Você está na página 1de 8

A SAXS and Swelling Study of Cured Natural Rubber/ StyreneButadiene Rubber Blends

W. SALGUEIRO,1 A. SOMOZA,1 A. J. MARZOCCA,2 I. TORRIANI,3 M. A. MANSILLA2


1

IFIMAT- UNCentro and Comision de Investigaciones Cientcas de la Provincia de Buenos Aires, Pinto 399, B7000GHG Tandil, Argentina LPMPyMC, Departamento de Fsica, Facultad de Ciencias Exactas y Naturales, Universidad Nacional de Buenos Aires, Ciudad Universitaria, Pabellon I, C1428EGA Buenos Aires, Argentina Instituto de Fsica Gleb Wataghin, Universidade Estadual de Campinas, CP 6165, 13084-971 Campinas, Sao Paulo, Brazil

Received 3 June 2009; revised 15 July 2009; accepted 7 August 2009 DOI: 10.1002/polb.21828 Published online in Wiley InterScience (www.interscience.wiley.com).

ABSTRACT: A small-angle X-ray scattering (SAXS) and swelling study of natural rubber and styrenebutadiene rubber blends (NR/SBR) is presented. To this aim, specimens of NR and SBR and blends with 75/25, 50/50, and 25/75 NR/SBR ratios (in phr) were prepared at a cure temperature of 433 K and the optimum cure time (t100). This time was obtained from rheometer torque curves. The system of cure used in the samples was sulfur/n-t-butyl-2-benzothiazole sulfenamide. From swelling tests of the cured samples, information about the molecular weight of the network chain between chemical crosslinks was obtained. For all cured compounds, in the Lorentz plots built from SAXS scattering curves, a maximum of the scattering vector q around 0.14 A1 was observed. However, the q position shows a linear-like shift toward lower values when the SBR content in the SBR/NR blend increases. In pure NR or SBR the q values show a different tendency. The results obtained are discussed in terms of the existence of different levels of vulcanization for each single phase forming the blend and the existence C of a third level of vulcanization located in the interfacial NR/SBR layer. V 2009 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys 47: 23202327, 2009

Keywords: blends; NR; rubber; rubber blends; SAXS; SBR; vulcanization

INTRODUCTION
Blends of elastomers are used in rubber products to improve physical properties, increased service life, easier processing, and reduced product cost. It must be considered that the complete miscibility of polymers requires that the free energy of the mixing be negative, which implies exothermic mixing or large entropy of mixing.1 In fact, most
Correspondence to: A. Somoza (E-mail: asomoza@exa. unicen.edu.ar)
Journal of Polymer Science: Part B: Polymer Physics, Vol. 47, 23202327 (2009)
C V 2009 Wiley Periodicals, Inc.

blends of elastomers are immiscible because mixing is endothermic and the entropic contribution is small because of the high molecular weights. Fortunately, miscibility is not a requirement for most rubber applications. However, adhesion between the polymer phases is necessary. In the cured state of completely miscible elastomers, the glass transition process is characterized by a single narrow peak. On the other hand, for immiscible or partially compatible elastomers, two (or only one broad) glass transition distributions represent the cured state.27 In this scenario, covulcanization of elastomers containing different phases and their respective

2320

A SAXS AND SWELLING STUDY OF NR/SBR BLENDS

2321

Table 1. Compound Formulations (in Parts per Hundred Rubber) Materials SBR1502 NR (SMR20) Stearic acid ZnO Antioxidant Sulfur TBBS A 0 100 2 5 1.2 2.3 0.7 B 25 75 2 5 1.2 2.3 0.7 C 50 50 2 5 1.2 2.3 0.7 D 75 25 2 5 1.2 2.3 0.7 E 100 0 2 5 1.2 2.3 0.7

imental technique, swelling tests were used to determine the crosslink density of the cured compounds. The results obtained are discussed in terms of the existence of different levels of vulcanization for each single phase forming the blend. Special attention was given to the study of the layer between NR and SBR developed during the vulcanization process of the blends. Specically, we have idealized this layer as a zone, where the NR has higher crosslink density than the SBR phase due to the migration of the curatives from the SBR phase, which is normally overcured.

