You are on page 1of 9

www.acsami.

org Research Article

Enhancing the Performance of Rubber with Nano ZnO as Activators


Xuan Qin, Haoshu Xu, Ganggang Zhang, Jiadong Wang, Zhao Wang,* Yuqi Zhao, Zongyu Wang,*
Tianwei Tan, Michael R. Bockstaller, Liqun Zhang, and Krzysztof Matyjaszewski*
Cite This: https://dx.doi.org/10.1021/acsami.0c15114 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information

ABSTRACT: The vulcanization of rubber is a chemical process to


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.
Downloaded via AUCKLAND UNIV OF TECHNOLOGY on October 12, 2020 at 20:25:39 (UTC).

improve the mechanical properties by cross-linking unsaturated polymer


chains. Zinc oxide (ZnO) acts as an activator, boosting the rubbers’ sulfur
vulcanization. Maintaining the level of ZnO content in the rubber
compounds as low as possible is desirable, not only for economic reasons
but also to reduce the environmental footprint of the process. In this
contribution, octylamine (OA) capped ZnO nanoparticles (5 nm
diameter), prepared through a thermal decomposition method, were
demonstrated to be efficient activators for the sulfur vulcanization of
natural rubber, enabling the reduction of the required amount of ZnO as
compared to commercial systems. The effect of different ZnO activators
(OA capped ZnO/commercial indirect process ZnO) on the curing
characteristics, cross-linking densities, and mechanical performance, as
well as the thermal behavior of rubber compounds, were investigated. Compared to the commercial indirect process ZnO, OA
capped ZnO nanoparticles not only effectively enhanced the curing efficiency of natural rubber but also improved the mechanical
performance of the composites after vulcanization. This was interpreted as, by applying the OA capped ZnO nanoparticles, the ZnO
levels in rubber compounding were significantly reduced under the industrial vulcanization condition (151 °C, 30 min).
KEYWORDS: ZnO, vulcanization, natural rubber, styrene−butadiene rubber, mechanical performance

■ INTRODUCTION
“Rubbers” are a member of a class of polymers known as
all different activators, ZnO is believed to be the most
efficient.14 By improving the kinetics of the curing process and
“elastomers”, i.e., polymers that are characterized by large promoting the short sulfide cross-link formation, activator
strain, low-modulus extensibility, and instantaneous and (ZnO) plays a prominent role in the first steps of vulcanization
complete recovery after removal of the load. After stretching to achieve higher cross-linking densities.10 As a powerful
and deformation, they can return to the original shape.1−4 The activator, ZnO reacts with accelerators and forms particularly
rubber retains an exclusive place in modern technology, due to active zinc chelates, which subsequently interact with sulfur,
its exceptional strength and tack in the prevulcanized state, and resulting in zinc polysulfide complexes, and then further bond
advanced crack-growth resistance and excellent mechanical to the rubber precursor chains to generate cross-linked
performance once vulcanized.5,6 The vulcanization process by structures.15−17 Only the Zn2+ attached to the surface of
which rubber is heated with sulfur and other additives to create ZnO nanoparticles can attend and accelerate the reac-
a chemically cross-linked network was invented in 1839 by tions.18−20 The contact and interaction between the activator
Charles Goodyear.7−9 The unaccelerated vulcanization proce- particles (ZnO) and the accelerators are greatly affected by the
dure applied elemental sulfur at 8 parts per 100 parts of rubber size, specific surface area, and crystal structure of the ZnO
(8 phr) and an essential temperature of 143 °C for 6 h. It is particles, as well as their dispersibility in the polymer
desirable to accelerate the rate of vulcanization to lower matrix.21,22 To optimize the performance of the vulcanized
energy/resource consumption and reach higher productiv- rubber, usually 3−5 phr ZnO nanoparticles are used in the
ity.10,11 By adding a few parts of the organic accelerators, the formula. One rising concern regarding the potential ecological
vulcanization time was reduced massively.12 Further, cross-
linking formed a 3-D network structure, mainly coming from
single, bisulfide, and polysulfide bonds, the yield of which Received: August 21, 2020
depended on the quantities and forms of the vulcanization Accepted: October 2, 2020
reagents and additives, such as the rubber precursors, activators
(ZnO), fatty acids, accelerators, and antioxidants.13
Nowadays, ZnO is widely employed in global rubber
commercial manufacturing, for sulfur vulcanization. Among

© XXXX American Chemical Society https://dx.doi.org/10.1021/acsami.0c15114


A ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Scheme 1. Schematic Presentation for the Preparation Procedure of OA Capped ZnO Nanoparticles and OA Capped ZnO
Nanoparticles/Rubber Composites

