Sei sulla pagina 1di 8

Journal of Colloid and Interface Science 331 (2009) 335342

Contents lists available at ScienceDirect

Journal of Colloid and Interface Science


www.elsevier.com/locate/jcis

Gelling nature of aluminum soaps in oils


Xiaorong Wang , Mindaugas Rackaitis
Bridgestone Americas, Center for Research and Technology, 1200 Firestone Parkway, Akron, OH 44317, United States

a r t i c l e

i n f o

a b s t r a c t
Aluminum soaps are notable for their ability to form soap-hydrocarbon gels of high viscosity. For more than half a century, it has been believed that the gelling mechanism is due to a formation of polymeric chains of aluminum molecules with the aluminum atoms linking along the axis and with the fatty acid chain extended sideways. Here we report results from an investigation using high-resolution electron microscopy and rheology measurements that clearly resolve the ambiguity. Our results reveal that the gelling mechanism stems from the formation of spherical nano-sized micelles from aluminum soap molecules, and those colloidal micelle particles then aggregate into networks of highly fractal and jammed structures. The earlier proposed polymer chain-like structure is denitely incorrect. The discovery of aluminum soap particles could expand application of these materials to new technologies. 2008 Elsevier Inc. All rights reserved.

Article history: Received 11 July 2008 Accepted 13 November 2008 Available online 20 November 2008 Keywords: Aluminum soap Soap-hydrocarbon gels Rheology Electron microscopy Fractal network Colloidal micelle particle Jamming

1. Introduction Soap is a surfactant used with water for washing and cleaning. The earliest record of producing such materials dates back to 2800 BC, when a process known as saponication [1] was used to produce sodium or potassium-fatty carboxylates. At modern time, soaps are usually manufactured via metathesis processes in aqueous solutions. A systematic investigation of their chemical and physical properties only started from the late of 1930s [2], when metal carboxylates in groups 1 to 8 of the periodic table were well prepared in pure state for the rst time. Many metal carboxylates, as we know them today, cannot be used for washing and cleaning, though they are still called as soaps. Examples are metal carboxylates formed from groups 3 to 8 in the periodic table. They are mostly covalently bonded and typically waterinsoluble, however, are rather soluble in organic solvents. Among them, aluminum soaps have received the most noticeable attention because they can effectively thicken many organic solvents and oils [26]. The excellent thickening power of the aluminum soaps can be illustrated by the fact that with 0.3% aluminum soap, the relative viscosity of hydrocarbons can rise as high as 30. In addition, a rigid gel can be obtained with the soap concentration as low as 3% [3]. Because of chemically inert, odorless and nontoxic, aluminum soaps have been used extensively in manufacture of greases, paints, gels, cosmetics, drugs and foods [7,8]. During World War II, they were used as the thickening agents for making stick fuels for ame-throwers and incendiary gels for napalm

Corresponding author. E-mail address: wangxiaorong@bfusa.com (Xr. Wang).

bombs [9]. They were also used in petroleum industrials for closing pores in water-sensitive zones [10]. Aluminum soaps of industrial grade usually contain various amounts of aluminum hydroxide, free fatty acids, aluminum monoand di-carboxylates [4]. Among them the effective components are the aluminum di-carboxylates (or di-soaps) [11]. Many studies [2] have shown that aluminum di-soaps from dibutyrate to distearate have very similar properties in organic solutions, differing in degree but not in kind. All of them can swell and dissolve in hydrocarbons to form transparent gels. Addition of small amount of polar chemicals, such as alcohol, cellosolves, or xylenols, can change the gel to jelly or even to viscous sol [1220]. It would thus appear that the aluminum soaps might exist as polymers of high molecular weight. Sheffer [3] measured both viscosity and osmotic pressure in diluted solutions (104 molar) in order to determine apparent molecular weights. He found that the aggregation number of aluminum di-soaps, from the octanote to the octadecanoate, in benzene solution was about between 500 and 1000 molecules. Gray [12] proposed a polymeric structure of aluminum molecules in explanation of their ability to thicken hydrocarbons based on an octahedral coordination. This concept was later extended by McGee [13] with adjacent chains held together by van der Waals forces and hydrogen bondings. Bauer and co-workers [14,15] based upon their infrared studies advanced this proposal in which they suggested that aluminum had a coordination of six, and the octahedral was jointed by coordinate bonding through both hydroxyl groups and carboxylate groups of the fatty acid from which the soap is derived. Leger et al. [18] conducted a light scattering study of aluminum distearate in diluted benzene. They found the ratio of molecular weight to the square of the size of the soap aggre-

