Sie sind auf Seite 1von 9

Polymer 54 (2013) 685e693

Contents lists available at SciVerse ScienceDirect

Polymer
journal homepage: www.elsevier.com/locate/polymer

Confocal Raman imaging, FTIR spectroscopy and kinetic modelling of


the zinc oxide/stearic acid reaction in a vulcanizing rubber
Pellegrino Musto a, *, Domenico Larobina b, Salvatore Cotugno c, Paolo Straf c, Giuseppe Di Florio a, d,
Giuseppe Mensitieri d
a

Institute of Chemistry and Technology of Polymers, National Research Council of Italy, Via Campi Flegrei 34, Olivetti Buildings, 80078 Pozzuoli (NA), Italy
Institute of Composite and Biomedical Materials, National Research Council of Italy, P.le Tecchio 80, 80125 Naples, Italy
Bridgestone Technical Center Europe S.p.A., Via del Fosso del Salceto 13/15, Rome, Italy
d
Department of Materials and Production Engineering, University of Naples Federico II, P.le Tecchio 80, 80125 Naples, Italy
b
c

a r t i c l e i n f o

a b s t r a c t

Article history:
Received 18 September 2012
Received in revised form
5 December 2012
Accepted 6 December 2012
Available online 16 December 2012

The reaction of zinc oxide (ZnO) with stearic acid (StA) to form zinc stearate (ZnSt) has been investigated
experimentally in a model matrix (unvulcanized styreneebutadiene rubber) by using confocal Raman
microscopy and FTIR transmission spectroscopy. The heterogeneous nature of the reacting system has
been conrmed. The Raman analysis has revealed the coreeshell structure of the product, which is
formed via the gradual shrinkage of the ZnO core and the concurrent formation of a surrounding ZnSt
shell of increasing thickness. FTIR spectroscopy has provided information about the molecular state of
aggregation of StA when dissolved in the rubber, as well as quantitative information on the reaction
kinetics. The kinetic behaviour of the system has been interpreted using a semi-quantitative heterogeneous reaction model grounded on the Raman imaging results, which was able to catch the essential
features of the phenomenon and to simulate reliably the experimental conversion vs time data at three
different temperatures.
2012 Elsevier Ltd. All rights reserved.

Keywords:
Zinc oxide
SBR rubber
Vibrational spectroscopy

1. Introduction
Vulcanization (i.e. the cross-linking reaction by sulphur) is one of
the oldest and best developed processes in the rubber industry.
Since its discovery by Goodyear in 1839, it has been continuously
improved by the introduction of new ingredients (accelerators,
activators, retarders), in order to improve mechanical properties.
Nowadays, in the tyre industry, where the process represents the
core technology, a vulcanizing rubber formulation is a very complex
system comprising tens of components. Most of these formulations
have been developed and optimized by skilful and lengthy techniques based on trial-and-error but, owing to the complexity of the
chemistry involved and the multiplicity of ingredients, the mechanism of vulcanization has not been fully elucidated yet. It is generally recognized that the vulcanizing rubber is a heterogeneous
system which, in turn, gives rise to a heterogeneous vulcanizate at
the end of the process. How to precisely measure the degree of
heterogeneity and to control it for a rational product design,
remains an open issue [1]. Another relevant problem encountered in

* Corresponding author.
E-mail address: musto@ictp.cnr.it (P. Musto).
0032-3861/$ e see front matter 2012 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.polymer.2012.12.021

rubber processing is represented by undesired side reactions,


whose occurrence increases with the number and reactivity of the
formulation components. For instance, zinc oxide is still the best
known activator for sulphur vulcanization.
Zinc oxide (ZnO) is in many vulcanisation systems a precursor to
zinc-derived accelerators; it reacts with most accelerators to form
the highly active zinc salt [1,2] Complex formation of the zinc ion
with different accelerators is critical to get efcient curing. In
addition, there is also evidence that the inclusion of ZnO in
a compound reduces heat-build-up and improves abrasion resistance. Due to its high thermal conductivity, ZnO acts as a heat sink,
which accepts frictional energy dissipating local heat concentrations without a large increase in internal temperature. It has also
been found that ZnO improves heat resistance of the vulcanisates
and their resistance to the action of dynamic loading [2e6].
Moreover, several fatty acids are employed in rubber formulations
as plasticizers, but for a few of them (for instance stearic acid, StA) it
has been postulated that, beside acting as an internal lubricant [7],
they play a further role as co-activators [4]. By reacting with stearic
acid, zinc oxide forms water and the hydrocarbon-soluble zinc
stearate (ZnSt), which has been claimed to be the actual activator
species [7]. In fact, when the zinc coordinates the carboxylate
ligands of stearic acid, the solubility and reactivity is strongly

