Molecular mechanisms associated with adhesion and contact angle

Szu-Ying Chen , Yair Kaufman , Alex M. Schrader , Dongjin Seo , Dong Woog Lee , Steven H. Page , Peter H. Koenig , Sandra Isaacs , Yonas Gizaw , and J...
0 downloads 0 Views 2MB Size
10736

J. Phys. Chem. 1991, 95, 10736-10747

of this type of bond, a decrease in the wavenumber of the S-H stretching vibration, an increase of the half-width of the band, and an increase of the integral intensity were observed. However, the shift of the band to lower wavenumbers (AvSH20-30 cm-I) relative to the band of the S-H stretching vibration of the mercaptan molecule adsorbed on SiOH groups suggests a very weak interaction through the S-H group.

Conclusions Ethanethiol and butanethiol act as electron pair donor molecules upon adsorption on Si02, H-ZSMS, and Na-ZSMS. Judged from the perturbation caused by the adsorption, the thiol molecules act as considerably weaker EPD than alcohols, but, except for part of the butanethiol molecules, they did not act as Bronsted acid donating protons to the surface. In contrast to alcohols, the larger sulfur seems to stabilize noncyclic hydrogen bonding at hydroxyl groups. The form of adsorption of the thiols on alkali-metal cation

exchanged ZSMS (Le., via the lone electron pair of the sulfur atom) was similar to that previously reported for alcohols. On H-ZSMS three forms of adsorption were determined: (i) linear (asymmetric) hydrogen bonding to the SiOHAl site, (ii) cyclic hydrogen bonding to the SiOHAl site, and (iii) protonization at the SiOHAl site. At the higher equilibrium pressure (1 mbar) clusters of thiol molecules were observed which were held together by weak hydrogen bonding. The unusual strong hydrogen bonding in these clusters, as seen with protonated water14J5and alcohol,*6 was not observed. Acknowledgment. This work was supparted by the Fonds zur Forderung der Wissenschaftlichen Forschung under project P 73 12 CHE. C.L.G. acknowledges partial support from CONICET (Argentina). We are grateful to Mobil Oil Corp. for supplying us with the H-ZSMS sample. Registry No. HSEt, 75-08-1; HSBu, 109-79-5.

Molecular Mechanisms Associated with Adhesion and Contact Angle Hysteresis of Monolayer Surfaces Y. L. Chen, C. A. Helm,+ and J. N. Israelachvili* Department of Chemical and Nuclear Engineering, and Materials Department, University of California, Santa Barbara, California 93106 (Received: April 5, 1991)

Experiments were carried out on a variety of surfactant-coated mica surfaces using the surface forces apparatus technique and contact angle measurements. The experiments were designed to clarify the molecular mechanisms underlying adhesion hysteresis (during loading-unloading cycles) and contact angle hysteresis (of advancing/receding liquids), and to explore any possible relationship between these two energy-dissipating phenomena. We found that hysteresis effects are not simply due to surface imperfections,such as roughness or chemical heterogeneity. Even surfaces that are initially smooth and chemically homogeneous can exhibit large adhesion and contact angle hysteresis effects. Our results indicate that, for such surfaces, hysteresis arises because of molecular rearrangements occurring at solidsolid or solid-liquid interfaces afer they have come into contact. This results in a lower surface free energy during the approach of two surfaces (or during spreading) than during separation (or retraction). We have studied a number of factors that enhance hysteresis: (i) increasing the freedom of the surface molecules to reorder, (ii) increasing the load and time surfaces are allowed to remain in contact, and (iii) increasing the rate of separation (or retraction). These findings highlight the inherent nonequilibrium nature of most loading-unloading and wetting-dewetting cycles and suggest ways for reducing the energy-dissipating hysteresis associated with such processes. Our results further indicate that the adhesion or pull-off force F between two curved surfaces of radius R is related to the surface energy y by the Johnson-Kendall-Roberts theory, for example, F = 3uRy for a sphere on a flat surface, but only when the separation occurs under equilibrium conditions. Preliminary results also indicate a correlation between adhesion hysteresis and friction/stiction.