0021-9797/$ see front matter doi:10.1016/j.jcis.2008.11.032

2008 Elsevier Inc. All rights reserved.

336

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

Fig. 1. Earlier proposed polymer chain-like structure [318,20] of aluminum soap molecules. The jelly-like transparent material in the bottle is a toluene solution containing 2.5 wt% aluminum di-laurate.

gate is approximately constant and then interpreted the results as formation of linear chains of random coil shapes in the solution. Since then, for more than half a century, such a centipede chain-like structure with the aluminum atoms along the axis and with the fatty acid chain extended sideways, as shown in Fig. 1, has received wide acceptance among theorists and experimentalists [2022], though a rigorous prove is lacking. To the best of our knowledge, at present there is no conclusive evidence that aluminum soaps in hydrocarbons do take this polymeric chain-like structure. In this contribution we report an experimental investigation of aluminum di-laurate soap in an attempt to verify the proposed gelling mechanism. Here we report our new discovery based on this study and briey consider their implication on the nature of aluminum-soap solutions. 2. Materials preparation Lauric acid (C12 H24 O2 , 98% purity), sodium hydroxide (NaOH, 99+% purity), aluminum potassium sulfate dodecahydrate (AlK(SO4 )2 12H2 O, 99+% purity) and squalane (C30 H62 , bp 285 C/ 25 mm, 99% purity) were purchased from Aldrich. Hexane (C6 H14 , bp 69 C/760 mm, 99.6% purity) and toluene (C7 H8 , bp 110 C/ 760 mm, 99.8% purity) were used as supplied under nitrogen from the Firestone Polymer Company. Squalane was purged with nitrogen before use. In literature there are two ways to synthesize aluminum dilaurate. One is the wet method in which aluminum di-laurate was prepared from the corresponding fatty acid by a metathesis reaction of its sodium soap with an aluminum salt in aqueous solutions. This method has been used by a number of investigators in the past, including Sheffer [3], McBrain et al. [4,5], Gray [12], McGee [13], Bauer and co-workers [14,15] and Leger et al. [18], and has been well documented [218]. There is also a dry method suggested by Mehrotra [19] in which aluminum dilaurate was prepared by reacting aluminum isopropoxide with a fatty acid in an organic solvent under nitrogen atmosphere, then subsequently distilling off the isopropanol under reduced pressure to yield aluminum-soap materials. This method is more effective in making aluminum tri-soaps. In this study, we adopted the wet method instead of the dry method to make aluminum