and human health impacts is releasing zinc complexes and their the control experiments, the commercial indirect process ZnO
derivatives into the ecosystem. Releasing of zinc from rubber nanoparticles (average size 210 ± 10 nm by TEM, specific
into the environment occurs during manufacturing, decom- surface area ∼9.6 m2/g), termed ZnO-C, were used. Their
posing, and recycling of rubber commodity, as well as during properties and performance were compared with ZnO-OA.
service conditions like leaching in landfill sites and the wear of
tires.23−29 To judiciously reduce ZnO levels in rubber
compounding, a fundamental requirement is to get a deeper
■ RESULTS AND DISCUSSION
Many previous works have demonstrated that nanosized ZnO
insight into the mechanistic details of sulfur vulcanization and particles can effectively reduce the ZnO filling content in the
to achieve both a smaller pollution and also an economic rubber compound and meanwhile cut the curing time of
profit. vulcanization.23,44 However, due to the physical attraction to
One factor impacting the activity of the activator is its align along with the specific crystal patterns and easy
specific surface area. Nano ZnO is an outstanding substitute aggregation, the large-scale synthesis of homogeneously
for conventional ZnO due to its defined crystal structure and isolated and uniform-sized ZnO nanoparticles is challeng-
high specific surface area.30−33 Nano ZnO nanoparticles are ing.34,45 Herein, the OA capped ZnO nanoparticles were
typically prepared with specific surfaces ranging from 30 to 70 prepared via a controlled thermal decomposition method of
m2/g, compared to ∼6.0 m2/g for traditional commercial ZnO. zinc(II) 2-ethylhexanoate (Zn(EH)2) in the presence of
Since nano ZnO particles have a higher specific surface area, a surfactant octylamine.43,46,47 The purified OA capped ZnO
larger amount of zinc is available. It was reported that the nanoparticles were dispersed in THF and retained stable up to
comparable substitution of micron-sized ZnO by nano ZnO months. The product solutions were studied by TEM and DLS
particles results in an improvement of the rubber compound (Figure S2), which presented nanocrystals with an average
properties, in particular the abrasion resistance and tear diameter of ca. 5 nm with low dispersity. Meanwhile, the
strength.33−36 weight percent of ZnO in nanoparticles was studied by TGA,
We have recently prepared hybrid materials with ZnO which yielded an 83 wt % inorganic content (Figure S2). The
nanoparticles (with a size down to 5 nm) by block copolymer XRD diffraction patterns for the commercial indirect process
templating and by the evaporative ligand exchange proc- ZnO nanoparticles and OA capped ZnO nanoparticles are
ess.37−43 In this contribution, 5 nm ZnO nanoparticles capped shown in Figure S4, which confirmed the excellent crystallinity
with octylamine (surface area ∼40.3 m2/g), termed ZnO-OA, of ZnO nanoparticles in agreement with the wurtzite ZnO
were prepared and used as activators in the sulfur vulcanization crystallinity pattern (ICSD-01-079-0206).48 Peak broadening
of natural rubber and styrene−butadiene rubber (see Scheme that can be discerned in the spectra of OA capped ZnO
1). The octylamine ligands effectively prevented the agglom- nanoparticles suggests that the size of individual crystalline
eration of nanoparticles, forming the homogeneous dispersion nanoparticles was fairly small.
of OA capped ZnO nanoparticles in the rubber matrix. The The dispersion state of the ZnO activator particle fillers in
crystallinity and pore features of the OA capped ZnO the rubber compound has a great impact on the activation of
nanoparticles were investigated with X-ray diffraction (XRD) the sulfur vulcanization as well as the mechanical performance
and N2 adsorption/desorption (BET test), and the morphol- of the materials. The aggregation of activator particles (ZnO)
ogy and structure were characterized by transmission electron diminishes their contact surface and interactions with other
microscopy (TEM) and dynamic light scattering (DLS). The modules in the cross-linking system, resulting in a decrease of
dispersion state of the fillers in the rubber matrix was the vulcanization efficiency. To investigate the composites’
investigated using scanning electron microscopy (SEM) and morphology, SEM and TEM were utilized to evaluate the
TEM. After adding OA capped ZnO into the rubber dispersion degree of additives and ZnO-OA/ZnO-C fillers in
compounds, the curing efficiency and the performance of the the rubber compounds. SEM was applied to image the brittle
product after vulcanization were significantly improved. For surfaces (which were triggered by liquid nitrogen) of ZnO-C/
B https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 1. SEM images of ZnO nanoparticles/rubber mixtures after vulcanization: (a, b) commercial indirect process ZnO nanoparticles with
natural rubber, (c, d) OA capped ZnO nanoparticles with natural rubber, (e, f) commercial indirect process ZnO nanoparticles with styrene−
butadiene rubber, (g, h) OA capped ZnO nanoparticles with styrene−butadiene rubber. ZnO contents: 2 phr. Scale bars: (a, c, e, g) 50 μm; (b, d, f,
h) 2 μm.

natural rubber, ZnO-OA/natural rubber, ZnO-C/styrene−


butadiene, and ZnO-OA/styrene−butadiene composites (2
phr ZnO contents), the images are shown in Figure 1. ZnO
nanoparticle agglomerates are evident in materials that were
fabricated using the commercial indirect process ZnO
nanoparticles (Figure 1a,b and e,f), which is attributed to its
larger particle size nature. Based on the proposed mecha-
nism,23,49 the activator’s (ZnO) surface works as a platform
and template in the catalytic reactions to activate and join the
reactants. Additives (sulfur, fatty acid, antioxidant, and
accelerator) diffuse through the mixture of the compounds
and are adsorbed onto the surface of ZnO nanoparticles,
forming reactive intermediate Zn chelate complexes. Con-
sequently, the contact and interaction between activator
particles (ZnO) and other additives in the rubber compounds
should be maximized to increase the vulcanization efficacy.
Generally, it is believed that this contact is highly affected by
the particle size and shape of the activator. Measured by BET,
the surface areas of the commercial indirect process ZnO and
OA capped ZnO nanoparticles are 9.6 and 40.3 m2/g (Figure Figure 2. TEM images of ZnO nanoparticles/rubber mixtures after
S5), respectively. With much larger surface areas and smaller vulcanization: (a) commercial indirect process ZnO with natural
rubber, (b) OA capped ZnO nanoparticles with natural rubber, (c)
particle size, OA capped ZnO nanoparticles are expected to commercial indirect process ZnO nanoparticles with styrene−
have better dispersion in both rubber matrixes. As shown in butadiene rubber, (d) OA capped ZnO nanoparticles with styrene−
Figure 1c,d and g,h, ZnO-OA/rubber composites have butadiene rubber. ZnO contents: 2 phr. Scale bars: 500 nm.
smoother surfaces, which indicates a more uniform dispersion
compared to commercial indirect process ZnO composites. easier zinc ion release and the following formation of the zinc−
More detailed morphological structures of the ZnO/rubber accelerator complex along with the curing procedure.
composites were studied by TEM. In agreement with SEM Figure 3 displays the vulcanization curves for the ZnO/
results, the TEM images demonstrate the presence of large rubber composites. Table 1 summarizes the curing parameters
agglomerates of particles in ZnO-C/rubber composites, sized acquired from the vulcanization curves of the ZnO/rubber
from 500 nm to several microns, Figure 2a and c. The OA composites. In the ZnO/natural rubber system, as shown in
capped ZnO nanoparticles can be rather evenly distributed in Figure 3a, with a low ZnO content (0.5 phr OA capped ZnO
natural rubber matrix; slight agglomerates 50−100 nm in size nanoparticles and 0.5/1.0 phr commercial indirect process
consisting of nanosized (∼5 nm) primary nanoparticles were ZnO nanoparticles), after reaching the maximum torque (MH),
formed. Besides, styrene−butadiene rubber exhibits better the vulcanization curves decline sharply as the curing time
compatibility with OA capped ZnO nanoparticles than natural increases. It is generally believed that it is triggered by the
rubber matrix (Figures 1d,h and 2b,d). These characteristics of formation and successive dissociation of certain polysulfide
the small dimensions of particles and improved dispersion in cross-links; this phenomenon is termed reversion.50,51 Besides,
the matrix of OA capped ZnO nanoparticles should favor an Table 1 and Figure 3a show the comparison between
C https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 3. Vulcanization curves of the ZnO nanoparticles/rubber composites: (a) ZnO/natural rubber, (b) ZnO/styrene−butadiene rubber. Black
curves (a) and blue curves (b) represent commercial indirect process ZnO nanoparticles, and red curves (a) and orange curves (b) represent OA
capped ZnO nanoparticles. The symbols ■, ●, ▲, and ◆ represent 0.5, 1, 2, and 3 phr for ZnO, respectively.

