the GO content can affect the loading of the particles. The pure Al–CuO (Figure 5a)
are spherical, which is consistent with the TEM images (Figure 4b). It is obvious that with
the increase in GO, the load of the particles on the GO layer decreased. This is because in
the water–isopropanol system, GO is dispersed, the same content of particles are dispersed
on more GO, and the corresponding particles on the GO layer will be reduced. This reduces
the contact area between Al and CuO. The results show that the content of GO had an
important effect on the distribution of Al and CuO particles on the lamella. The effect of
GO content on the heat release of the composites needs to be further analyzed using a DSC.
with the increase in GO, the load of the particles on the GO layer decreased. This is be-
cause in the water–isopropanol system, GO is dispersed, the same content of particles are
dispersed on more GO, and the corresponding particles on the GO layer will be reduced.
This reduces the contact area between Al and CuO. The results show that the content of GO
had an important effect on the distribution of Al and CuO particles on the lamella. The effect
Molecules 2022, 27, 7614 6 of 17
of GO content on the heat release of the composites needs to be further analyzed using a
DSC.

Figure 6. SEM images of Al–CuO/GO with (a) 0 wt%, (b) 0.5 wt%, (c) 1.0 wt%, (d) 3.0 wt%, and (e)
Figure 6. SEM images of Al–CuO/GO with (a) 0 wt%, (b) 0.5 wt%, (c) 1.0 wt%, (d) 3.0 wt%, and
5.0 wt% of GO.
(e) 5.0 wt% of GO.

Figure 77 shows
Figure shows the
the DSC
DSC curves of the
curves of the composites
composites withwith different
different GO GO contents,
contents, andand the
the
parameters are shown in Table 2. This can be observed in the range
parameters are shown in Table 2. This can be observed in the range 25–1000 C, with one25–1000 ◦°C, with one
heat-absorbing peak
heat-absorbing and two
peak and two exothermic
exothermic peaks.
peaks. The exothermic peak
The exothermic peak atat around
around 190 190 ◦°C
C
disappeared, unlike the peak for GO content of 5% in Figure 5. For thermite
disappeared, unlike the peak for GO content of 5% in Figure 5. For thermite (GO: 0wt%), (GO: 0wt%),
there were
there were two
two exothermic
exothermic peaks
peaks atat 576.3
576.3 ◦°C
C and
and 727.2
727.2 °C.
◦ C. Aluminum
Aluminum melts melts at at 660
660 °C.
◦ C.
There was
There was aa solid–solid
solid–solid reaction
reaction peak
peak between
between Al Al NPs
NPs and
and CuO CuO NPs
NPs before
before thethe melting
melting
point of
point of aluminum.
aluminum.There Therewas
wasa solid–liquid
a solid–liquid reaction peak
reaction peakafter the the
after melting point
melting of alu-
point of
minum. Specifically, the solid–solid reaction peak occured between 550~580
aluminum. Specifically, the solid–solid reaction peak occured between 550~580 C, while °C, while
◦ the
solid–liquid
the reaction
solid–liquid reactionpeak occured
peak occuredbetween
between 700~900
700~900 ◦ C.
°C. It Itcan
canbebeseen
seenfrom
fromthe the Table
Table 22
that the lower the GO content, the greater the heat release of the composite. When the GO
was 0.5
content was 0.5 wt%,
wt%, thethe heat
heatrelease
releasereached
reached1567
1567J/g.
J/g. This
This waswas 1166
1166 J/g
J/g higher than that
thermite (GO:
of pure thermite (GO: 0wt%).
0wt%). There
There was
was aa 128
128 J/g
J/g increase
increase in heat release compared with
the nanocomposites prepared using the the single
single self-assembly
self-assembly method.
method. According to the
analysis results
SEM analysis results(Figure
(Figure5),
5),the
thelower
lowerthetheGO
GOcontent,
content,the themore
more particles
particles loaded
loaded onto
onto it.
This is why the heat release increased with the decrease in GO content. Moreover, with the
addition of GO, the exothermic peak of thermite reaction was advanced to varying degrees,
and the maximum advance was about 27 ◦ C.
The DSC curves indicate that the exothermic and exothermic efficiency of the com-
posite was much higher than that of pure thermite. For the purpose of understanding the
reaction mechanisms of other reactions of Al–CuO/GO MICs, a series of control experi-
ments were carried out.
Molecules 2022, 27, 7614 7 of 18

it. This is why the heat release increased with the decrease in GO content. Moreover, with
Molecules 2022, 27, 7614 7 of 17
the addition of GO, the exothermic peak of thermite reaction was advanced to varying
degrees, and the maximum advance was about 27 °C.

Figure 7.
Figure 7. DSC
DSC results of composites
results of composites with
with 0.5
0.5 wt%,
wt%, 1.0
1.0 wt%,
wt%, 3.0
3.0 wt%,
wt%, and
and 5.0
5.0 wt%
wt% of
of GO
GO at
at aa heating
heating
rate of 10 °C/min.

rate of 10 C/min.

Table 2. DSC parameters of Al–CuO/GO with 0 wt%, 0.5 wt%, 1.0 wt%, 3.0 wt%, and 5.0 wt% of
Table 2. DSC parameters of Al–CuO/GO with 0 wt%, 0.5 wt%, 1.0 wt%, 3.0 wt%, and 5.0 wt% of GO.
GO.
Sample Tpk1 ◦ C) (◦ C) (◦ C) ∆H (J/g)
Sample Tpk1( (°C) TT
pk2
pk2 (°C)
TTpk3
pk3 (°C) ΔH (J/g)
Al–CuO
Al–CuO // 576.3
576.3 727.2 400.4
Al–CuO/GO
Al–CuO/GO0.5wt%
0.5wt% 278.3
278.3 568.4
568.4 756.2
756.2 1567
1567
Al–CuO/GO
Al–CuO/GO 1wt%
1wt%
282.2
282.2 549.9
549.9 782.7
782.7 1235
1235
Al–CuO/GO
Al–CuO/GO 3wt%
3wt%
278.4
278.4 567.1
567.1 770.9
770.9 709.3
709.3
Al–CuO/GO5wt% 279.9 570.6 789.1 726.6
Al–CuO/GO5wt% 279.9 570.6 789.1 726.6