686

P. Musto et al. / Polymer 54 (2013) 685e693

enhanced. It has been reported [8aeh] that, in sulphur vulcanization of rubbers, the crosslink density increases with increasing zinc
stearate concentration, while the effect of stearic acid consists in
increasing the initial cure and in lowering the level of polysuldic
cross-links, thus enabling a more efcient use of sulphur [9,10].
It has also been suggested that both these effects are caused by
stearic acid allowing a better dispersion of ZnO. In another study
[11], it was observed that, if the vulcanizing medium contains zinc
stearate but no stearic acid, the end-properties of the nal products
worsened.
From the above arguments it follows that a quantitative characterization of the zinc oxide/stearic acid reaction, not only in terms
of overall conversion but also with respect to the heterogeneity of
the system, is of paramount importance towards a deeper understanding of the vulcanization mechanism. Previous literature
reports have shown that, depending on the mixing conditions, the
two reactants can already form zinc stearate during the compounding of the rubber. The calorimetric (DSC) results of Kruger
and McGill [12] indicated the formation of zinc stearate and its
subsequent reaction with sulphur, but the phase separation of the
system components was not investigated.
In the present contribution we examine the interaction between
zinc oxide and stearic acid in a medium (uncrosslinked styrenee
butadiene rubber, SBR) suitable to simulate the actual vulcanizing
system. The dispersion of the components before and after the
reaction is investigated by confocal Raman microspectroscopy. This
technique offers unique advantages in the characterization of
heterogeneous polymer systems: it provides good contrast due to
the specicity of the spectrum, molecular-level information and the
possibility to implement quantitative or semi-quantitative analysis.
The limiting spatial resolution (0.8e1.0 mm in the focal plane) is
adequate for the systems under scrutiny. Furthermore, the state
of aggregation of stearic acid in the rubber matrix and its
reaction with zinc oxide are investigated by FTIR spectroscopy,
which provides information on the H-bonding interaction of the
stearic acid dimer and kinetic data at different temperatures.
These are used to test the reliability of a semi-quantitative kinetic
model developed on the basis of the spectroscopic evidences and
based on a simple heterogeneous reaction scheme.
2. Experimental
2.1. Materials
Styreneebutadiene rubber (SBR) had a styrene/butadiene molar
ratio of 20/80 and a trans/vynil/cis C]C ratio of 24/56/20. Zinc oxide
with a purity >99% was supplied by Bridgestone Technical Center
Europe (TCE), Pomezia, Italy. The particle size distribution of ZnO, as
determined by static light scattering (SLS), is reported in Fig. 1. The
number average equivalent diameter (Dn ) was 1.775 mm. Reagent
grade stearic acid and zinc stearate were supplied by Aldrich
Co., Milan (Italy) and were used without further purication.
2.2. Sample preparation
The adopted composition of the reactive mixture was representative of a commercial formulation. To prepare the ternary reactive
mixture, 85 g of SBR were dissolved in cyclohexane (5% wt/vol) at
ambient temperature for 24 h under vigorous mechanical stirring.
After complete dissolution, 9.0 g of stearic acid were added to the
highly viscous solution, together with a suspension containing 4.0 g
of Zinc oxide in 50 mL of cyclohexane. The stirring was continued
until a homogeneous and stable dispersion of zinc oxide was
attained, as demonstrated by the appearance of a uniform milky
colour. Films with a thickness of 20e30 mm were prepared by

Fig. 1. Particle size distribution (in terms of equivalent particle diameter) of ZnO as
determined by SLS measurements.

solution casting on Petri dishes, followed by slow solvent


evaporation (24 h at ambient temperature). Residual traces of
solvent were eliminated under vacuum at 40  C. Complete
elimination of the solvent was assessed by the absence of solvent
bands in the FTIR spectrum of the lm.
The sample with only SBR and stearic acid, used to investigate
the StA monomer-to-dimer equilibrium, and several samples containing SBR and zinc stearate, used as standards for analytical
purposes, were prepared in the same way as the ternary mixture,
with the appropriate amount of the respective components.
2.3. Techniques
Raman spectra were collected by a confocal Raman spectrometer (Horiba-Jobin Yvon Mod. Labspec Aramis) equipped with
a HeeNe laser source emitting at 632 nm. The sample was placed
on a piezo-electrically driven microscope stage with a x,y resolution
of 10  0.5 nm and a z resolution of 15  1 nm and scanned at
a constant stage speed with a 2.0 mm step size. The 180 backscattered radiation was collected by an Olympus metallurgical
objective (MPlan 50, NA 0.75) with confocal and slit apertures
both set to 200 mm. A grating with 1200 grooves/mm was used
throughout. The radiation was focused onto a Peltier-cooled CCD
detector (Synapse Mod. 354308) operating in the Raman-shift
range 3200e100 cm1 for single point measurements or in the
range 1600e200 in the case of the mapping experiments, to
reduce the acquisition time. Exposure times were 30 s for the single
point acquisition and 25 s for the mapping experiments. The data
gathered by the instrument were converted to ASCII format and
transferred to the MATLAB (Mathworks, Natick, MA) computational
platform, for further processing. The Raman images were elaborated by programs written in our laboratory, making use of the
image processing facilities and the surface interpolating algorithms
of the MATLAB environment.
FTIR spectra were collected in the transmission mode using
a System 2000 spectrometer from PerkineElmer. The instrument
was equipped with a Germanium/KBr beam splitter and
a deuterated tryglycine sulphate (DTGS) wide-band detector. The
instrumental parameters adopted for the spectral collection were
as follows: resolution 4 cm1, scan speed [i.e. Optical Path
Difference (OPD) velocity] 0.20 cm s1, spectral range 4000e
400 cm1. For the kinetic measurements, a single data collection
was performed for each spectrum, which took 6 s to complete. It
was found that, even with a single acquisition, the signal-to-noise
ratio of the spectra (5000:1) was suitable for quantitative
analysis.
FTIR measurements were performed at different temperatures
on SBR containing only StA and on the reactive system comprising

