You are on page 1of 14
FILLER-ELASTOMER INTERACTIONS. PART IV. THE EFFECT OF THE SURFACE ENERGIES OF FILLERS ON ELASTOMER REINFORCEMENT* SizGFRIED WOLFF AND MENG-J1a0 WANG Deaussa AG, INORGANIC CiEMICAL PRODUCTS DIVISION, APPLED RESEARCH AND TECHNICAL SERVICE FOR FILLER AND Ruste CHEMICALS, KALSCHEURENER SreaGE 11, W-5030 HORTH, FEDERAL REPUBLIC OF GERMANY ABSTRACT Carbon black N210 and a precipitated silica, which have comparable surface area and structure, were selected as model filers to study the effect of filler surface energies on rubber reinforcement. In comparison with carbon black, the surface energies of silica are characterized by a lower dispersive component, 7, and higher specific component, ‘yf. Iewas found that the high 9” of silica leads to strong interaguregate interaction, resulting in higher viscosity of the compounds, higher ay, and higher moduli of the vuleanizates at small strain. The higher 7 of carbon black, in contrast, causes strong filler-polymer interaction, which is reflected in a higher bound rubber content of the compounds. and higher moduli of the vulcanizates at high elongation INTRODUCTION The previous papers of this study on filler~polymer interaction’ reported on surface energy measurements carried out on a wide range of fillers by means of inverse gas chro- matography. It was found that both the dispersive and the specific (polar) components of surface energy (yi and 1%”) vary considerably for the different fillers, depending on their microstructure and chemical constitution. It would now be highly interesting to link filler surface energy to in-rubber properties. This study therefore attempts 1. Investigate the effects of surface energy, both of the dispersive and the specific polar component, on rubber properties, 2. Promote the general understanding of the influence of filler-filler and filler-polymer interaction on rubber reinforcement. To achieve this objective, two model fillers, a carbon black, N110, and a precipitated silica, P1, were used. Both fillers have comparable surface areas and structures but differ largely in their surface characteristics (Table I). The carbon black is characterized by a high dispersive component of surface energy, yi, and high adsorption energies, AG’, for hydro- carbons. On the other hand, the specific component of surface energy, estimated by the specific (polar) component of adsorption energy per unit surface of filler with a polar ad- sorbent, /*", seems to be greater for the silica. The high value of J*” observed for the adsorption of nitrile on the silica involves not only strong dipole-dipole interaction but also hydrogen bonding. Generally speaking, whereas fillers have polar surfaces, hydrocarbon rubbers are nonpolar (or low-polar) materials. The higher adsorption energy of the nonpolar adsorbent, ¢.g., hep- tane, which is related to the high 72 would be indicative of stronger interaction between hydrocarbon rubber and filler. On the other hand, the higher adsorption energy of the highly polar probe, e.g., acetonitrile, which is associated with a high 7;”, reflects interaction between the filler aggregates. Based on this consideration, a specific interaction factor S; is defined as the adsorption energy of a given probe, 4G’, divided by that of an alkane (real or hypo- thetical) whose surface area is identical with that of the given adsorbent, AG2x: AG? AGax a) * Presented at a mecting of the Rubber Division, American Chemical Society, Toronto, Ontario, Canada, May 21- 24, 1991. G. 329 330 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 Taste 1 ANALYTICAL PROPERTIES OF CARBON BLACK AND SILICA Carbon black N10 Silica PL Surface area: CTAB, m?/g 127.0 140.0 BET(N2), m'/g 134 DBPA (2414), em?