Você está na página 1de 11

Acta Materialia 55 (2007) 6372–6382

www.elsevier.com/locate/actamat

Improved mechanical and functional properties of elastomer/graphite


nanocomposites prepared by latex compounding
a,b
Jian Yang , Ming Tian a, Qing-Xiu Jia b, Jun-Hong Shi a, Li-Qun Zhang a,b,*
,
Szu-Hui Lim c, Zhong-Zhen Yu c, Yiu-Wing Mai c,*
a
Key Laboratory for Nano-materials, Beijing University of Chemical Technology, Ministry of Education of China, Beijing 100029, China
b
Key Laboratory on Preparation and Processing of Novel Polymer Materials, Beijing University of Chemical Technology, Beijing 100029, China
c
Centre for Advanced Materials Technology (CAMT), School of Aerospace, Mechanical and Mechatronic Engineering (J07),
The University of Sydney, Sydney, NSW 2006, Australia

Received 21 April 2007; received in revised form 25 June 2007; accepted 26 July 2007
Available online 12 September 2007

Abstract

The facile latex approach has been adopted to finely incorporate graphite nanosheets into elastomeric polymer matrix to obtain high-
performance elastomeric nanocomposites with improved mechanical properties and functional properties. Scanning electron microscopy,
transmission electron microscopy and X-ray diffraction experiments show that the nanostructures of the final nanocomposites exhibit a
high degree of exfoliation and intercalation of graphite in the nitrile-butadiene rubber (NBR) matrix. Mechanical and dynamic-mechan-
ical tests demonstrate that the NBR/graphite nanocomposites possess greatly increased elastic modulus and tensile strength, and desir-
ably strong interfaces. The unexpected self-crosslinking of elastomer/graphite nanocomposites was discovered and then verified by
oscillating disc rheometry and equilibrium swelling experiments. After critically examining various polymer types by X-ray photoelectron
spectroscopy, electron spin resonance and Fourier transform infrared spectroscopy, a radical initiation mechanism was proposed to
explain the self-crosslinking reaction. These NBR/graphite nanocomposites possess significantly improved wear resistance and gas bar-
rier properties, and superior electrical/thermal conductivity. Such versatile functional properties make NBR nanocomposites a promising
new class of advanced materials.
 2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.

Keywords: Nanocomposites; Elastomers; Graphite; Nanostructure; Latex compounding

1. Introduction gest materials per unit weight and has unique functional
properties (e.g. good electrical and thermal conductivities,
Nanocomposites have been shown to afford remarkable and good lubricating properties) compared to layered
property enhancements compared to conventional micro- silicates [13]. In addition, graphite is cheap compared to
composites [1–3]. Polymer nanocomposites with layered sil- carbon nanotubes. Recently, polymer/graphite nanosheet
icates [4–9] and carbon nanotubes [10–12] have attracted composites have made a great impact in nanocomposite
great interest for the improvement of structural properties research [14–27].
and the development of new materials with different func- To date, some polymer/expanded graphite (EG) nano-
tional properties. Graphite is a layered material with a high composites with good electrical conductivity have been
aspect ratio in its exfoliated state; it is also one of the stron- prepared (e.g. via in situ polymerization) [14–22]. But EG
does not significantly enhance the mechanical properties
*
Corresponding authors.
of the polymers due to poor dispersion of the graphite,
E-mail addresses: zhanglq@mail.buct.edu.cn (L.-Q. Zhang), y.mai@ voids trapped in the composites and weak interfacial adhe-
usyd.edu.au (Y.-W. Mai). sion [19–26]. Indeed, the functional performance of the

1359-6454/$30.00  2007 Acta Materialia Inc. Published by Elsevier Ltd. All rights reserved.
doi:10.1016/j.actamat.2007.07.043
J. Yang et al. / Acta Materialia 55 (2007) 6372–6382 6373