CuO–GO composites
The DSC curves werethat
indicate prepared using theand
the exothermic same method. The
exothermic raw GOofand
efficiency theCuO–
com-
GO composites
posite was muchwerehighertested
thanusing
that ofa DSC
pure (Figure 8).For
thermite. The results
the show
purpose that the peak tem-
of understanding the
peratures of the exothermic peaks of the GO and CuO–GO were 217.3 ◦ C and 270 ◦ C,
reaction mechanisms of other reactions of Al–CuO/GO MICs, a series of control experi-
respectively. This corresponds
ments were carried out. to peak 1 and peak 2 of Figure 5. It can be observed from
the DSC parameters that the peak
CuO–GO composites were prepared temperature of the
using the sameAl–CuO/GO
method. The MICs
raw was advanced
GO and CuO–
to ◦
GOa certain extent.
composites There
were is nousing
tested GO exothermic peak at8).about
a DSC (Figure The 190 C inshow
results Figure 7. This
that may
the peak
be due to the fact
temperatures that
of the not all of the
exothermic peaksGOofparticipates
the GO andinCuO–GO
the reactionwereof 217.3
forming composites
°C and 270 °C,
when the GO This
respectively. content is too high.toFigure
corresponds peak 19 and
shows the2XRD
peak pattern
of Figure 5. of the CuO–GO
It can be observedreaction
from
product. The diffraction peaks of the graphite, CuO, Cu O, and
the DSC parameters that the peak temperature of the Al–CuO/GO MICs was advanced Cu can be observed. It
Molecules 2022, 27, 7614 2
8 of 18to
can be inferred that some of the CuO reacted with the GO, while the other
a certain extent. There is no GO exothermic peak at about 190 °C in Figure 7. This may be part did not
participate
due to the in the
fact reaction.
that not all of the GO participates in the reaction of forming composites
when the GO content is too high. Figure 9 shows the XRD pattern of the CuO–GO reaction
product. The diffraction peaks of the graphite, CuO, Cu2O, and Cu can be observed. It can
be inferred that some of the CuO reacted with the GO, while the other part did not
participate in the reaction.

(a) (b)
Figure
Figure8.8.The
TheDSC
DSCcurves
curvesofof(a)
(a)GO
GOand
and(b)
(b)CuO–GO.
CuO–GO.


• • Cu2O
Cu
▼ CuO
.)
Molecules 2022, 27, 7614 (a) (b) 8 of 17

Figure 8. The DSC curves of (a) GO and (b) CuO–GO.


• • Cu2O
Cu
▼ CuO

Intensity(a.u.)
♦ C



• •
▼ ▼

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
2θ(°)

XRDpattern
Figure9.9.XRD
Figure patternofofthe
thereaction
reactionproducts
productsofofCuO–GO.
CuO–GO.

Accordingtotothe
According theprevious
previousanalysis,
analysis,we
wecan
canconclude
concludethat
thatthe
thethermal
thermalreaction
reactionofofthe
the
nanocomposites at about 570 ◦ C was not a single thermite reaction. Figure 10 shows the
nanocomposites at about 570 °C was not a single thermite reaction. Figure 10 shows the
reactionproducts
reaction productsofofthe
thenanocomposites
nanocomposites(CuO:Al
(CuO:Al= =1.5:1,
1.5:1,GO:
GO:0.5
0.5wt%)
wt%)asascharacterized
characterized
by XRD. Figure 10 shows the peaks of the CuAlO
by XRD. Figure 10 shows the peaks of the CuAlO2, Cu , Cu O, Cu, α-Al O
2 2O,2 Cu, α-Al2O32, and3 and
, γ-Alγ-Al O .
2O3. 2In 3
In addition, most alumina is likely to be amorphous, with wide peaks such
addition, most alumina is likely to be amorphous, with wide peaks such as the one dis- as the one
displayed in Figure 10. Depending on the products observed, the entire reaction can be
played in Figure 10. Depending on the products observed, the entire reaction can be de-
described by Equation (1). In addition, CuAlO can be created by the reactions of Cu2 O
scribed by Equation (1). In addition, CuAlO2 can 2be created by the reactions of Cu2O with
with Al2 O3 (as illustrated in Equation (2)) [39,40].
Al2O3 (as illustrated in Equation (2)) [39,40].

Molecules 2022, 27, 7614


+ 9CuO → 6Cu + Cu
5Al5Al+9CuO→6Cu+Cu O2 O2Al O 2 O3CuAlO
+ 2Al + CuAlO2 9 (1)
(1)
of 18

Cu O +Al
Cu 2 O →2CuAlO
Al2OO3 → 2CuAlO2 (2)(2)

■ CuAlO2

▼ Cu2O

◆ α-Al2O3
Intensity(a.u.)


● γ-Al2O3

▲ Cu





● ■

● ◆ ◆ ■ ◆●

5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80
2θ(degrees)

Figure
Figure10.
10.XRD
XRDpattern
pattern of the
the reaction
reactionproducts
productsofof
thethe nanocomposites
nanocomposites (CuO:Al
(CuO:Al = 1.5:1,
= 1.5:1, GO: GO: 0.5
0.5 wt%).
wt%).
2.2. Optimization of Experimental Conditions
2.2. Optimization of Experimental
There are many Conditions
factors affecting the crystalline growth of CuO nanoparticles, among
which
There are many factors affecting thethe
the amount of added water and holding time
crystalline areof
growth theCuO
most important conditions
nanoparticles, among
affecting the crystal growth. In this experiment, the composites were
which the amount of added water and the holding time are the most importantpreparedconditions
with differ-
ent holding
affecting times (1growth.
the crystal h) and deionized water content
In this experiment, (10 mL), which
the composites werewere singlewith
prepared variables.
dif-
ferent holding times (1 h) and deionized water content (10 mL), which were single varia-
bles. The crystalline growth state and morphology of the nanoparticles were investigated
and compared with the nanocomposites made under the previous conditions.
The XRD patterns of the nanocomposites with a holding time of 1 h and 10 mL of
deionized water are presented separately in Figure 11a,b. In Figure 11a, the XRD patterns
Molecules 2022, 27, 7614 9 of 17