Introduction Adhesion and Wetting (Contact Angle) Hysteresis. Most real processes involving adhesion and wetting are hysteretic or energy-dissipating even though they are usually described in terms of (ideally) reversible thermodynamic functions such adhesion free energy, reversible work of adhesion, surface tension, interfacial tension, etc. [Throughout the text the following convention is adopted:” the surface energy per unit area of the solid-vapor or liquid-vapor interface is denoted by ys or yL,which by definition is half the adhesion energy or cohesion energy, W, this beiig the (positive) work done on separating two unit areas of the surfaces from contact in vapor. For two different surfaces in contact, such as a solid-liquid or a solid l s o l i d 2 interface, the interfacial energy is defined by y I 2= y I + y2- W, where the adhesion energy between the two surfaces is related to the surface energies of the two media by the approximate relation: W = In cases where subscripts A and R are used, e.g., WA, these refer to advancing (loading) and receding (unloading) values, respectively.] For example, the energy change, or work done, on

’ Permanent address: Institut fur Physikalische Chemie, Johannes Gutenberg Universitat, Jakob-Welder-Weg 1 I , D-6500,Mainz, Germany. 0022-3654/91/2095-lO736$02.50/0

separating two surfaces from adhesive contact is generally not fully recoverable by bringing the two surfaces back into contact again. This may be referred to as adhesion hysteresis and expressed as wR

receding (separating)

>

wA

advancing

(approaching)

or A W = (WR- WA) > 0

(1)

where WRand WAare the adhesion energies for receding (separating) and advancing (approaching) two solid surfaces, respectively. Adhesion hysteresis is responsible for such phenomena as “rolling” friction and “elastoplastic” adhesive contacts2 during loading-unloading and adhesion-decohesion cycles. Hysteresis effects are also commonly observed in wetting/ dewetting phenomena (Figure l).3 For example, when a liquid (1) Bowden, F. P.; Tabor, D. Friction and Lubricution; Methuen: London, 1967. (2) Greenwood, J. A.; Johnson, K. L. Philos. Mug. A 1981, 43, 697. Michel, F.; Shanahan, M. E. R. C. R.Acad. Sci. Paris 1990, 310 (11). 17. Maugis, D. J . Mufer. Sci. 1985, 20, 3041.

0 1991 American Chemical Society

Adhesion on Monolayer Surfaces

The Journal of Physical Chemistry, Vol. 95, No. 26, 1991 10737

Hacroscopic jumps

Holecular jumps (peeling)

Approach

Figure 1. Examples of wetting and contact angle hysteresis. (A) Solid surface in equilibrium with the vapor phase. On wetting the surface, a contact angle e, is observed, on dewetting it decreases to eR. (B) A liquid droplet resting on a flat solid surface cannot be described as a true equilibrium situation: at the three-phase contact boundary the vertical component of the liquid stress, yL sin 8, is balanced by high local stresses on the solid which induce elastic or plastic deformations (inset) and/or chemical rearrangements to relax these stresses. (C) and (D) Contact angle hysteresis is usually explained by the inherent roughness (left side) and/or chemical heterogeneity (right side) of surfaces. (E) Interdiffusion and interdigitation at an interface may induce roughness and chemical heterogeneity even though initially both surfaces are perfectly smooth and homogeneous. Such surface restructuring effects can occur both on the macroscopic, microscopic, and molecular scales. Here we will mainly be concerned with effects occurring at the molecular level.

spreads and then retracts from a surface the advancing contact angle 8 A is generally larger than the receding angle 8 R (Figure 1A). Since the contact angle, 8, is related to the liquid-vapor surface tension, yL,and the solid-liquid adhesion energy, W,by the DuprE equation (Figure 1B) (1 + cos =w (2) we may conclude that wetting hysteresis or contact angle hysteresis (8, > 8,) implies that either yL:A > y L , R or that W R > W A . Since there is no reason or any previous evidence for a liquid surface having different surface tensions for increasing and decreasing areas (so long as the liquid is pure), we may conclude that wetting or contact angle hysteresis actually implies adhesion hysteresis, viz. eq 1. In all the above cases at least one of the surfaces is always a solid. In the case of solidsolid contacts, the hysteresis has generally been attributed to viscoelastic bulk deformations of the contacting materials or to plastic deformations of locally contacting asperities.’I2 In the case of solid-liquid contacts, hysteresis has usually been attributed to surface roughness or to chemical heterogeneity (Figure 1C,D),3though there have been reports of significant hysteresis on molecularly smooth chemically homogeneous surface^.^ This paper focuses on two other, less studied but possibly more important, mechanisms that can give rise to hysteresis. These may be conveniently referred to as (i) mechanical hysteresis, arising from intrinsic mechanical irreversibility of many adhesion/decohesion processes (Figure lB, inset; and Figure 2), and (ii) chemical hysteresis, arising from the intrinsic chemical irreversibility at the surfaces associated with rearrangements of (3) Miller, C. A.; Neogi, P. Interfacial Phenomena; Marcel Dekker: Basel, 1985. (4) Schwartz, A. M. J. Colloid Interface Sci. 1980, 75, 404.