interfaces are an important subject of study. As mentioned in the literature, blends possess an interfacial layer of rubber with a state of cure different from that corresponding to the individual phases.6 When dealing with practical applications and also from the theoretical point of view, the interfaces studies become an important goal of investigation. Thus, the NR/SBR system is an example of not completely miscible elastomers and its use in automobile industry is well known.1,2 Most of the authors of this work have been involved in the study of vulcanization processes in natural rubber (NR) and styrenebutadiene rubber (SBR) by means of different experimental techniques.8,9 In a recent article, we used smallangle X-ray scattering (SAXS) to analyze the structural changes in NR vulcanized samples as a function of the cure temperature.10 On the other hand, the same technique, and its variants, has successfully been used to investigate elastomers (see, e.g., refs. 3, 4). According to the results reported therein, a correlation could be established between the crosslinking developed during vulcanization and changes in the broad intensity peaks of the X-ray scattering patterns. In this work, SAXS was used once more as one of the main experimental techniques to study structural details in vulcanized blends prepared with a NR matrix by addition of SBR using n-tbutyl-2-benzothiazole sulfenamide (TBBS) as an accelerator and sulfur. As a complementary exper-

EXPERIMENTAL
Preparation and Characterization of the Compounds Natural rubber (NR-SMR20) and styrenebutadiene rubber (SBR1502) were used to prepare the blends studied in this research. The molecular weights of the polymers were measured by GPC using a Shimadtzu L-6A liquid chromatograph system with THF as elute. Thus, the values of Mn 91,350 g/mol and Mn 149,910 g/mol were obtained for SBR and NR, respectively. The densities of the polymers used were q (NR) 0.917 g/cm3 and q (SBR) 0.935 g/cm3. Five compounds were prepared using the formulations given in Table 1. For more details of the compounds preparation, see ref. 7. The mixes were characterized at 433 K by means of torque curves measured with an Alpha MDR2000 rheometer.7 The time to achieve the maximum torque, t100, was calculated for each sample and is given in Table 2. Samples sheets of 150 150 2 mm3 were cured in a mold at 433 K in a press up to time t100 to make sure that the vulcanization reaction was completed. At the end of the curing cycle, the samples were nally cooled rapidly in an icewater mixture. The density q of each cured compound measured by Archimedes procedure is also given in

Table 2. Optimum Time Cure t100 (MDR2000, 433 K), Density q, and Tg of the Studied Samples A t100 (min) q (g/cm3) Tg (K) 10.7 0.9536 213.1 B 17.0 0.9566 214.3/226.7 C 24.7 0.9648 214.5/226.4 D 34.4 0.9722 228.2 E 42.3 0.9786 229.6

In the case of the blend (samples B and C), two values of Tg were detected.7

Journal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

2322

SALGUEIRO ET AL.

Table 3. Equilibrium Volume Fraction in the Swollen State (v2m) and Molecular Weight Between Crosslinks Mcs of the Studied Samples A v2m Mcs (g/mol) 0.198 5,451 B 0.178 7562 C 0.180 8,080 D 0.173 9,653 E 0.166 11,958

Table 2. As can be seen, q increases when the SBR content in the formulation increases. On the other hand, the glass transition (Tg) of the cured samples shown in Table 2 are those reported in a previous work.7 As can be seen in the table, for the samples prepared with only SBR or NR, one Tg was obtained. In the blends B and C, two glass transition temperatures were found. Swelling Tests The molecular weight of the network chain between chemical crosslinks, Mcs, was determined from swelling tests. To this aim, the method described by Cunneen and Russell was applied11 and the FloryRehner relationship12,13 Mcs
1=3 q1 2=/V1 v2m ln 1 v2m vv2 v2m 2m

precisely for this work this functionality, / 4, was used in eq 1. To obtain the experimental Mcs, 17-mm diameter discs of the cured sheets were cut with a die. These discs remained in pyridine for 16 h at room temperature. Then, they were continuously extracted in acetone for 24 h (ASTM D 297-93 (2006)) and dried. One probe of each sample was completely immersed in pure toluene, in a sealed glass bottle, at room temperature until equilibrium swelling occurred. Normally, this process takes more than 48 h. When this step was completed, the samples were removed from the bottles, the excess toluene from the surface of the samples was wiped off and the swollen weight immediately measured using an electronic balance with an accuracy of 104 g. The samples were dried at 333 K and weighted again when all the solvent was evaporated. The volume fraction v2m was calculated based on the following equation v2m bWd Wf =qc ; Wd Wf =q Ws Wd =qs (2)