P. Musto et al. / Polymer 54 (2013) 685e693

SBR, ZnO and StA. Tests were carried out in an environmental


chamber constructed in house by modifying the commercially
available SPECAC 20100 cell. This unit was controlled by a Eurotherm PID temperature regulator mod. 2416, with an accuracy of
0.5  C. The chamber was directly mounted in the spectrometer
and the spectra were taken at selected time intervals to monitor insitu the resulting chemical changes. All measurement was carried
out under a continuous ux of N2 (60 cm3 min1).
3. Results and discussion
3.1. Raman spectroscopy
In Fig. 2AeD are reported the Raman spectra of the system
components.
Zinc oxide displays two sharp peaks at 437 and 98 cm1
(E2 modes) [13] which occur in an interference-free region and
provide the required spectroscopic contrast. Stearic acid also
displays intense inelastic scattering and a series of peaks in a region
where the rubber matrix does not interfere (see inset of Fig. 2C).
The intense and sharp components at 1296, 1130 and 1062 cm1
have been assigned, respectively, to the CH2 twist, and to the
symmetric and anti-symmetric CeC stretching modes [14,15]. The
prominent intensity and the sharpness of these peaks are associated with the ordered crystalline structure, as demonstrated by
the inset of Fig. 2C which compares the spectrum in the 1600e
800 cm1 range taken at 30  C (blue trace in web version) and
above the melting temperature, (80  C, red trace). In particular,
the sharp triplet at 1130, 1101, 1062 almost vanishes on melting,
thus providing a reliable spectroscopic signature of the crystalline
phase.
Being the Raman spectrum of stearic acid strongly dominated by
CH and CC modes, no major differences are expected between the
latter and the spectrum of Zinc stearate. In fact, the region between

687

1300 and 1000 cm1 is essentially coincident (compare Fig. 2C and


D), which, in the present context, prevents the possibility to
discriminate the two molecular species from their Raman signals.
A preliminary Raman analysis was carried out to test the state of
aggregation and the degree of dispersion of stearic acid in SBR
achieved with our solution casting procedure. The mixture was
initially prepared without ZnO, to prevent the possible reaction
between the components. The Raman image collected on
a 34  34 mm2 area is reported in Fig. 3. There are regions of the map
where the peaks characteristic of the crystalline StA phase are clearly
discernible, while in other areas these peaks are undetectable (see
inset of Fig. 3). The signals intensity was converted to relative
concentration of scattering species making use of a general
formalism whereby the sample and laser variables are factorized
with respect to the collection variables [16]. Thus, we may dene
a specic scattering intensity L, expressed in photons sr1 cm2 s1

L PD bD[

(1)

where
PD laser power density (photon s1 cm2)

b ds=dU=N Raman cross section

cm2 molecules1 sr1


D density of scatterers (molecules cm3)
[ path length illuminated by the laser beam (cm).
In the denition of b, s F=J, where F is the total scattered
power and Jis the irradiance of the incident laser beam. s refers to
the whole solid angle U around the scattering center, while
b represents the fractional intensity scattered into a given direction
within an innitesimal element of the solid angle dU. N is the
number of scattering molecules. Furthermore, an instrumental
collection function C (cm2 sr e photon1) may be dened, which

Fig. 2. Raman spectra of SBR rubber (A); ZnO (B); stearic acid (C); zinc stearate (D).

688

P. Musto et al. / Polymer 54 (2013) 685e693

Fig. 3. Confocal Raman image of a SBR/StA mixture 91/9 wt/wt. Image reconstruction performed by use of the intensity ratio I1126/I620. The inset compares the spectra collected at
the positions indicated.

expresses the fraction of the total scattered light that is collected,


analysed and detected by the instrument:

C AD UD TQ

(2)

where
AD sample area monitored by the spectrometer (cm2)
UD the collection solid angle of the spectrometer
(steradians, st)
T instrumental transmission (unitless)
Q quantum efciency of the detector (e photons1)
Finally, the observed intensity I(n), in photoelectrons, is

In LCt PD bD[,AD UD TQ t

(3)

with t being the exposure time in seconds.


