You are on page 1of 15

Advances in Colloid and Interface Science 123 126 (2006) 33 47 www.elsevier.

com/locate/cis

The influence of surface-active media on the mechanical properties of materials


Eugene D. Shchukin
Institute for Physical Chemistry of the Russian Academy of Sciences, Moscow 117915, Russia Moscow State University, Department of Chemistry, Moscow 119899, Russia Johns Hopkins University, Department of Geography and Environmental Engineering, Baltimore, MD 21218, USA Available online 11 July 2006

Abstract This survey summarizes studies dealing with the influence of surface-active media on the mechanical behavior of materials that have been carried out at the Institute of Physical Chemistry of the Russian Academy of Sciences, at the Department of Chemistry of Moscow State University and at the Department of Geography and Environmental Engineering of the Johns Hopkins University, and presented partially at the SIS Symposia. The phenomena of the environment-sensitive mechanical behavior take place for all kinds of solids under the effect of adsorption layers and due to the contact with liquid phases containing some specific components which are physicallychemically akin with respect to a given solid and cause lowering of the surface energy of the solid and weakening of the cohesive forces in the surface (interfacial) layer. Studies of these effects and their atomic-molecular mechanisms provide basis for control, prevention or utilization of these phenomena. 2006 Elsevier B.V. All rights reserved.
Keywords: Adsorption; Surfactants; Mechanical properties

Contents 1. Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 2. Influence of adsorption on the strength of solids with hydrophilic and hydrophobic surfaces 3. Influence of adsorption on the strength of particle contacts . . . . . . . . . . . . . . . . . 4. Influence of adsorption on the mechanical properties of metals, liquid metal embrittlement . 5. Influence of adsorption on surface damage . . . . . . . . . . . . . . . . . . . . . . . . . . 6. Strength decrease of catalysts in reactive media . . . . . . . . . . . . . . . . . . . . . . . 7. Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Acknowledgements . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33 34 36 40 43 44 45 46 46

1. Introduction Among the various effects caused by adsorption, a special effect is the adsorption induced decrease in strength of solid
Johns Hopkins University, Department of Geography and Environmental Engineering, Baltimore, MD 21218, USA. Tel.: +1 410 358 6270; fax: +1 410 516 8996. E-mail address: shchukin@jhu.edu. 0001-8686/$ - see front matter 2006 Elsevier B.V. All rights reserved. doi:10.1016/j.cis.2006.05.013

bodies (ADS) commonly known as the Rehbinder effect. It was first described by Peter A. Rehbinder in 1928 as the facilitation of crystal cleavage and decrease in hardness of crystals caused by surfactant solutions, and was explained by him as the weakening of bonds in the crystal surface layer due to adsorption and lowering of the specific surface free energy, [1,2]. Since that time, a universal character of this phenomenon has been demonstrated in numerous experimental and theoretical

34

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

studies, for solids and materials of all kinds and nature, with ionic, covalent, metal, molecular structure, compact and porous, without exceptions, up to refractory metals, alumina, quartz, diamond, under the influence of some corresponding surfaceactive media; the general validity of its explanation has also been confirmed [311]. Systematic studies show that the level of these effects caused by external environment, and even the very form of their manifestation (strength decrease, embrittlement, hardness reduction, plasticizing), depend upon many physical and chemical factors, including the characteristics of the stressed state caused by applied forces, deformation rate, presence of various defects in the real structure of material, temperature, amount of active component in the surrounding medium and conditions of its delivery into the fracture zone. However, a general precondition of these phenomena is the specific physicalchemical interaction between the solid and the medium manifested in decrease. An elementary physicalchemical approach to the quantitative interpretation of the Rehbinder effect can be illustrated in the following way [6]. In accordance with the Gibbs-Volmer thermodynamic theory of the new phase nucleation, the critical nucleus size cc is determined by the balance between the release of some energy excess associated with the metastability of the initial phase, and the work necessary for the formation of a new surface (interface). For a special case of cavity with size c in the mechanically stressed solid, the first terms can be expressed as the release of free energy of elastic deformation with the density of p2/2E per unit volume (papplied stress, Eelasticity modulus) in the surrounding cavity volume, proportional to c3. This corresponds to a decrease in free energy of the system equal to a(p2/E)c3. At the same time, an increase in free energy takes place described by the term bc2, i.e. the product of cavity surface area proportional to c2, and . Here a and b are dimensionless coefficients. Combination of these two terms gives the critical cavity size, cc E/p2, and the critical stress, i.e. the strength of the body pc (E/c)1/2. This simple approach is valid both for 2- and 3-dimensional cases. Thus, one obtains the famous Griffith equation Pc aEr=c1=2 : The dimensionless parameter depends on the geometry of the cavity and elastic characteristics of the body (Poisson coefficient for an isotropic body). For brittle solids, is generally close to 1. A similar dependence Pc = Pc(E,,c) (E/c)1/2 also holds for a semi-brittle (and even ductile) fracture. In this case, the true thermodynamic surface (interfacial) energy can be replaced by the effective work of the new surface formation, , which includes dissipation of energy by plastic deformation, sound waves, etc. The value of can be orders of magnitude greater than true , but still depends on the energy necessary for bond rupture during the new surface appearance, i.e., after all, upon [6,10]. This paper presents a survey of studies on the Rehbinder effect that were carried out at the Institute of Physical Chemistry

of the Russian Academy of Sciences, at the Chemistry Department of Moscow State University, and at the Department of Geography and Environmental Engineering of the Johns Hopkins University. In the next section, the influence of adsorption on the strength of solids with hydrophilic and hydrophobic surfaces is described, and a common qualitative thermodynamic approach is proposed. In the third section, the adsorption effects on the strength of contacts between particles of various natures in different liquids and surfactant solutions are considered, and compared with the corresponding quantitative data on the free energy of interaction. Section 4 is devoted to the influence of adsorption on the mechanical properties of metals: plasticizing effect as a basis of surfactant use in lubrication, and the embrittlement caused by the surface-active metal melts. Section 5 illustrates the solid surface damage in surfactant solutions. All these sections concern mainly the reversible physical chemical manifestations of the Rehbinder effect, under the common influence of the decrease in the surface (interfacial) energy caused by adsorption, and mechanical stresses. Finally, Section 6 is devoted to the catalyst strength decrease and facilitation of wear caused by chemisorption effects in reactive media. A special attention is paid to examples illustrating the role and importance of studies of trends of these phenomena and their atomic-molecular mechanisms as physicalchemical means to control, prevent or utilize effects of strength decrease. 2. Influence of adsorption on the strength of solids with hydrophilic and hydrophobic surfaces Experiments with fine porous magnesium hydroxide described in [8,11] can serve as an example of quantitative correlation between the decrease in strength and in the surface energy of solids in the case of brittle fracture. Cylindrical samples with size 2 2 cm were prepared by pressing followed by hydration of the finely dispersed magnesium oxide. Water vapor adsorption was performed in desiccators at various vapor pressures, pH2O, with samples placed above the corresponding water/sulfuric acid mixtures. Two independent series of measurements were carried out. The mechanical strength of samples, P, was measured by compression tests, and the adsorption was determined as the increase in weight (for such, very high specific surface, increase in weight due to the presence of a monolayer of water is about 1%). Let 0 and P0 correspond to the initial values for dry samples and A and PA to these values after adsorption. Following the Griffith equation, one can expect a proportionality, namely
2 2 2 P0 PA =P0 r0 rA =r0 Dr=r0 :

Such proportionality is shown in Fig. 1, and the slope of the graph gives a reasonable estimation of 0, 300 mJ/m2. The following point deserves special attention. The water molecules, adsorbing from vapor phase, act here like a monolayer of any other surface-active substance. In this series

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

35

Fig. 1. Relation between the decrease in mechanical strength and in surface energy of highly porous magnesium hydroxide specimens shown in terms of variables of the Griffith equation.