di-laurate. This is because in industry all aluminum di-soaps are currently manufactured via a similar process in aqueous solutions [7,8]. Their structures and their gelling nature in oils are the most interesting because they are used daily in manufacture of greases, paints, gels, cosmetics, drugs and foods. In preparation, the sodium soap solution of 10 wt% solute was made by reacting a stoichiometric amount of the sodium hydroxide with the fatty acid in presence of distilled water. The aluminum salt solution of 10 wt% solute was made by dissolving weighted aluminum potassium sulfate in distilled water. Both solutions were then heated to 75 C. Under vigorous stirring, the aluminum salt solution was then added slowly into the sodium soap solution until the precipitate was completed [11]. The precipitate was then washed with water many times until free from sulfate and dried in vacuum at 90 C. This raw product was then soaked in dried acetone for 24 h to extract out the free fatty acid [3]. The nal product was then washed with acetone 2 times, and dried at vacuum at 50 C. On the basis of ash values the aluminum di-laurate prepared in this manner gave ratios of the moles of aluminum to the moles of fatty acid of 1:2 within 2% error bar. The aluminum di-laurate synthesized according to present method was analyzed by 1 H-NMR, 13 C-NMR, and 27 Al-NMR techniques. The experiments were carried out in CDCl3 solutions. 1 H-NMR experiments showed that the content of free fat acid (chemical shift at 9.8 ppm) in the compound was less than 1%. 13 C-NMR spectrum showed that the ratio of free carbonyl group (chemical shift at 180.74 ppm) to hydrogen-bonded carbonyl group (chemical shift at 179.99 ppm) was 1:1 within 1% error bar, which was a result of that the disoap molecule had two carbonyl groups and one hydroxyl group in its structure. 27 Al-NMR experiments were little bit complicated, because the spectrum was sensitive to the viscosity of the solution, particularly the gel formation. To signicantly reduce the viscosity, we added a small amount of isopropanol into the Al-soap/CDCl3 solution. The resultant spectrum showed that the aluminum dilaurate synthesized according to present method displayed only a single peak at the chemical shift of 6.9 ppm. Based on these results, we estimated that the purity of the product was about 95+%, and concluded that the product prepared was aluminum dilaurate.

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

337

Fig. 2. EM photographs of the morphological structure of aluminum di-laurate from a toluene gel containing 2.5 wt% solute.

Solutions of the aluminum soap in hydrocarbon solvents of various ratios were prepared by dissolving the soap into hydrocarbon solvents of measured weights in sealed bottles. Aluminum di-laurate swells in cold hydrocarbons, however, complete solution can only be made by heating the mixture to an elevated temperature for some time. Hexane solutions were prepared under vigorous stirring for 2 h at 70 C. Toluene solutions were prepared at 100 C and squalane solutions at 140 C. Once the solutions were clear, they were placed at room temperature (23 C) for at least 2 days prior to use. Depending on the soap concentration, a solution at 23 C could be the gel, jelly or viscous sol. 3. Experiments Electron microscopy (EM) observations were carried out with a Hitachi S-4800 electron microscope in scanning electron microscopy (SEM) and scanning transmission electron microscopy (STEM) modes. This equipment has the ability to visualize objects of 1 nm sizes. To investigate the morphological structures of dried aluminum soaps, thin lms of samples were made by cutting or by casting the corresponding soap solutions on a carbon-coated copper micro-grid. In the case of casting, the aluminum soap solution was rst diluted to about 102 wt%. Then, a small drop of the diluted solution was deposited on a carbon coated micro-grid. After the solvent was evaporated, the grid was directly examined under the electron microscopy. To observe the morphological structures of aluminum soaps in solution, we selectively used squalane for this investigation because it is non-volatile and can survive in the

high vacuum compartment of an electron microscopy. In sample preparation, we rst make an aluminum soap/squalane solution containing about 2.5 wt% the soap. The solution was then diluted by adding hexane to about 102 wt%. The diluted solution was deposited on a carbon coated micro-grid. After the hexane evaporation, the squalane formed a thin liquid lm on the grid. The grid was examined under the microscopy. Measurement of dynamic moduli ( G and G ) of the solutions were carried out at various strain amplitudes using a Rheometrics ARES strain-controlled rheometer equipped with dual 200 and 2000 g cm force rebalance transducers. Strain sweeps were made at 1 Hz. A cone and plate geometry was used to ensure a homogeneous strain eld. The plate diameter/cone angle combination used was 25 mm/0.02 radian. An aluminum soap gel was rst heated to above its melting point, and then was loaded in the cone/plate xture. The setup was then cooled down to a desired temperature and was allowed to equilibrate for 1 h. After the normal force relaxed to zero, a strain sweep from 0.01% to 30% was performed in logarithmic increments. The strain during oscillatory shear varied as sin t , where is the strain amplitude and is the angular frequency. The experiments were only conducted on squalane solutions because hexane and toluene are too volatile for testing. Turbidity measurements of the aluminum soap solutions were carried out with a turbidimeter (DRT-15CE) from USEPA & HF Scientic. The turbidity was recorded in Nephelometric Turbidity Units (NTU). Standard formazin solutions certied by HF Scientic were used for the calibration. Optical glass vials (Liquid Scintillation) with 28 mm outside diameter and 55 mm height from