Table 1. Curing Parameters of the ZnO/Rubber 3b and Table 1, with the same amount of ZnO nanoparticles,
Composites OA capped ZnO composites exhibit a much shorter curing
time t90. Generally, short curing time means low energy
min max curing
torque torque time consumption and high production efficiency. On the other
(ML)c (MH)c (t90)c ΔM = MH − MLc hand, OA capped ZnO composites showed lower MH,
entrya (kPa) (kPa) (min) (kPa) especially under low ZnO content conditions (0.5/1.0 phr).
Natural Rubberb This indicates a lower degree of rubber reticulation in the
ZnO-C 0.5 phr 0.39 7.75 7.33 7.36 ZnO-OA/styrene−butadiene rubber composites.
ZnO-C 1.0 phr 0.57 9.08 8.72 8.51 To determine and compare the physical properties of the
ZnO-C 2.0 phr 0.71 10.98 11.4 10.27 prepared composites, which will govern the ultimate
ZnO-C 3.0 phr 0.83 11.40 12.45 10.57 application, the mechanical property analyses of cured ZnO-
ZnO-OA 0.5 phr 0.71 8.34 6.35 7.63 C and ZnO-OA/rubber composites were conducted to
ZnO-OA 1.0 phr 0.74 10.55 9.82 9.81 investigate the impact of the different curing activators on
ZnO-OA 2.0 phr 0.83 12.25 11.55 11.42 the mechanical performance of the materials. The effect of the
ZnO-OA 3.0 phr 0.56 13.11 11.45 12.55 commercial indirect process ZnO and OA capped ZnO
Styrene−Butadiene Rubberb nanoparticle contents on the mechanical properties of the
ZnO-C 0.5 phr 1.31 11.80 24.13 10.49 natural rubber/styrene−butadiene rubber composites is
ZnO-C 1.0 phr 2.18 13.09 23.1 10.91 illustrated in Figure 4. From Figure 4c, it can be seen that
ZnO-C 2.0 phr 2.01 12.42 24.35 10.41 the tensile strength increased with the content of ZnO in
ZnO-C 3.0 phr 1.99 11.95 25.75 9.96 natural rubber composites. This can be interpreted as a
ZnO-OA 0.5 phr 1.82 7.16 15.35 5.34 consequence of the higher cross-linking density of the network.
ZnO-OA 1.0 phr 1.67 6.14 19.33 4.47 Compared with the ZnO-C/natural rubber composites, the
ZnO-OA 2.0 phr 2.06 10.82 15.95 8.76 elongation at the break and the tensile strength of ZnO-OA/
a
ZnO-OA 3.0 phr 2.14 11.02 13.27 8.88
natural rubber composites were higher at all filling levels (0.5/
Rubber (natural rubber/styrene−butadiene rubber) 100 phr, ZnO 1.0/2.0/3.0 phr). This is attributed to the improved dispersion
nanoparticles (commercial indirect process ZnO nanoparticles/OA of the ZnO with a smaller size in the natural rubber matrix,
capped ZnO nanoparticles) 0.5/1/2/3 phr, sulfur 1.1 phr, stearic acid which results in the enhancement of cross-linking density. The
2 phr, carbon black N330 45 phr, antioxidant 4020 1.5 phr,
accelerator NS 1.5 phr. bThe vulcanization conditions for natural
enhancement of vulcanization efficiency from uniform OA
rubber and styrene−butadiene rubber are 143 °C with curing time t90 capped ZnO can contribute to the improvement of the
+ 3 min and 151 °C with curing time t90 + 3 min, respectively. cML, mechanical performance. Additionally, although styrene−
the minimum torque; MH, the maximum torque; t90, the optimum butadiene rubber has better compatibility with OA capped
curing time. ZnO nanoparticles, no clear trends between the ZnO contents
and mechanical properties (elongation at break, tensile
commercial indirect process ZnO nanoparticles and OA strength, modulus) of the ZnO-OA/styrene−butadiene rubber
capped ZnO nanoparticles in the vulcanization of natural composites were observed (Figure 4d, Table S2). Compared
rubber. A clear trend of MH increasing with increased ZnO with commercial indirect process ZnO nanoparticles, at low
contents is observed in ZnO/natural rubber composites, which ZnO content filling level (0.5 phr), there is a significant
is contributed by the activator effect of ZnO. OA capped ZnO increase in the tensile strength in OA capped ZnO nano-
composites evidence a curing time very similar to t90 and a particle composites (Figure 4d). However, after considering
higher MH than commercial indirect process ZnO ones the MH decrease and random fluctuations at higher ZnO
through holding the same amount of ZnO, which indicates a content filling level, there is no evidence to support OA capped
higher curing efficiency and more rubber reticulation in ZnO- ZnO nanoparticles have a more positive effect on the
OA/natural rubber composites. These results indicate the vulcanization of styrene−butadiene rubber compared with
faster curing mechanism with the incorporation of OA capped commercial indirect process ZnO nanoparticles. One possible
ZnO nanoparticles in the natural rubber compounds. On the explanation is the large difference in the number of allylic
contrary, in the styrene−butadiene system, as shown in Figure hydrogens in each repeat unit between natural rubber and
D https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 4. (a) Stress−strain curves of ZnO nanoparticles/natural rubber nanocomposites vulcanized at 143 °C with curing time t90 + 3 min and (b)
stress−strain curves of ZnO nanoparticles/styrene−butadiene rubber nanocomposites vulcanized at 151 °C with curing time t90 + 3 min. (c) Plot
of the tensile strength of ZnO nanoparticles/natural rubber vs ZnO filling content after vulcanization. (d) Tensile strength of ZnO nanoparticles/
styrene−butadiene rubber vs ZnO filling content after vulcanization. The symbols ■, ●, ▲, and ◆ represent 0.5, 1, 2, and 3 phr for ZnO,
respectively, the black and blue columns represent commercial indirect process ZnO nanoparticles, and the red and orange columns represent OA
capped ZnO nanoparticles, respectively.