(1)

was used, where q is the density of the rubber network, / the functionality of the crosslinks, v2m the polymer volume fraction at equilibrium (maximum) degree of swelling, and V1 the molar volume of solvent. v is an interaction parameter between the polymer and the swelling agent. The interaction parameter polymersolvent v was evaluated using a mixture law, starting from the v for the systems NR/toluene and SBR/toluene. These values were v(NR) 0.43 0.05v2m,14 v(SBR) 0.524 0.285v2m,15 and V1 (toluene) 106.29 mL/mol.16 Finally, the values of Mcs obtained for each vulcanized blend are given in Table 3. It is important to point out that a crosslink is considered as a small region in a macromolecule from which at least four chains emanate. Specically, a crosslink behaves as a central core in which macromolecules are attached and their number denes the crosslink functionality /. From the structural point of view, there is a higher electron density at the site of the crosslinks. Besides, sulfur cured elastomers are usually considered a four-functional network,17 and

where Wd is the weight of the sample after swelling and drying, Ws the weight of the swollen sample, and qs the density of the solvent (0.8669 g/cm3 for toluene16). Wf is the weight of the nonextractable ller in the sample. The compounds used in this research do not contain ller as carbon black or silica, but as it can be observed in Table 1, there is ZnO in the composition and we have used the ASTM D 297-93 (2006) method to evaluate Wf. The volume fractions v2m for each blend composition are given in Table 3 and Figure 1. The values of v2m for samples 90NR/10SBR and 10NR/90SBR were also measured and included in Figure 1. These cured samples were prepared for a previous research work to study the thermal diffusivity on these blends.7 SAXS Measurements From 2-mm thick sheets, 8-mm diameter samples were cut with a die. The samples were set up in the sample holder device of the D11A-SAXS1 beamline of the Brazilian National Synchrotron Laboratory (LNLS, Campinas, Brazil) at room temperature. A wavelength k 1.608 A was selected for the monochromatic beam used in the experiments. The scattering intensity distributions I(q) as a function of the scattering vector q (q (4p/k) sin h; 2h being the scattering angle) were obtained in the q-range between $0.02 and 0.40
Journal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

A SAXS AND SWELLING STUDY OF NR/SBR BLENDS

2323

Figure 1. Equilibrium volume fraction in the swollen state (v2m) as function of the SBR content in the cured samples. Dash line represents the mixture law obtained from eq 5. Dash dot line is only an eye guide.

A1. A typical spectrum was obtained after measuring for 15 min. This time was enough to get reliable results for this type of measurements. A onedimensional position sensitive detector was used with a 729.8-mm sample-detector distance. To reduce parasitic scattering, the samples were kept under vacuum during the exposures. Data obtained were normalized by the integrated intensity and detector response curve. Background subtraction was subsequently performed for every sample measurement. Because of the strong scattering signal, the data were acquired using a slit in front of the detector window for the full q-range, and in a second step, blocking the low q region with a shield to allow a longer time exposure and improve the statistics in the region around q 0.14 A1. The two sets of data were later spliced together (as shown in Fig. 2).

mum is directly proportional to the SBR content in the blend. Besides, the difference in the I(q) versus q curves corresponding to 100 phr NR and 100 phr SBR indicates that the scattering properties of the two main components in the blend have some similarities, but present differences in the intensity and maximum position. As usual, to get a more precise localization of the angular position of the intensity maximum, Lorentz plots (I(q)q2 vs. q)18 for the different compounds are presented in Figure 3. In agreement with the aim of this work, we have focused our attention in the region in which the different maxima are present, that is, a region for the scattering vector q between 0.10 and 0.18 A1. It is well accepted that for dense (bulk) polymer networks, like the compounds studied in this work, that in the experimental SAXS curves the angular position of the intensity maximum (qmax) in the Lorentz plots is frequently used to better dene a structural characteristic correlation distance between the center of the scatterers, n, by means of   2p (3) n qmax As can be seen in Figure 3, only the qmax values corresponding to the blends B, C, and D present well-dened systematic changes. The experimental data corresponding to each compound were well tted by using a single Gaussian distribution (see solid lines). From the rst-order statistical momentum of each Gaussian function, the qmax was obtained for all curves.