/100 g 92 yi at 150°C, md/m? 22.9 AG* of n-heptane at 150°C, kJ/mol 92 AG of heptene® at 150°C, kJ/mol 116 1” of acetonitrile at 150°C, mJ/m* 212.6 252.0 S,of acetonitrile at 160°C 1.64 3.58 * The heptene is trans-3-heptene In this definition, the reference for the calculation of AG” is the AG” of a hypothetical alkane with zero surface area which was extrapolated from AG? of a series of n-alkanes®, The factor S; is related to the ratio between the specific and the dispersive components of surface energy. The higher the S, values for the adsorption of a highly polar probe, the stronger is the interaggregate interaction. With regard to filler surface energy, two parameters are therefore of importance as far as hydrocarbon rubbers are concerned: y;, which reflects filler-polymer interaction, and S,, which is a measure of the ability of the filler to form a network in the polymer matrix. ‘The data in Table I suggest that while the high y; of carbon blacks should lead to strong filler-polymer interaction, the high S, of silicas would result in developed agglomeration of the aggregates. EXPERIMENTAL Natural rubber (NR) was selected for this investigation, Since it has been demonstrated that the radical reaction of peroxide vulcanization is not inhibited by the presence of silica or carbon black*, dicumyl peroxide (DCP) was chosen as a curing agent to obtain constant crosslink density. The formulation used in this investigation is given in Table II. A two-stage mixing procedure was employed for all compounds, using a laboratory BR Banbury. In the first stage, the polymer, filler, and other ingredients (except curatives) were mixed in the Banbury with a batch size adjusted to a fill factor of 0.67. The curatives were then added to the compound on an open miill at a batch temperature not exceeding 60°C. All compounds were cured at 155°C to ty determined with a Monsanto ODR rheometer. Standard test procedures as prescribed by ISO, DIN, or ASTM were used for the deter- mination of compound and vulcanizate properties. Dynamic properties were measured using an MTS 831 machine and applying the test conditions specified in DIN 53 513 (sample di- Tame 1 Compounn Formutation® NR (RSS1) 100 Filler 0-50 Zine stearate 05 Dicumyl peroxide variable * In parts by weight. FILLER-ELASTOMER INTERACTIONS 331 mensions: 10 X 10 mm, preload: 50 N, amplitude: 25 N, frequency: 16 Hz). Payne Effect measurements were conducted at 5 Hz at 23°C. For the measurement of bound-rubber content, the samples were conditioned at room temperature for seven days, then placed into a stainless-steel cage, and immersed in toluene. The solvent was renewed after three days. After seven days, the steel cage with the swollen sample was removed from the solvent, the samples dried in air for 24 h, then dried to constant. weight in an oven at 115°C, and the bound-rubber content was calculated. Swelling measurements of the vulcanizates were performed in benzene at room temper- ature. For the ammonia treatment, the uncovered bottle containing the toluene and the vulcanizate sample were placed in an ammonia atmosphere. The samples were allowed to equilibrate for seven days, and again the solvent was renewed after three days to remove soluble materials from the vulcanizates. Equilibrium swell was expressed as the volume fraction of rubber, V,, in the swollen matrix. For mixed solvent (acetonitrile-toluene), equi- librium swell was expressed as the swelling index, Z.e., the weight of the solvent divided by the weight of the sample. RESULTS AND DISCUSSION BOUND RUBBER The formation of bound rubber has been attributed mainly to physical adsorption and chemisorption™”’. Recent investigations® by xylene extraction of carbon-black-containing NR and SBR compounds at high temperature and the extraction of silica-filled rubber in an ammonia atmosphere showed that bound rubber is essentially caused by physical adsorption. The surface energy should therefore play a significant role in the formation of bound rubber. The results of the bound-rubber measurements by toluene extraction for N110 and the silica Pl are listed in Table III. These results show that the bound-rubber content is much higher for the carbon black than for the silica. The high value for the carbon black is certainly related to its high dispersive component of surface energy, which indicates strong filler- polymer interaction. VISCOSITY OF COMPOUNDS Figure 1 shows the dependence of Mooney viscosity measured at 100°C on filler loading. In order to illustrate the effect of the degree of filler loading, the relative viscosity, ML", was used for the plot. It is defined as Ml Me Ma a @) where ML, is the Mooney viscosity of the filled compound, and ML is the Mooney viscosity of the gum. In both cases, the relative viscosity increases with filler loading. At small con- centrations, there is no significant difference in ML® between the two fillers. However, at high loadings, considerable increases in relative viscosity are observed for the silica. This is due, mainly if not entirely, to the agglomeration of silica aggregates, which is caused by the high 5; value. The abrupt rise of ML" indicates that agglomeration is so strong that it cannot be dispersed during shearing. With similar conceptions, Lee” and Sircar"? defined a "Tanta IT Bounp Runaen, % Loading, phr Carbon black N110 374 43.2 52.0 Silica PL 20.5 25.8 318 40 50. 