338

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

Fig. 3. Morphological structure of aluminum di-laurate casted on an amorphous carbon lm surface from a toluene solution containing 0.02 wt% the solute.

Wheaton Scientic were used for holding the samples. Before the vial was placed into the equipment, the outside surface of the vial was carefully cleaned with a lint free wiper moistened with tetrahydrofuran and hexane. A measurement was taken from an average of three to ve readings. Reported data are within a 5% error bar. 4. Results and discussion For a toluene solution that is loaded with 2.5 wt% aluminum dilaurate, the solution at 23 C is jelly-like transparent material (see Fig. 1). The gel can withstand its own weight without ow, but it can be easily broken upon shaking. Increasing the soap content can improve the gel strength. Nevertheless, if the aluminum soap molecules indeed line up to form a chain of centipede structure as that shown in Fig. 1, the diameter of the chain due to the existence of fatty centipedes should be larger than 1 nm. Under current microscopic technology it might be possible to visualize the aluminum soap structure. Fig. 2 shows EM photographs of the morphological structure of the aluminum di-laurate from the toluene gel. The sample was cut from the gel. At rst glance, the results shown in the picture are very surprising. On micro-scale, the material cannot collapse completely after solvent evaporation and the lm is of highly porous nature. On nano-scale, the aluminum soap molecules do not organize themselves into linear chains as that proposed in Fig. 1. Instead, they form mono-dispersed and spherical shaped micelles with size of about 68 nm. Generally speaking, organic materials are composed of low atomic number materials and thus they exhibit little variation in electron density. That makes dicult to get contrast in conven-

tional TEM or STEM images. Staining with electron dense metal atoms is a technique commonly used. During the staining metal atoms are incorporated into the polymer material and thus increase contrast of different areas depending on their reactivity. In our case the aluminum soap already has metal atom that in itself indicates increased concentration of aluminum atoms in the image. Therefore images did not involve any additional staining in order to detect structure of the particles. The darker spots in the images show increased concentration of aluminum atoms and thus inner structure of particles. If the aluminum atoms would be evenly distributed across all sample there would not be noticeable contrast increase in any placeelectron scattering from evenly dispersed aluminum atoms would just increase opaqueness of the overall sample. In this study images do show that there is a denite structure of particles. To conrm this result, we dilute the gel solution to concentration of 0.02 wt% of the solute. This dilution was carried out at 23 C and homogenized using a mechanical shaker. The reason is that, according the Leger et al. [18], performing dilution at cold condition does preserve the characteristic structure in the starting solution. After the dilution, a small drop of the solution was deposited on a carbon-coated micro-grid. After the solvent was evaporated, the grid was examined under the electron microscopy. Fig. 3 shows the microscopy pictures for the morphological structure of aluminum di-laurate soap from the diluted toluene solution. Again, we nd that the aluminum soap molecules selfassemble themselves into micelles of size about 68 nm. Those primary particles then link together as colloidal aggregates to form a percolated network. Clearly, the network is not formed by chainlike objects, rather by nano-sized particles.

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

339

Fig. 4. Morphological structure of aluminum di-laurate casted an amorphous carbon lm surface from a hexane solution containing 0.02 wt% the solute. The bottom panel on the right with magnication of 300 nm was taken on a sample prepared under a restricted evaporation of hexane for 4 h. Other panels were taken on a sample prepared under open evaporation condition, in which completing the evaporation took about 5 min.