of experiments water was not present as an individual liquid phase in magnesium hydroxide samples, as was demonstrated by NMR. This means that a decrease in the mechanical strength was not a consequence of the condensation of adsorbate molecules in contacts between particles of the solid, and was not connected with any processes of mass transfer and contact dissolution. The nature of this effect of weakening of interparticle contacts was simply a facilitation of bond rupture and formation of the new surface due to the adsorption of surface active molecules in statu nascendi (at the moment of creation) of these new surface cells, i.e. a decrease in in the course of water molecule migration into the mechanically stressed zone. The next step in this series of experiments was the transition to the multilayer adsorption and to the presence of liquid phase, which is shown in Fig. 2a. The appreciable decrease in the strength of magnesium hydroxide samples was induced by covering the surface with 14 adsorbate monolayers. The principal contribution to the decrease in mechanical strength (by a factor of 1.52) occurs in the range of moisture content of 1 1.5%, which corresponds to the formation of a saturated monolayer of molecules of the medium on the external and internal surfaces of these highly porous specimens. The maximum decrease in the mechanical strength found for a 3 4% moisture content is associated with further decrease in the

surface free energy upon multilayer adsorption. A subsequent increase in mechanical strength was observed in these experiments in the presence of high content of adsorbate, concluded then by a final continued drop; this has been explained as follows. In contrast to compact bodies, in dispersed porous media the decrease in mechanical strength may be partially compensated by capillary contraction forces which appear upon condensation of the vapor in the pores. For comparison, Fig. 2b shows that in the presence of ethyl alcohol the rise in the curve is not so pronounced as in the presence of water owing to weakening of capillary contraction forces due to its lower surface tension (at the boundary with air) than for water. An example of the whole spectrum of the effects of active media on mechanical strength is presented in experiments with KCl polycrystalline wire. In this case, a continuous transition took place from hydrocarbon medium, inactive towards an ionic solid, to dioxane, and further to water, first adsorbed from dioxane, and then as a phase, at 100%, as presented in Fig. 3. This figure shows the continuous transition from adsorption to liquid phase of dioxane, and then again, from adsorption to liquid phase of water. Since fracture is close to the brittle one, it is possible to use this strength isotherm for calculating the surface energy isotherm (at small water concentrations in dioxane), by applying the Griffith equation. This gives estimates for max and s1min (area occupied by a water molecule at the KCl surface), yielding a reasonable value for s1min 0.3 nm2. A monotonic continuous transition from the effect caused by dilute monolayers to that of a liquid phase may be viewed as a characteristic one for purely reversible physical adsorption. The role of water as a surface-active factor causing the decrease of cohesion forces is typical for various hydrophilic solids, compact and porous bodies and disperse structures in numerous natural, technological, and geological processes. One can include here rock disintegration, and soil formation, wet comminution and grinding of ores and other materials, swelling and creep of bentonite clays [1215]; the possible loss of ground stability after secondary oil recovery and filling oil collector with water, the need to protect the glass surface from water traces in fiber optics, etc., can also be mentioned here. This behavior of water as a surface-active and, sometimes, strongly surface-active substance with respect to ionic solids

Fig. 2. Compressive strength of magnesium hydroxide specimens after adsorption of water (a) and ethyl alcohol (b); concentration Wm corresponds to formation of an adsorption monolayer.

Fig. 3. Dependence of the strength of polycrystalline samples of potassium chloride on composition of the medium in heptanedioxane and dioxanewater solutions.

36

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

and other hydrophilic materials can be explained by its high physicalchemical affinity towards such solids and high value of the work of adhesion, i.e. the high level of compensation of unsaturated bonds within the interfacial layer at the solid surface. All these factors contribute to essential decrease in the surface (interfacial) energy caused by adsorption, and particularly by the contact of solid with a corresponding (related, akin) liquid phase. Indeed, interfacial = sl is related to the work of adhesion, Wa, and to the surface energies of the solid, s, and the liquid, l, by the Dupre equation rsl rs rl Wa 1=2Wcs 1=2Wcl Wa : Here Wcs and Wcl are the cohesion energies of the solid and of the liquid, respectively. Chemical affinity between a solid and a liquid is dictated by the similarity in Wc and Wa values and can cause significant decrease in sl relative to s. Similar but opposite trends in behavior are observed for various solids with hydrophobic surfaces under the influence of nonpolar substances in vapor and liquid states. Fig. 4 presents a comparison of independently found values of relative reduction in surface energy, , and in strength, P, of naphthalene samples in various media: from benzene, a very related phase, which has the maximal similarity to naphthalene, and to alcohol aqueous solutions, where the difference in polarity is the highest [10,11]. Analogues features of the environment-sensitive mechanical behavior were observed in studies of mechanical strength decrease of pressed calcium carbonate, zinc oxide, sodium chloride, charcoal and silica gel samples in the presence of water, benzene, carbon tetrachloride vapors, carbon dioxide, and nitrogen. The effects of mechanical strength decrease were always the greater, the closer were the polarities of the adsorbent and the medium (e.g. water on silica gel, benzene on charcoal). This phenomenon of the adsorption-induced decrease in strength was observed also with more complicated porous structures, such as specimens of NaA and NaX zeolites
Fig. 5. Effect of alcohols of the saturated series CnH2n+1OH on (a) the strength, Pc, of polyethylene, and (b) on the limiting stress, Pt, corresponding to the time before fracture tf = 100 s for poly(methyl methacrylate).

granulated with different clays (kaolinite, bentonite), aluminosilicate catalysts of cracking filled with zeolite, etc. [8]. Fig. 5 illustrates a different character of effects caused by a series of alcohols, from highly polar ones, such as methanol and ethanol, to octanol on the strength of hydrophobic polyethylene and poly(methyl methacrylate) with relatively hydrophilic surface. On the one hand, it is essential to stress that only reversible physicalchemical interactions between the medium and the solid surface have been considered here as examples of the Rehbinder effect, in the absence of any dissolution or corrosion processes taking place in an aggressive environment. On the other hand, it should be pointed out that all the described phenomena are caused by a simultaneous action of surrounding medium and mechanical stresses applied to the body. 3. Influence of adsorption on the strength of particle contacts A direct quantitative estimation of changes in the surface energy of solids is quite complex. These difficulties are partially overcome in studies of contact interactions in coagulates with globular structure, and particularly those between individual particles in various polar and nonpolar media and surfactant solutions. These primarily are mechanically reversible coagulation contacts (without bridging) between molecularly smooth particles. Such studies were carried out using various experimental techniques in a number of scientific centers, including our scientific school at the Moscow State University, in collaboration with the Institute for Physical Chemistry of the Russian Academy of Sciences and the Johns Hopkins University [1622]. The disjoining pressure, (h), which is the universal thermodynamic characteristic of interaction between particles in a given medium, is defined as the force of interaction per unit area between two plane-parallel surfaces separated by a gap (film), h. The first integration of (h) with respect to h yields the work, F(h), that one needs to be performed in order to decrease the gap between the surfaces from infinity to h; this value is commonly called the specific free energy of interaction. Similarly to , the latter is positive for repulsion, and negative for attraction. For the case of a gap between two planar solid

Fig. 4. Comparison of independently found values of the relative reduction in surface energy and in strength of naphthalene samples in various media: 1, benzene; 2, heptane; 3, methylene chloride; 4, chloroform; 5, carbon tetrachloride; 6, methyl alcohol; 7, ethyl alcohol; 8, propyl alcohol; 9, butyl alcohol; 10, tertiary butyl alcohol; 11, 0.2 N solution of butyl alcohol in water; 0 and P0 refer to tests in air; straight line corresponds to the Griffith equation.