RESULTS AND DISCUSSION


In Figure 2 the SAXS curves I(q) as a function of the scattering vector q for the different compounds studied are shown. The curve obtained for the sample named A (100 phr NR) is in very good agreement with that previously reported by the authors in ref. 10 for the same elastomer. As can be seen, the curves presented in the gure show a wide band for scattering vector values between $0.12 and $0.18 A1. When SBR is added to the NR matrix, increasing SBR contents (samples B to D) correspond to a higher I(q) value reaching a maximum for 100% SBR (sample E). Summarizing, the systematic increase of the intensity maxiJournal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

Figure 2. Scattering curve I(q) as a function of the scattering vector q for all compounds studied. Sample A: 0SBR/100NR; B: 25SBR/75NR; C: 50SBR/50NR; D: 75SBR/25NR; and E: 100SBR/0NR.

2324

SALGUEIRO ET AL.

Figure 3. A selected region from the Lorentz representation for scattering vector between 0.10 and 0.185 A1 is presented to evidence the curves tted (solid lines). The results corresponding to different compounds are labeled (Sample A: 0SBR/100NR; B: 25SBR/75NR; C: 50SBR/50NR; D: 75SBR/25NR; and E: 100SBR/0NR). The arrow evidences the shift of the maxima qmax obtained from the tted curves for each blend.

tribute to the SAXS patterns in the q-range analyzed in this experiments. In fact, as the sulfur groups localize heavier elements in the rubber matrix, they produce a local increase in the electron density, creating specic inhomogeneities for NR and SBR, as discussed further later. In the compounds studied in this work, the crosslinks are formed with sulfur as a crosslink agent; therefore, the local density uctuations at the crosslink sites should be different from the surrounding. Such differences would produce differences in the size of the sulfur clusters, which are detected in our SAXS experiments. In fact, as it can be seen in Table 3, Mcs measured in all the compounds analyzed by swelling test are different. In Figure 5, the correlation distance n obtained using eq 3 is presented as a function of the SBR content in the elastomers. A preliminary analysis of the data presented in this gure can be divided into two groups:

In Figure 4 the angular positions of the intensity maxima as a function of the SBR content in the compounds are presented. As can be seen, the analysis of the data can be divided in two parts: one corresponds to the elastomers 100 phr NR and 100 phr SBR, in which a increase difference in qmax from 0.1415 A1 for NR to 0.1435 A1 for SBR is observed. The other one corresponds to the blend compounds. In this case, the qmax values sys tematically decrease from 0.1425 A1 for a SBR content of 25 phr to 0.140 A1 when the SBR content is 75 phr. This three angular positions corresponding to the intensity maxima for the blends B, C, and D could be satisfactorily tted by using a linear function (with a correlation coefcient of 0.995), which is represented by a solid line in the gure. We want to point out that the assumption regarding the use of a linear tting on only three data points must be considered as a rst approximation to the analysis of the experimental data. However, the most important result obtained analyzing this gure is that qmax decreases when the SBR content in the blends increases. For a general discussion of the experimental results reported earlier, it must be taken into account that all the compounds studied in this work contain zinc oxide and sulfur in their formulation (see Table 1). As was previously discussed,10 the ZnO and carbon particles do not con-

Pure elastomers, 100 phr NR and 100 phr SBR (samples A and E) In this case, it can be observed that n is lower (43.8 A) in the cured 100 phr SBR sample than in the 100 phr NR one (44.4 A). This last value of n is different than that estimated from Figure 5 of ref. 10 for the same vulcanization temperature T 433 K (n % 43.9 A). But, this disagreement can be easily explainable on the basis that in both cases the sulfur and accelerant contents of the 100 phr NR samples were different (S

Figure 4. Position of the maximum qmax (A1) as a function of the SBR content in the compounds. A linear t to experimental points obtained in blends is depicted in solid line. Journal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

A SAXS AND SWELLING STUDY OF NR/SBR BLENDS

2325

Figure 5. Correlation distance n (A) as a function of SBR content in the compounds. Dash dot line is only an eye guide. Mcs (in g/mol) values are given in brackets beside each point (see text).