nanocomposites may be curtailed. Polymer/graphite are strongly reinforced and the functional properties of
nanocomposites with both enhanced mechanical and the corresponding NBR nanocomposites become quite
functional properties derived from the unique multifunc- prominent.
tional nature of graphite nanosheets have rarely been More importantly, an unexpected self-crosslinking of
reported, probably because of the great difficulties in simul- graphite/NBR is found, resulting in strong interfacial inter-
taneously achieving good exfoliation and dispersion of action. Thus, several elastomer-based nanocomposites with
graphite and strong interfacial adhesion between graphite EG were prepared by latex compounding in order to study
and polymer. the self-crosslinking behavior, which was mainly investi-
Elastomeric compounds are a class of materials widely gated via oscillating disc rheometry and equilibrium swell-
used not only in general products, but also in specialized ing experiments. This self-curing process may help to
fields such as aerospace, biomedical, micro-electrome- establish a desirable covalently bonded interface in the
chanical systems and shape memory polymers [28–31]. nanocomposites, hence eliminating the use of conventional
In many of these applications, high strength, superior additives such as curing agents, accelerators and activators,
electrical and thermal conductivities, and improved tribo- and thereby making it possible to design simple composi-
logical and barrier properties are required. Various types tions for potential applications.
of filler have been incorporated into rubber matrices in
order to provide reinforcement or functional properties. 2. Material and methods
However, most of those filled composites do not combine
enhanced mechanical properties with superior multifunc- 2.1. Materials
tional performance. Thus, graphite may be an ideal nano-
filler to impart these mechanical [32,33] and functional Styrene butadiene rubber (SBR) latex (styrene content
properties to rubber materials, provided that fine disper- 23%) was obtained from Jilin Petrochemical Company
sion of the filler and strong interfacial interaction between (China). Natural rubber (NR) latex was provided by Bei-
filler and rubber are achieved. To the best of our knowl- jing Latex Factory (China). NBR latex (AN content 24–
edge, few papers have been published on elastomer/ 26%) was supplied by Lanzhou Petrochemical Company
graphite nanocomposites. (China). Graphite oxide was provided by Pingdu Huadong
Even though conventional direct blending techniques Graphite Processing Factory (China). Other chemicals and
usually provide sufficient force for filler dispersion during rubber curing additives were purchased from stores and
rubber processing, due to the high viscosity of the material, used as received.
it cannot ensure satisfactory dispersion of graphite layers.
As most rubbers exist in latex forms, in which latex parti- 2.2. Preparation procedure
cles with sizes of 50 nm are uniformly stabilized in water,
the latex compounding method (LCM), which offers a EG with an expansion ratio along the c-axis of 250
good performance/cost ratio, has been developed. Using was prepared by microwave irradiation of graphite oxide
this method, we have prepared various rubber nanocom- for 30 s in a microwave oven (Sanyo EM-183MS1) with
posites with layered silicates, fibrillar silicates and starch a power of 700 W and a frequency at 2.45 GHz [43]. Then
[34,35]. Karger-Krocsis et al. have also fabricated nano- the EG was mixed and saturated with deionized water with
composites via this latex technology [36,37]. the aid of the surfactant SDS. In a typical procedure, an
By exploiting the ‘‘partially ionic’’ nature of the graphite EG/SDS/H2O ratio of 1 g/5 g/1.5 l was used. The mixture
layers and the functional groups (epoxide and hydroxyl) on was subjected to ultrasonic irradiation with a power of
the exfoliated graphite oxide, as well as the ease of interca- 100 W for 8 h, which resulted in the formation of a com-
lating organic compounds [38–42], we firstly modified paratively stable aqueous suspension of graphitic nano-
in situ graphite layers by introducing sodium dodecylsulfo- sheets. The NBR latex was added to the suspension
nate (SDS) in an aqueous suspension of graphite to stabi- slowly under vigorous stirring, and subsequently treated
lize the material. After mixing the suspension with rubber with ultrasonic irradiation. CaCl2 aqueous solution
latex, SDS began to build bridges between the graphite (1 wt.%) was then added to co-coagulate rubber and graph-
nanosheets and rubber macromolecules during the fast ite. In direct blending, which is the commonest rubber
co-coagulation, which prevented the aggregation of graph- blending technique, EG powder pre-treated with ultrasonic
ite nanosheets and led to the rubber-intercalated structure irradiation was dried and pulverized, and then added
and nanoscale dispersion. Nitrile-butadiene rubber directly into NBR in a Haake mixer.
(NBR), a widely used elastomer with a polar cyanide group For regular tests, a sulfur curing system was added. The
(C”N), was chosen as the matrix in order to obtain strong NBR/EG compound was further mixed in the mixer with
interfacial adhesion between graphite nanosheets and elas- the curing ingredients according to the recipe in Table 1.
tomer through the interaction between the C”N group and To prepare samples for mechanical tests, the curing time
polar groups of the nanosheets. It was found that latex was determined using an oscillating disc rheometer
compounding is very effective in achieving the nanoscale (ODR) at 160 C. The compounds were then vulcanized
dispersion of graphite in NBR. The graphite nanosheets for testing in a press at a pressure and temperature of
6374 J. Yang et al. / Acta Materialia 55 (2007) 6372–6382

Table 1 lnð1  V r Þ þ V r þ vV 2r
Formulation of the sulfur-curing system for NBR/EG compounding Ve ¼ ð1Þ
a
V s ðV r1=3  V r =2Þ
Materials Amounts (phr )
NBR 100 where Vr is the polymer volume fraction in the vulcanizate
Sulfur 1.5 swollen to equilibrium and Vs is the solvent molar volume
ZnO 5 (106.5 cm3 mol1 for toluene). v is the rubber–toluene
Stearic acid (SA) 1 interaction parameter and is taken as 0.435 and 0.446 for
Anti-aging reagent 4010NAb 2 NBR and SBR, respectively [45,46].
Accelerator DMc 1.5
Graphite Variable Nitrogen permeation tests were performed with a gas
a permeability-measurement apparatus described elsewhere
Parts per hundred rubber parts in weight.
b
N-isopropyl N 0 -phenyl 1,4-phenylenediamine. [9]. The thermal conductivity of nanocomposites was mea-
c
Dibenzothiazyl disulfide. sured by a FOX50 heat flow meter from Laser Comp., Inc.
USA. Bulk electrical conductivity tests were conducted
using a Zheng-Yang Ohmmeter (Model QJ84) with a
15 MPa and 160 C, respectively, using the optimum cure four-terminal measurement feature to obtain the bulk
time (T90) determined from the ODR results. conductivity. Wear properties were evaluated using a
ring-on-disc tester (manufactured by Jinan Shijin Co.,
2.3. Characterization and testing China) [47,48].