I

F i m e 2. Origin of mechanical adhesion hysteresis during 1 le approach

and separation of two solid surfaces. Top: In all realistic situations the force between two solid surfaces is never measured at the surfaces themselves, at S, but at some other point, say S’, to which the force is elastically transmitted via the backing material supporting the surfaces. Center (left): Macroscopic “magnet” analogy of two approaching surfaces, where the lower is fixed and where the other is supported at the end of a spring of stiffness Ks. Bottom: Force-distance curve for two attracting surfaces (or magnets), showing the path taken by the upper surface on approach and separation. On approach, an instability occurs at D = DA,where the surfaces spontaneously jump into “contact” at D = D,,. On separation, another instability occurs where the surfacesjump apart from -Do to DR. Center (right): On a molecular level, the separation of two surfaces is accompanied by the spontaneous popping of molecular bonds, which is analogous to the jump apart of two macroscopic surfaces.

chemical groups occurring during the necessarily finite time it takes to go through any adhesion/decohesion or wetting/dewetting process (Figure 1E). Henceforth we shall use the term approach-separation to refer quite generally to any cyclic process, such as adhesion-decohesion, loading-unloading, advancing-receding, and wetting-dewetting cycles. As will be argued below, because of natural constraints offinite time and the finite elasticity (or compliance) of materials most approachseparation cycles are thermodynamically irreversible, and therefore energy dissipating. By thermodynamic irreversibility we simply mean that over any reasonable time period one cannot go through the approachseparation cycle via a continuous series of equilibrium states because some of these are connected via spontaneous-and therefore thermodynamically irreversibletransitions. During such transitions there is an absence of mechanical and/or chemical equilibrium. In many cases the two will be intimately related and occur at the same time (and there may also be an absence of thermal equilibrium), but the above distinction is nevertheless a useful one since there appears to be two fairly distinct molecular processes that give rise to them. These will now be considered in turn. Mechanical Hysteresis. Consider two solid surfaces a distance D apart (Figure 2) interacting with each other via an attractive potential with a hard-wall repulsion at some cutoff separation, Do. Let the materials supporting the surfaces have a bulk elastic modulus K (in units of N m-2) so that depending on the system geometry the surfaces may be considered to be supported by a simple spring of effective “spring constant” Ks (in units of N m-1).5 (5) Pethica, J. B.; Sutton, A. B. J. Vac. Sci. Technol. A 1980,6 (4), 2490. Landman, U.; Luedtke, W. D.; Burnham, N. A.; Colton, R. J. Science 1990, 248,454.

10738 The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

When the surfaces are brought toward each other, a mechanical instability occurs at some finite separation, D,, from which the two surfaces will jump into contact. This instability occurs when the gradient of the attractive force, dF/dD(also in units of N m-I), exceeds KS.’ Likewise, on separation from adhesive contact, there will be a spontaneous jump apart from Do to DR (Figure 2). Separation jumps are generally greater than approaching jumps. Such spontaneous jumps occur at both the macroscopic and atomic levels. For example, they occur when two macroscopic (I? = 1 cm) surfaces are brought together in surface forces experiments;6 they occur when STM or AFM tips ( R = 10 nm) and they occur when individual bonds approach a flat are broken (or ”popped”) during fracture and crack propagation in solids.* But such mechanical instabilities will not occur if the attractive forces are weak or if the backing material supporting the surfaces is very rigid (high K ) . However, in many practical cases these conditions are not met and the adhesion-decohesion cycle is inherently hysteretic regardless of how smooth the surfaces or how perfectly elastic the materials. Due to the spontaneous jumps in and out of contact, the process cannot be reversible in a thermodynamic sense, and the adhesion (free) energy needed to separate the surfaces from contact must be greater than that on approach. It is important to note that this irreversibility does not imply that the surfaces become damaged or changed in any way, or that the molecular configuration is different at the end from what it is at the beginning of the cycle. Energy is always dissipated in the form of heat when two surfaces or molecules impact each other. Before continuing with other types of hysteresis, it is worth mentioning what the true equilibrium situation is in the presence of mechanical instabilities. Referring to Figure 2, if the two surfaces are brought to some arbitrary separation, D’, between D, and DR, and left to equilibrate at a finite temperature for an infinite length of time, the equilibrium separation will actually be a bimodal Boltzmann distribution of distances, peaking at D = D’(the separated state) and D = Do (the contact state). Every now and then a spontaneous thermally induced fluctuation will occur, taking the surfaces from the separated to the contacting state, or vice versa. Over a long enough period of time the surfaces will have moved back and forth, sampling the whole of “ergodic” space. The resulting distribution defines the true thermodynamic equilibrium “state” of the system. It is also worth noting that (i) there is no one unique thermodynamically equilibrium separation, and (ii) if D’lies between D, and DR,the time-averaged distribution will center around two discrete separations D’and Do which is entirely analogous to a two-phase (e.g., solid-gas) system even though only two surfaces, or particles, are involved. Chemical Hysteresis. When two surfaces come into contact, the molecules at the interfaces relax and/or rearrange to a new equilibrium configuration that is different from that when the surfaces were isolated (Figure 1E). These rearrangements may involve simple positional and orientational changes of the surface molecules, as occurs when the molecules of two homopolymer surfaces slowly intermix by interdiffusion9 and reptation.I0 In more complex situations, new molecular groups that were previously buried below the surfaces may appear and intermix at the interface. This commonly occurs with surfaces whose molecules have both polar and nonpolar groups, for example, copolymer surfaces,” surfactant s ~ r f a c e s , ~and * ~ ’protein ~ surfaces.14 All (6) Horn, R.; Israelachvili, J. N. Chem. Phys. Letf. 1980, 71, 192. (7) Weisenhorn, A. L.; Hansma, P. K.; Albrecht, T. R.; Quate, C. F. Appl. Phys. Lett. 1989, 54, 26. Hansma, P. K.; Elings, V . B.; Marti, 0.;Bracker, C. E. Science 1988, 243, 1586. (8) Lawn, B. R.; Wilshaw, T. R. Fracture of Brittle Solids; Cambridge University Press: London, 1975. Sahimi, M.; Goddard, J . G. Phys. Reo. B 1986, 33, 7848. (9) E M , M. D.; Gent, A. N. J . Polym. Sci.: Polym. Phys. 1984.22, 1953; 1985, 23, 1823. Shanahan, M. E. R.; Schreck, P.;Schultz, J. C.R . Acad. Sei. Paris 1988, 306 (II), 1325. Okawa, A. BSc. Thesis, Department of Materials Science, University of Utah, June 1983. Holly, F. J.; Refojo, M. J. J . Biomed. Mater. Res. 1975, 9, 3 15. (IO) Klein, J. J . Chem. Soc., Faraday Trans. I 1983, 7 9 , 99; Makromol. Chem. Macromol. Svmp. 1986, I , 125.