If we consider two fully resolved peaks at frequencies ni and nj,
characteristic of the scattering species i and j, the intensity ratio
Ini =Inj will be:

b ,D ,Tn ,Q ni
Ini
D
n
  i i  i
  K i K i
Dj
nj
bj Dj ,T nj ,Q nj
I nj

(4)

where K bi ,Tni ,Q ni =bj Tnj ,Q nj depends only on the


choice of the analytical peaks and is invariant with the very many
instrumental parameters. In Equation (4), ni, nj represent moles of
the respective species.
The image reconstruction was performed by use of the intensity
ratio I1126/I620 which, according to Equation (4), is proportional to
the concentration ratio CStA/CSBR. This analysis reveals that, in the

rubber matrix, StA segregates in the form of lenticular crystalline


domains having a size from 2 to 5 mm. The observation that, even
when focussing on a StA domain, the SBR spectrum is still dominant
(see inset of Fig. 3, blue trace in web version) is an indication that
the thickness of the crystallites is lower than the spatial resolution
along the z axis (2.5 mm).
Fig. 4A and B display the Raman images of a sample containing
4.0 wt % of ZnO and 9.0 wt % of StA. In particular, Fig. 4A was obtained by considering the intensity ratio I437/I620 and is therefore
representative of the concentration ratio CZnO/CSBR, while Fig. 4B
was reconstructed in the same way as Fig. 3 and denotes the spatial
distribution of StA.
The image analysis reveals a heterogeneous system in which
domains of ZnO coexist with a StA crystalline phase broadly
distributed within the sampled area. In the central region several StA
crystallites coalesced forming a larger aggregate. The ZnO particles
display an elongated shape with main dimensions around 4e5 mm.
The particle size as measured by Raman imaging compares well
with the particle size distribution of ZnO as revealed by SLS, thus
conrming the efciency if the dispersion processes. It is noted that
in several places (i.e. the upper right and the lower left corners) the
domains of the two species are localized at approximately the
same positions, while in other regions (the central area) their
spatial distribution is largely different. Recalling that the analytical
signals corresponding to StA and ZnSt are coincident, this
observation may be interpreted assuming that, during the sample
preparation procedure, although the temperature was purposely
kept as low as possible, a partial reaction occurred between ZnO
and StA, with subsequent precipitation of solid ZnSt crystallites
around the ZnO particles. In other words, during the solution
casting and subsequent drying, StA may undergo two competing
processes, crystallization and reaction with ZnO: the former is

Fig. 4. Confocal Raman images of a SBR/StA/ZnO mixture 87/9/4 by wt. Image reconstruction performed by use of the intensity ratio I437/I620 (A) and of the intensity ratio
I1126/I620 (B).

P. Musto et al. / Polymer 54 (2013) 685e693

strongly predominant but the latter plays a non-negligible role. This


interpretation is conrmed by the FTIR spectra of the as-prepared
solution cast lms which display low intensity peaks characteristic
of ZnSt (vide infra).
In a further experiment, the reactive mixture was allowed to
react at 100  C for 1 h and then cooled at ambient temperature prior
to be analysed by confocal Raman spectroscopy. FTIR analysis
demonstrated that, in these conditions, the StA conversion reached
its maximum value (80%). The Raman images of the above sample
are reported in Fig. 5A, B.
Firstly, it is explicitly noted that, although the thermal treatment
has been carried out in-situ, the region sampled in Fig. 5 is different
from that sampled in Fig. 4. This is because the sample undergoes
a volume expansion when heated, which prevents to localize the
specic area imaged before rising the temperature. When the
reaction is complete the position of most residual ZnO domains
coincides with that of the ZnSt crystallites. It is inferred that at
100  C the StA melts (Tm 60  C, as evaluated by FTIR) and migrates
towards the ZnO domains where it reacts forming ZnSt which
immediately crystallizes around the oxide particles being its
melting point (120e130  C) well above the reaction temperature.
Some of the ZnSt domains (i.e. those denoted 1 and 2) appear
considerably larger than the corresponding ZnO particles, while
in one case (domain denoted as 3) no Zinc oxide is detectable at
the corresponding position. It is likely that, in these cases, the
ZnO particles have been largely depleted by the reaction, leaving
an inner core much smaller than its original size.
3.2. FTIR spectroscopy
As for the Raman analysis, we started with the molecular
characterization of StA when dispersed in the rubber matrix with
no ZnO added. For a 91/9 wt % SBR/StA mixture the peaks due to StA
appear in the form of a doublet centred at 1755 and 1712 cm1. This
is shown in Fig. 6, where the 1820e1650 cm1 range is reported for
spectra taken at different temperatures in the 60e120  C interval.
It is noted that, at ambient temperature (data not reported) both
carbonyl peaks are more complex, exhibiting an unresolved ne
structure with strong asymmetry. This is a consequence of crystal
eld effects which vanish upon melting. The spectral data in Fig. 6
conrm the complete melting of StA at 60  C. The component at
1755 cm1 displays a complex prole which turns out to be due to
the presence of underlying rubber peaks. This is demonstrated by
difference spectroscopy (see inset of Fig. 6), whereby removal of the
SBR spectrum reveals the symmetric, fully resolved shape of the
1755 cm1 peak. The prominent absorption at 1712 cm1 is due to
the n(C]O) vibration of the carboxyl groups self-associated to form
dimers. The relatively low frequency of this mode reects the