The crystalline growth state and morphology of the nanoparticles were investigated and
compared with the nanocomposites made under the previous conditions.
The XRD patterns of the nanocomposites with a holding time of 1 h and 10 mL of
deionized water are presented separately in Figure 11a,b. In Figure 11a, the XRD patterns
demonstrate intense diffraction peaks at 38.5◦ , 44.7◦ , 65.1◦ , and 78.2◦ , indicating Al in the
nanocomposite. In addition, the diffraction peaks at 35.4◦ and 38.5◦ indicate CuO in the
nanocomposite. Compared with Figure 1, for which the holding time was 30 min, there was
no significant difference in the XRD patterns. By contrast, the XRD patterns in Figure 11b
demonstrate intense diffraction peaks at 36.4◦ , 42.3◦ , 61.3◦ , and 73.5◦ , indicating Cu2 O in
this nanocomposite. This may be due to the chemical reaction between the Al nanoparticles
and the excessive water forming Al(OH)3 and H2 , which reduces CuO to Cu2 O under
heating conditions. Aluminum hydroxide is unstable under heating conditions and further
decomposes into alumina. This process is represented by Equations (3)–(5). These results
suggest that the amount of water added has a significant influence on the formation of
the composites.
2Al + 6H2 O→2Al(OH)3 + 3H2 ↑ (3)

Molecules 2022, 27, 7614 2CuO + H2 →Cu2 O + H2 O 10 of 18(4)

2Al(OH)3 →Al2 O3 + H2 O (5)

(a) (b)
Figure 11.11.
Figure XRD patterns
XRD forfor
patterns (a)(a)
a holding time
a holding of of
time 1 h1 and (b)(b)
h and 10 10
mLmL
of of
deionized water.
deionized water.

From
From thethe
SEMSEManalysis
analysis(Figure
(Figure12a),
12a),ititcan
can be
be seen that the
the CuO
CuOparticles
particlesloaded
loadedon onthe
theGOGOlamella
lamella areare
obviously
obviously larger than
larger those
than those in Figure 2f, and
in Figure thatthat
2f, and the distribution uniformity
the distribution uni-
is poor.
formity is The
poor.reason for thisfor
The reason phenomenon
this phenomenonmay bemay thatbe
CuOthatparticles were loaded
CuO particles onto the
were loaded
active sites of the GO lamellae, forming nuclei and crystallizing growth
onto the active sites of the GO lamellae, forming nuclei and crystallizing growth in the in the reaction.
With the
reaction. increase
With in holding
the increase time, time,
in holding the CuO particles
the CuO grewgrew
particles up gradually, resulting
up gradually, in an
resulting
in uneven
an unevendistribution
distribution of CuO
of CuOparticles onon
particles thethe
GOGO surface.
surface.InInaddition,
addition,asascan
canbebeseen
seen in
in Figure
Figure 12b, the GO
12b, the GO lamellae
lamellaewerewerealso
alsothin,
thin,butbutthey
theywere
wereloaded
loadedwith
with bulk
bulk crystals
crystals that
that
agglomerated on the GO lamellae. From the XRD characterization of Figure 11b, it can bebe
agglomerated on the GO lamellae. From the XRD characterization of Figure 11b, it can
inferred
inferred that
that these
these bulk
bulk crystals
crystals may
may bebe
CuCu O crystals.
2O2 crystals.

Figure 12. SEM images for (a) a holding time of 1 h and (b) 10 mL of deionized water.
formity is poor. The reason for this phenomenon may be that CuO particles were loaded
onto the active sites of the GO lamellae, forming nuclei and crystallizing growth in the
reaction. With the increase in holding time, the CuO particles grew up gradually, resulting
in an uneven distribution of CuO particles on the GO surface. In addition, as can be seen
in Figure 12b, the GO lamellae were also thin, but they were loaded with bulk crystals that
Molecules 2022, 27, 7614 10 of 17
agglomerated on the GO lamellae. From the XRD characterization of Figure 11b, it can be
inferred that these bulk crystals may be Cu2O crystals.

Figure12.
Figure 12. SEM
SEM images
images for
for(a)
(a)aaholding
holdingtime
timeof
of11hhand
and(b)
(b)10
10mL
mLof
ofdeionized
deionizedwater.
water.

The crystallinity,
The crystallinity,grain
grainsize,
size,and
andpattern
patternof ofthe
thesample
samplewere
werefurther
furtherinspected
inspected using
using
TEM.
TEM. The
The TEM
TEM images
images ofof the
the nanocomposite
nanocomposite with with aa holding
holding time
time of
of 11 hh and
and 10
10 mL
mL of of
deionized
deionizedwater
waterarearedisplayed
displayedinin Figure
Figure13.13.
It can be be
It can observed
observedclearly thatthat
clearly the particles are
the particles
no
arelonger regular
no longer circles
regular (Figure
circles 13b),13b),
(Figure which
whichis due to the
is due to reaction between
the reaction between partpart
of the Al
of the
and the excessive water. Moreover, the particle size of the nanocomposites
Al and the excessive water. Moreover, the particle size of the nanocomposites in Figure in Figure 7a can
Molecules 2022, 27, 7614 be
7a estimated to be on to
can be estimated thebemicroscale, with particle
on the microscale, withsizes of approximately
particle 0.6 µm to 1.5
sizes of approximately 11 µm.
0.6 of
μm 18
This
to 1.5means that an
μm. This increase
means thatinanholding
increasetime will increase
in holding time the
willdiameter
increaseofthethediameter
inert particles.
of the
inert particles.

Figure13.
Figure 13.TEM
TEMimages
imagesfor
for(a)
(a)aaholding
holdingtime
timeof
of11hhand
and(b)
(b)10
10mL
mLof
ofdeionized
deionizedwater.
water.

Figure 14
Figure 14 shows the the DSC
DSCcurves
curvesobtained
obtainedatata aheat-up
heat-up rate of of
rate 10 10 ◦ C/min
°C/min in argon, and
in argon,
and the corresponding
the corresponding parameters
parameters are listed
are listed in Table
in Table 3. Compared
3. Compared with Figure
with Figure 7, the
7, the compo-
composites
sites with awith a holding
holding time of time
1 hof 1 h (Figure
(Figure 14a) showed
14a) showed no obvious
no obvious difference
difference in peakin peak
tem-
temperature. However,
perature. However, thethe
heatheat decreased
decreased with
with the increased
the increased holding
holding time.time. Asholding
As the the holding
time
time increased
increased fromfrom
30 min 30 to
min to the
1 h, 1 h,heat
the heat changed
changed fromfrom 1045.03
1045.03 J/g toJ/g to 652.34
652.34 J/g.means
J/g. This This
means
that thethat the increase
increase in crystalinsize
crystal
was notsizeconducive
was not conducive to the
to the reaction reaction
at this stage. at this stage.
Furthermore,
Furthermore, the composites with a dosage of deionized water of 10 mL
the composites with a dosage of deionized water of 10 mL had no peak 2 and peak 3, mean-had no peak 2 and
peak 3, meaning that there was no thermite reaction. This is the same as the
ing that there was no thermite reaction. This is the same as the findings for the XRD patterns findings for
the XRD11b).
(Figure patterns (Figure 11b).
sites with a holding time of 1 h (Figure 14a) showed no obvious difference in peak tem-
perature. However, the heat decreased with the increased holding time. As the holding time
increased from 30 min to 1 h, the heat changed from 1045.03 J/g to 652.34 J/g. This means
that the increase in crystal size was not conducive to the reaction at this stage. Furthermore,
the composites with a dosage of deionized water of 10 mL had no peak 2 and peak 3, mean-
Molecules 2022, 27, 7614 11 of 17
ing that there was no thermite reaction. This is the same as the findings for the XRD patterns
(Figure 11b).