The existence of a network of aggregated colloidal particles in aluminum soap gels is very surprising. In order to determine solvent inuence, we changed the toluene with hexane and made a solution that has approximately the same concentration. The morphological structure of aluminum di-laurate in hexane is shown in Fig. 4. Again, we see the same structure: a highly porous network composed of nano-sized particles. Further, we restricted the evaporation rate of the solvents by cover the wet with a lid that had only a small hole over the center through which the solvent vapor could escape. Completing the evaporation of the solution at room temperature could take from 0.1 to 4 h, depending on size of the hole. However, there is no signicant change in the morphological structure of the aluminum di-laurate as that shown in Fig 4. Another possible reason of formation of this structure could be due to drying artifacts, because a phase separation of solvent and solute could occur on the micro-grid surface if the two components have different anities to the surface during the evaporation. To conrm that our observation is actually occurring in hydrocarbon solutions, squalane was selected for this investigation because it is non-volatile and can survive in the high vacuum of the electron microscope. In this case, we rst made a squalane solution that contains about 2.5 wt% soap. Then we diluted the solution by adding hexane to about 102 wt% at room temperature. After that, a small drop of the diluted solution was deposited on a carbon coated micro-grid. After the hexane was evaporated, the squalane formed thin lms on the carbon-coated grid and was examined under the electron microscopy. Fig. 5 displays the morphological structure of aluminum di-laurate inside the squalane

solution. Apparently, the squalane cannot completely wet carbon surface and has broken up as small at droplets of various thicknesses. For droplets that are thin enough or of thin edges, one can see the aggregated structures of the aluminum di-laurate in the liquid media. To the best of our knowledge, this is the rst time that experiments under well-dened conditions provide such clear picture that aluminum soap molecules form spherical nano-sized micelles in hydrocarbon solutions, and those colloidal micelle particles then aggregate into networks of highly fractal and jammed structures. Experiments on rheological measurement also support this conclusion. The dynamic moduli ( G and G ) of the aluminum disoap/squalane solutions as functions of the di-soap concentration are presented in Fig. 6. Similar to many gel forming systems, the sol-gel transition of the system is characterized by a crossover between G and G . At low concentrations (e.g., when C < C gel = 2.5 wt%), G of the material is larger than the storage modulus G . The materials in this regime behave like viscous liquids. Above the critical concentration C gel , the storage modulus G is always larger that the loss modulus G , and the materials in this regime typically act like elastic gels. Beyond the gel point, the storage modulus G increases powerly with increasing the departure from the critical point (C C gel ), as shown in the insert of Fig. 6. The relationship between G and ( C C gel ) may be described by a simple scaling relation G (C C gel )m , with the scaling exponent m = 1.25 0.03. It is worthy to note that in dealing with exible chain and rigid rod networks, an exponent of 2 to 3 is usually observed [23]. The exponent = 1.25 for aluminum di-soap solutions, as observed here,

340

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

Fig. 5. Morphological structure of aluminum di-laurate inside a squalane solution. Aggregated structure of the aluminum di-laurate can be seen in the liquid droplets.

Fig. 6. Dynamic moduli ( G and G ) of the aluminum di-soap/squalane solutions as functions of the di-soap concentration. The insert shows that the relationship between G and ( C C gel ) may be described by a simple scaling relation G (C C gel )m .

suggest that the gel structure may have remarkably different nature from any chain-like structure. In addition, the dynamic storage modulus G for those gels is very sensitive to the change of strain amplitude, as shown in Fig. 7. At low strains (when < 1%), a strain-independent state is usually observed. Upon further increasing the strain amplitude, the system becomes nonlinear, and the modulus falls off beyond a sort of yielding strain ( y 3%). At still higher strains the drop in modulus is enormous and all gels apparently approach their liquid val-

ues. It is well known that for polymer solutions or melts dynamic strains up to 50% ordinarily fall well within the range of linear viscoelasticity. Here the presence of a much lower yielding strain (i.e., 3%) for aluminum di-soap solutions also suggest that the solution contains transit networks of particle aggregations, as it can be broken up easily by strain or stress. Similar mechanical responses have been observed in oil suspensions of colloidal silica particles [24], aqueous suspensions of boehmite alumina powder [25], latex dispersions of polystyrene particles [26], carbon black or silica lled