Chen et al. these effects act to enhance the adhesion or cohesion of the contacting surfaces. What distinguishes chemical hysteresis from mechanical hysteresis is that during chemical hysteresis the chemical groups at the surfaces are different on separation from on approach. However, as with mechanical hysteresis, if the cycle were to be carried out infinitely slowly it should be reversible. Previous Work. A number of have studied how molecular reorientations at surfaces correlate with the contact angle hysteresis of water on a polymer surface and showed that the increased watersurface adhesion energy on receding is related to the exposure of polar groups following the initial spreading (advance) of the liquid. Qualitatively similar findings were reported by Chen and co-workers13 on the adhesion of two surfactant-coated mica surfaces. The surfaces were initially hydrophobic on approach (high 6) but had become significantly more hydrophilic on separation (low 0) due to the reorientation of the surfactant molecules (flip-flop). As might be expected, these effects depended on the contact time and the presence of water (relative humidity) as indeed are most adhesion processes. Present Approach. While interdigitation effects appear to be well-known in polymer and protein science, to our knowledge the adhesion has not been studied systematically as a function of contact time, contact pressure, separation speed, polymer type, temperature, etc. [In this paper terms such as “interdigitation”, “interdiffusion”, “interpenetration”, etc., are used somewhat loosely to cover any molecular restructuring occurring at or across an interface (see, for example, Figure 1E) but without reference to any one particular mechanism. This may involve one or more of the following (i) diffusive interpenetration of chain segments via reptation or Rouse motion, (ii) reorientations of bipolar molecular groups at surfaces without involving any translational displacement, and (iii) exchange of different molecular species with others that were previously buried in the bulk.] In our studies we have chosen to study these effects using molecularly smooth mica surfaces onto which well-characterized surfactant monolayers were adsorbed, either by self-assembly or by the Langmuir-Blodgett technique. Different types of surfactants and deposition techniques were used to provide monolayers with a wide variety of properties such as surface coverage, surface roughness, phase state, etc. The surface forces apparatus technique was used for measuring adhesion forces, adhesion energies (by two independent methods), and the geometry of adhering surfaces. We also investigated factors which favor hysteresis. These include the length of time two surfaces were in contact before they were separated, the temperature and load, the rates of loading and unloading, the surface coverage of the deposited monolayers (or chains), the effects of different headgroups, etc. Quantitative studies of all these effects provided a fairly consistent picture of the mechanisms underlying adhesion hysteresis. Complementary contact angle measurements were conducted to correlate the adhesion data with 6. Organic molecules are known to have a high affinity for hydrocarbon surfaces (with small contact angles, 6). Low molecular weight hydrocarbons (below dodecane) are also known to penetrate into the hydrocarbon regions of surfactant monolayers and lipid bilayers,I5 where they alter the packing of the molecules and the structure of the chains. It was therefore interesting to see how such penetration might affect the adhesion properties of monolayers exposed to organic vapors. If the separation processes were all occurring at thermodynamic equilibrium, then the values expected for the surface energies y should be the same on approach and separation, and correspond ( 1 1) Yasuda, H.; Sharma, A. K.; Yasuda, T. J . Polym. Sci., Polym. Phys. 1981, 19, 1285. (12) Langmuir, I. Science 1938,87,493. Wasserman, S . R.; Tao, Y.-T.; Whitesides, G . M.h n g m u i r 1985, 5 , 1074. (13) Chen, Y . L. E.; Gee, M. L.; Helm, C. A,; Israelachvili, J. N.; McGuiggan, P. M. J . Phys. Chem. 1989, 93, 7057. (14) Andrade, J . D.; Smith, L. M.; Gregonis, D. E. In Surface and Interfacial Aspects of Biomedical Polymers; Andrade, J. D., Ed.: Plenum: New York, 1985; Vol. 1, p 249. (15) Gruen, D. W. R.; Haydon, D. A. Biophys. J . 1981, 33, 167.