332 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 5 ML f, OCP - 1.62 phr 4 3 Pt 2 N110 1 ¢ ° 0 0.04 0.08 012 016 02 0.24 0.28 Fic. 1.—Relative Mooney viscosity as a function of volume fraction for carbon black N110 and silica Pl critical loading, at which point an index L, i.e., the difference between relative viscosity and relative modulus (modulus of the filled compound divided by that of the gum), increases sharply. They assumed that the concentration had then reached a point where there was not enough rubber to fill all available voids in the filler. This may be true for carbon blacks which, on account of their high surface energies, possess a high affinity to hydrocarbon rubbers, resulting in good wetting of the surfaces. This does not seem to be the case for silicas, since the strong aggregate-aggregate interaction may cause agglomeration to occur at lower concentrations" where there is still more than enough rubber to fill all the voids of the filler. R ran DCP - 1.62 phr N110 0 0.04 008 O12 016 02 0.24 0.28 Fic, 2.—Relative minimum torque of curometer measurement as @ function of volume fraction for carbon black N110 and silica Pl FILLER-ELASTOMER INTERACTIONS 333 ‘The minimum torque obtained with a Monsanto Oscillating-Dise Rheometer is also a mea- sure of compound viscosity. Figure 2 gives the relative minimum torque, which is defined, similar to relative Mooney viscosity, as a function of loading. The change pattern in relative minimum torque is similar to that of relative Mooney viscosity, although the difference in viscosity in this case is greater than that observed in Mooney measurements. This, of course, should be related not only to the different test temperatures, but also to the different patterns of strain, Whereas a biconical rotor oscillates in the Monsanto Rheometer, a shearing disc rotates continually in one direction in the Mooney Viscometer. Therefore, the shear involved in the Mooney measurement can be considered as infinite strain, and the filler agglomerates can be broken down to a greater extent. This would reduce the difference in viscosity between compounds filled with carbon blacks or silicas, respectively. Moreover, the effect of the interaction between polymer and filler on compound viscosity should be taken into account. The high adsorption capacity of N10, on the one hand, would result in a slightly larger immobilized shell of polymer on the filler surface which, in turn, would result in a somewhat greater hydraulic volume of the filler. On the other hand, the higher content of bound rubber in the carbon-black-filled compound would also cause a considerable resistance to the shearing strain, since the movement of the whole molecule in the flow field would be restricted once a segment is anchored on the filler surface. In addition, the high content of bound rubber would decrease the portion of free-flow polymer in the medium. Consequently, the strong interaction of N110 with the polymer would partially compensate for the difference caused by the less developed agglomeration of the carbon-black aggregates. This effect could be more pronounced for the high-strain viscosity. a When a filler is incorporated into a compound, the maximum change in curometer torque during vulcanization increases. The ratio between the torque increase of the loaded compound and that of the gum was found to be directly proportional to filler loading, The slope of the linear plot showing the relative torque increase to be a function of filler loading was defined by Wolff??? as ay: Daze = Darin Dima — Dias where Dac ~ Dmis is the maximum change in torque during vulcanization for filled rubber, Dirae ~ Din is the maximum change in torque during vuleanization for gum, mp