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

37

surfaces, direct measurements of this quantity are practically impossible. However, in the case of a curved gap, such as that between two identical molecularly smooth spheres with radius r, the integration of (h) over the surface of these particles (double integration) yields the force of interaction between the particles, p(h), which is related to the free energy of interaction via Derjaguin's theorem [23] ph prFh: This fundamental relationship was confirmed in precise experiments with various surfaces, in various media, described in a series of classical works [2331]. For an equilibrium gap, h = h0, in an immediate contact, one can write ph0 p1 prFh0 prF: The free energy of interaction in an immediate equilibrium contact, F, can also be referred as the free energy of cohesion. It is related to the interfacial energy, 12, and to the specific free energy of the residual equilibrium film, , in immediate contact zone as follows g 2r12 F: (In other notions, f = 212 + f, where f = , and f = F.) For the case of a perfect contact with no traces of the medium (small ), 12 F/2. This opens a way for a direct measurement of the surface energy of a solid, the possibility of which was doubted in some earlier studies. In other cases, one needs for estimation conditions in the gap to combine the previous equation containing three variables with some additional equation, established by an independent method [11,1719]. In the following text, both the cohesive force, p1, and the free energy of cohesion, F, are given by modulus. The corresponding measurements of p1 were performed using the contact, or cohesive force (CF) apparatus described in [11,16,22,32]. In the case considered, one is interested only in estimating the cohesive strength (in contrast to the known fundamental studies of the long-range interactions which include both the force and distance measurements) [23,25,26,28,31]. For such measurements of cohesive strength, an extremely pliable dynamometer system can be used here allowing very high sensitivity in measuring forces, alongside with relatively simple construction, on the basis of magnetoelectric dynamometer. Fig. 6 illustrates the schematic of this apparatus. The magnetoelectric, i.e. direct current microampermeter (galvanometer) 1 serves as a highly sensitive dynamometer, which is the principal element of this device. Samples, a and b, are attached to holders 2 and 3. The bent holder 2 is attached to the frame (coil) of microampermeter, as extension of its axis. Holder 2 can thus rotate around this vertical axis when electric current is applied to the coil. The other holder, 3 is connected to a manipulator 4. With the help of the manipulator the samples are brought into contact. The moment of direct contact is observed with a longfocus microscope as the first deviation of sample b from the
Fig. 6. The scheme of the device for particle contact force measurements: a, b samples, solid or liquid ones, 1magnetoelectric dynamometer, 2,3holders, 4manipulator, 5vessel with liquid; the compressive ( p) and tensile ( f ) forces are determined by the current passing through the coil of galvanometer.

initial position. By passing current through the microampermeter coil, one applies a compressive load, f, to the samples. When the direction of the current is changed, this results in a tensile (rupture) force p. The minimum value of current necessary to cause rupture of contact is measured. The contact rupture is also observed by a microscope. Depending on the sensitivity of the microampermeter, this technique allows one to measure forces down to 10 nN, and even to a few nanoneutrons (0.0001 mg of force), which is of the order of strength of an individual interatomic bond. The CF apparatus allowed us to study cohesive forces between particles of any shape and nature. The systems studied included solid samples and small liquid droplets in air, as well as contact formation and rupture in various polar and nonpolar media and surfactant solutions. These studies covered a broad range of the contact interactions of particles responsible for rheological and mechanical properties of various concentrated dispersions and other materials [11,1622]. In this context, the terms cohesion and adhesion are used interchangeably with respect to coagulation contacts and structures. For the phase contacts arising due to particle bridging, the term cohesion is more preferable. The term adhesion describes also the interaction between a solid and a liquid at their interface. Cohesive force, or contact force, has the same meaning as the strength of an immediate (direct) equilibrium contact between particles. Experiments with the CF technique cover a very broad spectrum of the free energy of interaction (and of 12) values. In the case of dissimilar phases such as nonpolar surfaces (e.g. glass modified with methylation and fluorination) in aqueous medium, or polar surfaces (glass or quartz) in nonpolar hydrocarbons, as well as in the case of contact between these and other surfaces in air, values of F/2 are measured in tens of mJ/m2, thus characterizing a lyophobic character of the interface (Table 1). In contrast to this, in the case of nonpolar

38

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

Table 1 Free energy of interaction (cohesion), 1/2F (mJ/m2), in contacts between molecularly smooth hydrophilic and hydrophobic spherical particles a Particles Glass Methylated glass Fluorinated glass
a

Air (40) 22 28

Heptane 25 0.01 5

Water 0.01 40 50

Data for particles with radius R of 12 mm, in polar and non-polar media (extreme cases), presented as 1/2F=1/2Rp (mJ/m2). For methylated surfaces in air this value is of 1/2FS=22 mJ/m2 which corresponds to the surface tension of hydrocarbon.

particles in hydrocarbons or a hydroxylated glass surface in water, the interaction is weakened by 34 orders of magnitude, down to 0.01 mJ/m 2 , which characterizes a lyophilic character (lyophilicityity) of the interface. The intermediate cases are presented in Figs. 710 as the isotherms of the free energy of cohesion between hydrophobic (methylated) surfaces in liquids of various polarities and surfactant aqueous solutions [1820]. In Fig. 7, these data are given for a series of aqueous alcohol solutions. The figure shows that these isotherms follow the Ducleaux-Traube rule. For ethanol solutions, the 1/2F values are compared with changes in the interfacial energy calculated from the data for the work of wetting. In Fig. 8, the isotherm 1/ 2F(C) for methylated surfaces in sodium dodecyl sulfate (SDS) aqueous solution is accompanied by the data on interfacial energy changes obtained from adsorption measurements in a similar (but highly disperse) system. In both cases such a comparison is consistent with low residual f, i.e. corresponds to expulsion of adsorbed layer from the interparticle gap upon particles compression. Indeed, in both of these cases, the work of adsorption from aqueous phase is determined only by dispersion interactions between hydrophobic parts of surfactant molecules and the hydrophobic surface of the solid. The F(C) isotherm for the interaction of hydrophobic particles in a cationic surfactant solution (Fig. 9) is similar to

Fig. 8. Isotherm of the free energy of cohesion (1) between methylated particles in aqueous solutions of sodium dodecylsulfate with concentration C, in comparison with the interfacial energy 12 between methylated surface and SDS solutions (2) estimated as 1/2F0 s, where s is the two-dimensional pressure derived from the adsorption isotherm (a), and isotherm of the surface tension of SDS aqueous solutions (b).

that in a solution of an anionic surfactant. Both isotherms reach the final (non-zero) level of minimal F value. However, in the special case of a more lyophilic system, where one of the particles is acetylated, the final experimental F(C) values approach zero. At the same time, the calculated curve F(C) based on changes in the work of wetting intersects with the horizontal axis, which is indicative of the transition from negative disjoining pressure (attraction) to positive disjoining pressure corresponding to real disjoining. The 1/2F(C) isotherm for hydrophobic surfaces in a solution of non-ionic surfactant is presented in Fig. 10. This isotherm is similar to those for solutions of ionic surfactants. However, in this case the final minimal 1/2F value depends significantly on

Fig. 7. Isotherms of the free energy of cohesion between methylated particles in aqueous solutions of (1) methyl, (2) ethyl, (3) propyl, (4) butyl alcohols, with volume fraction ; (2) is the isotherm of the interfacial energy 12 between solid paraffin and ethanol solutions estimated as 1/2F0 L cos.

Fig. 9. Isotherms of the free energy of cohesion (a) between two methylated particles (1), and between one methylated and one acetylated particles (2) in aqueous solutions of cetylpyridinium bromide with concentration C, in comparison with values of F0132 1 2 (b). are the changes in energies of wetting; subscripts 1, 2, 3 refer to methylated and acetylated surfaces, and medium, respectively, superscript 0 refers to pure water.

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

39

Table 2 Free energies of interaction, Fsl/2 (mJ/m2), of hydrocarbon surfaces (methylated glass) in various nonpolar liquids a Nonpolar liquid PFTBA PFMCP C7F16 PFD n-C7H16 n-C8H18 Fsl/2 7.0 6.8 5.2 4.8 <0.01 <0.01

a The liquids are PFTBA, perfluorotributylamine; PFMCP, perfluoromethylcyclohexylpiperidine; PFD, normal fluorocarbon: perfluorodecaline (FL) and normal hydrocarbons (HL).