The Journal of Physical Chemistry, Vol. 95, No. 26, 1991 10739

Adhesion on Monolayer Surfaces to the well-known literature values for hydrocarbon surfaces, obtained from surface tension, critical wetting and contact angle measurements.16 These vary from about y = 23 mJ/m2 for surfaces composed mainly of CH3 groups, to about y = 31 mJ/m2 for surfaces composed mainly of CH2groups. Note, however, that the phase state of hydrocarbon chains has little effect on y, as can be ascertained from the similar values for liquid hexadecane (27 mJ/m2) and solid paraffin wax (25-30 mJ/m2) at the same temperature of 25 O C . 1 6 Thus, we would expect the equilibrium values of y to fall within the range 23-31 mJ/m2.

Theoretical Background Here we briefly describe the two standard continuum theories underlying wetting and adhesion, starting with the former. The theoretical background for the contact angles of a liquid on a solid surface (Figure 1B) is based on a straightforward balance of forces or minimization of surface energies” and leads to the Dupre equation, eq 2, or the Young-Dupre equation: 7s = YSL + Y L cos @

(3)

Note, however, that the normal component of the tension, yL sin 0 in Figure l B , must be balanced by high local stresses on the solid at the three-phase boundary, a matter that is not straightforward to resolve within a continuum framework. Turning now to adhesion, modern theories of the adhesion mechanics of two contacting solid surfaces are based on the theory of Johnson, Kendall, and Roberts (known as the JKR theory).’*Jg In the JKR theory two spheres of radii R I and R2,bulk elastic moduli K,and surface energy y per unit area will flatten when in contact. The contact area will increase under an external load or force, F, such that at mechanical equilibrium the contact radius a is given by a3 = ![F

K

+ 6 r R y + (127rRyF + ( 6 ~ R y ) ~ ) ~ (’ 4~) l

where R = R , R 2 / ( R l+ R 2 ) . Note that R remains unchanged for a single sphere of radius R ( R , = R ) on a flat surface (R2 = m), as shown in Figure 3A. Equation 4 is the basic equation of the JKR theory and leads to a number of important predictions. First, the deformation of the sphere, i.e., the flattening, increases on increase of the surface energy. One can demonstrateZothat, by increasing the adhesive energy y, the total energy of the system can fall in spite of a further deformation of the adhering sphere. Second, eq 4 shows that even under zero external force ( F = 0) there is a finite contact radius of a

= a. = ( 1 2 ~ R ~ y / K ) l / ~

(5)

In the absence of any adhesion (y = 0) one obtains a = 0 at F = 0 (“Hertzian” contact), as expected. Equation 4 further shows that two solids continue to adhere even under negative loads ( F < 0) until at some critical negative force-the adhesion force, Fs-the surfaces abruptly separate while the contact radius is still finite. For a sphere on a flat (Figure 3A) the contact radius on separation is given by US

= ( 6 r R 2 y / 2 K ) ‘ l 3= 0.63~0

(6)

and the adhesion force measured on separation is given by

Fs = -3rRys = -3/2aRW

(7)

(16) Zisman, W. A. Ind. Eng. Chem. 1963, 55, 19. Fowkes, F. M. Ind. Eng. Chem. 1964, 56, 40. Zisman, W. A.; Fox, J. J . Colloid Sci. 1952, 7,