Figure
Figure 14.
14. DSC
DSC plots
plots for
for (a)
(a) aa holding
holding time
time of
of 11 hh and
and (b)
(b) 10
10 mL
mL of
of deionized
deionized water.

Table 3. DSC
Table 3. DSC parameters
parameters of
of the
the nanocomposite at different
nanocomposite at different conditions
conditions in
in argon.
argon.

Sample
Sample TTpk1
pk1 (°C)
(◦ C) Tpk2T(°C) ◦ Tpk3 (°C)
Tpk3 (◦ C) Tpk4 (°C)(◦ C) ΔH
∆H(J/g)
pk2 ( C) Tpk4 (J/g)
Holding time is 1 h 191.4 278.1 575.3 898.4 652.3
Holding time is 1 h 191.4 278.1 575.3 898.4 652.3
Water content
Water content is is
1010
mLmL 181.6
181.6 / / / / 740.5740.5 115.2
115.2

2.3. Thermokinetics Study


2.3. Thermokinetics Study
Compared with other multiple heating rate methods, the American Standard Testing
Compared with other multiple heating rate methods, the American Standard Testing
Society (ASTM) E-689 method [37] is more commonly applied to identify the activation
Society (ASTM) E-689 method [37] is more commonly applied to identify the activation
energy, as in Equaiton (6):
energy, as in Equaiton (6):
 β⁄Tp = ln AR⁄E - E⁄RTP
2
ln (6)
2
ln β/Tp = ln(AR/E)−E/RTP (6)

where β is the rate of heating, TP is the peak temperature obtained from the DSC param-
eters, A is the pre-exponential factor, E is the activation energy, and R is the gas constant
(8.314 J/mol. K).
The ASTM E-689 method involves a free kinetics model, where the model necessitates
DSC data for at least three experiments at various heating rates. By constructing and
computing a least square “best fit” line using these points, the plot lgβ against 1/Tp was be
obtained, where Tp was the calibrated peak temperature in K. According to the designation
of E698-11, E and A are accessible from Equations (7) and (8), separately. The slope of the
“best fit” line was taken as the value of dlg(β)/d(1/Tp ).
" #
dlg(β)
E = − 2.19R  (7)
d 1/Tp
 
βE E
A= exp (8)
RT2p RTp

The thermal behavior of the Al–CuO/GO under various heating rates is depicted
in the DSC curves shown in Figure 15. As the heating rate grew, the DSC curves of the
Al–CuO/GO gradually moved towards higher temperatures. As is shown in Table 4, with
increasing heating rates, all characteristic temperatures will be raised, including the start
temperature (Te ), the peak temperature (Tp ), and the final temperature (Tf ).
RT2p RTp

The thermal behavior of the Al–CuO/GO under various heating rates is depicted in
the DSC curves shown in Figure 15. As the heating rate grew, the DSC curves of the Al–
CuO/GO gradually moved towards higher temperatures. As is shown in Table 4, with
Molecules 2022, 27, 7614 12 of 17
increasing heating rates, all characteristic temperatures will be raised, including the start
temperature (Te), the peak temperature (Tp), and the final temperature (Tf).

552.5℃ 5K/min
10K/min
15K/min
20K/min

DSC(mW/mg)
558.3℃
565.3℃

550.2℃

500 550 600 650

Temperature (°C)

Figure 15. DSC results of Al–CuO/GO at various heating rates.


Figure 15. DSC results of Al–CuO/GO at various heating rates.
Table 4. The characteristic temperatures and exothermic enthalpy of Al–CuO/GO at various heating rates.
Table 4. The characteristic temperatures and exothermic enthalpy of Al–CuO/GO at various heating
rates. β (K·min−1 ) Te (◦ C) Tp (◦ C) Tf (◦ C) ∆H (J/g)
5 −1 534.3 550.2 575.1
β (K·min ) Te (°C) Tp (°C) Tf (°C) ΔH 427.8
(J/g)
10 545.2 569.3 602.1 415.3
5 15 534.3
557.1 550.2
592.5 575.1
618.4 427.8
480.7
1020 545.2
569.0 569.3
598.3 602.1
624.1 415.3
598.5
15 557.1 592.5 618.4 480.7
20 process of applying
The 569.0 the ASTM E-689
598.3method to the624.1 598.5
DSC data by the use of software
Molecules 2022, 27, 7614 13 of 18
is depicted in Figure 16. The activation energy of the Al–CuO/GO obtained from the slope of
theThe
plotprocess
was 150of±applying
17 KJ/mol the ASTM
and E-689ofmethod
the value to the
ln(A) was alsoDSC data by
calculated asthe
6.85use
minof−soft-
1.
ware is depicted in Figure 16. The activation energy of the Al–CuO/GO obtained from the
slope of the plot was 150 ± 17 KJ/mol and theAl/CuO@GO
value of ln(A) was also calculated as 6.85 min−1.
1.4
log(Heating rate/(K/min))

1.2

1.0

0.8 Activation Energy(KJ/mol):150.35


log(Α/s-1):6.85

0.6
1.14 1.16 1.18 1.20 1.22
-1
1000/T(K )

FigureFigure
16. ASTM−698 −698 thermal
16. ASTMthermal decomposition
decomposition fit diagram
fit diagram of Al–CuO/GO
of Al–CuO/GO at four at four heating
heating rates. rates.