Fig. 10. Isotherms of the free energy of cohesion between two methylated particles in aqueous solutions of a nonionic surfactantthe oxyethylated ether C12E20.

time: relatively large molecules need more time for achieving an equilibrium structure of adsorbed layer. The previous data relate to a reversible (liquid-like) behavior of medium in the contact zone, due to non-specific adsorption from an aqueous solution at the hydrophobic surface. In the opposite case of adsorption from nonpolar phase at the hydrophilic surface, some irreversibility can take place manifesting intrinsic strength of the modifying (chemisorption) layer [19,22]; its finite resistance to rupture is then observed as the critical compressive force fc, i.e. as a jump-like transition from a very weak (lyophilic) contact to a stronger (lyophobic) one. The probability of this transition, W, depends on both the characteristics of adsorption and compressive force, f (Fig. 11). The systems with fluorinated substances deserve special attention. Of particular interest is a comparison of properties and behavior of adsorption layers of common, hydrocarbon (HS) and fluorinated (FS) surfactants at the interface between their aqueous solutions and various nonpolar liquid phases, such as common, hydrocarbon (HL) and fluorinated (FL) ones [33,34].

Direct measurements of the free energy of cohesion, F, by the same method show that for two methylated (solid) surfaces in FL and fluorinated surfaces in HL, cohesion is much stronger than for methylated surfaces in HL and for fluorinated ones in FL (see Tables 1 and 2). This means weak adhesion, i.e. low Wa energy, and relatively high 12 = S/L value at the S/L interface for the two first mentioned couples (with different natures of components) in comparison with the two last couples. Thus, it appears that the Fowkes' [35] approach (using geometric average in evaluating combined Hamaker constant and the work of adhesion) significantly overestimates Wa and underestimates F and 12 values for fluorocarbon/hydrocarbon interfaces, while Antonow's rule (12 = 1 2) gives much better agreement (Table 3). All these data explain why the stability of adsorption layers and of corresponding emulsions with respect to coalescence is often significantly higher for the HS/ FL and FS/HL systems than for the HS/HL and FS/FL systems [34]. A low affinity between nonpolar liquid phase and hydrophobic groups of surfactant molecules results in their expulsion from the nonpolar liquid into a thin interfacial layer where they build mechanically strong structure-rheological barrier, in accordance with the Rehbinder's doctrine. [36,37]. The observed values of the free energy of cohesion F span four to five orders of magnitude from 10 3 to 102 mJ/m2 (for molecular surface forces in mechanically reversiblecoagulation contacts between particles). For spherical particles with radii, r, in the range of 10 610 1 cm, the spectrum of contact attraction forces, p1 rF, extends over many orders of magnitude, say, from 10 4 nN to tens of 104 nN.

Table 3 Interfacial energy, 12 (mJ/m2), and the free energy of interaction (cohesion), F/ 2 (mJ/m2), for solid hydrocarbon surfaces (methylated glass) in various nonpolar liquids Nonpolar liquid Calculations 12 Fowkes Fig. 11. Probability W of breaking the adsorption layer between particles with polar surfaces (glass spheres) depending on the contact compressive force f in heptane solutions of (a) cetyl alcohol, 10 3 mol/L, (b) palmitic acid, 10 5 mol/L, and (c) octadecylamine, 10 5 mol/L. The W value is estimated as the share (%) of measurements in which the transition from weak to more strong cohesion occurs. PFTBA nC8F18 PFD nC7H16 nC8H18 0.3 1 <0.1 0 0 Antonow 4.5 6.3 2.3 0.1 0.1 6.4 6.4 3.3 0.1 0.1 7.0 5.5 4.7 <0.01 <0.01 F/2 = 1 2 Wa Experiment F/2

40

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

In combination with corresponding models of disperse structures, these data form the basis for evaluating the rheological (structure-mechanical) properties of various disperse systems [7,20]. As an initial approach, the model of additive summation of contact cohesive forces can be used, i.e. the estimation of the structure strength as Pc p1
av

v;

is the average cohesive strength of an individual where contact, and is the number of contacts in a unit cross-section [8,20,38]. In the same simple approach, the number of contacts per unit cross-sectional area in the primitive cubic globular structure can be estimated as 1/(2r)2. This gives an illustrative evaluation of the strength (critical shear stress) of a globular coagulation structure with only two parameters, namely, the free energy of cohesion and dispersity Pc prF=2r2 F=r: For the above ranges of F and r values, this shows extremely broad, but realistic spectrum of rheological properties of such globular structures with molecular interparticle forces: in this approach, the critical shear stress can be as low as 10 3 Pa, and as high as 107 Pa. For structures with phase contacts, arising due to bridging of particles during various processes of pressing, sintering, hydration hardening, etc., the strength characteristics can achieve the level of GPa and more, approaching properties of compact construction materials [11,39,40]. These evaluations provide for broad opportunities in manipulating the strength and rheological characteristics of various dispersed systems by selecting appropriate media and surfactant additives [4144]. 4. Influence of adsorption on the mechanical properties of metals, liquid metal embrittlement As mentioned above, the systematic investigations of adsorption influence on the mechanical properties of materials have shown that these effects can take place for all kinds of solids. Some data for metals are presented in [310]. In Fig. 12, the tension diagrams (dependence of applied stress on deformation) are given for single crystal samples of tin immersed in a pure vaseline oil (the situation in air is the same), and in the solutions of oleic acid in vaseline oil. In all cases the addition of oleic acid decreases the sample resistance to plastic deformation. This decrease is most significantly seen for a concentration of 0.2 wt.%. The same optimal concentration of 0.2 wt.% for the maximum value of this effect was also observed for polycrystalline samples of tin, lead and copper. This characteristic concentration corresponds to the equilibrium saturated adsorption layer, and does not depend on the nature of the metal, but depends only on the nature of surfactant molecules. By analogy with the Ducleaux-Traube rule, this concentration decreases with the increase in the length of hydrocarbon chain for a homologous series. In the present case, the medium adsorption influence manifests itself as plasticizing effect, i.e. the facilitation of plastic deformation. Also, similarly to the above considered
Fig. 12. Tension curvesthe applied stress, P, dependence on deformation, , for single crystals of tin in solutions of various concentrations of oleic acid in vaseline oil: 1vaseline oil without oleic acid, 20.1% oleic acid, 31% oleic acid, 40.5% oleic acid, 50.2% oleic acid.

(av) p1

cases, the lowering of the surface energy caused by adsorption, and weakening of cohesive forces in the surface layers of material occur as primary effects. However, as mentioned earlier, a significant strength decrease is possible only if a number of factors are simultaneously met, namely, a combination of large decrease in the surface energy, the hard stressed state, kinetic opportunity for surfactant molecules to penetrate into the fracture zone, etc. In the described case, such a combination of these factors is not sufficient for promoting the sample strength decrease, for nucleation and propagation of fracture cracks. In particular, the adsorption of a common organic surfactant at a metal surface does not cause a strong (relative) decrease in the surface energy. The adsorption effect is manifested then as a facilitation of the formation of new surface cells accompanying the exit of dislocations at the surface (linear defects providing the plastic flow process) [4750]. As a result, the essential plasticization (softening) takes place just in the vicinity of the surface layers. The profound fundamental and applied studies of this effect have been carried out in the PhysicalMechanical Institute (Lviv, Ukraine), in collaboration with the Institute for Physical Chemistry of the Russian Academy of Sciences, and other scientific centers [45,46]. This phenomenon of plasticization is a defining one for the role that surface-active additions play in lubricants for friction nodes, for metal treatment by pressure, and in cooling-lubricant liquids for treatment by cutting. Inactive liquid media with low viscosity have no essential influence on the ability of a metal to undergo the treatment by pressure. In Fig. 13, data are presented for the drawing of aluminum bars. This figure shows changes in the tangential component, , of the stress tensor in the surface layer in various media: in pure octane (about the same as in air), octyl alcohol

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

41

[3,10,45,46,5154]. In Fig. 14, photos of a bar of aluminum alloy are shown: first, after bending in air, when the material is plastic and strong; then, deposition on the same metal of a small droplet of the surface-active metal melt containing tin and gallium; and last, brittle fracture after small, elastic only deformation, under much less applied stresses, due to the presence of liquid metal wetting in status nascendi the surface of a crack [3,45,51]. In Fig. 15, the stressstrain diagrams for single crystals of cadmium in the presence of very thin layers of tin, which has the lower melting point, are given, at a room temperature and at sufficiently high temperature, when the covering tin layer becomes liquid. Only in the last case the liquid metal embrittlement (LME) effect and its kinetic conditions can take