1.8 phr and TBBS 1.2 phr for the samples studied in ref. 10; and S 2.3 phr and TBBS 0.7 phr for the samples prepared for this work). The difference in sulfur/accelerator ratio produces different network structure in the cured NR.19,20 For four-functional networks, the crosslink density is dened by (see ref. 10 and references therein): ! q 1 1 ; (4) lc 2 Mcs Mn where q is density of the polymer. From eq 4, it is clear that the crosslink density is related to the inverse of the molecular weight between crosslinks Mcs; therefore, from Table 3 results that the cured SBR sample has a lower crosslink density than that of the NR elastomer. Both NR as well as SBR have many allylic hydrogens in its structure and, during the vulcanization process, the free radicals can react with these hydrogens much easier than with other hydrogens present in the molecules. It is usual to expect a higher reactivity in NR than in SBR, because statistically the rst one has more allylic hydrogens in its structure.21 Then, it would be expected a higher crosslink density in NR than in SBR when the same formulation of curatives is used in both elastomers. The relationship between network structure and vulcanization chemistry is very complex for elastomers cured with a sulfur/acJournal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

celerator system. The amount of monosulde, disuldes, and polysuldes crosslinks and cyclic structures formed during vulcanization at a given temperature are not necessary the same in cured NR and cured SBR.22,23 Qualitatively, it could be thought that the values of n are directly related to the crosslink density. However, in a recent SAXS study reported by the authors10 on cured NR at different temperatures, a reduction in the correlation distance when the crosslink density decreases was found. This behavior was attributed to a different ratio between the number of polysuldic and disuldic crosslinks as well as the number of monosuldic crosslinks, which produce a modication in the correlation distance. Following the ideas described earlier, it must be considered that the slight difference in n between the cured samples of 100 phr NR and 100 phr SBR seems to be specically related to the crosslinks type present in the vulcanized network created during the vulcanization process. Blends (samples B, C, and D) As can be seen in Figure 5, the correlation distance systematically increases when the SBR content increases. Taking into account the Mcs values reported in Table 3, the total crosslink density decreases for higher SBR content in the blend. Then, there exists an inverse correlation between n and the total crosslink density for all blends studied. Based on the elements discussed in the previous item, this behavior should be considered as consistent. In addition to the general discussions given earlier, we consider that it is necessary to go deeper into the interpretation of the data presented in Figure 5. To this aim, aspects linked to the role of the interphase stability in the blends and the level of vulcanization of each elastomer will be discussed later. When dealing with blends, as it can be seen in Table 2, we note that the t100 value for the sample with 100 phr NR is about four times lower than that measured for the sample with 100 phr SBR. On the other hand, a systematic increase is observed for t100 when the SBR content increases in the different compounds. In other words, we know that t100 (NR) \ t100 (NR/ SBR) \ t100 (SBR). This fact implies that the NR phase is always overcured in the blends, indicating the existence of two vulcanization levels,

2326

SALGUEIRO ET AL.

one in each phase of the different blends. Such a behavior is clearly deduced from Figure 1 of ref. 7: for all the times t100 obtained measuring each blend, the rheometer curve corresponding to the cured 100 phr NR already shows reversion. This assertion could be reinforced considering the change of the values of both glass transition temperatures detected in the blend compounds (see Table 2). Specically, the Tg associated with each phase of the blends depends on the blend composition and the curing level of each phase into the blend. Furthermore, it is well accepted that the NR phase is less stable than the SBR phase19 and mainly polysuldic linkages change to disuldic and monosuldic linkages in overcured NR.2427 Assuming that when preparing the compounds the distribution of chemicals is the same in both NR and SBR phases, and considering the different reactivities of NR and SBR in the vulcanization process the solubility of sulfur and TBBS is different in both elastomeric phases. Curative diffusion between the phases of an elastomer blend takes place during the vulcanization process but not during mixing.28 It is known that NR crosslinks more rapidly than SBR, and a rapid depletion of the curatives in the NR phase of the blend and a replenishment of the curatives due to the diffusion from the SBR phase were reported.6 The consequence of these mechanisms is the development of a zone of high crosslink density (i.e., layer) in the NR phase close to the interphase, higher than in the case of the NR sample and with different state of cure than the bulk.1,6 The increase in the Tg values associated with the NR phase in the NR-rich blends would conrm this hypothesis. This behavior is observed in Table 2. At the same time, the migration of curatives from SBR to NR produces a diminution in the crosslink density into the SBR phase when compared with that corresponding to the 100 phr SBR sample. Therefore, a lower Tg value associated with the SBR phase would be expected. In fact, this behavior is precisely reported in Table 2. Under the aforementioned assumptions, it can be concluded that there exists an interphase layer between the two elastomeric phases. This interphase should have a different vulcanization level in comparison with those corresponding to the samples of pure NR or pure SBR. The same behavior was reported by Mallon and McGill6 from the study of polyisoprene/SBR blends. On the other hand, it must be considered that a blend of two elastomers 1 and 2 can theoretically