The morphologies of the freeze-fractured surfaces of the 3. Results and discussion


compounds were determined by scanning electron micros-
copy (SEM) (Philips, The Netherlands). Transmission elec- 3.1. Morphology of rubber/graphite nanocomposites
tron microscopy (TEM) observations were performed with
a Hitachi H-800 (Japan) at an acceleration voltage of Graphite nanosheets were obtained by thermal decom-
200 kV. X-ray diffraction (XRD) data were acquired with position of graphite oxide and subsequent ultrasonic
a Rigaku D/Max 2500VBZt/PC X-ray diffractometer irradiation [49] (see Fig. 1a). When the graphite nanosheets
(Japan). Expanded graphite was characterized by X-ray were directly mixed with NBR in a Haake mixer, their
photoelectron spectroscopy (XPS) and electron spin reso- dispersion was poor as evidenced by many large EG
nance (ESR). The XPS spectra of the samples were aggregates and a large amount of rubber matrix without
recorded by using an ESCA LAB 250 (VG Scientific graphite filler in Fig. 1b. In contrast, LCM gives a much
Ltd., UK) and Al Ka radiation (1486.6 eV) was used as finer and more uniform dispersion of graphite nanosheets
an X-ray source. ESR was conducted at room temperature in NBR as shown in Fig. 1c–f. It can be seen that the sur-
with a Bruker ESP300E instrument equipped with a micro- face dimensions of the dispersed graphite layers are quite
wave counter (microwave power 10 mW). The Fourier large (several to tens of microns) with nanometer-scale
transform infrared spectroscopy (FT-IR) spectra of the thickness. This causes some difficulties in clearly observing
samples were recorded by a Perkin–Elmer System 2000 the dispersion quality with TEM, especially in distinguish-
FT-IR spectrometer using the attenuated total reflection ing individual graphite sheets. By comparing Fig. 1d with
(ATR) technique with a resolution of 2 cm1. Fig. 1b, it can clearly be seen that the interface between
The Shore A hardness of the vulcanizates were mea- the graphite sheets and NBR is greatly improved using
sured according to ASTM D2240 using an XY-1 type A LCM. This will definitely enhance the reinforcement effi-
durometer (No. 4 Chemical Machinery Plant of Shanghai ciency of graphite nanosheets.
Chemical Equipment Co. Ltd., China) and three different
locations of a sample (>6 mm in thickness) to obtain an 3.2. XRD analysis of nanocomposite structures
average value. Tensile tests of dumbbell specimens of the
vulcanizates were carried out on a CMT4104 testing The microstructure of the NBR/EG nanocomposites
machine (Shenzhen SANS Testing Machine Co. Ltd., prepared by LCM was further studied by XRD, which
China) at a speed of 500 mm min1 according to ASTM reveals the coexistence of intercalated and exfoliated
D412. Storage modulus, E 0 , and dynamic loss, tan d, were microstructures (see Fig. 2), as indicated by the shift in
measured as a function of temperature on a Dynamic the well-defined (0 0 1) reflections.
Mechanical Thermal Analyzer DMTA V (Rheometrics Fig. 2a shows the XRD patterns of EG powder before
Science Corporation) under tension mode at 1 Hz and and after the ultrasonic treatment, with and without surfac-
3 C min1. tant SDS. The original EG displays only one diffraction
The crosslinking density, i.e. the elastically active net- peak at 2h = 26.6, corresponding to a basal spacing of
work chain density, Ve (mol cm3), was determined by 0.335 nm, which is the lamellar characteristic of graphite.
the Flory–Rehner equation [44], through equilibrium After ultrasonic treatment without SDS, the XRD pattern
swelling experiments using toluene as a solvent at ambient of EG powder is unchanged. However, after ultrasonic
temperature: treatment with SDS, a new (0 0 1) diffraction peak emerges
J. Yang et al. / Acta Materialia 55 (2007) 6372–6382 6375

Fig. 1. SEM micrographs showing: (a) EG powder pretreated with ultrasonic irradiation; (b) NBR composite with 5 phr EG prepared by direct blending;
(c,d) NBR nanocomposite with 5 phr EG prepared by latex compounding. (e,f) TEM images of NBR nanocomposite with 5 phr EG prepared by latex
compounding.

at 2h = 2.55, corresponding to a basal spacing of 3.47 nm, many factors, such as graphite loading, solid content in
which is due to the intercalation of the SDS surfactant. suspension, type of flocculant, surfactant type and content.
However, the peak at 2h = 26.6 still remains, though with After latex compounding, the primary diffraction peak of
a reduced intensity. The reason is that not all of the carbon graphite at small angles has shifted to a very low angle of
sheets were opened after the thermal expansion of graphite 1.02 (Fig. 2, curve c3), indicating that a portion of EG
oxide. However, with ultrasonic treatment, nanometer- became intercalated by NBR macromolecules. Another
thick graphite sheets could be obtained in the aqueous sus- peak at 2.89 is thought to be due to the change of SDS
pension, and those sheets were covered by surfactant intercalation structure during LCM and hot pressing, e.g.
molecules. conformational change of SDS molecules between graphite
The XRD pattern in the small angle range (0.5–10) of layers, or contraction of some SDS molecules.
the EG powder pretreated with ultrasonic treatment and In the wide-angle range, the diffraction of graphite at
SDS is shown in Fig. 2 (curve c2), and again displays a 2h = 26.6, corresponding to a basal spacing of 0.335 nm,
well-defined (0 0 1) diffraction pattern for the SDS is much weaker in the nanocomposite prepared by LCM
intercalated structure. After adding NBR, the latex parti- than that by direct blending with the same loading of
cles enter the sheet network in suspension and graphite EG, as shown in Fig. 2b. The existence of a peak at
sheets are separated by strong ultrasonic irradiation. Dur- 2h = 26.6 for composites prepared by LCM can be
ing subsequent flocculation, the flocculant causes the NBR ascribed to both the unopened nanosheets during the
macromolecules and graphite sheets to flocculate simulta- thermal expansion and consequent treatment, and the
neously and a nanoscale dispersion of graphite nanosheets restacking of graphite layers during co-coagulation. How-
is hence maintained in the NBR matrix. This is similar to ever, the thickness of aggregate is generally nanoscale due
the process of preparing rubber/clay nanocomposites by to the separation effect of the latex particles [9]. These
LCM [9]. It is noted that there is competition between observed results indicate that LCM gives a greater extent
the separation of graphite sheets by coagulated rubber of exfoliation, intercalation and separation of graphite in
and the re-aggregation of the sheets, which depends on the NBR matrix.
6376 J. Yang et al. / Acta Materialia 55 (2007) 6372–6382