Fig. 13. Drawing of aluminum samples: tangential stress as a function of deformation for various concentrations of octyl alcohol in octane: 1pure (100%) octane, 20.25% solution of octyl alcohol in octane, 30.5% solution, 43% solution, 55% solution, 610% solution, 718% solution, 825% solution, 9pure (100%) octyl alcohol.

solutions in octane, and in pure octanol. In the case of drawing of a metal sample with the clean, juvenile surface, in air or inactive medium, the shear stress in the metal layer contacting the tool approaches the shear resistance (strength) of the treated material itself, i.e., the drawing cannot be done without damaging the sample. The rupture already occurs at the reduction of a sample (draft) of only 78%. Immersion into a surface-active medium changes the situation dramatically. Impressive effects of the lowering of drawing forces and of increase in deformability (the limit reduction) are primarily caused by plasticization of the surface layers of a metal. This plasticized layer acts as a non-extrudable lubricant. The role of the organic lubricant layer rheology is much less than that of surface layer plasticization. In the last case, under conditions of purely physical, reversible adsorption of alcohol, its optimal concentration is 100% liquid phase (the same as in the case of water adsorption on magnesium hydroxide discussed earlier). In the previous case of adsorption of organic acid on metal, the chemisorption can be predominant, and the formation of a saturated monolayer occurs as an optimal one (like in flotation, without the so-called overoiling and formation of bilayers). Let us return now to the notion of the defining role of the high physicalchemical affinity of the solid and the medium in the strong manifestation of the medium effect. With respect to solid metals, some liquid metals can serve as highly akin ones. Indeed, the presence of such an affinity and strong lowering of interfacial energy rather often lead to a catastrophic decrease in strength and to embrittlement of metal (provided that other conditions are met, such as sufficiently hard stressed state with tensile components, opportunity for the atoms of an active component to penetrate into the zone of pre-fracture, etc.)

Fig. 14. Deformation of an aluminum alloy bar in air is plastic and needs significant force (top picture), but in the presence of a small (several mg) droplet of the surface-active metal melt containing gallium and tin (middle picture), a similar bar undergoes brittle fracture at low applied stresses (bottom picture).

42

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

Fig. 15. Deformation curvestrue tensile stresses, P, as a function of elongation, , of a single crystal of cadmium without coating (dashed lines) and with tin coatings (solid curves) at two characteristic temperatures: below and above the tin melting point.

their place, being inseparably connected with the mobility of the active component atoms allowing them to come to the zone of fracture. The LME effect has a very widespread character. Some data for LME effect are presented in Table 4, where they are compared to the values of mixing energy of the corresponding, solid and liquid metals. The affinity discussed above and a sharp decrease in surface energy (provided that the interface in the two-phase system is stable) are observed for small positive values of the energy of mixing. In terms of binary diagrams, these values correspond predominantly to a simple eutectic diagram with limited mutual solubility of components in the solid phase. (High positive values of energy of mixing are typical for the absence of substantial interaction, those close to zero correspond to infinite mutual solubility, and large negative onesto the chemical interaction of elements). This statement can be considered as an essence of a common (approximate) thermodynamic approach to the description of possibility, selectivity, and predictability of the LME. Similar effects of the strong decrease in strength were also observed for some minerals under the influence of appropriate salt melts, e.g. for silicates in contact with melted sulfides [6,12,13]. Using metals as candidates for the investigations of this effect was definitely advantageous for the explanation of its atomic-molecular mechanism. In the molecular dynamics simulation of such systems, one can be confined in some cases to a relatively simple Lennard-Jones [612] interaction potential for spherically symmetric particles [10,5558]. In Fig. 16, the results of molecular dynamics simulation are presented for such crystalline structure (two-dimensional systems): (a) plastic deformation at high temperature and brittle cracking at low temperature, and (b) simultaneous processes of crack nucleation under the influence of active foreign atoms influence, and propagation of these atoms to the tip of cracks. Even such simple molecular dynamics approach helps to illustrate the essence of the LME effect at a microlevel. The facilitation of bond rupture in a solid body in the presence of active atoms represents substitution of solidsolid bonds by solidliquid bonds which are similar in their nature, with high values of the work of adhesion Wa. At the same time, mobility

of active atoms in the adsorption layer (and particularly in the liquid phase) promotes filling the crack with these atoms and further propagation of a crack. As a rule, only relatively weak or moderate effects of surrounding media can be utilized in practical applications. Particularly, such effects are caused by organic and elementorganic surfactants in lubricants as mentioned above. Some attention was paid also to using these phenomena for enhancing rock drilling, grinding (i.e., dispersion) of various materials (e.g., see Table 5). The strong LME effects were considered mostly in the prevention of failure of stressed materials upon their contact with metal melts, e.g. in the known accidents of melting of an antifriction alloy in railway car bearings, in metal welding and soldering, and in particular in units utilizing liquid metal heat-transfer agents. However, it has been demonstrated that some liquid metals can be used to facilitate mechanical treatment of extremely hard materials. Fig. 17 compares the drilling effectiveness (depth vs. time) of highly tempered steel in the presence of a small amount of ZnSn eutectic melt to that in regular coolant liquids. However, the presence of a liquid metal phase hinders broad practical applications of the LME effect because of the need to maintain elevated temperatures, as well as difficulties in obtaining perfect wetting and reliable contact of substrate with the melt, and removing residual films or traces of melt. Excess of active metal left after treatment is inert as a solid phase, but occurs dangerous if being melted once. Nevertheless, these technological complications inspired a principally new approach to use the LME effects both in studies and applications in the absence of a liquid metal phase. This is achieved by substituting the liquid phase with a minimal, down to a monolayer, and controlled amount of active component, by means of electrochemical reduction of the corresponding cations on the treated surface at room temperature, as shown schematically in Fig. 18 [5961]. This electrochemical mechanical treatment (ECMT) involved in grinding of
Table 4 Calculated values of the enthalpy of mixing, Hm (eV), and interfacial energy, 12 (mJ/m2), for some solid and liquid metals, and qualitative data on the liquid metal embrittlement (LME) effect: (+) experimentally observed, or () absent in a given system Solid metalliquid metal pair AlZn CoBi CuBi MoAu MoHg NbCd NbSn TiCd TiGa TiHg TiPb WHg WPb ZnGa ZnSn ZrBi Hm 0.16 0.01 0.17 0.74 3.14 2.57 0.22 0.28 1.79 0.06 2.73 3.75 0.39 0.04 0.08 0.94 12 60 110 220 580 1980 1470 240 260 (860) 140 (1380) 2380 420 110 150 500 LME + + + + + + +

+ + +

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

43

Fig. 16. Molecular dynamics simulation of deformation and fracture in a two-dimensional crystal: (a) plastic deformation and formation of a dislocation AB at high temperature (above), and brittle crack at low temperature (below); (b) crack nucleation under the influence of surface-active atoms, at high temperature in both upper and lower cases. Arrows show the direction of applied stress.

the highly tempered carbon steel (hardness 6062 HRC) in the presence of active cations, supplied by the surrounding electrolyte, increases the grinding effectiveness (the amount of removed metal) by a factor of 2 for Zn2+ ions and by a factor of 3 for Cd2+ ions. The calculations of consumed electric current density indicate formation of an extremely thin film of the active component, down to monolayer, in which the active atoms (or ions) still possess the necessary mobility. The embrittlement caused by liquid metal takes place also for some solids with covalent bonds, e.g. germanium under the influence of thin films of gallium, copper, and particularly gold, as well as for various ceramics. 5. Influence of adsorption on surface damage Among all the phenomena considered, essentially dealing with surface, a special attention should be paid to the effect of adsorption on stability and damage of the surface itself. Traditional studies in this direction are performed by the dislocation etching technique revealing changes in the nearsurface dislocation patterns after indentation, and by a direct monitoring of the appearance of microcracks in the course of microscratching [62,63]. The latter is particularly informative in studies of surface damage caused by small loads under conditions of surface-active substance adsorption. Significant and reproducible media effects were identified for glasses, ionic
Table 5 Mechanical tests on sandstone samples from several coal fields: coefficient KIc = (3Fl/h2b)(c)1/2(c/h) MPa m1/2 for four-point bending samples of l h b size, with a notch of c depth, under the load of 2F/2 Coal field KIc In water In surfactant solutions Nonionic Donetsk Vorkuta Kuzbass 0.54 1.05 1.10 0.37 0.61 0.85 Anionic 0.41 0.58 0.80 Cationic 0.48 1.47 ? 1.04

crystals, semiconductors, quartz, limestone, and other materials by this and other methods. The decrease in surface damage resistance up to 50% and in same cases even highest was found. The data for the brittle damage probability for a fused silica surface in microscratch tests in aqueous solutions of cetyltrimethylammonium bromide are shown in Fig. 19. The resistance

Fig. 17. Effectiveness of drilling: the depth, h, of the drill bit (hard alloy WC Co) penetration into highly tempered steel versus time, t, in ZnSn eutectic (a), and in various commonly used cooling liquids, with the immediate blunting of a drill bit (b).