is the mass of polymer in the compound, and mp is the mass of filler in the compound. This parameter has been used to characterize the filler structure existing in vulcanizates, in particular for carbon blacks. Figure 3 shows the linear relationship between Daz ~ Dai and N110 loading, which was also observed in numerous measurements for other carbon-black-filled compounds. This sug- gests that carbor-black structure remains constant over the range of loadings investigated. Equation (3) furthermore shows that, based on a simple test, ay can be calculated from the change in torque which occurs during the vulcanization of two compounds, one gum and one loaded. In Figure 4, the values of a, are plotted for the two fillers versus filler concen- tration, yielding two curves with completely different patterns, The following conclusions can be drawn from these results: 1. ay is higher for silicas than for carbon blacks. 2. Whereas a, is independent of carbon-black loading, it increases with filler concentration in the case of silicas. 3. A sharp upward curvature at high loadings is observed for silica-filled compounds. Since a, is a measure of the in-rubber structure of fillers, the high ay values for silicas would be another evidence of additional, i.e., secondary structures formed in the vulcanizates due to their high S; factors. In the case of silicas, these secondary structures are so strong that they are not destroyed by oscillating shear strain in the rheometer. The increase of ay with filler loading can therefore be attributed to the decrease in the distance between ag- Me mp @) oy: B34 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 20 OCP - 1.62 phr 0 02 04 0.6 08 Maximum change in curometer torque during vulcanization as & function of my/me for earbon black NLLO and silica PL gregates, rendering the secondary network more developed. Evidence supporting this inter- pretation can be obtained by surface modification of silicas. When the high specific component of surface energy is shielded by long alkane chain grafts, a, decreases drastically and no longer shows a concentration dependence". This implies that the silica network no longer exists at this point. ‘The above discussion seems to suggest that no agglomerates are formed in carbon-black- filled materials. This contradicts a widely accepted view. Indeed, numerous studies concerning the strain dependence of modulus'*"” and electrical conductivity"™' have demonstrated that an agglomerate network is formed in carbon-black-filled vulcanizates. This network is 10 94% DCP - 1.62 phr a4 74 6 : pr 4 4s 34 24 N110 0 0.04 0.08 012 016 02 0.24 0.28 Fic. 4.—ay as a function of volume fraction for earbon black N110 and silica PL FILLER-ELASTOMER INTERACTIONS 335 probably so weak, as suggested by the low S, values of carbon blacks, that it may be considered destroyed by the shear deformation. ‘This is understandable, since an angular movement of the rotor of the Oscillating-Disc Rheometer of 3° is equivalent to approximately 20% strain”. In this region of strain, most of the carbon-black agglomerates may no longer exist. This applies even more so at high temperatures (150°C). SWELLING Whereas uncrosslinked rubber can be dissolved in a good solvent, vulcanized rubbers can only be swollen in solvent to an extent determined by crosslink density and the nature of the solvent. After incorporation of the filler, the volume fraction of the filled rubber in the swollen gel, V,, (corrected for the filler volume, since the filler is assumed not to swell in the solvent), should differ from that of the pure gum, V;.. In the case of reinforcing fillers, the strong interaction caused by the strong physical and/or chemical adsorption can have the same effect as crosslinks in the composites. The ratio V,,/V,, would thus decrease with increases in filler loading. On the other hand, the addition of the filler may change the crosslink density in the polymer matrix, since it affects the vuleanization reaction. For per- oxide vulcanization of natural rubber, however, Hummel’ and Wolff” have shown that the crosslink density in the rubber matrix is not affected by fillers. In cases where the crosslink density of the matrix is independent of the presence of the filler, Kraus™ derived the following