The thermal
The thermal reactions
reactions of energetic
of energetic materials
materials are interactions
are interactions thatthat occur
occur in solid
in solid hetero-
het-
geneous systems. The reaction progress for energy-containing composites
erogeneous systems. The reaction progress for energy-containing composites is not nor-is not normally
mallyobtained
obtainedininone
onestep,
step,but
butin
inmultiple
multiplesteps.
steps. There
There is
is aa complex relationship between
complex relationship between each
step. Each step may have a distinct dynamic model function. Hence, the conventional
each step. Each step may have a distinct dynamic model function. Hence, the conventional
methods are unable to accurately characterize the kinetics of this complicated system.
methods are unable to accurately characterize the kinetics of this complicated system.
Non-linear multiple regression is an important method for deriving kinetic models. It is a
Non-linear multiple regression is an important method for deriving kinetic models. It is a
unique way to determine between various reaction models and to obtain a global model
unique way to determine between various reaction models and to obtain a global model
that provides credible results over the entire range of parameters.
that provides credible results over the entire range of parameters.
As indicated in Figure 17, the simulation of the kinetic model using NETZSCH Ther-
mokinetics was well suited to the experimental data. The optimized values of the kinetic
parameters are presented in Table 5. The value of the correlation coefficient was 0.974.
Based on the calculated outcomes, the kinetic model for Al–CuO/GO was f(α) = (1 − α) n(1
erogeneous systems. The reaction progress for energy-containing composites is not nor-
mally obtained in one step, but in multiple steps. There is a complex relationship between
each step. Each step may have a distinct dynamic model function. Hence, the conventional
methods are unable to accurately characterize the kinetics of this complicated system.
Molecules 2022, 27, 7614 Non-linear multiple regression is an important method for deriving kinetic models. It13isofa17
unique way to determine between various reaction models and to obtain a global model
that provides credible results over the entire range of parameters.
AsAsindicated
indicatedin Figure
in Figure17, the
17, simulation of theofkinetic
the simulation model model
the kinetic using NETZSCH
using NETZSCHTher-
mokinetics was well suited to the experimental data. The optimized values
Thermokinetics was well suited to the experimental data. The optimized values of the of the kinetic
parameters are presented
kinetic parameters in Table 5.inThe
are presented value
Table 5. of thevalue
The correlation
of thecoefficient
correlationwas 0.974.
coefficient
Based on the calculated outcomes, the kinetic model for Al–CuO/GO
was 0.974. Based on the calculated outcomes, the kinetic model for Al–CuO/GO was f(α) = (1 − α) nwas
(1
+ f(α)
kcat·α),
= (1where
− α) k(1
n iskthe
cat + autocatalytic kinetic rate constant. The reaction order (n) of Al–
cat ·α), where kcat is the autocatalytic kinetic rate constant. The reaction
CuO/GO
order (n)was 1.20, and lgkcatwas
of Al–CuO/GO was 1.67.
1.20, and lgkcat was 1.67.

6
5K/min
5K/min
5 10K/min
10K/min
15K/min
Heat flow rate(W/g)

15K/min
4
20K/min
20K/min

525 550 575 600 625


T(℃)
Figure 17. DSC curves of Al–CuO/GO at four heating rates compared between experiments
Figure 17. DSC curves of Al–CuO/GO at four heating rates compared between experiments and
and simulations.
simulations.

Table 5. The optimal values of the kinetic factors of Al–CuO/GO obtained using the nonlinear
multivariate regression approach.

Kinetic Parameters Al–CuO/GO


lg A (s−1 ) 4.97
Eα (kJ·mol−1 ) 134.92
lgkca 1.67
Reaction order (n) 1.20
Correlation coefficient 0.974

2.4. Combustion Performance


Figure 18 shows the photographs of the combustion process of the Al–CuO/GO
with different GO contents. All samples were successfully ignited, and self-sustained
combustion occurred after laser irradiation. The burning rate of the samples was defined as
the average combustion rate from the beginning to the end of combustion (r = ∆x/∆t). Each
sample was measured at least three times, and the average value was taken as the burning
rate of the sample to ensure the reliability of the data. The results are shown in Figure 19. It
can be seen that the Al/CuO sample burned most vigorously and had the highest burning
rate. With increasing GO content, the burning rate of the nanocomposite tended to decrease.
The burning rate of the composite at 5wt% GO content was 80.6 m/s, which was 35.5%
lower than that of pure Al–CuO. This indicates that GO has a negative effect on the burning
rate of thermite. This result is very interesting. Because the high thermal conductivity
of carbon materials can improve the thermal conductivity of composites, this may be
an important factor affecting the combustion properties of the material. However, the
obtained experimental results showed the opposite. This may be due to the poor thermal
conductivity of rGO, a combustion product of GO. During the combustion process, rGO
decrease. The burning rate of the composite at 5wt% GO content was 80.6 m/s, which was
35.5% lower than that of pure Al–CuO. This indicates that GO has a negative effect on the
burning rate of thermite. This result is very interesting. Because the high thermal conduc-
tivity of carbon materials can improve the thermal conductivity of composites, this may
Molecules 2022, 27, 7614 be an important factor affecting the combustion properties of the material. However, the
14 of 17
obtained experimental results showed the opposite. This may be due to the poor thermal
conductivity of rGO, a combustion product of GO. During the combustion process, rGO
is
is deposited
deposited on
on the
the combustion
combustion end
end face,
face, which
which hinders
hinders the
the subsequent
subsequent convective
convective heat
heat
transfer process.
transfer process.

Molecules 2022, 27, 7614 Figure


Figure 18.
18. Flame
Flame propagation
propagation images
images of
of (a)
(a) Al–CuO,
Al–CuO, (b)
(b) Al–CuO/GO
Al–CuO/GO0.5wt% 15 ,of(d)
, (c) Al–CuO/GO1.0wt% 18
0.5wt% , (c) Al–CuO/GO1.0wt% ,
Al–CuO/GO 3.0wt%, and (e) Al–CuO/GO5.0wt%. The dotted line represents the burning end face.
(d) Al–CuO/GO3.0wt% , and (e) Al–CuO/GO5.0wt% . The dotted line represents the burning end face.

Figure 19. Burning rate of Al–CuO/GO


Al–CuO/GO MICs
MICswith
withdifferent
different GO
GO contents.
contents.