styrene−butadiene rubber molecules. Because of the resonance


stabilization effect, hydrogens of allyl and benzyl groups are
very reactive. Both natural rubber and styrene−butadiene
rubber have numerous allylic hydrogens, and sulfur can react
with the allylic hydrogen much easier than with other
hydrogens in the rubber polymer chains.52 Therefore, with a
sufficient amount of sulfur, the vulcanization rate is affected by
the number of allylic hydrogens. Specifically, the repeat units of
natural rubber and styrene−butadiene rubber have 7 and 3
allylic hydrogens, respectively. The overall rate of vulcanization
of natural rubber was higher than that of styrene−butadiene
rubber.52 This is consistent with the total amount of allylic
hydrogens in the rubber molecules. Conversely, with much less
allylic hydrogens, styrene−butadiene rubber shows a less Figure 5. Cross-linking density of ZnO nanoparticles/natural rubber
composites. Black and red columns represent commercial indirect
sensitive response to the change of ZnO contents.
process ZnO nanoparticles and OA capped ZnO nanoparticles,
Additionally, to the quantities discussed above, the cross- respectively. The curing time is t90 + 3 min.
linking density affects the vulcanizates’ mechanical perform-
ance. The torque is directly related to the cross-linking density.
The low-field nuclear magnetic resonance (LF-NMR) was specimens.53 The dynamic mechanical performance of the
applied to determine the cross-linking density of the natural ZnO/natural rubber composites was analyzed over a broad
rubber composites. As shown in Figure 5, the cross-linking temperature range from −100 to 100 °C using a dynamic
density depends critically on the ZnO filling level, while the mechanical analyzer. Figure S14a,b and d,e displays the
OA capped ZnO composites exhibit higher cross-linking obtained storage mechanical modulus E′ and loss factor tan δ
density. This could be attributed to the more homogeneously as a function of temperature collected by DMA for each ZnO/
distributed OA capped ZnO nanoparticles in the natural natural rubber composite. Figure S14c and f shows the tan δ
rubber compounds, which develops a more extended particle− value at 60 °C of the ZnO/natural rubber composites vs ZnO
rubber polymer chain and is homogeneously spread through- contents, that the tan δ value at 60 °C is coupled to the cross-
out the rubber compounds. Besides the torque difference, the linking density.54 Combined with cross-linking density
modulus at 300% strain would increase with the increasing measurements, as the ZnO content increases, the cross-linking
cross-linking density as well, as shown in Figure S13a. Dynamic density of the products increases, which is associated with a
mechanical analysis (DMA) is especially sensitive to the decrease in the tan δ values. This is attributed to the cross-
physical and chemical cross-linking density of the testing linking of the rubber matrixes that resulted in a better
E https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Figure 6. Characterization of ZnO/natural rubber composite under industrial vulcanization conditions: 151 °C, curing time of 30 min. (a) Plot of
the torque difference (MH − ML) of ZnO nanoparticles/natural rubber vs ZnO filling content. (b) Plot of the tensile strength of ZnO
nanoparticles/natural rubber vs ZnO filling content. Black and red columns represent commercial indirect process ZnO nanoparticles and OA
capped ZnO nanoparticles, respectively.

synergistic motion, which diminished the friction between the nonsensitive response to the change of ZnO content. This can
rubber polymer chains.55 be attributed to the lack of presence of the allylic hydrogens in
The effect of OA capped ZnO nanoparticles as a the repeat units of styrene−butadiene rubber compared with
vulcanization activator and accelerator on promoting the natural rubber. Additionally, under a practical vulcanization
vulcanization of natural rubber has been studied and compared condition (151 °C, 30 min) applied to the ZnO/natural rubber
with commercial indirect process ZnO nanoparticles applied systems, the OA capped ZnO nanoparticle composites gave a
under “ideal” vulcanization conditions (143 °C with the curing significant enhancement in comparison with commercial ZnO
time t90 + 3 min) as above; OA capped ZnO has shown an nanoparticles. We believe OA capped ZnO nanoparticle
excellent vulcanization performance. However, in the practical activators are good substitutes for commercial ZnO for the
rubber curing process, to compromise the various types of sulfur vulcanization of natural rubber without any disadvanta-
rubber elastomers and additives and maintain the fabrication geous influence on the vulcanization temperature and time.
efficiency, the vulcanization condition is set with the curing This achievement has a potentially great impact on the
time 30 min at 151 °C instead of 143 °C. The performance of reduction of ZnO content in the commercial rubber
the commercial indirect process ZnO and OA capped ZnO manufacturing.