expression relating the ratio V,,/V,-to filler loading: Vio Vio (for nonadhering filler) 4 and (for adhering filler), 6) with m = 3CUL — Vie") + Vyo ~ 6) where C is a parameter which is characteristic of the filler and which is related to filler- polymer interaction and the filler structure in the vulcanizate. For fillers having the same structure, higher values of C suggest stronger interaction between filler and the rubber matrix. Figure 5 shows the change in V,./V;y with filler loading for vuleanizates, using benzene asa solvent. For the two fillers investigated, V,»/V;, decreases with increases in filler loading, suggesting a swelling restriction of the rubber matrix due to the presence of the filler. Fur- thermore, the swelling restriction caused by the carbon black seems to be smaller than that of the silica. The C values calculated from Kraus plots are given in Figure 6. It was observed that C does not change significantly with the DCP dosage until 1.89 phr, beyond which the C values of Pl and N110 decrease. This confirms that C depends only on filler characteristics. The low values of C at high DCP concentrations may be due to the degradation of polymer chains. ‘The question now arises as to why the swelling restriction (or C) of carbon black is less than that of silica. The reason is not quite clear. It is certainly not caused by weak polymer- filler interaction, since the adsorption energy of the carbon black for model compounds of NR is very high. It may be associated with the filler structure in the swollen vulcanizates, since C is a structure-related parameter. Rigbi and Boonstra* have also shown that high filler structure also leads to a considerable restraint on swelling. The more developed filler structure in the silica-filled vulcanizates, especially the secondary structure formed by strong interaggregate interaction, might therefore still exist, at least partially, in the swollen vul- canizates. This may be understandable, since a swelling index of 2 corresponds to an ap- proximate expansion of 50% in each direction, 336 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 12 DCP - 1.62 phr WwW 09 08 o/(1-9) 07 ° 0.1 0.2 03 04 Fia, 6.—Kraus plots of swelling in benzene for N110- and PI-filed vuleanizates Figure 7 shows Kraus plots of toluene swelling data for silica-filled vulcanizates, treated with ammonia and untreated. After ammonia treatment, swelling increases considerably, even the Vio/V,y ratio is slightly greater than unity, which is actually characteristic of 2 nonadhering filler. Since the ammonia treatment is only able to cleave the physical linkages between filler and polymer and/or filler and filler and leaves the crosslink density of the matrix unaffected”, the increase in the V,./V.y ratio suggests that a rich solvent layer has formed around the filler aggregates. This illustrates that silica-rubber interaction in vul- canizates is a physical phenomenon. In addition, swelling measurements were conducted with a viscous mixture of toluene and acetonitrile. The former is a good solvent for NR, the latter a poor one. As shown in 09 4 i PI o7 4 ON 0.6 4 os 4 N110 04 4 0.3 4 DCP, phr 02 — ™— + Os 09 1.3 17 24 25 Fio. 6.—C as a function of the DCP dosage for N110- and Pl-filled vuleanizates, PILLER-ELASTOMER INTERACTIONS 337 14 4 WeoVer 124 Pi OCP - 2.03 phr 1 NHsg treated Untreated 0.8 06 4 gi(t-9) 0.4 - 0 0.04 0.08 012 016 02 0.24 0.28 Fic, 7.—Kraus plots of swelling of N10- and P1-filled vulcanizates in toluene, with and without ammonia treatment. Figure 8, in which the swelling index is defined as the amount of solvent absorbed per unit weight of vulcanizate, the total portion of solvent in the swollen network is lower than the sum of the two solvents used, indicating a strong deswelling ability of acetonitrile with regard to toluene swelling. Furthermore, the high swelling indices of carbon-black-filled rubber decrease more rapidly with increases in the acetonitrile concentration in the mixed solvent than those of silica-filled rubber. At high acetonitrile concentrations, the indices of the N110 vulcanizate are even lower. This may have been caused either by stronger inter- action of the silica surface with acetonitrile (as expected from the high specific interaction Swelling index DCP-2.03phr NHs treated __ © NT10 a Pt Solvent volume, % Toluene 0 20 40 60 80 100 MeCN 100 80 60 40 20 oO Fi. 8.