3. Experimental
3. Experimental
3.1. Materials and Reagents
3.1. Materials and Reagents
Nano aluminum (nano-Al, 50 nm, 99.9%, Aladdin, Shanghai, China) and GO (Aladdin,
Nano aluminum (nano-Al, 50 nm, 99.9%, Aladdin, Shanghai, China) and GO (Alad-
Shanghai, China) were purchased for this research. The active aluminum content was
din, Shanghai, China) were purchased for this research. The active aluminum content was
65.7%. The Isopropanol (C3 H8 O), ethanol (C2 H6 O), and Cupric Acetate Monohydrate
65.7%. The Isopropanol (C3H8O), ethanol (C2H6O), and Cupric Acetate Monohydrate
(Cu(CH3 COO)2 ·H2 O) were of analytical purity, so there was no need to reprocess them, and
(Cu(CH3COO)2·H2O) were of analytical purity, so there was no need to reprocess them,
they were bought from Sinopharm Chemical Reagent (Shanghai, China). Sodium dodecyl
and they
sulfate wereLing
(SDS, bought from
Feng Sinopharm
Chemical Chemical
Reagent, Reagent
Shanghai, (Shanghai,
China) servedChina).
as theSodium do-
dispersant.
decyl sulfate (SDS, Ling Feng Chemical Reagent, Shanghai, China)
Distilled water was utilized during the entire working process. served as the disper-
sant. Distilled water was utilized during the entire working process.
3.2. Preparation of the Al–CuO/GO
3.2. Preparation
In order toofobtain
the Al–CuO/GO
the GO-modified Al–CuO with the best performance, we investi-
gatedIna order
series to
of obtain the GO-modified
conditions Al–CuO
and their effects on thewith the best performance,
Al–CuO/GO we investi-
nanocomposites. These
gated a series
included of conditions
the equivalent ratioand their effects
of CuO:Al on2:1,
(2.5:1, the1.5:1,
Al–CuO/GO
and 1:1)nanocomposites.
and the content ofThese
GO
included
(0 wt%, 0.5thewt%,
equivalent
1 wt%,ratio of CuO:Al
3 wt%, (2.5:1,in
and 5 wt%), 2:1, 1.5:1,the
which andratio
1:1) and the content
of CuO and Al of GO to
refers (0
wt%,
the 0.5 wt%,
molar ratio.1At
wt%,
the 3same
wt%,time,
and since
5 wt%),
theinmorphology
which the ratio ofCuO
of the CuOparticles
and Al refers to the
was closely
molar ratio.
related Atexperimental
to the the same time, since the morphology
conditions, the effects ofofthetheholding
CuO particles wasthe
time and closely
amountre-
lated to the experimental conditions, the effects of the holding time and the amount of
added deionized water were studied. These were single variables during the experiment.
Typically, the copper acetate monohydrate (Cu(OAc)2•H2O), Al, and GO were dissolved
in 10 mL of isopropanol by sonicating for 30 min. Then, the Cu(OAc)2•H2O and Al dis-
persions were dispersed ultrasonically for 20 min. After this, the GO dispersion was
Molecules 2022, 27, 7614 15 of 17

of added deionized water were studied. These were single variables during the exper-
iment. Typically, the copper acetate monohydrate (Cu(OAc)2 •H2 O), Al, and GO were
dissolved in 10 mL of isopropanol by sonicating for 30 min. Then, the Cu(OAc)2 •H2 O and
Al dispersions were dispersed ultrasonically for 20 min. After this, the GO dispersion was
placed in a round-bottom flask (part of the reflux equipment) and the mixed dispersion of
Cu(OAc)2 •H2 O and Al was dropped into the GO dispersion under stirring. The mixture
was heated to about 80◦ C under vigorous stirring and held at that temperature for 40 min.
Then, 5 mL of deionized water was quickly poured into the above boiling solvent and
heated at 80 ◦ C for another 30 min. The solution was cooled to room temperature and
scoured with ethanol. After this, it was dried overnight under vacuum at 55 ◦ C.

3.3. Characterization of the Al–CuO/GO


The crystal structure was ascertained by X-ray diffraction (XRD) employing a Bruker
D8 Advance X-ray diffractometer with Cu-Kα radiation. The grazing angle was kept at 1◦
and the 2θ collection angle ranged between 5◦ and 85◦ in 0.02◦ steps with length of stay of
1 s per point. The samples were measured by transmission electron microscopy (TEM) with
a JEOL JEM-2100 microscope at 200 kV. One drop of the sample dispersion was deposited
onto a 300 mesh copper grid coated with a carbon layer. The SEM images were acquired
using a Quanta 250 field emission scanning electron microscope (FEI, America) outfitted
with an energy-dispersive X-ray spectroscopy (EDX) detector.
Differential scanning calorimetry (DSC) analyses were performed on a TA-DSC-Q20
in the range 25–1000 ◦ C at a heating rate of 10 ◦ C/min under a flow of argon. Based on
the DSC date, the thermokinetic properties of material were studied using NETZSCH
Thermokinetics 3. The kinetic parameters were calculated using the Kissinger Method. It
was feasible to determine the kinetic parameters, such as the pre-exponential factor (A),
activation energy (Ea), and kinetic model (f(α)), to completely describe the entire reaction.

3.4. Burning Rate Measurements


In order to study the effect of GO on the combustion performance of Al–CuO, the
sample was pressed into a 2 mm diameter quartz glass tube and ignited with a pulsed
laser (Nd: YAG, 60 mJ, 6.5 ns). The combustion process was recorded by high-speed
photography at 500,000 frames per second (fps). To ensure the reliability of the data, each
sample underwent the experimental process three times.

4. Conclusions
In summary, Al–CuO/GO nanocomposites were prepared successfully using a combina-
tion of the self-assembly and in-suit synthesis methods. When the equivalence ratio of CuO to
Al was 1.5:1 and the mass fraction of GO was 0.5%, the Al and CuO nanoparticles loaded uni-
formly on the GO sheets. The DSC results confirm that the heat release of the nanocomposite
with this formulation was 1566.7 J/g, which is approximately four times the heat released
by the pure Al–CuO, and the peak temperature dropped by 7.9–26.4 ◦ C. As the GO content
increased, the exothermic heat of the composites decreased, indicating that the excess GO
played a negative role in the exothermic reaction of the composites. The introduction of GO
changed the conventional Al–CuO reaction mechanism, and the reaction produced CuAlO2 .
The burning rate measurement results show that GO inhibits the combustion of Al–CuO. The
nanocomposite with 5wt% GO content had a 35.5% lower burning rate than pure Al–CuO.
The results show that the self-assembly structure of Al/CuO nanocomposites with GO sheets
has the advantages that its exothermic and combustion performances can be modified, which
may facilitate the practical application of nanothermites.