nanoparticles on the vulcanization of natural rubber was
evaluated, and the curing characteristics are shown in Figure ASSOCIATED CONTENT
S10 and Table S6, and the mechanical properties are presented
in Figure 6, Figure S11, Figure S13b, and Table S2. Compared * Supporting Information

with the ideal vulcanization condition, a more pronounced The Supporting Information is available free of charge at
enhancement is seen between the ZnO-C/natural rubber and https://pubs.acs.org/doi/10.1021/acsami.0c15114.
ZnO-OA/natural rubber composites with the same ZnO filling Materials and characterization methods; TGA curves for
content under the industrial vulcanization condition. OA capped ZnO/rubber masterbatch composites; TGA
We believe that 2 phr OA capped ZnO nanoparticles and DLS curves, TEM images, and XRD pattern of OA
enhance the activation during the vulcanization process and capped ZnO nanoparticles; TEM images and XRD
boost the formation of the cross-linking of the natural rubber pattern of the commercial indirect process ZnO
network structure. The amount of OA capped ZnO nano- nanoparticles; the comparison of N2 adsorption/
particles was lower than the amount of commercial indirect desorption isothermal and pore size distribution of OA
process ZnO nanoparticles (3 phr). OA capped ZnO capped ZnO nanoparticles and the commercial indirect
nanoparticles have a favorable effect on reducing the ZnO process ZnO nanoparticles; DSC curves of ZnO/natural
filling content used in rubber manufacturing. rubber composites after vulcanization under different

■ CONCLUSIONS
In summary, by investigating the torque increment values
vulcanization conditions; vulcanization curves/stress−
strain curves/cross-linking density of the ZnO/natural
rubber composites; and tables of mechanical properties
through the vulcanization process and the cross-linking of ZnO/natural rubber composites and ZnO/styrene−
densities of natural rubber compounds, it was concluded butadiene rubber composites (PDF)


that, with the same amount of ZnO filling contents introduced
into the rubber, 5 nm OA capped ZnO nanoparticle AUTHOR INFORMATION
composites exhibited both higher curing efficiency and better
mechanical performance. This is due to the improved Corresponding Authors
dispersion of the ZnO with a smaller size in the matrix. In Zhao Wang − State Key Laboratory of Organic−Inorganic
spite of that, with the same formula and vulcanization Composites and Key Laboratory of Beijing City for Preparation
conditions, the styrene−butadiene rubber composites showed and Processing of Novel Polymer Materials, Beijing University of
a less obvious change between OA capped ZnO nanoparticles Chemical Technology, Beijing 100029, China;
and commercial indirect process ZnO nanoparticles and a Email: wangzhao@mail.buct.edu.cn
F https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Zongyu Wang − Department of Chemistry, Carnegie Mellon Foundation (2019M650451), NSF (DMR 1501324 and DMR
University, Pittsburgh, Pennsylvania 15213, United States; 1410845), and the Department of Energy (DE-EE0006702), as
Email: zongyuw@alumni.cmu.edu well as the Scott Institute for Energy Technologies at Carnegie
Krzysztof Matyjaszewski − Department of Chemistry, Carnegie Mellon University. X.Q. gratefully acknowledges financial
Mellon University, Pittsburgh, Pennsylvania 15213, United support from the China Scholarship Council.