—Swelling index of N110- and P1-flled rubbers in a mixture of taluene and acetonitrile asa function of the volume ratio of the solvents 338 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 between silica and nitriles") or by the cleavage of physical linkages between polymer and filler surface by nitrile adsorption (as in the case of ammonia treatment). However, it is not clear whether the low swelling indices of carbon-black-filled rubber in the mixed solvent and the nonsignificant change of swelling in pure toluene after ammonia treatment (Figure 8) are caused by strong physical interaction between polymer and carbon black surface or by chemisorption. If the attachment of the polymer molecules on the carbon-black surface is of a physical nature, it could be strong enough not to be cleaved by ammonia adsorption. MODULUS It has been widely accepted that the laws governing the change in elastic modulus are determined by compounding, processing and, especially, vulcanization conditions. As far as the filler is concerned, the modulus of the vulcanizate should furthermore reflect the effect of filler morphology (particle size and structure) and surface activity which, as stated earlier, determines filler-polymer and aggregate-aggregate interaction. With regard to morphology, structure is the most important parameter, since it determines the occlusion of rubber in the filler aggregates. This again noticeably increases the effective filler volume fraction and, hence, the modulus of the vulcanizate”, When two fillers with the same surface area and structure but different surface characteristics are compared with regard to the modulus of the filled rubber, the observed difference can be attributed to the surface activity, in partic- ular, the surface energies of these fillers. Tensile modulus. —The tensile moduli at 100% and 300% elongation (a,99 arid 990), Pe- spectively, are presented in Figures 9 and 10. Since these vulcanizates were cured with the same DCP dosage, i.., since their crosslink densities are virtually identical, the difference in modulus and the strain dependence can only be explained by filler characteristics. At relatively small elongations (100%), the moduli of silica-filled rubber are much higher and increase more rapidly with increases in filler loading than for the carbon black. On the other hand, the higher 300% moduli of the black vulcanizate suggest a different mechanism of the filler effect on modulus. At small and moderate deformation under nonequilibrium conditions, the strong interaggregate association of silicas does not seem to have been completely broken down, which could contribute to the higher moduli. This is confirmed not only by the virtually identical 100% moduli of the two fillers at small loadings at which the aggregates hardly OCP - 1.62 phr O 0.04 0.08 012 016 02 0.24 0.28 Fla. 9.—oim a8 a function of volume fraction for earbon black N110 and silica PL FILLER-ELASTOMER INTERACTIONS 339 Ox9 , MPa DCP - 1.62 phr N110 Pt a ote ee ee ee io, 10.—-e5c0 modulus as a function of volume fraction for carbon black N110 and silica PL agglomerate, but also by the low loading dependence of the material filled with N110, which shows only a little tendency toward agglomeration. At high elongations, such as 300%, the effect of the interaggregate association should disappear, and the interaction between polymer and filler may now play an important role in modulus. In the case of the silica, the poor interaction with rubber leads to slippage and deattachment (dewetting) of rubber molecules on the filler surface, resulting in low modulus. The higher moduli of black vulcanizates are certainly related to the high localized bonding between polymer and filler, caused either by strong physical adsorption or by chemisorption. Effectiveness factor for modulus.