3.3. Mechanical properties

The effects of EG content on both mechanical and phys-


ical properties are summarized in Table 2. Fig. 3a and b
shows typical tensile stress–strain curves of the vulcanizates
filled with different amounts of EG by LCM, and compar-
isons of reinforcement efficiency of different preparation
routes (i.e. LCM and direct blending) and different fillers
(i.e. carbon black N330, silica) at the same loading of

Table 2
Properties of NBR/EG nanocomposites prepared from LCM
EG content (phr) 0 3 5 10 10a
Shore A hardness 50 61 74 89 56
Modulus at 100% elongation (MPa) 1.1 2.5 5.7 11.5 1.8
Tensile strength (MPa) 4.0 4.6 10.5 11.8 5.8
Elongation at break (%) 410 180 180 110 610
Permanent tensile set (%) 4 4 2 2 10
a
By direct blending.

Fig. 2. X-ray diffraction patterns: (a1) expanded graphite (EG); (a2) EG


powder pretreated with ultrasonic irradiation without adding surfactant
SDS; (a3) EG powder pretreated with ultrasonic irradiation and dried,
with SDS intercalated; (b1) unfilled NBR vulcanizate at wide angle (3–
30); (b2) NBR/EG (5 phr) nanocomposite prepared by latex compound-
ing at wide angle; (b3) NBR/EG (5 phr) composite prepared by direct
blending; (c1) unfilled NBR vulcanizate at small angle (0.5–10); (c2) EG Fig. 3. Mechanical properties: (a) tensile stress vs. strain curves of NBR/
powder pretreated with ultrasonic irradiation with SDS intercalated; (c3) EG nanocomposites prepared by latex compounding with different EG
NBR/EG (5 phr) nanocomposite prepared by latex compounding at small content; and (b) comparisons of efficiency of reinforcement of different
angle. fillers (i.e. carbon black N330 and silica) at the same loading of 10 phr.
J. Yang et al. / Acta Materialia 55 (2007) 6372–6382 6377

10 phr, respectively. For nanocomposites prepared by


LCM, the tensile strength increases substantially with EG
content in the range 0–10 phr. Also, the stress at small
strain (under 100%) increases more remarkably with
increasing EG, when compared to NBR/EG (10 phr) com-
posites by direct blending (Fig. 3b). The different tensile
properties of NBR/EG vulcanizates prepared by LCM
and direct blending can be attributed to the different dis-
persion quality of graphite sheets. The fine dispersion of
graphite sheets achieved by LCM generates more interfa-
cial interaction, resulting in better reinforcement.
It can be seen from Fig. 3b that NBR/EG (10 phr) nano-
composite prepared by LCM exhibits the highest reinforce-
ment efficiency compared to those of composites reinforced
with other traditional reinforcements for rubbers, such as
carbon black and silica. The high aspect ratio and large
surface areas of graphite nanosheets offer much stronger
reinforcement. In addition, the polar groups on the EG
surface, e.g. –OH and –COOH, might interact physically
and chemically with rubber macromolecules, leading to
strong interfacial interactions, which are further substanti-
ated by the increased glass transition temperatures of the
nanocomposites.

3.4. Dynamic mechanical properties

Fig. 4 shows the temperature dependence of the storage


modulus (E 0 ) and tan d of NBR/EG nanocomposites. It can
be seen that the storage modulus at a given temperature
above Tg increases with increasing EG content, resulting
in an increased E 0 compared with unfilled NBR. In
addition, the glass transition is strongly shifted to higher Fig. 4. Dynamic-mechanical properties: (a) storage modulus and (b) loss
temperature as EG content increases, when compared to factor of NBR/EG nanocomposites with different EG content.
the clay-reinforced rubber nanocomposites reported else-
where [50–52], which is attributed to the markedly reduced
mobility of the NBR chains around the EG sheets. Also, conducted on the pressed EG sheets: an O/C ratio of about
the area of the tan d peak is greatly reduced with increasing 7–8% was found together with an asymmetric C1s peak cen-
EG content, indicating significantly reduced damping. tered at 284.5 eV with a shoulder to higher binding energy
Conversely, for NBR/EG composites prepared by direct (see Fig. 5a), indicating the existence of epoxide and hydro-
blending, Tg changes little with increasing EG content. xyl groups on the EG surface [38,39]. The physical interac-
The Tg of NBR/EG (10 phr) composite prepared by direct tion between these active polar groups and NBR molecules
blending is only 5 C higher than that of neat NBR. Sim- (with C”N) would enhance the reinforcement of graphite
ilarly, the enhancement of storage modulus in NBR/EG sheets and lead to a greatly increased glass transition tem-
(10 phr) composite prepared by direct blending is inferior perature of the NBR/EG nanocomposites. In addition, the
to that of NBR/EG nanocomposites prepared by LCM. electron spin resonance (ESR) spectrum of EG (Fig. 5b)
With 10 phr EG, the storage modulus at 40 C is enhanced shows one strong, intense signal, indicating the existence
by 1.8 times only in the former composite compared to that of free radical intermediates [39,53]. After thermal decom-
of neat NBR, but the enhancement in the latter nanocom- position of EG precursor or heat treatment in air, free rad-
posite is 22.4 times. icals form as a consequence of pyrolytic bond cleavage (e.g.
evolution of CO2, CO), and these can be stabilized either
3.5. Self-crosslinking of nanocomposites by the formation of a graphene sheet with free edges or
by the saturation of residual reactive sites, e.g. by O2
In addition to the dispersion of graphite nanosheets, the chemisorption, as reported previously for carbon species
interaction between graphite and rubber should also be [39,53–56]. These radicals could react with the C@C bonds
considered in explaining the reduced mobility of rubber and adjacent a-H of the rubber macromolecules at elevated
molecules and the strong reinforcement of EG fillers. To temperature, enhancing the interfacial interaction and
study the nature of the interfacial interaction, XPS was facilitating the reinforcement of graphite sheets.
6378 J. Yang et al. / Acta Materialia 55 (2007) 6372–6382