produce different types of vulcanizates such as: Type I, a dispersion of a cured 1-phase in an uncured 2-matrix; Type II, a dispersion of an uncured 1-phase in a cured 2-matrix; Type III, with both 1 and 2 phases cured.29 In the last case, many different situations can be established depending on the crosslink level of each phase. Considering the samples B, C, and D of this study, we could assume that we are dealing with blends of Type III. Under conditions of free swelling, the equilibrium swelling behavior of the blend should be expressed by v2m f1 v1 f2 v2 2m 2m 2 ; 1 v2m f2 v2m v1 2m (5)

v2m

where v1 and v2 are the corresponding volume 2m 2m fractions of two components phases equilibrium (maximum) degree of swelling. f1 and f2 are the volume fractions of the two elastomers present in the blend29 and f1 f2 1. The comparison of the experimental data of v2m with eq 5 for the blend compositions analyzed is given in Figure 1. It can be observed that eq 5 does not t well the experimental data, especially for those blends in which NR is the main phase. This behavior points out that the existence of an interphase that behaves as a third phase should be taken into account in the analysis given in this work. Unfortunately, SAXS results give only an average information regarding the bulk structure of the elastomers studied. The nonmonotonic behavior of the correlation distance when the SBR content increases (see Fig. 5) is hard to explain in terms of the presence of two or more different phases in the blend compounds. To analyze the scattering curves, the contribution of each phase should be identied. In our experiments, the correlation distance n increases at higher SBR content in the blend, and as can be seen in Figure 3, there is a noticeable change in the peak broadening, which can be attributed to additional electron density contrast between the phases in the blend. According to a previous work,10 n changes with the type of crosslinks generated in the cured samples. In fact, this value would decrease when a reduction of polysulde linkages followed by a moderate increase of the monosulde crosslinks is produced. It is interesting to notice that the region of SBR near to the interface has also an increase in the crosslink level compared with the bulk of the SBR phase. In this layer, because of the migration of curatives, it is expected a higher concentration of
Journal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

A SAXS AND SWELLING STUDY OF NR/SBR BLENDS

2327

polysulde linkages at higher concentration of SBR in the blend. Recent studies on vulcanized SBR, with the same system of cure as the one used in this research, conrm this assertion.15 Thus, the increased presence of polysulde linkages and a less diffuse interphase boundary may result in a narrower distribution of the interdomain correlation distance in the blend.

CONCLUSIONS
A SAXS and swelling study of vulcanized NR and SBR blends (NR/SBR) with 75/25, 50/50, and 25/ 75 phr ratios and 100 phr NR and 100 phr SBR was carried out. Samples were cured at 433 K at the optimum time. Swelling of cured samples shows that the crosslink density increases when the SBR phase increases in the blend. However, the volume fraction of the blend at equilibrium (maximum) degree of swelling does not follow a mixture law considering only the two elastomeric phases, NR and SBR, in the blend. Then, it is clear that the interphase between the two elastomeric domains has to be considered as another phase into the analysis. Evidences reported in the literature have conrmed that the diffusion of curatives from SBR to NR takes place during vulcanization. In this work, we have associated the shift of the scattering vector qmax in the SAXS experiments toward lower values when the SBR content increases in the formulation of the blend with the curatives diffusion. Furthermore, this change would be associated with changes in the type of crosslinks in the interphase layer due to the different level of cure of the interphase, depending on the cure time of the samples.
The authors acknowledge the support of the LNLS, Brazil (Project: D11A-SAXS1 # 4181/05). Agencia Nacional de Promocion Cientca y Tecnologica (PID 2003-0435 and PICT 2006-1650), Comision de Investigaciones Cientcas de la Provincia de Buenos Aires and SECAT (UNCentro), Argentina. The authors are very grateful to F Queiruga Rey and Tomas Plivelic for the technical support during measurements at LNLS. This work was partially supported by the Universidad de Buenos Aires, Argentina (Research Project 2006-2009 X808).