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

341

Fig. 7. Nonlinear behavior of the aluminum di-soap/squalane solutions as functions of strain amplitude. All gels yield beyond a sort of critical strain ( y 3%).

rubbers [2730], where particle (or ller) networks are prevailing in these systems. Fig. 8 shows the response of a squalane solution to shear at various temperatures. The solution was loaded with 15.5 wt% of aluminum di-laurate. As expected, the modulus G of solution decreases with increasing temperature. The melting point of this gel is about 130 C. For viscosity of the solution, however, it takes a very different dependence: as the temperature increases initially it also increases, then it decreases as the temperature increases further. This phenomenon was previously observed only in some particle-lled systems [27], but not in pure polymer solutions. An explanation for this phenomenon is based on particle networks. Before the thermal energy is able to break the particle network, increasing temperature will expand the network and thus may loosen the jammed structure a little bit, which will result in lower G . However, as long as this increased distance between the particles is not enough to uidize the particle aggregates, under shear the expanded aggregates will jam together easily, which as a result will increase the viscosity. There will come a point when the thermal energy is large enough to move particles away from the network, then further temperature increase melts the network structure and reduces the viscosity. Measurements of critical micelle concentrations (CMC) of aluminum di-soap in hexane, toluene, and squalane were carried out by measuring the turbidity. The CMC was found by plotting the excess turbidity (the turbidity of the solution minus the turbidity of the solvent) versus the concentration. The concentration at which the excess turbidity is zero is the values of CMC. Fig. 9 presents such a plot. The critical micelle concentrations for di-laurate aluminum soap in hexane, toluene and squalane are 6.3 104 , 5.5 104 , and 1.5 105 wt%, respectively. All of the solutions investigated here have concentrations higher than those critical values. The present work demonstrates that aluminum di-laurate forms micelles in the oils and that the aggregation of these micelles forms a network that gives rise to the gel formation at concentrations above approximately 2.5 wt%thereby refuting the long-held belief that the gel formation was the result of polymeric chains of aluminum association. This micelle colloidal network mechanism also explains previously observed phenomena in such systems that are hardly explained by a polymeric chain model, such as why the viscosity or the modulus of a gel decreases with prolonged storage time [3]. It is now understood that these micelle particles aggregate into highly fractal structures that are transiently jammed and

Fig. 8. Temperature dependence of the nonlinear behavior of a 15.5 wt% aluminum di-soap/squalane solution. Top is the modulus G plot and bottom is the viscosity plot.

Fig. 9. Critical micelle concentration (CMC) of aluminum di-laurate in hexane, toluene, and squalane.

not stable in nature. Given suciently long time, they will tend to minimize their conguration energy by re-arranging the particles into more close packed forms. As a result, the hydrodynamic volume of these aggregates will decrease and the number of chains that link the network will decrease. Thus, the modulus of a gel

342

Xr. Wang, M. Rackaitis / Journal of Colloid and Interface Science 331 (2009) 335342