Author Contributions: Conceptualization, J.S. and Y.H.; methodology, J.S. and B.Z.; Methodology,
software, validation and formal analysis, J.S. and Y.Y.; writing—original draft preparation, J.S. and
R.S.; writing—review and editing, J.S. and Y.H. All authors have read and agreed to the published
version of the manuscript.
Molecules 2022, 27, 7614 16 of 17

Funding: This research was funded by the Natural Science Foundation of Jiangsu Province
(Grant No. SBK2014043634).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Acknowledgments: We thank the Micro-Nano Energetic Devices Key Laboratory.
Conflicts of Interest: The authors declare that they have no conflicts of interest.
Sample Availability: Samples of the compounds are available from the authors.

References
1. He, W.; Ao, W.; Yang, G.; Yang, Z.; Guo, Z.; Liu, P.-J.; Yan, Q.-L. Metastable energetic nanocomposites of MOF-activated aluminum
featured with multi-level energy releases. Chem. Eng. J. 2019, 381, 122623. [CrossRef]
2. Rossi, C.; Zhang, K.; Esteve, D.; Alphonse, P.; Tailhades, P.; Vahlas, C. Nanoenergetic Materials for MEMS: A Review. J. Microelec-
tromechanical Syst. 2007, 16, 919–931. [CrossRef]
3. Shende, R.; Subramanian, S.; Hasan, S.; Apperson, S.; Thiruvengadathan, R.; Gangopadhyay, K.; Gangopadhyay, S.; Redner, P.;
Kapoor, D.; Nicolich, S.; et al. Nanoenergetic Composites of CuO Nanorods, Nanowires, and Al-Nanoparticles. Propellants Explos.
Pyrotech. 2008, 33, 122–130. [CrossRef]
4. Yan, Q.-L.; Cohen, A.; Petrutik, N.; Shlomovich, A.; Burstein, L.; Pang, S.-P.; Gozin, M. Highly insensitive and thermostable
energetic coordination nanomaterials based on functionalized graphene oxides. J. Mater. Chem. A 2016, 4, 9941–9948. [CrossRef]
5. He, W.; Tao, B.; Yang, Z.; Yang, G.; Guo, X.; Liu, P.-J.; Yan, Q.-L. Mussel-inspired polydopamine-directed crystal growth of
core-shell n-Al@PDA@CuO metastable intermixed composites. Chem. Eng. J. 2019, 369, 1093–1101. [CrossRef]
6. Tao, Y.; Zhang, J.; Yang, Y.; Wu, H.; Hu, L.; Dong, X.; Lu, J.; Guo, S. Metastable intermolecular composites of Al and CuO
nanoparticles assembled with graphene quantum dots. RSC Adv. 2017, 7, 1718–1723. [CrossRef]
7. Yan, Q.-L.; Zhao, F.-Q.; Kuo, K.K.; Zhang, X.-H.; Zeman, S.; DeLuca, L.T. Catalytic effects of nano additives on decomposition and
combustion of RDX-, HMX-, and AP-based energetic compositions. Prog. Energy Combust. Sci. 2016, 57, 75–136. [CrossRef]
8. Liu, Y.; Gao, B.; Qiao, Z.; Hu, Y.; Zheng, W.; Zhang, L.; Zhou, Y.; Ji, G.; Yang, G. Gram-Scale Synthesis of Graphene Quantum Dots
from Single Carbon Atoms Growth via Energetic Material Deflagration. Chem. Mater. 2015, 27, 4319–4327. [CrossRef]
9. Lyu, J.-Y.; Chen, S.; He, W.; Zhang, X.-X.; Tang, D.-Y.; Liu, P.-J.; Yan, Q.-L. Fabrication of high-performance graphene oxide doped
PVDF/CuO/Al nanocomposites via electrospinning. Chem. Eng. J. 2019, 368, 129–137. [CrossRef]
10. 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. [CrossRef]
11. Pantoya, M.L.; Granier, J.J. Combustion Behavior of Highly Energetic Thermites: Nano versus Micron Composites. Propellants
Explos. Pyrotech. 2005, 30, 53–62. [CrossRef]
12. Piercey, D.G.; Klapooke, T.M. Nanoscale Aluminum-Metal Oxide (Thermite) Reactions for Application in Energetic Materials.
Cent. Eur. J. Energ. Mat. 2010, 7, 115–129.
13. Patel, V.K.; Bhattacharya, S. High-Performance Nanothermite Composites Based on Aloe-Vera-Directed CuO Nanorods. ACS
Appl. Mater. Interfaces 2013, 5, 13364–13374. [CrossRef] [PubMed]
14. Kim, J.H.; Cha, J.K.; Cho, M.H.; Kim, H.; Shim, H.-M.; Kim, S.H. Thermal reactions of nitrocellulose-encapsulated Al/CuO
nanoenergetic materials fabricated in the gas and liquid phases. Mater. Chem. Phys. 2019, 238, 121955. [CrossRef]
15. Marín, L.; Gao, Y.; Vallet, M.; Abdallah, I.; Warot-Fonrose, B.; Tenailleau, C.; Lucero, A.T.; Kim, J.; Esteve, A.; Chabal, Y.J.; et al.
Performance Enhancement via Incorporation of ZnO Nanolayers in Energetic Al/CuO Multilayers. Langmuir 2017, 33, 11086–
11093. [CrossRef] [PubMed]
16. Son, S.; Yetter, R.A.; Yang, V. Introduction: Nanoscale Composite Energetic Materials. J. Propuls. Power 2007, 23, 643–644.
[CrossRef]
17. Sanders, V.E.; Asay, B.W.; Foley, T.J.; Tappan, B.C.; Pacheco, A.N.; Son, S. Reaction Propagation of Four Nanoscale Energetic
Composites (Al/MoO3, Al/WO3, Al/CuO, and B12O3). J. Propuls. Power 2007, 23, 707–714. [CrossRef]
18. Rossi, C.; Esteve, A.; Vashishta, P. Nanoscale energetic materials. J. Phys. Chem. Solids 2010, 71, 57–58. [CrossRef]
19. Salvagnac, L.; Assie-Souleille, S.; Rossi, C. Layered Al/CuO Thin Films for Tunable Ignition and Actuations. Nanomaterials 2020,
10, 2009. [CrossRef]
20. Huang, C.; Yang, Z.; Li, Y.; Zheng, B.; Yan, Q.; Guan, L.; Luo, G.; Li, S.; Nie, F. Incorporation of high explosives into nano-aluminum
based microspheres to improve reactivity. Chem. Eng. J. 2019, 383, 123110. [CrossRef]
21. Kwon, J.; Ducéré, J.M.; Alphonse, P.; Bahrami, M.; Petrantoni, M.; Veyan, J.-F.; Tenailleau, C.; Estève, A.; Rossi, C.; Chabal, Y.J.
Interfacial Chemistry in Al/CuO Reactive Nanomaterial and Its Role in Exothermic Reaction. ACS Appl. Mater. Interfaces 2013, 5,
605–613. [CrossRef] [PubMed]
22. Chen, Y.; Ren, W.; Zheng, Z.; Wu, G.; Hu, B.; Chen, J.; Wang, J.; Yu, C.; Ma, K.; Zhou, X.; et al. Reactivity adjustment from the
contact extent between CuO and Al phases in nanothermites. Chem. Eng. J. 2020, 402, 126288. [CrossRef]
Molecules 2022, 27, 7614 17 of 17