428. (1 7) Israelachvili, J. N. Intermolecular and Surface Forces; Academic Press: New York, 1985. (18) Johnson, K. L.; Kendall, K.; Roberts, A. D. Proc. R. SOC.London A 1971, 324, 301. (19) Pollock, H. M.; Barquins, M.; Maugis, D. Appl. Phys. Lett. 1978, 33 (9), 798. Barquins, M.; Maugis, D. J . Mic. ThPor. Appl. 1982, I , 331. (20) Helm, C. A.; McGuiggan, P. M.; Israelachvili, J. N. Biochemistry, submitted for publication.

brnnch (YA)

Figure 3. (A) Reversible contact radius ( n ) vs load (F)curve of nonadhesive Hertzian contact and adhesive JKR contact under ideal conditions. No hysteresis. (B) Irreversible a-F curves and the hysteresis loops they give rise to during an advancing-receding cycle (also commonly referred to as loading-unloading, compression-decompression, and bonding-debonding cycles).

where, by definition, the reversible work of adhesion, W, is related to the surface energy, ys, of the solids by

w = 2ys

(8)

However, there is one unresolved issue concerning the relation between the adhesion force, Fs, and the surface energy, W. Other theories of adhesion mechanics predict different relations than that given by eq 7. For example, the so-called DMT theory2’ differs from the JKR theory in that it includes the effects of the attractive forces between the two surfaces just outside the contact zone, but it assumes that the elastically deformed shapes of the surfaces are not modified by these forces. This leads to a different relationship between the adhesion force and the surface energy, viz. Fs = -2rRW = -4rRys

(9)

which differs from eq 7 by 25%. Note that according to both JKR and DMT theory a finite elastic modulus, K,while having an effect on how the contact area varies with the externally applied force, has no effect on the adhesion force, Fs, an interesting and unexpected result that has nevertheless been verified experimentally to within about 25%.17-22-24 It was found that the JKR theory provides a better description of the deformed shapes of elastic bodies than the DMT theory.25 But the continuum JKR theory predicts infinitely high stresses at the contact boundary. This matter cannot be resolved within a continuum framework and requires a molecular analysis of the local forces operating at the bifurcation zone. Recent theoretical (21) Derjaguin, B. V.; Muller, V. M.; Toporov, Yu. J . Colloid Znterface Sci. 1975, 53, 314. Muller, V. M.; Yushchenko, V. S . ; Derjaguin, B. V. J . Colloid Interface Sci. 1980, 77, 91; 1983, 92, 92. (22) Christenson, H. K.; Claesson, P. J . Colloid Interface Sci. 1990, 139, 589. (23) Moy, E.; Neumann, A. W. J . Colloid Interface Sci. 1990, 139. 591. (24) Israelachvili, J. N.; Perez, E.; Tandon, R. K. J . Colloid Interface Sci. 1980, 78, 260. (25) Horn, R. G.;Israelachvili, J. N.; Pribac, F. J . Colloid Interface Sci. 1987, 115, 480.