particle networks [24,27]. The discovery of the mechanism and the aluminum soap nanoparticles could expand application of these materials to new technologies. Supporting material The online version of this article contains additional supporting information: the remarks on microscopy method used and the remarks on NMR experiments. Please visit DOI: 10.1016/j.jcis.2008.11.032. References
[1] M. Willcox, in: H. Butler (Ed.), Pouchers Perfumes, Cosmetics and Soaps, 10th ed., Kluwer Academic, Dordrecht, 2000, p. 453. [2] N. Pilpel, Chem. Rev. 63 (1963) 221. [3] H. Sheffer, Can. J. Res. B 26 (1948) 481. [4] J.W. McBrain, W.L. McClatchie, J. Am. Chem. Soc. 54 (1932) 3266. [5] J.W. McBrain, E.B. Working, J. Phys. Colloid Chem. 51 (1947) 974. [6] N. Pilpel, Trans. Faraday Soc. 51 (1955) 1307. [7] Encyclopedia of Chemical Technology, vol. 8, Interscience, New York, 1979. [8] E.S. Peattison, Fatty Acids and Their Industrial Applications, Dekker, New York, 1968. [9] L.F. Fieser, G.C. Harris, E.B. Hershberg, M. Morgana, F.C. Novello, S.T. Putnam, Ind. Eng. Chem. 38 (1946) 768. [10] C.H. Phelps, E.T. Strom, USP 5,150,754, 1992. [11] W.W. Harple, S.E. Wiberley, W.H. Bauer, Anal. Chem. 24 (1952) 635. [12] V.B. Gray, Trans. Faraday Soc. B 42 (1946) 196. [13] C.G. McGee, J. Am. Chem. Soc. 71 (1949) 278. [14] W.O. Ludke, S.E. Wiberley, J. Goldenson, W.H. Bauer, J. Phys. Chem. 59 (1955) 222. [15] N. Weber, W.H. Bauer, J. Phys. Chem. 60 (1956) 270. [16] W.H. Bauer, N. Weber, S.E. Wiberley, J. Phys. Chem. 62 (1958) 106. [17] W.H. Bauer, N. Weber, S.E. Wiberley, J. Phys. Chem. 62 (1958) 1245. [18] A.E. Leger, J.C. Hyde, H. Sheffer, Can. J. Chem. 36 (11) (1958) 1584. [19] R.C. Mehrotra, Nature 172 (1953) 74. [20] A.S. Lodge, in: Elastic Liquids, Academic Press, London, 1964, pp. 236239. [21] R.B. Bird, R.C. Armstrong, O. Hassager, in: Dynamics of Polymer Liquids, vol. 1, Fluid Mechanics, second ed., Wiley, New York, 1987, pp. 7879. [22] F.G. Gaskins, J.G. Brodnyan, W. Philippoff, Trans. Soc. Rheol. 13 (1) (1969) 17. [23] Xr. Wang, J. Polym. Sci. Part B Polym. Phys. 37 (1999) 667. [24] C.G. Robertson, Xr. Wang, Phys. Rev. Lett. 95 (2005) 075703. [25] W.-H. Shih, W.Y. Shih, S.-I. Kim, J. Liu, I.A. Aksay, Phys. Rev. A 42 (1990) 4772. [26] R. de Rooij, D. van den Ende, M.H.G. Duits, J. Mellema, Phys. Rev. E 49 (1994) 3038. [27] Xr. Wang, C.G. Robertson, Phys. Rev. E 72 (2005) 031406. [28] Xr. Wang, M. Rackaitis, Europhys. Lett. 75 (2006) 590. [29] Xr. Wang, C.G. Robertson, Europhys. Lett. 79 (2007) 18001. [30] C.G. Robertson, Xr. Wang, Europhys. Lett. 76 (2) (2006) 278.

Fig. 10. Gelling nature of aluminum soaps in oils: spherical nano-sized micelles from aluminum soap molecules aggregate into networks of highly fractal and jammed structures.

will decrease and the viscosity of a solution will decrease with increasing storage time. 5. Conclusion We conclude that the gelling mechanism stems from the formation of spherical nano-sized micelles from aluminum soap molecules, and those colloidal micelle particles then aggregate into networks of highly fractal and jammed structures, as shown in Fig. 10. Earlier proposed polymer chain-like structure is denitely incorrect. Our observed mechanism also allows for better explanation and understanding of phenomena observed earlier though hardly explained by a polymeric chain model. One is that the viscosity of a solution or the modulus of a gel decreases with prolonged storage time [3]. The other is that the relative viscosity of a solution can sometimes increase with increasing temperature [4,5]. Now we understand that all those are the consequence of jammed

Potrebbero piacerti anche