44

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

Fig. 18. Experimental scheme of the ECMTelectro-chemo-mechanical treatment: (1) sample, (2) electrolyte, (3) loading, (4) grinding wheel, (5) treated surface, (6) power supply, (7) motor, (8) anode.

of samples to damage drops monotonically as the CTAB concentration increases up to the critical micelle concentration; at higher concentrations, some new moderate increase takes place. These results are in agreement with direct measurements of strength decrease of sandstone samples from several coal basins caused by surfactant adsorption (Table 5). 6. Strength decrease of catalysts in reactive media Up to this moment, the influence of adsorption on the mechanical properties of solids was related, with a few exceptions, only to the case of completely reversible and purely physical adsorption. Now let us consider some reactive environment effects caused by particular chemical interactions. These effects include chemisorption and mutual interaction

between the solid phase and the medium in the course of heterogeneous catalysis. The phenomenon of accelerated disintegration of a catalyst's structure in the course of catalytic reaction has long been known. Rehbinder and Shchukin proposed for the first time to treat this effect as one of the manifestations of the environmentsensitive mechanical behavior, with the special role of chemisorption [8,6466]. Indeed, all the necessary conditions of the strong influence of adsorption on the mechanical properties of a solid are present here. The facilitation of bond rupture in catalyst granules can be caused by the adsorption effects of the initial, final, and particularly intermediate products, at various non-equilibrium stages of reaction. Contacts between particles in granules, which are the only carriers of mechanical strength, are open to the action of the medium. Mechanical stresses are present in granules during catalysis as thermoelastic or/and residual ones (inherited from the granule formation), weight of a layer, and particularly as a result of collisions in fluidized processes. The influence of intermediate products of the soft oxidation of propylene, such as carbon dioxide, acetaldehyde, and especially acrolein, on the mechanical behavior of cobaltmolybdenum catalyst is shown in Fig. 20, along with data for an inert (nitrogen) medium [8,66]. The results in this figure are presented as the dependence of durability (estimated as the logarithm of the time to fracture, tf, under a constant load) on the applied stress, P. The chosen coordinates are consistent with the well known, universal equation established by Zhurkov for a broad spectrum of various inorganic and organic crystalline and amorphous materials [6,67] tf t0 expU0 mP=kT ; where U0 is the activation energy, while t0 and are constants. The adsorption lowers the energy barrier for bond rupture, U0, and may also lead to the increase of the structure factor (activation volume), . Indeed, the data presented for the effect

Fig. 19. Probability of brittle damage, R = l/l0, for a fused silica surface which was subjected to microscratch tests with small indenter loads, F, in aqueous solutions (pH 6.4) with different cetyltrimethylammonium bromide concentrations: 1.2510 3 M/L (1), 1.2510 4 M/L (2), 1.2510 5 M/L (3), and 1.2510 6 M/L (4). The probability, R, is obtained from microscopic study of microcracks arising along the microscratch made by indenter, and evaluated as the ratio of the share of the indenter path, l, filled with microcracks with respect to full length of the path, l0.

Fig. 20. Durabilitythe time to fracture, tf, (in logarithm scale) of a cobaltmolybdenum catalyst under constant compressive stress, Pc, at 200 C, in various media: 1nitrogen, 2carbon dioxide (partial pressure 350 mm Hg), 3acetaldehyde (60 mm Hg), 4acrolein (60 mm Hg).

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

45

of chemisorption on granule durability obey the Zhurkov equation, yielding a straight line in coordinates lgtf P. The chemisorption causes a 3040% decrease in the strength and enormous, up to several orders of magnitude, decrease in durability. The experiments with MgO, CoMo, CaNiP, AlCrK and other catalysts show that during catalysis their strength and durability both decrease significantly, sometimes even dramatically, in comparison with identical tests in inert media. The influence of catalysis on mechanical strength and durability of magnesium oxide catalysts in isopropyl alcohol dehydrogenation reaction may serve as a typical example of this [65,66]. Already after 2030 s had elapsed after the start of the reaction, prior to the establishment of stationary conditions, the compressive strength of samples (tablets) diminished by a factor of 1.52. In this series of tests, the measurements were carried out after the cessation of catalytic reaction and drying of the tablets. Besides the Rehbinder approach, these effects can be considered as manifestation of the chemo-mechanical effect described by Westwood et al. [4] and Latanision et al. [5,47]. The most direct data are obtained by measurements performed in the course of a reaction. A special technique allowing estimation of time to fracture under constant loading of samples, controlled temperature and gas throughput rate has been developed for such experiments [8,66]. The same technique was also used in the experiments described with chemisorption. Fig. 21 illustrates a significant decrease in the mechanical strength, and particularly a dramatic one in the time to fracture, i.e. acceleration of wear and disintegration of magnesium oxide specimens in the course of dehydrogenation of propyl alcohol reaction. These effects provided the basis for the idea of a general phenomenon of the mutual, reciprocal influence of solid phase and the medium in catalytic processes. The new bonds arising between the solid phase and the medium components cause bonds weakening and rupture both in adsorbed molecules and in solid surface. This may be viewed as the manifestation of the Rehbinder effect, i.e. the decrease in surface energy due to adsorption (chemisorption), under the conditions of deviations from equilibrium and reversibility. There also are other reasons for various changes in the solid catalysts in catalysis. These include the so-called catalytic corrosion [68], reconstruction of surface morphology, phase transformation in the subsurface area, and others. However, besides these, the decrease in the surface mechanical stability and strength during catalysis can be treated as a universal and thermodynamically predictable effect. The catalyst's porous structure, whose strength is fully determined by the rather weak interparticle contacts, may collapse due to the influence of strongly active medium. In this respect, catalyst occurs a victim of its destination. However, the resistance of catalyst granules to wear can be substantially enhanced by improving the technology of catalyst production: by increasing the number of particle contacts with a selection of optimal size grading of primary particles forming granules, using finely dispersed inactive fillers, reducing residual internal

Fig. 21. Durabilitythe time to fracture, tf, (in logarithm scale) of a MgO catalyst under constant compressive stress Pc, at 395 C: 1in an inert medium, 2during catalytic reaction.

stresses, and, in particular, by strengthening the contacts themselves. The last can be achieved by substituting weak coagulation contacts existing in pressed powders with stronger and more stable phase contacts obtained by hydration hardening of corresponding constituents [65]. At the same time, the weakening of the bonds between the atoms at the surface of a catalyst leads to a sharp increase in the number of adatoms having a high mobility, and thus to the intensification of surface self-diffusion. Direct observations by the method involving healing of the scratch indicate an enormous (by 45 orders of magnitude) increase in the coefficient of surface self-diffusion. This effect has been observed with iron catalysts during the synthesis of ammonia and with nickel catalysts during hydrogenation of benzene [69,70]. The intensification of self-diffusion may, in turn, intensify the sintering process, making it possible at rather low temperatures. The sintering temperatures of iron and nickel powders in the process of CO oxidation may be reduced by 300 C for achieving the same mechanical strength of samples, due to development of strong contacts between the particles. This peculiar, secondary effect of porous structure strengthening (catalysis enhanced sinteringCES) caused by the primary, basic phenomenon of surface bonds weakening under the influence of medium adsorption, has been recently established also for some ceramic powders, such as alumina, zirconia, and yttria [70,71]. 7. Conclusion Numerous experimental studies show that the Rehbinder effect, i.e. the change in mechanical properties under the influence of adsorption from ambient medium is a very general physical chemical phenomenon. Without any exception, it takes place for all solids and materials with ionic, covalent, metal, molecular bond types, and, depending on a variety of factors, can be manifested at various levels, and in different forms (strength decrease, embrittlement, plasticization). These factors include the nature and intensity of stressed state, temperature, deformation rate, initial structure of defects in the solid, mobility of molecules