10740 The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

analyses26 have concluded that the problem is complex and model-dependent and that the correct expression for the pull-off force probably lies between eqs 7 and 9. Unfortunately, this difference of 25%has not been able to be resolved experimentally in spite of a number of a t t e n ~ p t s . ~ ~We - * ~return to address this issue again later. As we shall see, eq 4 of the JKR theory offers a convenient way of testing the effects of surface conditions and time on adhesion energy hysteresis. It has been tested before both for adhering and nonadhering surfaces.26 For nonadhering surfaces, y = 0, and eq 4 reduces to the “Hertzian” limit: a3 = R F / K . The inverse cubic dependence of a on F was verified by Horn and co-workersZ5 for two molecularly smooth curved mica surfaces in aqueous salt solution, for which y = 0. Moreover, the measured contact radius was reversible (nonhysteretic) for increasing and decreasing loads. However, in the case of adhering mica surfaces in air (y > 0), there was a significant hysteresis in the curves of a versus F. This was attributed to viscoelastic effects in the glue supporting the mica sheets. Recently, Chaudhury and WhitesideP tested the JKR theory for contacting elastomeric surfaces (lenses) of poly(dimethylsi1oxane) (PDMS) in various liquids and vapors and obtained good agreement between values obtained for Wand those determined independently from contact angle measurements. They also measured large hysteresis effects with functionalized PDMS but not with unfunctionalized PDMS. We return to reconsider all these matters in the light of the new results presented here. Experimental Section Materials. Several types of surfactants were chosen for the present study. A single-chained positively charged surfactant hexadecyltrimethylammonium bromide (CTAB) was obtained from Sigma (St. Louis, MO) and recrystallized in methanol: acetone. A surfactant with a similar headgroup, but two chains, dihexadecyldimethylammoniumacetate (DHDAA) was bought from Sogo Pharmaceutical Co., LTD, Japan, as the bromide salt. The bromide ion was ion-exchanged by a ~ e t a t e . ~ ’ A calcium alkylbenzenesulfonate surfactant (CaABS) was supplied by Exxon Chemical Co. (Annandale, NJ). It is a randomly branched alkyl-chained surfactant with an average of 12 carbons per chain. Its headgroup is a calcium benzenesulfonate. The point of attachment of the alkyl chain to the ring is random, too. The CaABS was purified as previously described.28 The zwitterionic phospholipids L-a-dimyristoylphosphatidylethanolamine (DMPE) and L-a-dipalmitoylphosphatidylethanolamine (DPPE) were purchased from Avanti (Pelham, AL) and used without further purification. Octadecyltrichlorosilane (OTS) was obtained from Sigma (St. Louis, MO). Water was filtered in a Milli-Q+ unit. All salts and alkanes used were p.a. grade. Surface Preparation. Two molecularly smooth rectangular mica sheets, with sides 10 mm X 10 mm and thickness 1-3 pm,were cut, silvered, and glued with as little glue as possible onto the cylindrically curved surfaces of two silica disks. The glues used were the thermosetting epoxies: Shell Epon 1004 and 1007. A surfactant monolayer was then deposited or adsorbed onto each mica surface by one of the following techniques. I . Lungmuit-Blodgett deposition, in which insoluble doublechained surfactants such as DMPE or DPPE at the air-water interface were deposited at well-controlled areas per molecule onto the mica. The molecular areas given in the figures were determined by independent measurements of the transfer ratios.29 The CaABS was deposited from a 0.01 M CaC12 solution. 2. Adsorptionfrom solution, in which soluble surfactants such as CTAB or DHDAA were dissolved in dilute solutions below the cmc ( 5 X and 7 X M, respectively) and allowed to equilibrate for 1 day with the surrounding air environment to allow

-

(26) Hughes, B. D.; White, L. Q.J . Mech. Appl. Math. 1979, 32, 445. (27) Pashley, R. M.; McGuiggan, P. M.; Ninham, B. W. J . Phys. Chem. 1986, 90, 5841. ( 2 8 ) Gee, M. L.; Israelachvili, J . N. J . Chem. Soc., Faraday Trans. 1990, 86 (24). 4049. (29) Marra, J. J . Colloid Interface Sci. 1985, 107, 446; 1986, 109, 1 1 ,

Chen et al. Solid Crystalline (condensed)

1

DPPE 4 2 i 2 DMPE 43;‘

Solid Crystalline (expanded)

Amorphous

Fluid (liquid-like)

DMPE 53-67i2 DHDAA 75;‘ CTAB SUA2

CaABS 59;‘

Figure 4. Likely chain configurations for monolayers in the crystalline, amorphous, and fluid states (schematic). The first two phases are solid, the third is glassy or gellike, and the last is liquidlike.

for pH adjustment. Then, the mica sheets were immersed into the solution for 30 min, after which the surfaces were withdrawn and dried. The molecular areas of these monolayers were later determined by ESCAe30 OTS (4 X M) was dissolved in a solution of 8:12:80CHC13, CC14, and hexadecane. Mica sheets were immersed for 100 min in the solution and subsequently rinsed several times with CHC13, CCl,, and ethanol and later on with water. The surfactant-coated surfaces were then installed into the sealed chamber of the surface forces apparatus (SFA)31 and thoroughly dried by equilibrating the surfaces in a dry nitrogen atmosphere in contact with P205for 12 h. The temperature was controlled via the room temperature and was stable within 0.2 OC. Organic vapors were introduced by inserting a small beaker filled with liquid alkane into the SFA chamber. Each time the temperature and/or vapor pressure of the environment was changed the monolayer was allowed to equilibrate for 12 h before new measurements were taken. If transferred at high pressure, both DMPE and DPPE were in the crystalline condensed state, with the molecules either oriented vertically to the surface or slightly tilted (Figure 4).32 Another indication that DMPE in the frozen state forms a hard, solid surface comes from force measurements between two such surfaces across liquid hydrocarbons where oscillatory short-range forces are observed, indicative that the surfaces must be both rigid and atomically smooth.28 The glassy amorphous state was formed from saturated surfactants, both single and double chained, occupying an area greater than 22 AZper chain, but at a temperature below the chain melting temperature (Figure 4). Surfactants forming such monolayers were single-chained CTAB (at 60 AZper chain30),double-chained DHDAA (at 37 AZper chain30), and DMPE (at 27-34 A2 per chain, controlled by the deposition pressure3*). Molecules in the amorphous state are expected to be fairly rigid, but rough and disordered on the atomic scale. Fluid monolayers are those above the chain melting temperature. This was the case for the CaABS (at 29 A2 per hai in*^^^^) and most likely for CTAB above 25 OC (Le., above the Krafft temperature and melting point. of hexadecane). Adhesion Force Measurements. Using surface forces apparatuses (Mark I1 and 111),31 adhesion forces were measured from the “pull-off“ force, Fs, as previously described,25from which the N