23. Zhang, T.; Li, C.; Wang, W.; Guo, Z.; Pang, A.; Ma, H. Construction of Three-Dimensional Hematite/Graphene with Effective
Catalytic Activity for the Thermal Decomposition of CL-20. Acta Phys. -Chim. Sin. 2020, 36, 1905048.
24. Zhang, J.; Zhao, F.; Yang, Y.; Yan, Q.; Zhang, M.; Ma, W. Enhanced catalytic performance on the thermal decomposition of TKX-50
by Fe3O4 nanoparticles highly dispersed on rGO. J. Therm. Anal. 2019, 140, 1759–1767. [CrossRef]
25. Elbasuney, S.; El-Sayyad, G.S.; Yehia, M.; Aal, S.K.A. Facile synthesis of RGO-Fe2O3 nanocomposite: A novel catalyzing agent for
composite propellants. J. Mater. Sci. Mater. Electron. 2020, 31, 20805–20815. [CrossRef]
26. Tan, L.; Lu, X.; Liu, N.; Yan, Q.-L. Further enhancing thermal stability of thermostable energetic derivatives of dibenzotetraaza-
pentene by polydopamine/graphene oxide coating. Appl. Surf. Sci. 2020, 543, 148825. [CrossRef]
27. Liu, C.; Li, X.; Li, R.; Yang, Q.; Zhang, H.; Yang, B.; Yang, G. Laser ignited combustion of graphene oxide/nitrocellulose
membranes for solid propellant micro thruster and solar water distillation. Carbon 2020, 166, 138–147. [CrossRef]
28. Li, R.; Wang, J.; Shen, J.P.; Hua, C.; Yang, G.C. Preparation and Characterization of Insensitive HMX/Graphene Oxide Composites.
Propellants Explos. Pyrotech. 2013, 38, 798–804. [CrossRef]
29. Abdolmaleki, M.; Sari, M.G.; Rostami, M.; Ramezanzadeh, B. Graphene oxide nanoflakes as an efficient dispersing agent for
nanoclay lamellae in an epoxy-phenolic nanocomposite coating: Mechanistic approach. Compos. Sci. Technol. 2019, 184, 107879.
[CrossRef]
30. McCrary, P.D.; Beasley, P.A.; Alaniz, S.A.; Griggs, C.S.; Frazier, R.M.; Rogers, R.D. Graphene and Graphene Oxide Can “Lubricate”
Ionic Liquids based on Specific Surface Interactions Leading to Improved Low-Temperature Hypergolic Performance. Angew.
Chem. Int. Ed. 2012, 51, 9784–9787. [CrossRef]
31. Krishnan, D.; Kim, F.; Luo, J.; Cruz-Silva, R.; Cote, L.J.; Jang, H.D.; Huang, J. Energetic graphene oxide: Challenges and
opportunities. Nano Today 2012, 7, 137–152. [CrossRef]
32. Abdollahi, H.; Samkan, M.; Hashemi, M.M. Fabrication of rGO nano-sheets wrapped on Ni doped ZnO nanowire p–n heterostruc-
tures for hydrogen gas sensing. New J. Chem. 2019, 43, 19253–19264. [CrossRef]
33. Wang, S.; Zhao, W.; Xie, L.; Cao, H.; Huang, W. Solution-processable GO/rGO: Preparation, functionalization, self-assembly and
applications in smart information devices. Chin. Sci. Bull. 2019, 64, 2689–2702. [CrossRef]
34. Bartolucci, S.F.; Paras, J.; Rafiee, M.A.; Rafiee, J.; Lee, S.; Kapoor, D.; Koratkar, N. Graphene–aluminum nanocomposites. Mater.
Sci. Eng. A 2011, 528, 7933–7937. [CrossRef]
35. Zhang, W.; Yang, Y.; Ziemann, E.; Be’Er, A.; Bashouti, M.Y.; Elimelech, M.; Bernstein, R. One-step sonochemical synthesis of
a reduced graphene oxide—ZnO nanocomposite with antibacterial and antibiofouling properties. Environ. Sci. Nano 2019, 6,
3080–3090. [CrossRef]
36. Zeynizadeh, B.; Gilanizadeh, M. Synthesis and characterization of a magnetic graphene oxide/Zn–Ni–Fe layered double
hydroxide nanocomposite: An efficient mesoporous catalyst for the green preparation of biscoumarins. New J. Chem. 2019, 43,
18794–18804. [CrossRef]
37. Cohen, A.; Yang, Y.; Yan, Q.-L.; Shlomovich, A.; Petrutik, N.; Burstein, L.; Pang, S.-P.; Gozin, M. Highly Thermostable and
Insensitive Energetic Hybrid Coordination Polymers Based on Graphene Oxide–Cu(II) Complex. Chem. Mater. 2016, 28, 6118–6126.
[CrossRef]
38. Thiruvengadathan, R.; Chung, S.W.; Basuray, S.; Balasubramanian, B.; Staley, C.S.; Gangopadhyay, K.; Gangopadhyay, S.
A Versatile Self-Assembly Approach toward High Performance Nanoenergetic Composite Using Functionalized Graphene.
Langmuir 2014, 30, 6556–6564. [CrossRef]
39. Lei, H.; Gu, Q. Preparation of Cu-doped colloidal SiO2 abrasives and their chemical mechanical polishing behavior on sapphire
substrates. J. Mater. Sci. Mater. Electron. 2015, 26, 10194–10200. [CrossRef]
40. Shy, J.H.; Tseng, B.H. Characterization of CuAlO2 thin film prepared by rapid thermal annealing of an Al2O3/Cu2O/sapphire
structure. J. Phys. Chem. Solids 2005, 66, 2123–2126. [CrossRef]

You might also like