—It was found that the Kinstein-Guth-Gold equation could be applied to filled vulcanizates over a wide range of strains if the effective volume of the filler was introduced”, oy = ool 1 + 2.8/6 + 14.1F%9%), (7) with dan = Sr where o,is the modulus of the filled vulcanizate, oy is the modulus of the gum, ¢ is the volume fraction of the filler, ¢,y is the effective volume fraction of the filler, and fis the effectiveness factor for transforming ¢ into dey The effectiveness factor is illustrated in Figure 11 as a function of strain. In the low- strain region, the effectiveness factor decreases with increases in strain, probably due to the destruction of the filler network. At high strain, where crystallization and nonaffine deformation (non-Gaussian behavior) of the polymer chains are involved, stress increases, which is equivalent to an apparent rise in filler volume. This effect would be diminished by deattachment and slippage of the rubber molecules on the filler surface, related to poor filler-polymer interaction. Kilian® has shown that the slippage energy of polymer chains on the carbon-black surface at high strain is related to the dispersive component of filler surface energy, ys. Thus, the different patterns of change in the shift factors for silica and carbon black can be explained by their surface energies. Dynamic modulus E'.—Figure 12 shows the elastic modulus as a function of filler loading for N110 and the silica Pl, At small dosages, the elastic modulus, £’, is comparable for both fillers. However, in the practical range of filler loadings, the silica gives a three times higher 340 RUBBER CHEMISTRY AND TECHNOLOGY Vou. 65 DOP - 1.62 phr 1 3 5 7 Fig, 11.—Effectiveness factor f for modulus as a function of strain for N110- and Pi-filled vuleanizates. elastic modulus than the carbon black. Since the carbon black exhibits much stronger fller- polymer interaction, hence leads to a greater volume of immobilized rubber shell, the high E' of the silica can only be interpreted as a result of the strong interaggregate interaction. This secondary filler—filler structure cannot be destroyed by the small strain applied in the measurement of dynamic properties (<2%). However, increases in strain will reduce these differences. Further increases in strain will alter the picture as in the case of 300% modulus, where the carbon black leads to a higher modulus than the silica. This can be confirmed by measuring the “Payne Effect.” 30 OCP - 2.3 phr 20 10 0 0.04 008 0.12 O16 02 024 0,28 Fra, 12.—Elastie modulus £” as a function of volume fraction for carbon black N110 and silica PL FILLER-ELASTOMER INTERACTIONS. 341 Payne Effect.—The plots of E’ as a function of log amplitude for vulcanizates containing 50 phr N110 or P1, respectively, are presented in Figure 13. The following observations can be made concerning the relationship between £” and the strain amplitude: 1, At low strain amplitude, £” is much higher for Pl than for N10. This is related to the stronger interaggregate interaction of the silica. 2, E decreases with increases in strain amplitude. This phenomenon is generally termed “Payne Effect”®*', However, the strain amplitude at which £’ starts to decrease is smaller for the carbon black than for the silica, indicating stronger filler-filler interaction in silica- filled rubber, since the Payne Effect is mainly attributed to the breakdown of a filler network. 3. The rate of decrease in the elastic modulus is greater for the silica than for the carbon black, suggesting a more rapid destruction of the silica network beyond a level of certain dynamic deformation, 4. At high strain amplitude, the elastic modulus reaches a low plateau in both cases, from which it can be inferred that filler-filler interaction must have been largely eliminated at this point. ‘The difference in behavior between the two fillers in the above plot should also be related to the mode of formation of the filler network. In the case of the silica, which is poorly compatible with hydrocarbon rubber and possesses a highly polar surface, agglomeration may take place by direct contact between aggregates, perhaps even via hydrogen bonds®, giving a rigid construction. This type of structure would be rapidly destroyed above a certain level of strain. In the case of carbon black, which has a greater affinity toward hydrocarbon rubbers, the network could be formed by a joint shell mechanism. The polymer enclosed in the joint shell gives a higher modulus than the surrounding rubber matrix. This joint shell construction is much less rigid than the direct contact of aggregates observed in the case of the silica. Although this type of filler network begins to break down at lower levels of strain, the breakdown proceeds less rapidly. SUMMARY Two fillers with comparable surface areas and structures, carbon black N110 and pre- cipitated silica PL, were selected as model fillers for the study of the effect of surface energy 40 Filler loading: $0 phr 5 Hz, 23°C 35 30 25 20 15 10 5 0 3.5 2.5 AS 0.5 Fig, 19.