Fig. 6. Curing dynamics of NBR nanocomposites with 5 phr graphite.

the dispersion of EG filler is poor, as in the composite pre-


pared by direct blending without a curing agent, the torque
remains constant with time (curve d), indicating that the
dispersion quality of EG sheets plays a significant role in
establishing effective reactions between EG and NBR mac-
romolecules. For nanocomposites from LCM, the good
dispersion of EG sheets provides a much higher contact
area between EG sheets and rubber molecules, so the active
sites on EG (e.g. the free radicals indicated by ESR analy-
sis) can react with the NBR macromolecules, resulting in
covalent bonding between EG and rubber.

3.6. Temperature dependence of self-crosslinking


Fig. 5. Characterization of EG structures: (a) XPS, using XPSPEAK
software; and (b) ESR.
To further investigate the self-curing effect of the rubber
nanocomposites containing graphite nanosheets, SBR/EG
ODR is a well-established technique for studying the and NR/EG nanocomposites were also prepared using
dynamics of the rubber curing reaction by monitoring the LCM. As shown in Fig. 7, the dispersion quality of graph-
modulus change of rubber composites with time. This ite sheets in these rubber matrices is quite satisfactory. It is
method can also be used to investigate the evolution of also found that the self-curing effect of EG is temperature
interfacial interactions caused by chemical reactions that dependent. Fig. 8 shows the ODR results of the NBR
have occurred at the processing temperature. Therefore,
we performed ODR analysis on compounds without curing
agents (e.g. sulfur). As shown in Fig. 6, the initial torque
(e.g. within the first 1 min) of the NBR/5 phr EG nano-
composite (curve b) is much higher than the torques of
pure NBR (curve a) and directly blended NBR/5 phr EG
composite (curve d). This finding is attributed to the strong
interactions between polar groups of EG sheets and polar
CN groups of NBR macromolecular chains, in addition
to the fine dispersion of EG in rubber matrix. Interestingly,
the torque of the NBR/5 phr EG nanocomposite without
any curing reagents increases rapidly with time (curve b)
at 160 C, clearly confirming that a chemical crosslinking
reaction of NBR, which is very unusual, has occurred. In
contrast, the torques of neat NBR or NBR filled with
5 phr SDS remain almost unchanged (curves a and c).
These results strongly suggest the self-curing effect of Fig. 7. TEM images of (a) SBR and (b) NR nanocomposites with 5 phr of
NBR which is induced by EG. More importantly, when graphite.
J. Yang et al. / Acta Materialia 55 (2007) 6372–6382 6379

in NR/graphite nanocomposites in the temperature range


140–190 C (not shown). Hence, the torque of the NR com-
pound normally decreases with increasing temperature.
For self-crosslinked NBR/graphite nanocomposites
(1 h curing at 160 C), we conducted swelling experiments
to measure the crosslink density. The index for crosslinking
density of NBR/graphite nanocomposites is 1.18 · 102
mol cm3, which is very close to that of the same nanocom-
posites cured with sulfur, i.e. 1.22 · 102 mol cm3. This
again confirms the chemical self-crosslinking reaction of
the NBR matrix induced by graphite sheets. When SBR/
graphite nanocomposite was cured at 160 C, it could still
be partially dissolved in toluene. When the curing temper-
ature was higher (i.e. 190 C), the SBR/graphite plate could
no longer be dissolved in toluene, and its crosslinking den-
sity reaches 2.03 · 105 mol cm3. Due to the fact that the
vulcanization of NR/graphite nanocomposites is insuffi-
cient in the temperature range studied, the corresponding
plates dissolve in toluene, although at a much slower rate
than the neat component. This implies the possible forma-
tion of strong interfacial interactions between filler and
matrix.