689

strength of the H-bonding interaction [17]. The component at


1755 cm1 is the n(C]O) mode of the free carbonyls (i.e. the
monomeric form), whereby the considerable frequency enhancement is due to the stiffening of the C]O force constant as
a consequence of the H-bonding dissociation [17,18]. It is evident
that most of the StA is present in the rubber compound in the
dimeric form, albeit the concentration of free monomers is nonnegligible.
As expected, increasing the temperature causes a decrease of
the main carbonyl component and a concurrent increase of the
secondary component, that is, the concentration of self-associated
StA decreases with temperature. Fig. 7 shows a plot of the overall
n(C]O) band area as a function of temperature: from 60 to 120  C
the integrated absorbance remains essentially constant, while
dropping afterwards. A signicant decrease is also observed for the
remaining StA peaks, while, in the same temperature range, the
rubber bands remain essentially constant. This behaviour indicates
that above 120  C a signicant amount of StA leaves the mixture by
evaporation; thus, only the 60e120  C range can be reliably used to
investigate the monomer-to-dimer equilibrium.
The absorbance vs temperature curves relative to the peaks at
1712 and 1755 cm1 are reported in Fig. 8A.
From the above data the equilibrium value for ln A1712 =A21755
and, hence, the enthalpy of the dimer formation (see Scheme 1) can
be evaluated.
As usual, the DH value results from the slope of the
ln A1712 =A21755 vs 1/T diagram (see Fig. 8B) on account of the Vant
DM eDG=RT which, in terms of absorbance
Hoff equationKEQ

becomes ln A1712 =A21755 DH=RT DS=R ln 1712 =1755 .


The resulting value (7.6 kcal mol1) is in reasonable agreement
with literature data on carboxylic acids (DH 10.4 kcal mol1 for
acetic acid in CCl4 and 13.4  0.5 kcal mol1 for stearic acid in the
liquid state) [17].
To implement the kinetic model (see Section 3.3), it is advisable
to obtain an estimate of the equilibrium constant at the different
investigated temperatures. In order to transform the spectroscopic
DM values (whereK DM C =C 2 ), we made use
observables into KEQ
D
M
EQ
of a D =M ratio equal to 1.6  0.1 [19], which is representative of
a typical carboxylic acid forming a cyclic H-bonding structure as
that represented in Scheme 1. On this basis and using a mass
DM according to:
balance on StA, one can evaluate KEQ
Rubber;INI
CStA
CM 2CD
Rubber
It is explicitly noted that, here and in the following, CStA
represents the equivalent monomeric StA concentration, i.e. the sum
of the concentration of the monomeric form and of twice the

Fig. 5. Confocal Raman images of a SBR/StA/ZnO mixture 87/9/4 by wt reacted at 100  C for 1 h. Spectra taken at ambient temperature. Image reconstruction performed by use of
the intensity ratio I437/I620 (A) and of the intensity ratio I1126/I620 (B).

690

P. Musto et al. / Polymer 54 (2013) 685e693

110  C) are reported in Table 1 along with the values of the equiDM C =C 2 .
librium constant, KEQ
D
M
When ZnO is present in the mixture the reaction is evidenced by
a gradual lowering of the n(C]O) peaks of the monomer and the
dimer and the concurrent increase of a peak at 1538 cm1 due to
the asymmetric stretching vibration of the carboxylate anion. The
corresponding symmetric stretch occurs at 1395 cm1 and is
almost completely overlapped with intense rubber peaks. It has
a very limited analytical usefulness and was not considered further.
In Fig. 9 is shown the time evolution of the FTIR spectrum of
a reactive mixture comprising 9.0 wt % of StA and 4.0 wt % of ZnO,
treated isothermally at 80  C.
To monitor the extent of reaction, the 1538 cm1 peak is to be
preferred over the 1755 or 1712 cm1 peaks because it is fully
resolved, interference-free and avoids any ambiguity resulting from
monomer-to-dimer equilibrium.
In the following, we will describe the evolution of the system by
means of the relative conversion of equivalent monomeric StA, a,
dened as:
Fig. 6. FTIR spectra in the wavenumber range 1820e1650 cm1 of a SBR/StA mixture
91/9 wt/wt. Spectra taken at different temperatures as indicated. The inset displays the
difference spectra obtained by subtracting the plain SBR spectrum collected at the
same temperature.

concentration of the dimeric form. In the interconversion equilibrium experiment, this value is constant since no reaction of the
Rubber;INI
stearic acid occurs and it is equal to its initial value, CStA
.
Using the values of AM (i.e. A1712) and AD (i.e. A1755), which are
available from the spectroscopic analysis, one can write:

CM
CD

9
AM >
>
>

>
M [ =
A
0CD CM D M
>
AM D
AD >
>
>

[ ;

at

Rubber;INI
Rubber t
CStA
 CStA
Rubber;INI
CStA

It is assumed here, in dening a, that stearic acid is mostly


present within the rubber and that the amount dissolved in the
shell of the reacting ZnO particle is negligible for the purposes of
evaluating the total conversion. Accounting for the stoichiometry of
the reaction, which dictates that each ZnO mole reacts with 2 mol
of monomeric StA, the relative conversion of StA can also be
expressed in terms of concentration of ZnSt, i.e.:

at 2,

where [ represents the sample thickness. It follows that:

C Rubber;INI

CM  StA
A
12 D M
AM D
Rubber;INI
where, in the case at hand, CStA
0:291 mol=l. The values of
CM and CD at the three temperature of interest (i.e. 80, 100 and

(5a)

CZnSt t
Rubber;INI
CStA

(5b)

which can be written as:

at 2,

A1538 t
Rubber;INI
1538 ,[,CStA

and, in terms of absorbance:

at

A1538 t
a0
A1538 0

(5c)

The a(0) term in Equation (5c) arises from the fact that both
during the lm preparation procedure and the heating step to reach
the isothermal temperature, the reaction already starts. As
a consequence, at time t 0, (when the acquisition of isothermal

Fig. 7. Overall area of the carbonyl peaks (i.e. A1755 A1712) as a function of temperature for the difference spectra of the SBR/StA mixture 91/9 wt/wt.

Rubber t 0 < C Rubber;INI , i.e. a(0) > 0. The


kinetic data initiates) CStA
StA
a(0) value has been estimated from the normalized value of
A1538(0) by means of a calibration curve constructed with SBR/ZnSt
standards (see experimental).
At the investigated temperatures (i.e. 80, 100 and 110  C) the
stearic acid is assumed to be homogeneously dissolved within the
rubber.
The a vs time curves are collectively reported in Fig. 10. As expected, the reaction rate increases considerably with temperature
but the nal conversion is never complete: As shown in Fig. 10 and
conrmed by the presence of residual StA carbonyl bands at the end
of the reaction, the process seems to approach, in all cases,
a limiting value amax.

P. Musto et al. / Polymer 54 (2013) 685e693

691

Fig. 8. (A): Integrated absorbance area of the 1712 cm1 peak (B) and the 1755 cm1 peak (,); (B): Vant Hoff diagram obtained with the data of Fig. 8A.

2 CH3

CH2

CH3

C
16

O
CH2

O
C

C
16

OH

CH2

16

CH3

Scheme 1. Scheme of equilibrium between monomeric and dimeric forms of stearic acid.

3.3. Modelling the reaction kinetics


In the light of the results of Raman imaging and time-resolved
FTIR spectroscopy, a theoretical model has been formulated to
interpret the physicalechemical process of reaction of ZnO particles with stearic acid dispersed in the unvulcanized rubber matrix
making some relevant assumptions which afford a considerable
simplication of the problem without missing the main features of
the occurring phenomena.
Indeed, the Raman analysis shows that the reaction of ZnO
particles occurs in the form of an unreacted core surrounded by
a shell of zinc stearate. This experimental observation suggests that
a reaction front, within each particle, sets in at the beginning of the
reaction process and then moves towards the center of the particle
itself, gradually converting ZnO into ZnSt. Hence, in the course of
the reaction, the particles can be envisaged as a shrinking inner
core made of ZnO surrounded by an evolving shell of ZnSt.
A pictorial representation of the model is given in Fig. 11.
Rext(t)R(t) represents the growing thickness of the ZnSt layer,
while R(t) represents the time dependent radius of inner ZnO
core.
In formulating the model it has been assumed that the ZnO
reacting particles i) are evenly distributed throughout the rubber
domain, ii) have a spherical shape with mono-dispersed diameter,
and iii) do not interfere with each other, being considered as
independent reacting systems. Moreover, the stearic acid molecules
are taken as uniformly dispersed within the rubber matrix at the
reaction temperatures. As already discussed, the FTIR analysis

Table 1
DM as determined at the temperatures of interest for SBR/StA
Values for CD, CM and KEQ
system.
Temperature ( C)

CD [mol/l]

CM [mol/l]

DM [l/mol]
KEQ

80
100
110

0.131672
0.127078
0.123934

0.027655
0.036845
0.043132

172.1613
93.60877
66.61962

demonstrates that stearic acid is prevalently present in its dimeric


form, in equilibrium with the monomer in a much lower concentration. In the model view, both species are assumed to diffuse
towards the external surface of the reacting particle (whose initial
radius is indicated as R(t 0)), where, however, only the smaller
monomeric form is likely to dissolve, as ruled by a partition coefcient, K0 . Then, StA molecules diffuse through the ZnSt shell to
reach the surface of the unreacted ZnO core. Here the reaction
occurs, which is assumed to attain an instantaneous equilibrium,
and the product ZnSt precipitates, instantly increasing the thickness of the shell layer.
The evolution of molar concentration of monomeric StA within
Shell ,is ruled by the following mass balance:
the shell, i.e. CStA
Shell
vCStA
1
2
vt
r

Shell
v 2 Shell vCStA
r DStA
vr
vr

!
(6)

with boundary conditions:


Shell C EQ
CStA
StA

at r Rt :
at r Rext t :



Shell K 0 ,f C Rubber
CStA
StA

where

Rubber
CStA

1

q
DM C Rubber
1 8KEQ
StA
DM
4KEQ

DShell
StA is the stearic acid mutual diffusivity in the shell layer,
EQ
t stands for time and r is the radial coordinate. CStA
represents the
stearic acid molar concentration at chemical equilibrium at the
shellecore interface and determines an upper limit for the overall
conversion of stearic acid (i.e. amax < 1, see denition of amax in
Equation (14)). The boundary condition at r R(t) is dictated by the
assumed instantaneous reaction equilibrium, while the one at