States; orcid.org/0000-0003-1960-3402;
Email: matyjaszewski@cmu.edu REFERENCES
Authors (1) Suksaeree, J.; Pichayakorn, W.; Monton, C.; Sakunpak, A.;
Xuan Qin − State Key Laboratory of Organic−Inorganic Chusut, T.; Saingam, W. Rubber Polymers for Transdermal Drug
Delivery Systems. Ind. Eng. Chem. Res. 2014, 53 (2), 507−513.
Composites and Key Laboratory of Beijing City for Preparation
(2) Wu, W.; Cao, X.; Zou, J.; Ma, Y.; Wu, X.; Sun, C.; Li, M.; Wang,
and Processing of Novel Polymer Materials, Beijing University of N.; Wang, Z.; Zhang, L. Triboelectric Nanogenerator Boosts Smart
Chemical Technology, Beijing 100029, China; Department of Green Tires. Adv. Funct. Mater. 2019, 29 (41), 1806331.
Chemistry, Carnegie Mellon University, Pittsburgh, Pennsylvania (3) Wu, W.; Yang, T.; Zhang, Y.; Wang, F.; Nie, Q.; Ma, Y.; Cao, X.;
15213, United States Wang, Z. L.; Wang, N.; Zhang, L. Application of Displacement-
Haoshu Xu − State Key Laboratory of Organic−Inorganic Current-Governed Triboelectric Nanogenerator in an Electrostatic
Composites and Key Laboratory of Beijing City for Preparation Discharge Protection System for the Next-Generation Green Tire.
and Processing of Novel Polymer Materials, Beijing University of ACS Nano 2019, 13 (7), 8202−8212.
Chemical Technology, Beijing 100029, China (4) Lei, W.; Zhou, X.; Russell, T. P.; Hua, K. C.; Yang, X.; Qiao, H.;
Ganggang Zhang − State Key Laboratory of Organic−Inorganic Wang, W.; Li, F.; Wang, R.; Zhang, L. High Performance Bio-Based
Composites and Key Laboratory of Beijing City for Preparation Elastomers: Energy Efficient and Sustainable Materials for Tires. J.
Mater. Chem. A 2016, 4 (34), 13058−13062.
and Processing of Novel Polymer Materials, Beijing University of (5) Hernández, M.; Ezquerra, T. A.; Verdejo, R.; López-Manchado,
Chemical Technology, Beijing 100029, China M. A. Role of Vulcanizing Additives on the Segmental Dynamics of
Jiadong Wang − State Key Laboratory of Organic−Inorganic Natural Rubber. Macromolecules 2012, 45 (2), 1070−1075.
Composites and Key Laboratory of Beijing City for Preparation (6) Birkley, A. W. Rubbery Materials and Their Compounds J. A.
and Processing of Novel Polymer Materials, Beijing University of Brydson, Elsevier Applied Science, London, 1988. Pp. Xxii + 469,
Chemical Technology, Beijing 100029, China Price £68.00. Isbn 1-85166-21 5-4. Br. Polym. J. 1989, 21 (4), 361−
Yuqi Zhao − Department of Materials Science & Engineering, 361.
Carnegie Mellon University, Pittsburgh, Pennsylvania 15213, (7) Zhang, G.; Feng, H.; Liang, K.; Wang, Z.; Li, X.; Zhou, X.; Guo,
United States B.; Zhang, L. Design of Next-Generation Cross-Linking Structure for
Tianwei Tan − State Key Laboratory of Organic−Inorganic Elastomers toward Green Process and a Real Recycling Loop. Science
Bulletin 2020, 65 (11), 889−898.
Composites and Key Laboratory of Beijing City for Preparation (8) Zhang, G.; Zhou, X.; Liang, K.; Guo, B.; Li, X.; Wang, Z.; Zhang,
and Processing of Novel Polymer Materials, Beijing University of L. Mechanically Robust and Recyclable Epdm Rubber Composites by
Chemical Technology, Beijing 100029, China; orcid.org/ a Green Cross-Linking Strategy. ACS Sustainable Chem. Eng. 2019, 7
0000-0002-9471-8202 (13), 11712−11720.
Michael R. Bockstaller − Department of Materials Science & (9) Zheng, J.; Han, D.; Zhao, S.; Ye, X.; Wang, Y.; Wu, Y.; Dong, D.;
Engineering, Carnegie Mellon University, Pittsburgh, Liu, J.; Wu, X.; Zhang, L. Constructing a Multiple Covalent Interface
Pennsylvania 15213, United States; orcid.org/0000-0001- and Isolating a Dispersed Structure in Silica/Rubber Nanocomposites
9046-9539 with Excellent Dynamic Performance. ACS Appl. Mater. Interfaces
Liqun Zhang − State Key Laboratory of Organic−Inorganic 2018, 10 (23), 19922−19931.
Composites and Key Laboratory of Beijing City for Preparation (10) Krejsa, M. R.; Koenig, J. L. A Review of Sulfur Crosslinking
Fundamentals for Accelerated and Unaccelerated Vulcanization.
and Processing of Novel Polymer Materials, Beijing University of
Rubber Chem. Technol. 1993, 66 (3), 376−410.
Chemical Technology, Beijing 100029, China; orcid.org/ (11) Heideman, G.; Datta, R. N.; Noordermeer, J. W. M.; van
0000-0002-2103-6294 Baarle, B. Activators in Accelerated Sulfur Vulcanization. Rubber
Complete contact information is available at: Chem. Technol. 2004, 77 (3), 512−541.
https://pubs.acs.org/10.1021/acsami.0c15114 (12) Sjothun, I. J.; Alliger, G. Vulcanization of Elastomers: Principles
and Practice of Vulcanisation of Commercial Rubbers; Reinhold: New
York, 1964; p v, 410 p.
Author Contributions
(13) Wagner, A. M.; Spencer, D. S.; Peppas, N. A. Advanced
X.Q. and H.X. contributed equally. X.Q. and Z.W. synthesized Architectures in the Design of Responsive Polymers for Cancer
materials. H.X. performed characterization work. G.Z., J.W., Nanomedicine. J. Appl. Polym. Sci. 2018, 135 (24), 46154.
Z.W., Y.Z., and T.T. assisted in the characterization work. L.Z. (14) Duchácě k, V.; Kuta, A.; Pr̆ibyl, P. Efficiency of Metal Activators
and K.M. conceived and organized the project and together of Accelerated Sulfur Vulcanization. J. Appl. Polym. Sci. 1993, 47 (4),
with X.Q. and Z.W. wrote the manuscript. All authors 743−746.
discussed the results and commented on the manuscript. (15) Coran, A. Y. 7 - Vulcanization. In Science and Technology of
Rubber (Second ed.); Mark, J. E.; Erman, B.; Eirich, F. R., Eds.;
Notes Academic Press: San Diego, 1994; pp 339−385.
The authors declare no competing financial interest. (16) Ikeda, Y.; Yasuda, Y.; Ohashi, T.; Yokohama, H.; Minoda, S.;


Kobayashi, H.; Honma, T. Dinuclear Bridging Bidentate Zinc/
ACKNOWLEDGMENTS Stearate Complex in Sulfur Cross-Linking of Rubber. Macromolecules
2015, 48 (3), 462−475.
This work was supported by the National Science Foundation (17) Ikeda, Y.; Sakaki, Y.; Yasuda, Y.; Junkong, P.; Ohashi, T.;
for Young Scientists of China (51703007), Basic Science Miyaji, K.; Kobayashi, H. Roles of Dinuclear Bridging Bidentate Zinc/
Center Foundation (51988102), National Natural Science Stearate Complexes in Sulfur Cross-Linking of Isoprene Rubber.
Foundation of China (51790501), China Postdoctoral Science Organometallics 2019, 38 (11), 2363−2380.