3.7. Mechanism of self-crosslinking

The crosslinking of rubber/graphite nanocomposites


seems to be very similar to the curing process of rubbers
by peroxide in terms of the curing characteristics. Firstly,
the vulcanization rate is relatively low with a short scorch
time compared with the most widely used process, i.e.
accelerated sulfur vulcanization. Secondly, vulcanization
efficiency strongly depends on the rubber type, which is
Fig. 8. Temperature dependence of self-curing reaction: (a) NBR/5 phr
graphite nanocomposites, and (b) SBR/5 phr graphite nanocomposites.
similar to crosslinking by peroxide [57].
As indicated by XPS experiments, EG sheets contain C–
O and C–OH groups, which interact strongly with the
and SBR systems at different temperatures. As shown in polar groups of NBR, but do not seem to result in chemical
Fig. 8a, the torque of NBR/5 phr graphite nanocomposite bonding at the interface, especially for SBR, which lacks
increases distinctly and rapidly with time at temperatures polar groups. More importantly, this interaction happens
above 120 C, and gradually reaches a steady state. Below only at the interface and cannot lead to crosslinking of
100 C, no crosslinking reaction takes place. At 110 C and the whole rubber phase. The curing characteristics of
above, the rate of change of torque with time increases NBR/graphite and SBR/graphite nanocomposites strongly
markedly, particularly in the early stage of crosslinking. suggest that possible radicals or peroxide groups on EG
Generally, the torque of a rubber compound in an uncross- nanosheets act as initiators for the self-crosslinking behav-
linked state decreases with increasing temperature. How- ior of these rubber/graphite nanocomposites.
ever, this tendency is disturbed in NBR/EG system The radicals, as discussed in the ESR test on the EG fil-
because the crosslinking reaction induced by EG at higher ler, could react with the rubber macromolecules at elevated
temperatures becomes faster, and thus greatly and rapidly temperatures, either through the addition of the radical to
increases the torque of the NBR compound. a double bond, or the abstraction of a hydrocarbon atom
The temperature-dependent self-crosslinking behavior in the allylic position. Consequently, crosslinking could
of SBR/graphite nanocomposites is also evident in form either by the combination of two polymeric free rad-
Fig. 8b. As the temperature approaches 160 C, the SBR/ icals or by a chain reaction involving the addition of poly-
EG compound starts to show distinct curing dynamics. meric free radicals to double bonds. A chain reaction is
Similarly, pure SBR cannot be cured without a curing more likely in butadiene rubbers (i.e. NBR and SBR)
agent. Compared with the NBR system, the self-crosslink- because the double bonds are quite accessible and radical
ing rate of SBR compound is lower but the torque change addition gives highly reactive radicals that are easily added
in ODR tests increases more quickly with temperature. to other polymer chains [57,58]. More significantly, cross-
However, based on the ODR test, no curing is observed linking points could be formed without loss of free radicals.
6380 J. Yang et al. / Acta Materialia 55 (2007) 6372–6382

The slower curing rate and lower vulcanization efficiency of


SBR nanocomposites compared with NBR could be caused
by a higher stability of SBR to radicals. In the case of NR
nanocomposites, when the radicals on EG react with rub-
ber, the abstraction route predominates over radical addi-
tion, so that crosslinking occurs mainly via the coupling
of two polymeric radicals [57–59]. However, the abstrac-
tion reaction is quite limited since it happens only at the
contact points of the reactive radicals on EG and allylic
hydrogen atoms on the NR chains. Before two polymeric
radicals can unite, unproductive reactions would happen,
e.g. heterolytic cleavage, rearrangement, transfer, scission,
degradation, to the exclusion of crosslinking [60]. More-
over, even if the coupling does take place, the process
wastes two radicals. Therefore, the curing efficiency of
NR nanocomposites could be quite low, as displayed in
Fig. 9. FT-IR spectra of NBR/5phr graphite nanocomposites, before and
the ODR tests. However, this mechanism may still generate after self-curing at 160 C for 1 h.
chemical bonding at the interface between EG and NR,
which is reflected by the swelling experiment results.
More evidence of the self-crosslinking phenomenon is romolecular chain does not take part in the curing reac-
provided by the strong shift of the glass transition of the tions, and the IR peak is at m = 2236 cm1. For the
nanocompounds compared with neat rubber, as shown in uncured nanocomposite, the peak at m = 1644 cm1 corre-
Table 3. After flocculation from the latex mixture and dry- sponds to C@C stretch vibration. Compared to the change
ing in a vacuum oven at 50 C for 24 h, uncured NBR in peak at m = 2236 cm1 after self-crosslinking, the inten-
nanocomposites (before any heat treatment at high temper- sity of the peak at m = 1644 cm1 is distinctly reduced. This
ature) already show higher Tg than neat rubber. This is indicates that part of the C@C participates in the curing
because of the fine dispersion of the sheets and the strong reaction, which is consistent with the previous discussion
interfacial interaction between the polar groups (C”N) of on radical curing mechanisms. The absorption peak at
NBR and the polar groups of EG sheets, which greatly m = 1713 cm1, which usually corresponds to a carboxyl
restricts the mobility of the rubber chain segments. The group in EG, is capable of producing a physical interaction
Tg of the SBR and NR compounds does not change much with the C”N group.
because no polar groups exist in their polymer chains. The self-cured nanocomposites also possess better
After curing (1 h heat treatment at high temperature), the mechanical properties. For example, NBR/EG (5 phr)
Tg of all nanocomposites shifts towards higher tempera- nanocomposite, which is highly crosslinked in the presence
ture, especially for NBR and SBR compounds. Since the of filler (1 h curing at 160 C), has a tensile strength and an
chemical bonding for the two compounds is established initial modulus that is respectively 2 and 5 times higher
at the interface, and NBR and SBR matrices are also cross- than that of sulfur-cured gum.
linked to a certain extent after curing, the mobility of rub-
ber chain segments is further restricted. For the NR 3.8. Functional properties
compound, a widely used sulfur system was introduced
via a two-roll mill in order to obtain necessary crosslinking, With a fine dispersion of EG in NBR and a strong inter-
which also partly contributed to the increase in Tg. face between EG/NBR by LCM, EG filler should confer
The IR spectra of NBR/EG (5 phr) nanocomposites good functional properties on the rubber nanocomposites.
before and after self-crosslinking are shown in Fig. 9. It Similar to layered silicates, EG sheets also show superior
is reasonable to assume that the C”N group on NBR mac- gas barrier properties. As given in Table 4, the nitrogen
permeability of NBR/10 phr EG nanocomposite is reduced
by 62% compared to that of unfilled NBR vulcanizate;
Table 3 however, the permeability of the composite prepared by
Glass transition temperatures of rubber/graphite nanocomposites by
DMTA
direct blending is only decreased by 43%. EG sheets also
offer the advantage over layered silicates, that they make
Uncured Rubber/5 phr graphite nanocomposites
pure
NBR electrically and thermally conductive. NBR nano-
Uncured Cured Curing process composites become conductive when the EG loading
rubber
(C) (C)
(C) reaches 5 phr. The electrical conductivity of the nanocom-
NBR 14.1 0.7 3.7 160 C, 1 h, no curing agent posite with 10 phr EG is even higher than the rubber
SBR 42.4 41.4 29.5 190 C, 1 h, no curing agent composites filled with 30–40 phr conductive carbon
NR 57.8 56.7 53.0 143 C, 10 min, with curing black particles [61,62]. The thermal conductivity of NBR
agent is substantially improved by the addition of EG. The
J. Yang et al. / Acta Materialia 55 (2007) 6372–6382 6381