46

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47

(atoms) of the active medium, etc. However, the main factor of this effect is the weakening of cohesive forces in the surface layer of material, caused by both the adsorption of mono- and multilayer, and by the contact with the corresponding liquid phase. Within the scope of thermodynamic approach (the GibbsVolmer nucleation theory), these phenomena are characterized as a decrease in the surface (interfacial) energy, i.e. as physical chemical similarity (affinity) of contacting solid and liquid phases, and high value of the work of adhesion between them. This factor, in a mandatory combination with mechanical stresses, serves as a general precondition for the reversible physicalchemical effects of the environmentally sensitive mechanical behavior, in the absence of dissolution, corrosion, and disintegration in strongly aggressive media. A significant strength decrease occurs for ionic polar solids with hydrophilic surfaces due to the adsorption of polar substances, such as alcohols, and particularly water from the vapors, solutions in less polar media, as well as due to contact with water as a liquid phase. Some ionic salt melts may be strongly surface active with respect to the definite minerals. For nonpolar solids with hydrophobic surfaces such active medium components are the substances consisting of molecules with low or no polarity; a typical example is benzene with respect to naphthalene. For metals, with strong lowering of the surface energy in contact with a highly related (akin) liquid metal, catastrophic strength decrease and embrittlement can be observed. For the same metals and alloys, under the influence of organic surfactants, the plasticization phenomenon, i.e. softening of the surface layer, can take place hindering nucleation and propagation of cracks. The latter effect is the basis for using surfactants in lubricant formulations. The investigations of trends and atomic-molecular mechanisms of the phenomena of environment-sensitive mechanical behavior open ways for their practical applications and control. The latter include, on the one hand, preventing possible fracture, degradation, stability loss, surface damage, preventing or decelerating (in the case of catalysis) wear, and, on the other hand, oppositely, the useful application of these phenomena in technological processes for metal treatment by pressure, for the facilitation of comminuting, grinding, and cutting of the extremely hard solids, e.g. in the rock drilling and materials treatment. Acknowledgements The author wishes to pay tribute to his teacher and friend Peter A. Rehbinder, founder of this scientific direction, entitled by him as the PhysicalChemical Mechanics of Disperse Systems and Materials. The author is grateful to his colleagues and students who took part in carrying out experimental and theoretical studies described in this review. The author wishes to express sincerest gratitude to Dr. Kash Mittal who organized and chaired numerous SIS Symposia, at some of which the results of studies described here were presented. The author thanks Dr. Andrey Zelenev for valuable assistance in editing this manuscript.

References
[1] Rehbinder PA. On the influence of changes in surface energy on cleavage, hardness, and other properties of crystals. Proc. of the VI-th Congress of Russian Physicists, OGIZ, Moscow; 1928. p. 29. [2] Rehbinder PA. The hardness decrease due to surfactant adsorption. Sclerometry and physics of disperse systems. Z Phys 1931;72(34): 191202. [3] Rehbinder PA, Shchukin ED. Surface phenomena in solids during deformation and fracture processes. In: Davison SG, editor. Progress in Surface Science, vol. 3. Pergamon Press; 1972. p. 97188. Part 2. [4] Westwood ARC, Ahearn JS, Mills JJ. Developments in the theory and applications of chemomechanical effects. Colloids Surf 1971;2: 136. [5] Latanision RM, Pickens JR, editors. Atomistics of Fracture. New York: Plenum Press; 1983. [6] Shchukin ED, Pertsov NV, Osipov VI, Zlochevskaya RI, editors. Physicalchemical mechanics of natural disperse systems, Izd. Moscow: MGU Moscow State University; 1985. [7] Shchukin ED. Development of teaching of P.A. Rehbinder on surface phenomena in disperse systems. Trans (Izvestiya) Acad Sci USSR, Chem Sci 1990;10:242446. [8] Shchukin ED, Margolis LYa, Kontorovoch SI, Polukarova ZM. The influence of the medium on the mechanical properties of catalysts. Russ Chem Rev 1996;65:81323. [9] Kawagoe M, Doi Y, Fuwa N, Yasida T, Takata K. Effect of adsorbed water on the interfacial fracture between two layers of unsaturated polyester and glass. J Mater Sci 2001;36:51617. [10] Shchukin ED, Savenko VI, Kuchumova VM, Ivanova NI. Ambient media effects upon mechanical behavior and surface stability of minerals and rocks. Proc. of the 38th U.S. Rock Mechanics Symposium Rock Mechanics in the National Interest, Washington, DC (July 2001), separate reprint #1/405; 2001. p. 17. [11] Shchukin ED, Amelina EA. Surface modification and contact interaction of particles. J Dispers Sci Technol 2003;24:37795. [12] Pertsov NV. The physico-chemical influence of liquid phases on the rock fracture, in [6], 107117. [13] Pertsov NV, Traskin VYu. The Rehbinder effect in nature. In: Shchukin ED, editor. Advances in colloid chemistry and PhysicalChemical mechanics. Moscow: Nauka; 1992. p. 15565. [14] Pertsov AV. Spontaneous and mechanical dispersion and stability of obtained disperse systems, Dr. Sci. thesis, Moscow State University, Moscow, 1992. [15] Giese RF, van Oss CJ. Colloid and surface properties of clays and related materials. New York: Marcel Decker; 2001. [16] Shchukin ED, Amelina EA. Contact interactions in disperse systems. Adv Colloid Interface Sci 1979;11:23587. [17] Yaminsky VV, Pchelin VA, Amelina EA, Shchukin ED. Coagulation contacts in disperse systems. Moscow: Khimiya; 1982. [18] Shchukin ED, Yaminsky VV. Thermodynamic factors in the solgel transition, Colloids Surf 32 (1985) 1932, and 3355. [19] Shchukin ED. Some colloid-chemical aspects of the small particles contact interactions. In: Pelizzetti E, editor. Fine particles science and technology. Dortrecht, The Netherlands: Kluwer; 1996. p. 23953. [20] Shchukin ED, Pertsov AV, Amelina EA, Zelenev AS. Colloid and surface chemistry. Amsterdam: Elsevier; 2001. [21] Amelina EA, Shchukin ED, Parfenova AM, Pelekh VV, Vidensky IV, Bessonov AI, et al. Effect of cationic polyelectrolyte and surfactant on cohesion and friction in contacts between cellulose fibers. Colloids Surf A 2000;167:21527. [22] Shchukin ED. Surfactant effects on the cohesive strength of particle contacts: measurements by the Cohesive Force apparatus. J Colloid Interface Sci 2002;256:15967. [23] Derjaguin BV, Titijevskaia AS, Abrikosova II, Malkina AD. Investigations of the forces of interaction of surfaces in different media. Discuss Faraday Soc 1954;18:8598. [24] Tabor D, Wintertorn RHS. The direct measurements of normal and retarded van der Waals forces. Proc R Soc Lond A 1969;312:43550.