692

P. Musto et al. / Polymer 54 (2013) 685e693

Fig. 9. FTIR spectra in the 1800e1480 wavenumber range of a SBR/StA/ZnO mixture


87/9/4 by wt. Spectra taken at increasing reaction times under isothermal conditions
(80  C).

r Rext(t) is dictated by partition of StA between rubber and ZnSt


shell.
Rubber , although variable with time, is considered
The value of CStA
to be uniform across the rubber matrix surrounding the particles. In
fact, in developing the model, it has been assumed that the whole
process is controlled by StA diffusion through the shell, since, the
characteristic time of diffusion of monomeric and dimeric StA
through the rubber and the characteristic time of the surface
chemical reaction are considered to be much lower than the
characteristic time of StA diffusion within the shell layer. It is worth
to note that, this approach is obviously less accurate in the initial
stages of the process, when the shell is not sufciently thick to
guarantee the assumed diffusion control. In view of these considRubber can be evaluated using the
erations, the time evolution of CStA
following mass balance on equivalent monomeric stearic acid in the
bulk of the rubber matrix:

Rubber
Shell
dCStA
vCStA
Np 4pr 2 DShell
StA
dt
vr

(7)
rRext t

Fig. 11. Schematic representation of a reacting ZnO particle (assumed to be of spherical


shape) at a generic instant of time t. The original size of the ZnO particle (R), the
current size of the ZnO particle (R(t)) and the current size of the ZnO particle and of the
surrounding ZnSt shell (Rext(t)) are evidenced.

Based on the stoichiometry of the reaction, it is possible to relate


the consumption of StA to the dimension of the internal, R(t), and
external, Rext(t), radii of the shell, as follows:




4 
Rubber;INI
Rubber
CStA
 CStA
t brZnO Np p R3p  R3 t
3




R3ext t  R3 t rZnSt R3p  R3 t rZnO

(8)

(9)

where Rp represents the initial radius of a ZnO particle, before the


reaction starts (i.e. Rp R(t 0)), b is the stoichiometric coefcient
of the reaction, rZnO is the molar density of the ZnO particle and
rZnSt is the molar density of zinc stearate. After some algebra
Equations (8) and (9) give:
1
Rt
1  K1 a3
Rp

(10)

where Np represents the number of ZnO particles per unit volume


of the system.

Rext t

Rp

R3 t
R3p

r 1r

!1
3

r 1  r1  K1 a3

(11)

where:

K1

Fig. 10. Relative conversion of stearic acid in zinc stearate in the SBR/StA/ZnO mixture
87/9/4 by wt, monitored isothermally at the temperatures indicated. Symbols refer to
the measured values; continuous lines represent the kinetic model predictions.

Rubber;INI
CStA
4
brZnO Np pR3p
3

rZnO
rZnSt2

In summary, solution of the system of Equations (6), (7), (10) and


Rubber which can be used to
(11) provides the time evolution of CStA
evaluate a, from Equation (5a), to be compared with experimentally
determined values.
However, to simplify the mathematics of the model, we further
assumed a quasi-steady state regime for the diffusion process of StA
within the particles shell. This assumption consists in imposing

P. Musto et al. / Polymer 54 (2013) 685e693


Table 2
Values of best tting parameters of the kinetic model.
Temperature ( C)

amax

k (s1)

80.000
100.000
110.000

0.51
0.74
0.75

0.129
0.138
0.525

that the concentration prole inside the shell instantaneously


accommodates to the change of boundary conditions determined
by the evolution of the system. In other words, the system of
Equations (6) and (7), is replaced by:

h

i

Rubber  C EQ
Rubber
K 0 ,f CStA
dCStA
StA


Np DShell
StA 4p
1
1
dt

Rext t Rt

(12)

By substituting the relationships for R(t) and Rext(t) into Equation (12) and expressing a according to Equation (5a), we obtain the
following differential equation:

da

dt

interpreted theoretically by means of a kinetic model developed on


the basis of a simple heterogeneous reaction scheme.
Confocal Raman microspectroscopy revealed that, at ambient
temperature, both components are phase-separated in the form of
microcrystals. When the reaction temperature is reached (80  C
and above) only zinc oxide is present in the form of particles while
stearic acid melts and gets molecularly dispersed within the rubber
matrix. The analysis also points to a coreeshell structure of the
reacting system: stearic acid diffuses to the surface of the ZnO
domains causing the shrinkage of the ZnO core and the formation
of a shell of increasing thickness made of zinc stearate.
FTIR spectroscopy conrmed that, when molecularly dispersed
in the rubber matrix, StA is predominantly self-associated in
a dimeric form. Time-resolved measurements also yielded quantitative information on the kinetic behaviour of the reacting system.
In the light of the Raman imaging analysis a simple, yet physically sound, model was developed to interpret the FTIR spectroscopy results. The model is based on the assumption that the
controlling step of the reaction kinetics is the diffusion of the

0n 


o
4pNp DShell
Rubber;INI
Rubber;INI
StA Rp K
f CStA
1  amax  f CStA
1  a
Rubber;INI
CStA

1  1  rK1 a3 1  K1 a3


1

with
Rubber;INI
CStA
 f 1

amax h

EQ
CStA
K0

Rubber;INI
CStA

(14)

693

(13)

stearic acid within the zinc stearate shell surrounding the reacting
particles. The model is able to catch the main experimental features
and displays a reasonable tting capability of the conversion vs
time data at three different temperatures (80, 100 and 110  C).