G https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

(18) Ikeda, Y.; Higashitani, N.; Hijikata, K.; Kokubo, Y.; Morita, Y.; Carbon Hybrids Derived from Polymer Nanocomposite Precursor
Shibayama, M.; Osaka, N.; Suzuki, T.; Endo, H.; Kohjiya, S. Materials for Pseudocapacitor Electrodes with High Cycling Stability.
Vulcanization: New Focus on a Traditional Technology by Small- Polymer 2018, 137, 370−377.
Angle Neutron Scattering. Macromolecules 2009, 42 (7), 2741−2748. (38) Wang, Z.; Liu, S.; Zhang, J.; Yan, J.; Zhao, Y.; Mahoney, C.;
(19) Suzuki, T.; Osaka, N.; Endo, H.; Shibayama, M.; Ikeda, Y.; Asai, Ferebee, R.; Luo, D.; Pietrasik, J.; Bockstaller, M. R.; Matyjaszewski,
H.; Higashitani, N.; Kokubo, Y.; Kohjiya, S. Nonuniformity in Cross- K. Photocatalytic Active Mesoporous Carbon/Zno Hybrid Materials
Linked Natural Rubber as Revealed by Contrast-Variation Small- from Block Copolymer Tethered Zno Nanocrystals. Langmuir 2017,
Angle Neutron Scattering. Macromolecules 2010, 43 (3), 1556−1563. 33 (43), 12276−12284.
(20) Yasuda, Y.; Minoda, S.; Ohashi, T.; Yokohama, H.; Ikeda, Y. (39) Plath, L. D.; Wang, Z.; Yan, J.; Matyjaszewski, K.; Bier, M. E.
Two-Phase Network Formation in Sulfur Crosslinking Reaction of Characterization of Zno Nanoparticles Using Superconducting
Isoprene Rubber. Macromol. Chem. Phys. 2014, 215 (10), 971−977. Tunnel Junction Cryodetection Mass Spectrometry. J. Am. Soc.
(21) Cui, J.; Zhang, L.; Wu, W.; Cheng, Z.; Sun, Y.; Jiang, H.; Li, C. Mass Spectrom. 2017, 28 (6), 1160−1165.
Zinc Oxide with Dominant (1 0 0) Facets Boosts Vulcanization (40) Ding, H.; Yan, J.; Wang, Z.; Xie, G.; Mahoney, C.; Ferebee, R.;
Activity. Eur. Polym. J. 2019, 113, 148−154. Zhong, M.; Daniel, W. F. M.; Pietrasik, J.; Sheiko, S. S.; Bettinger, C.
(22) Lee, Y. H.; Cho, M.; Nam, J.-D.; Lee, Y. Effect of Zno Particle J.; Bockstaller, M. R.; Matyjaszewski, K. Preparation of Zno Hybrid
Sizes on Thermal Aging Behavior of Natural Rubber Vulcanizates. Nanoparticles by Atrp. Polymer 2016, 107, 492−502.
Polym. Degrad. Stab. 2018, 148, 50−55. (41) Wang, Z.; Mahoney, C.; Yan, J.; Lu, Z.; Ferebee, R.; Luo, D.;
(23) Heideman, G.; Datta, R. N.; Noordermeer, J. W. M.; Baarle, B. Bockstaller, M. R.; Matyjaszewski, K. Preparation of Well-Defined
V. Influence of Zinc Oxide During Different Stages of Sulfur Poly(Styrene-Co-Acrylonitrile)/Zno Hybrid Nanoparticles by an
Vulcanization. Elucidated by Model Compound Studies. J. Appl. Efficient Ligand Exchange Strategy. Langmuir 2016, 32 (49),
Polym. Sci. 2005, 95 (6), 1388−1404. 13207−13213.
(24) Adams, L. K.; Lyon, D. Y.; Alvarez, P. J. J. Comparative Eco- (42) Wang, Z.; Pan, X.; Li, L.; Fantin, M.; Yan, J.; Wang, Z.; Wang,
Toxicity of Nanoscale Tio2, Sio2, and Zno Water Suspensions. Water Z.; Xia, H.; Matyjaszewski, K. Enhancing Mechanically Induced Atrp
Res. 2006, 40 (19), 3527−3532. by Promoting Interfacial Electron Transfer from Piezoelectric
(25) Xia, T.; Kovochich, M.; Liong, M.; Mädler, L.; Gilbert, B.; Shi, Nanoparticles to Cu Catalysts. Macromolecules 2017, 50 (20),
H.; Yeh, J. I.; Zink, J. I.; Nel, A. E. Comparison of the Mechanism of 7940−7948.
Toxicity of Zinc Oxide and Cerium Oxide Nanoparticles Based on (43) Wang, Z.; Lu, Z.; Mahoney, C.; Yan, J.; Ferebee, R.; Luo, D.;
Dissolution and Oxidative Stress Properties. ACS Nano 2008, 2 (10), Matyjaszewski, K.; Bockstaller, M. R. Transparent and High Refractive
2121−2134. Index Thermoplastic Polymer Glasses Using Evaporative Ligand
(26) Franklin, N. M.; Rogers, N. J.; Apte, S. C.; Batley, G. E.; Casey, Exchange of Hybrid Particle Fillers. ACS Appl. Mater. Interfaces 2017,
P. S. Comparative Toxicity of Nanoparticulate Zno, Bulk Zno, and 9 (8), 7515−7522.
Zncl2 to a Freshwater Microalga (Pseudokirchneriella Subcapitata): (44) Maiti, M.; Basak, G. C.; Srivastava, V. K.; Jasra, R. V. Influence
The Importance of Particle Solubility. Environ. Sci. Technol. 2007, 41 of Synthesized Nano-Zno on Cure and Physico-Mechanical Proper-
(24), 8484−8490. ties of Sbr/Br Blends. Int. J. Ind. Chem. 2017, 8 (3), 273−283.
(27) Councell, T. B.; Duckenfield, K. U.; Landa, E. R.; Callender, E. (45) Ojha, S.; Dang, A.; Hui, C. M.; Mahoney, C.; Matyjaszewski,
Tire-Wear Particles as a Source of Zinc to the Environment. Environ. K.; Bockstaller, M. R. Strategies for the Synthesis of Thermoplastic
Sci. Technol. 2004, 38 (15), 4206−4214. Polymer Nanocomposite Materials with High Inorganic Filling