E.D. Shchukin / Advances in Colloid and Interface Science 123 126 (2006) 33 47 [25] Rabinovich YaI. Direct measurements of disjoining pressure in electrolyte solution as a function of distances between crossing filaments. Colloid J 1977;39:1094100. [26] Pashley R, Israelachvili J. A comparison of surface forces and interfacial properties of mica in purified surfactant solutions. Colloids Surf 1981;2:16987. [27] Churaev NV. Incorporation of structural forces in theory of stability of colloids and films. Colloid J 1984;46:30213. [28] Israelachvili JN. Intermolecular and surface forces. San Diego: Academic Press; 1985. [29] Derjaguin BV, Churaev NV, Muller VM. Surface forces. Moscow: Nauka; 1985. [30] Derjaguin BV. The theory of stability of colloids and thin films. New York: Plenum Press; 1989. [31] Claesson PM, Ederth T, Bergeron V, Rutland MW. Techniques for measuring surface forces. Adv Colloid Interface Sci 1996;67:11983; Christenson HK, Claesson PM. Direct measurements of the force between hydrophobic surfaces in water. Adv Colloid Interface Sci 1996;91: 391436. [32] Amel'yanets AA, Amelina EA, Pertsov NV, Shchukin ED, Yusupov RK. Apparatus for measurements of cohesive forces. The Certificates of Invention, The USSR State Committee on Affairs of Inventions and Discoveries, #767625 with priority on 14 July, 1979, and #877411 with priority on 19 December; 1978. [33] Amelina EA, Kumacheva EE, Chalykh AE, Shchukin ED. Structure of interfacial adsorbed layers in systems of aqueous solutions of block copolymers of ethylene and propylene oxides/fluorocarbon, Colloid J 58 (1996) 415419, and 420422. [34] Shchukin ED, Amelina EA, Parfenova AM. Influence of the nature of nonpolar phase on the mechanical stability of adsorption layers of hydrocarbon and fluorocarbon surfactants at the interface between their aqueous solutions and non-polar media. Colloids Surf 2001;A176:3551. [35] Fowkes FM. Additivity of intermolecular forces at interfaces: I. Determination of the contribution to surface and interfacial tensions of dispersion forces in various liquids. J Phys Chem 1963;67:253841. [36] Shchukin ED. Development of Rehbinder's doctrine on strong stabilization factors in disperse systems. Colloid J 1997;59:27084. [37] Izmailova VN, Yampolskaya GP, Tulovskaya ZD. Development of Rehbinder's concept on structure-mechanical barrier in stability of dispersions stabilized by proteins. Colloids Surf A 1999;160:89106. [38] Shchukin ED. The physicalchemical theory of strength of disperse structures and materials, in [6], 7290. [39] Shchukin ED, Amelina EA, Kontorovich SI. Formation of contacts between particles and development of internal stresses during hydration processes. In: Skalny J, editor. Materials Science of Concrete, vol. III. Westerville, OH: The American Chemical Society; 1992. p. 135. [40] Rybakova LM, Amelina EA, Kuksenova LI, Shchukin ED. Investigation of residual internal stresses of the I and II mode in cement-hardening structures. Colloids Surf A 1999;160:16370. [41] Korolev VA, Nikolaeva SK, Osipov VI. The rheology of clayey grounds, in [6], 222233. [42] Shchukin ED, Yaminskaya KV, Yaminsky VV. Interaction of particles and structuralmechanical properties of disperse systems. Dokl (Proc) Akad Nauk SSSR 1986;289:11869. [43] Churaev NV. Surface forces and the microrheology of concentrated dispersion systems. Colloid J 1988;50:10816. [44] Ur'ev NB, Potanin AA. Fluidity of suspensions and powders. Moscow: Khimiya; 1992. [45] Likhtman VI, Shchukin ED, Rehbinder PA. Physicochemical Mechanics of Metals, Izd. AN SSSR, Moscow, 1962; English translation: Israel Program for Scientific Translations, Jerusalem; 1964. [46] Rehbinder PA. Selected works: surface phenomena in disperse systems. Physicochemical mechanics. Moscow: Nauka; 1979. [47] Latanision RM, Fourie JF, editors. Surface effects in crystal plasticity, NATO Advanced Study Institutes Series. Series E: Applied Sciences. Leyden: Noordhoff; 1977. [48] Shchukin ED, Kochanova LA, Savenko VI. Surface and environment effects in the elasticplasticfracture transitions in metal crystals. In:

47

[49]

[50]

[51] [52] [53]

[54]

[55] [56] [57]

[58]

[59]

[60] [61]

[62]

[63] [64]

[65] [66]

[67]

[68] [69]

[70]

[71]

Lacombe P, editor. Physical chemistry of the solid state: applications to metals and their compounds. Amsterdam: Elsevier; 1984. Shchukin ED, Kochanova LA, Savenko VI. Electric surface effects in solid plasticity and strength. In: White RE, Conway BE, Bockris JO'M, editors. Modern Aspects of Electrochemistry, vol. 24. Plenum Press; 1993. p. 24598. Shchukin ED, Kochanova LA, Savenko VI. On the mechanism of environment-induced plasticizing under contact interactions. In: Latanision RM, Coutel RJ, editors. Advances in the mechanics and physics of surfaces. London: Harwood Academic Publishers; 1981. Rostoker W, McCaughey JM, Markus H. Embrittlement by liquid metals. New York: Reinhold; 1960. Kamdar MH, editor. Embrittlement by liquid and solid metals. Warrendale, PA: AIME; 1984. Shchukin ED, Kochanova LA, Pertsov AV. On the temperature transition from brittleness to plasticity under conditions of the adsorption effect of strength decrease. Crystallography 1963;8:6974. Maksymovych HH, Fedirko VM, Pavlyna VS. High temperature interaction of structural materials with ambient media. Mater Sci (PhysChem Mech Mater) 1996;32:2725. Shchukin ED. Environmentally induced lowering of surface energy and the mechanical behavior of solids, in [47], 701736. Shchukin ED, Yushchenko VS. Molecular dynamics simulation of mechanical behaviour. J Mater Sci 1981;16:31330. Yushchenko VS, Ponomareva TP, Shchukin ED. Environmental influence on the mechanical strength of chemical bonds in solidsab initio quantum calculations. J Mater Sci 1992;27:165962. Yushchenko VS, Edholm O, Shchukin ED. Molecular dynamics of the particle/substrate contact rupture in a liquid medium. Colloids Surf 1996;110:6373. Vidensky IV, Petrova IV, Shchukin ED. Electro-chemo-mechanical treatment: Facilitating steel grinding under electrochemical reduction of active cations. J Mater Res 1993;8:22247. Shchukin ED. Physicalchemical mechanics in the studies of Peter A. Rehbinder and his school. Colloids Surf A 1999;149:52937. Vidensky IV, Shchukin ED, Savenko VI, Petrova IV, Polukarova ZM. Liquid metal embrittlement in the absence of liquid metal phase: in studies of surface damageability and in hard materials machining. Colloids Surf A 1999;156:34955. Kochanova LA, Kuchumova VM, Savenko VI, Shchukin ED. Environmental effect on sclerometric brittleness of ionic crystals. J Mater Sci 1992;27:551622. Savenko VI, Shchukin ED. New applications of the Rehbinder effect in tribology: a review. Wear 1996;194:8694. Kontorovich SI. The colloid-chemical trends in strengthening of the highly disperse porous structurescatalysts and adsorbents, Dr. Sci. thesis, Inst. for Phys. Chem. of the USSR Acad. Sci., Moscow; 1989. Shchukin ED. Reciprocal influence of the solid phase and medium in processes of heterogeneous catalysis. Chem Ind 1997;6:4129. Shchukin ED, Bessonov AI, Kontorovich SI, Polukarova ZM, Savenko VI, Sokolova LN, et al. Physicalchemical mechanics of catalysts: strength and durability of the fine-porous materials in active media. Mater Sci (Phys-Chem Mech Mater) 2003;39(3):2843. Zhurkov SN, Sanfirova TP. Temperaturetime relation of the tensile strength of pure metals. Dokl (Proc) Akad Nauk SSSR 1955;101: 23740. Roginskii SZ, Tret'yakov II, Shekhter AB. Dokl (Proc) Akad Nauk SSSR 1953;91:8815. Shchukin ED, Kontorovich SI, Romanovsky BV. Porous materials sintering under conditions of catalytic reaction. J Mater Sci 1993; 28:193740. Romanovsky BV, Shchukin ED, Burenkova LN, Sokolova LN. The influence of catalytic reactions on the strength of porous materials with globular structure. Russ J Phys Chem 2002;76:9324. Abukais A, Burenkova LN, Zhilinskaya EA, Lamonier J-F, Murav'eva GP, Romanovsky BV, et al. Effect of catalytic alcohol conversion on the mechanical strength of porous ZrO2 and ZrO2 + Y2O3 materials. Inorg Mater 2003;39:5038.

You might also like