Table 4 (ARC) for the financial support of this project. Z.Z.Y.


Functional properties of NBR/EG nanocomposites prepared from LCM and Y.W.M. are Australian Postdoctoral Fellow and Aus-
EG content (phr) 0 3 5 10 10a tralian Federation Fellow, respectively, supported by the
N2 permeability 3.19 – 1.98 1.21 1.83 ARC and tenable at the University of Sydney.
(1017 m2 Pa1 s1)
Friction coefficientb 1.72 – – 1.36 1.38
Wear rate 5 · 102 – – 8 · 104 3.5 · 102 References
(mm3 N1 m1)
Electrical conductivity · · 2 · 103 3 · 101 · [1] Novak BM. Adv Mater 1993;5:422.
(S cm1)c [2] Mark JE. Polym Eng Sci 1996;36:2905.
Thermal conductivity 0.19 – – 0.30 0.23 [3] Vollath D, Szabó DV. Adv Eng Mater 2004;6:117.
(W m1 K1) [4] Messersmith PB, Giannelis EP. Chem Mater 1993;5:1064.
a [5] Giannelis EP. Adv Mater 1996;8:29.
By direct blending.
b [6] Zilg C, Thomann R, Finter J, Mulhaupt R. Macromol Mater Eng
Measured under a normal pressure of 0.1 MPa and sliding speed of
2000;280/281:41.
0.1 m s1.
c [7] Ray SS, Okamoto M. Prog Polym Sci 2003;28:1539.
Flocculated compounds.
[8] Usuki A, Hasegawa N, Kato M. Adv Polym Sci 2005;179:135.
[9] Wu YP, Wang YQ, Zhang HF, Wang YZ, Yu DS, Zhang LQ, et al.
nanocomposite prepared by LCM shows higher thermal Compos Sci Technol 2005;65:1195.
conductivity than that produced by direct blending. The [10] Ajayan PM, Zhou OZ. Topics Appl Phys 2001;80:391.
emergence of this greatly enhanced functional performance [11] Dai L, Mau AWH. Adv Mater 2001;13:899.
is mainly due to the fine dispersion of EG sheets and the [12] Coleman JN, Khan U, Gun 0 ko YK. Adv Mater 2006;18:689.
[13] Chung DDL. J Mater Sci 2002;37:1475.
formation of a conductive network (see Table 4). [14] Stankovich S, Dikin DA, Dommett GHB, Kohlhaas KM, Zimney EJ,
The addition of EG also greatly enhances the wear prop- Stach EA, et al. Nature 2006;442:282.
erties of NBR in terms of friction coefficient and wear rate. [15] Shen JW, Chen XM, Huang WY. J Appl Polym Sci 2003;88:1864.
Although the NBR/10 phr EG nanocomposite prepared by [16] Chen GH, Weng WG, Wu DJ, Wu CL. Eur Polym J 2003;39:2329.
LCM has a similar friction coefficient (l) to the NBR/ [17] Du XS, Xiao M, Meng YZ. J Polym Sci Part B Polym Phys
2004;42:1972.
10 phr EG composite prepared by direct blending, the for- [18] Lu JR, Weng W, Chen XF, Wu DJ, Wu CL, Chen GH. Adv Funct
mer shows a much lower wear rate (<103 mm3 N1 m1) Mater 2005;15:1358.
than the latter, and its wear rate is comparable to that of [19] Jia W, Tchoudakov R, Narkis M, Siegmann A. Polym Compos
NBR filled with at least 30–50 phr high abrasion furnace 2005;26:526.
[20] Chen GH, Wu DJ, Weng WG, Yan WL. J Appl Polym Sci
black (N330). The improved wear properties are significant
2001;82:2506.
and desirable. [21] Chen GH, Wu DJ, Weng WG, He B, Yan WL. Polym Int
2001;50:980.
4. Conclusions [22] Pan YX, Yu ZZ, Ou YC, Hu GH. J Polym Sci Part B Polym Phys
2000;38:1626.
To achieve high-performance rubber nanocomposites [23] Drzal LT, Fukushima H. Polym Prep 2001;42:42.
[24] Yasmin A, Daniel IM. Polymer 2004;45:8211.
that combine improved mechanical and functional proper- [25] Yasmin A, Luo JJ, Daniel IM. Compos Sci Technol 2006;66:1179.
ties, a facile approach has been developed to finely disperse [26] Liu DW, Du XS, Meng YZ. Polym Polym Compos 2005;13:815.
graphite nanosheets in NBR matrix by latex compounding, [27] Mack JJ, Viculis LM, Ali A, Luoh R, Yang GL, Hahn HT, et al. Adv
which results in a high extent of exfoliation and intercala- Mater 2005;17:77.
[28] Salkind M. Adv Mater 1989;1:157.
tion of graphite in NBR matrix and an unexpected self-
[29] Yang X, Grosjean C, Tai Y. J Microelectromech Syst 1999;8:393.
crosslinking of rubber/graphite nanocomposites. With very [30] Ha SM, Yuan W, Pei Q, Pelrine R, Stanford S. Adv Mater
fine dispersion of graphite nanosheets, their active sites 2006;18:887.
could well initiate reactions with rubber macromolecules [31] Liu L, Lu YL, He L, Zhang W, Yang C, Liu YD, et al. Adv Funct
and result in crosslinking of certain types of rubbers, e.g. Mater 2005;15:309.
NBR and SBR in this study, at elevated temperatures. [32] Hamed GR. Rubber Chem Technol 2000;73:524.
[33] Zhang LQ, Jia DM. The nano-reinforcing technique and science of
The reaction is very significant in establishing a strong, rubber. In: Symposium of international rubber conference 2004,
chemically bonded interface in these nanocomposites. Beijing, China, September 21–25, 2004, vol. A, p. 46.
The NBR/EG nanocomposites possess greatly increased [34] Zhang LQ, Wang YZ, Wang YQ, Sui Y, Yu DS. J Appl Polym Sci
modulus and strength, improved wear resistance, better 2000;78:1873.
gas barrier properties and superior electrical/thermal con- [35] Wu YP, Ji MQ, Qi Q, Wang YQ, Zhang LQ. Macromol Rapid
Commun 2004;25:565.
ductivity. These functional properties make the rubber/ [36] Varghese S, Karger-Kocsis J. Polymer 2003;44:4921.
EG nanocomposites a promising new class of advanced [37] Gatos KG, Martı́nez Alcázar JG, Psarras GC, Thomann R, Karger-
materials. Kocsis J. Compos Sci Technol 2007;67:157.
[38] Hontoria-Lucas C, López-Peinado AJ, López-González JdeD, Rojas-
Cervantes ML, Martı́n-Aranda RM. Carbon 1995;33:1585.
Acknowledgements
[39] Szabó T, Berkesi O, Forgó P, Josepovits K, Sanakis Y, Petridis D,
et al. Chem Mater 2006;18:2740.
We thank the National Natural Science Foundation of [40] Lerf A, He HY, Forster M, Klinowski J. J Phys Chem B
China (50403029) and the Australian Research Council 1998;102:4477.
6382 J. Yang et al. / Acta Materialia 55 (2007) 6372–6382