(28) Rhodes, E. P.; Ren, Z.; Mays, D. C. Zinc Leaching from Tire Fraction. Langmuir 2013, 29 (28), 8989−8996.
Crumb Rubber. Environ. Sci. Technol. 2012, 46 (23), 12856−12863. (46) Weber, D.; Botnaraş, S.; Pham, D. V.; Steiger, J.; De Cola, L.
(29) Smolders, E.; Degryse, F. Fate and Effect of Zinc from Tire Functionalized Zno Nanoparticles for Thin-Film Transistors: Support
Debris in Soil. Environ. Sci. Technol. 2002, 36 (17), 3706−3710. of Ligand Removal by Non-Thermal Methods. J. Mater. Chem. C
(30) Roy, K.; Alam, M. N.; Mandal, S. K.; Debnath, S. C. Sol−Gel 2013, 1 (18), 3098−3103.
Derived Nano Zinc Oxide for the Reduction of Zinc Oxide Level in (47) Epifani, M.; Arbiol, J.; Díaz, R.; Perálvarez, M. J.; Siciliano, P.;
Natural Rubber Compounds. J. Sol-Gel Sci. Technol. 2014, 70 (3), Morante, J. R. Synthesis of Sno2 and Zno Colloidal Nanocrystals from
378−384. the Decomposition of Tin(Ii) 2-Ethylhexanoate and Zinc(Ii) 2-
(31) Thomas, S. P.; Mathew, E. J.; Marykutty, C. V. Synthesis and Ethylhexanoate. Chem. Mater. 2005, 17 (25), 6468−6472.
Effect of Surface Modified Nano Zno in Natural Rubber Vulcan- (48) Decremps, F.; Pellicer-Porres, J.; Saitta, A.; Chervin, J. C.;
ization. J. Appl. Polym. Sci. 2012, 124 (4), 3099−3107. Polian, A. High-Pressure Raman Spectroscopy Study of Wurtzite Zno.
(32) Kim, J.-W.; Porte, Y.; Ko, K. Y.; Kim, H.; Myoung, J.-M. Phys. Rev. B: Condens. Matter Mater. Phys. 2002, 65 (9), 92101.
Micropatternable Double-Faced Zno Nanoflowers for Flexible Gas (49) Nieuwenhuizen, P. J. Zinc Accelerator Complexes.: Versatile
Sensor. ACS Appl. Mater. Interfaces 2017, 9 (38), 32876−32886. Homogeneous Catalysts in Sulfur Vulcanization. Appl. Catal., A 2001,
(33) Panampilly, B.; Thomas, S. Nano Zno as Cure Activator and 207 (1), 55−68.
Reinforcing Filler in Natural Rubber. Polym. Eng. Sci. 2013, 53 (6), (50) Loo, C. T. High Temperature Vulcanization of Elastomers: 2.
1337−1346. Network Structures in Conventional Sulphenamide-Sulphur Natural
(34) Susanna, A.; Armelao, L.; Callone, E.; Dirè, S.; D’Arienzo, M.; Rubber Vulcanizates. Polymer 1974, 15 (6), 357−365.
Di Credico, B.; Giannini, L.; Hanel, T.; Morazzoni, F.; Scotti, R. Zno (51) Li, Y.; Wu, J.; Zhang, Q.; Dong, F.; Xiong, Y. Novel
Nanoparticles Anchored to Silica Filler. A Curing Accelerator for Architecture of Zno Nanobundles Grown on Porous Silica as High
Isoprene Rubber Composites. Chem. Eng. J. 2015, 275, 245−252. Performance Vulcanization Accelerators That Reinforce Rubber
(35) Susanna, A.; D’Arienzo, M.; Di Credico, B.; Giannini, L.; Composites. Ind. Eng. Chem. Res. 2020, 59 (10), 4493−4503.
Hanel, T.; Grandori, R.; Morazzoni, F.; Mostoni, S.; Santambrogio, (52) Chough, S.-H.; Chang, D.-H. Kinetics of Sulfur Vulcanization
C.; Scotti, R. Catalytic Effect of Zno Anchored Silica Nanoparticles on of Nr, Br, Sbr, and Their Blends Using a Rheometer and Dsc. J. Appl.
Rubber Vulcanization and Cross-Link Formation. Eur. Polym. J. 2017, Polym. Sci. 1996, 61 (3), 449−454.
93, 63−74. (53) Qin, X.; Han, B.; Lu, J.; Wang, Z.; Sun, Z.; Wang, D.; Russell,
(36) Lin, Y.; Chen, Y.; Zeng, Z.; Zhu, J.; Wei, Y.; Li, F.; Liu, L. Effect T. P.; Zhang, L.; Liu, J. Rational Design of Advanced Elastomer
of Zno Nanoparticles Doped Graphene on Static and Dynamic Nanocomposites Towards Extremely Energy-Saving Tires Based on
Mechanical Properties of Natural Rubber Composites. Composites, Macromolecular Assembly Strategy. Nano Energy 2018, 48, 180−188.
Part A 2015, 70, 35−44. (54) Qin, X.; Wang, J.; Zhang, Y.; Wang, Z.; Li, S.; Zhao, S.; Tan, T.;
(37) Zhao, Y.; Wang, Z.; Yuan, R.; Lin, Y.; Yan, J.; Zhang, J.; Lu, Z.; Liu, J.; Zhang, L.; Matyjaszewski, K. Self-Assembly Strategy for
Luo, D.; Pietrasik, J.; Bockstaller, M. R.; Matyjaszewski, K. Zno/ Double Network Elastomer Nanocomposites with Ultralow Energy

H https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX
ACS Applied Materials & Interfaces www.acsami.org Research Article

Consumption and Ultrahigh Wear Resistance. Adv. Funct. Mater.


2020, 30 (34), 2003429.
(55) Zhang, G.; Liang, K.; Feng, H.; Pang, J.; Liu, N.; Li, X.; Zhou,
X.; Wang, R.; Zhang, L. Design of Epoxy-Functionalized Styrene-
Butadiene Rubber with Bio-Based Dicarboxylic Acid as a Cross-Linker
toward the Green-Curing Process and Recyclability. Ind. Eng. Chem.
Res. 2020, 59 (22), 10447−10456.

I https://dx.doi.org/10.1021/acsami.0c15114
ACS Appl. Mater. Interfaces XXXX, XXX, XXX−XXX

You might also like