References
The constant K1 corresponds to the molar defect of stearic acid
with respect to the stoichiometric ratio between stearic acid and
zinc oxide and, in the case at hand, takes a value of 0.32. The
Rubber;INI
nominal initial concentration of stearic acid, CStA
, is
0.310 mol/l. Moreover, based on the molar densities of zinc oxide
and zinc stearate, a value of 40 for 1  r was estimated. Inserting
these values into Equation (13) one obtains:

da
kff 0:311  amax   f 0:311  ag

1
1
dt
1 12:8a3 1  0:32a3

(15)

Rubber;INI
where k stands for 4pNp DShell
. Equation (15), in its
StA Rp K=CStA
numerically integrated form, has been used to t the experimental
data available from the FTIR analysis, with, amaxand k as adjustable
parameters. The initial condition for a has been taken as the
experimental value measured spectroscopically at t 0 (respectively equal to 0.15, 0.28 and 0.44 at 80, 100 and 110  C).
The results of the model tting are compared with the experimental FTIR data in Fig. 10, while the values determined for best
tting parameters, amaxand k, are reported in Table 2. The tting
quality is satisfactory, despite the adopted simplifying assumptions.
The dependence of the tting parameters on temperature is
reasonable in view of the expected increase of DShell
StA with T.

4. Conclusions
The interaction between zinc oxide and stearic acid in a medium
suitable to simulate a vulcanizing system has been investigated
experimentally using vibrational spectroscopy techniques and

[1] Ikeda Y, Higashitani, Hijikata K, Kokubo Y, Morita Y, Shibayama M, et al.


Macromolecules 2009;42:2741e8.
[2] Susamma AP, Mini VTE, Kuriakos AP. J Appl Polym Sci 2001;79:1e8.
[3] Moore CG, Saville B, Watson AA. Rubber Chem Technol 1961;34:795.
[4] Kresja MR, Koenig JL. The nature of sulfur vulcanization. In: Elastomer technology handbook. New Jersey: CRC Press; 1993. p. 475.
[5] Coran AY. Vulcanization in science and technology of rubber. In: Eirich ER,
editor. San Diego: Academic Press; 1978. p. 291.
[6] Domka L, Krysztafkiewicz A. Int Pol Sci Technol 1980;7. T/18.
[7] Sullivan AB, Hann CJ, Kuhls GH. Rubber division meeting. Toronto, Canada:
American Chemical Society; 1991.
[8] (a) Adams HE, Johnson BL. Ind Eng Chem 1953;45:1539;
(b) Duchacek V, Kuta A, Pribyl P. J Appl Polym Sci 1993;34:743e6;
(c) Bravar M, Jelencic J, Dabetic M. Kautschuk Gummi Kunststoffe 1986;39(4):
314e8;
(d) Coran AY. Rubber Chem Technol 1964;37(3):679e88;
(e) Xie HQ, Ao ZP, Guo JS. J Macromol Sci Phys 1995;B34(3):249;
(f) Campbell RH, Wise RW. Rubber Chem Technol 1964;37(3):650e67;
(g) Zaborski M, Slusarski, Donnet JB, Papirer E. Kautschuk Gummi Kunststoffe
1994;47(10):730e8;
(h) Powell CE, Preparation of rubber compositions with organoclays, US Patent
Appl. 2005, n. 2005/0090584 A1.
[9] Baldwin FP. Rubber Chem Technol 1972;45:1348.
[10] Saville B, Watson AA. Rubber Chem Technol 1967;40:100.
[11] Bridgestone TCE, Private communication.
[12] Kruger FWH, McGill WJ. J Appl Polym Sci 1991;42:2651.
[13] Damen TC, Porto SPS, Tell B. Phys Rev 1966;142:570e4.
[14] Ishioka T, Shibata Y, Takahashi M, Kanesaka I. Spectrochim Acta Part A 1998;
54:1811e8.
[15] Kobayashi M, Tadokoro H, Porter RS. J Chem Phys 1980;73:3635.
[16] Chalmers JM, Grifths PR, editors. Handbook of vibrational spectroscopy, vol.
1. Chichester, U.K.: Wiley; 2002. p. 921e32.
[17] Pimentel GC, McClellan L. The hydrogen bond. San Francisco, CA: Freeman and
Co.; 1960.
[18] Chalmers JM, Grifths PR, editors. Handbook of vibrational spectroscopy, vol.
3. Chichester, U.K.: Wiley; 2002. p. 1919e34.
[19] Lee JY, Painter PC, Coleman MM. Macromolecules 1988;21:346e54.

Das könnte Ihnen auch gefallen