REFERENCES AND NOTES


1. Hess, W. M.; Herd, C. R.; Vegvari, P. C. Rubber Chem Technol 1993, 66, 329375. Journal of Polymer Science: Part B: Polymer Physics DOI 10.1002/polb

2. Sircar, A. K. In Thermal Characterization of Polymeric Materials; Turi, E. A., Ed.; Academic Press: San Diego, 1981; pp 8871378. 3. Morn, I.; Ehrburger-Dolle, F.; Grillo, I.; Livet, F.; Bley, F. J. Synchrotron Rad 2006, 13, 445452. 4. Ehrburger-Dolle, F.; Bley, F.; Geissler, E.; Livet, F.; Morn, I.; Rochas, C. Macromol Symp 2003, 200, 157168. 5. Horston, D. J.; Song, M. J Appl Polym Sci 2000, 76, 17911798. 6. Mallon, P. E.; McGill, W. J. J Appl Polym Sci 1999, 74, 12641270. 7. Goyanes, S.; Lopez, C. C.; Rubiolo, G. H.; Quaso, F.; Marzocca, A. J. Eur Polym J 2008, 44, 15251534. 8. Marzocca, A. J.; Cerveny, S.; Salgueiro, W.; Somoza, A.; Gonzalez, L. Phys Rev E 2002, 65, 021801-1021801-5. 9. Salgueiro, W.; Marzocca, A. J.; Somoza, A.; Consolati, G.; Cerveny, S.; Quasso, F.; Goyanes, S. Polymer 2004, 45, 60376044. 10. Salgueiro, W.; Somoza, A.; Torriani, I. L.; Marzocca, A. J. J Polym Sci Part B: Polym Phys 2007, 45, 29662971. 11. Cunneen, J. I.; Russell, R. M. Rubber Chem Technol 1970, 43, 12151224. 12. Flory, P. J.; Rehner, J. J Chem Phys 1943, 11, 512520. 13. Flory, P. J.; Rehner, J. J Chem Phys 1943, 11, 521526. 14. Kraus, K. Rubber Chem Technol 1957, 30, 928951. 15. Marzocca, A. J. Eur Polym J 2007, 43, 26822689. 16. CRC Handbook of Chemistry and Physics; Lide, D. R., Ed.; CRC Press: New York, 1997; pp 355. 17. Queslel, J. P.; Mark, J. E. J Chem Phys 1985, 82, 34493452. 18. Balta-Calleja, F. J.; Vonk, C. G. Polymer Science Library 8; Elsevier: Amsterdam, 1989. 19. Akiba, M.; Hashim, A. S. Prog Polym Sci 1997, 22, 475521. 20. Cerveny, S.; Marzocca, A. J. J Appl Polym Sci 1999, 74, 27472755. 21. Chough, S. H.; Chang, D. H. J Appl Polym Sci 1996, 61, 449454. 22. Klei, B.; Koenig, J. L. Rubber Chem Technol 1997, 70, 231242. 23. Pelliccioli, L.; Mowdood, S. K.; Negroni, F.; Parker, D. D.; Koenig, J. L. Rubber Chem Technol 2002, 75, 112. 24. Loo, C. T. Polymer 1974, 15, 357365. 25. Mukhopadhyay, R.; De, S. K. Rubber Chem Technol 1978, 51, 704716. 26. Mori, M.; Koening, J. L. Rubber Chem Technol 1995, 68, 551562. 27. Marzocca, A. J.; Steren, C.; Raimondo, R.; Cerveny, S. Polym Int 2004, 53, 646655. 28. Gardiner, J. D. Rubber Chem Technol 1968, 41, 13121328. 29. Rehner, J.; Wei, P. E. Rubber Chem Technol 1969, 42, 985999.

Você também pode gostar