—Blastie modulus as a function of double strain amplitude for N110- and PI-flled vulcanizates (60 phr loading), 342 RUBBER CHEMISTRY AND TECHNOLOGY VoL. 65 on rubber reinforcement. Carbon black possesses a high dispersive component of surface energy, which is related to stronger interaction between filler and hydrocarbon rubber. The silica, on the other hand, is characterized by a higher specific component of surface energy, represented by a high S,, which is related to stronger interaggregate interaction. It is apparent from the investigation of the rheological properties of rubber compounds that a more developed interaggregate structure is formed with silica composites. The rapid increase in a, dynamic modulus, and 100% modulus with filler loading is indicative of the build-up of a strong filler network at high filler loadings in the case of the silica For N110, the higher bound-rubber content and higher 300% modulus are related to strong polymer-filler interaction This investigation shows that the surface energy of a filler plays a predominant role in rubber reinforcement and has a profound influence on the performance of rubber products. REFERENCES 'M.-J, Wang, $. Wolff, and J.-B Donnet, RuBaER CkeM, TRCHNOL. 64, 569 (1991). *M. J. Wang, S. Wolff, and J.-B, Donnet, Kantsch. Gummi Kunsist. 45(1), 11 (1992). *M.-J. Wang, S. Wolff, and J.-B. Donnet, Ruane Oxew. TECHNOL. 64, 714 (1991) *K. Hummel, Kautsch. Gummi Kunstst. 18, 1 (1962) *P.B. Stickney and RD. Falb, Runuer Cem, TEcHwou. $7, 1299 (1964), *C. M. Blow, Polymer 14, 309 (1973), * EM. Dannenberg, Rosser Citi, TecaNoL. 69, 512 (1986), *S. Woltf, M.-J. Wang, and E.-H. Tan, Paper 20, presented at a mecting of the Rubber Division, AC October 8-11, 1991 °B.L, Lee, RuonER CieM, Teruo. 52, 1019 (1979). A. K, Sirear, Rubber World 197(2), 30 (1987). 'B. Preund and S. Wolff, Paper 24, presented at a mecting of the Rubber Division, ACS, Mexico City, Mexico, May 9 12, 1089; abs. Rumen Cia, TecuNoL. 62, 770 (1989), "8, Wolff, Kautsch. Gummi Kunstst. 22, 367 (1969). ©, Woltt, Kautsch. Gummé Kunstst. 23, 7 (1970). 16, -B. Donnet, M.-J. Wang, E. Papirer, and A. Vidal, Kautsch. Gummi Kunstst. 39, 510 (1986). % A. R. Payne, RunnER CEM. TecHNL, 39, 365 (1966), ALR. Payne, in “Reinforcement of Elastomers,” G. Kraus Ba., Interscience, New York, 1985, Ch. 8 "G, Gerspacter, H.H. Yang, and J. M, Starita, Presented at the International Rubber Conference, Paris, France, June 12-14, 1900. ' G. Gerspacher, H. H. Yang, and C. P. O'Farrell, Paper 25, presented at a meeting of Rubber Division, ACS, Washington, D.C., October 9-12, 1990, '* J, -B. Donnet and A. Voet, "Carbon Black, Physies, Chemistry, and Elastomer Reinforcement,” Marcel Dekker, New York, 1976, Ch, 7. ® A, Voot and F. R. Cook, Runuen Chea, TECHNOL. 44, 1208 (1988). * A Voet, A. K. Sircar, and T, J. Mullens, RuuBER CHEM. TRCHNOL, 42, 874 (1969). #8, B, Boonstra, in “Rubber Technology and Manufacture,” Second Edition, C. M. Blow and. Hepburn, Eds, Butterworth Selentifie, London, 1982 © §, Wolff, Kautsch. Gummi Kunstst. 28, 7 (1970). %G. Kraus, J. Appl Polym. Sei. 7, 861 (1963), °7, Rigbi and B. B. Boonstra, Paper 3, presented at a meeting of the Rubber Division, ACS, Chicago, II, September 13-16, 1967, abs. RuonER Cima, TECHNOL, 41, 72 (1968). * KB Polmanteer and C. W. Lentz, RUBBER CHEM. TECHNOL. 48, 795 (1975). * AL Medalia, J. Colloid Interface Sci. 82, 118 (1970) AL Medalia, Rusoea Cum, TechNot. 45, 1171 (1972), ® AI, Medalia, Runuen Cum. Tecunot. 47, 411 (1974), © G. Kraus, Ruane Citea, TECHNOL. 44, 199 (1971), 218, Wolff and J. -B. Donnet, Rune Chea, TECHNOL. 63 32 (1990) °° H.-G. Kilian, unpublished A. R, Payne, J. Polym. Sci. 6, 57 (1962). 3A. R. Payne and B, E. Whittaker, Ruse Cita, PacuNoL, 44, 440 (1971), % 8. Woltf, Presented at the International Exhibition “Tires '91', Moscow, USSR, March 14-21, 1991. , Detroit, Michigan,

You might also like