[41] Wang WP, Pan CY. Eur Polym J 2004;40:543. [52] Shokri AA. Polym Int 2006, published online, DOI: 10.1002/pi.2076.
[42] Schniepp HC, Li JL, McAllister MJ, Sai H, Herrera-Alonso M, [53] Collins RL, Bell MD, Kraus G. J Appl Phys 1959;30:56.
Adamson DH, et al. J Phys Chem B 2006;110:8535. [54] Voudrias EA, Larson RA, Snoeyink VL. Carbon 1987;25:503.
[43] Chung DDL. J Mater Sci 1987;22:4190. [55] Manivannan A, Chirila M, Giles NC, Seehra MS. Carbon
[44] Flory PJ, Rehner J. J Chem Phys 1943;11:521. 1999;37:1741.
[45] Hwang W, Wei K, Wu C. Polym Eng Sci 2004;44:2117. [56] Hwang YL, Yang CC, Hwang KC. J Phys Chem A 1997;101:7971.
[46] Deng JS, Isayev I. Rubber Chem Technol 1991;64:296. [57] Coran AY. Vulcanization. In: Eirich RR, editor. Science and
[47] Bieliński M, Ślusarski L. Wear 1993;169:257. Technology of Rubber. New York: Academic Press; 1978. p. 325.
[48] Yang J, Tian M, Jia QX, Zhang LQ, Li XL. J Appl Polym Sci [58] Dluzneski PR. Rubber Chem Technol 2001;74:451.
2006;102:4007. [59] van der Hoff BME. Ind Eng Chem Prod Res Dev 1963;2:273.
[49] Chen GH, Weng WG, Wu DJ, Wu CL, Lu JR, Wang PP, et al. [60] Tobolsky AV, Mercurio A. J Am Chem Soc 1959;81:5535.
Carbon 2004;42:753. [61] Sau KP, Chaki TK, Khastgir D. J Appl Polym Sci 1999;71:887.
[50] Kim J, Oh T, Lee D. Polym Int 2003;52:1058. [62] Sau KP, Khastgir D, Chaki TK. Die Angew Makromol Chem
[51] Liu L, Jia D, Luo Y, Guo B. J Appl Polym Sci 2006;100:1905. 1998;258:11.

Você também pode gostar