(30) Chen, Y. L.; Chen, S.; Curtis, F.; Israelachvili, J . N., unpublished

results. ( 3 1 ) Israelachvili, J. N.; Adams, G. E. J . Chem. SOC.,Faraday Trans. I 1978, 74, 975. Israelachvili, J . N.; McGuiggan, P. M . J . Mater. Res. 1990, 5 (10). 2223-2231. (32) Fischer, A.; Sackmann, E. J . Phys. (Paris) 1984, 45, 517.

Tippmann-Krayer, P.; Steitz, R.; Mohwald, H., in press.

Adhesion on Monolayer Surfaces surface energy ys can be determined using eq 7 or 9. Note that, within the JKR theory, this value should be the same as the receding surface energy y R (Figure 3 ) which was also measured (independently) in these experiments. The measured adhesive forces had an error of about 1% (to eliminate errors from different spring calibrations, the same fixed double cantilever spring was used for all experiments). However, because the error in calibrating the radius, R , was about 10% the resulting errors in all our quoted values of y, W, etc., are also a bout 10%. As mentioned in the Introduction, an additional method was used to study both advancing (approaching) and receding (separating) adhesion energies. In this method the contact radius, a, was measured as a function of the externally applied force, F, both on increasing and decreasing F. Unless indicated otherwise, each point in the a3-F was measured at the end of a 1-min equilibration time (a full run takes about 20 min). On increasing F, the contact area increases as the surfaces are pressed together or "advance". By fitting the results to the JKR theory, eq 4, yA is obtained. On reducing the force, the area decreases, the surfaces peel apart or "recede", and yR is obtained. Under ideal (thermodynamically reversible) conditions y should be the same regardless of whether one is going up or down the JKR curve, as shown in Figure 3A. For some monolayers it was found that the advancing and receding branches were different (Figure 3B); however, each branch could usually be well fitted using eq 4 but with different values of y. Since the elastic modulus K is determined by the whole sandwiched system consisting of monolayer, mica, silver, and the glue, K was found to vary from experiment to experiment26over the range (2-7) X 1O'O N/m2. The optical technique associated with the SFA allows one to measure contact radii with a resolution of about 1 pm and forces with a resolution of about lo-' N. Since typical contact radii and forces were of the order of 20-40 pm and 10" N, this means that radii and forces were generally measurable with an error of less than *5%. All the results quoted are for fully dried monolayers. Exposing the monolayers to humid air had a drastic effect on the adhesion, in addition to loosening the headgroup binding to the surfaces, thus enhancing the lateral diffusion and flip-flop of the molecules within and acrms the monolayer^.'^ A full account of these "water penetration" effects will be reported in a later publication. Contact Angle Measurements. A sealed contact angle cell was built of glass, steel, and Teflon. A syringe needle, manipulated from the outside, could be made to approach different areas of a flat surface located at the center of the cell, and the plunger of the syringe could be moved in and out in a controlled way. In this way a number of droplets could be placed at different positions of the surface, and each could have its volume slowly increased or decreased with time. An external horizontal microscope with a protractor eyepiece was used to measure the contact angles of the advancing, receding, or stationary droplets. The atmosphere within the chamber could also be controlled. Directly after the monolayer deposition the mica sheet was placed into the contact angle cell. Then, the cell was purged with dry nitrogen for 1-2 h. It was found that this thorough purging increased the advancing contact angle by several degrees. Contact angle measurements were generally reproducible to within * O S 0 , so long as exactly the same procedure was repeated (including the previous history). First, the advancing contact angle was measured. Subsequently, the receding contact angle was measured at the three-phase line. For each measurement, a new 'virgin" part of the surface was used. The effects of different advancing/receding rates and waiting times between motions were also recorded. Results Figure 5 shows advancing and receding a3-F curves for various combinations of crystalline, glassy amorphous, and fluid monolayer surfaces in atmospheres of dry nitrogen gas, together with fits to the JKR theory, eq 4. The values for yA and yR obtained from

The Journal of Physical Chemistry, Vol. 95, No. 26, I991 10741

Figure 5. Measured advancing and receding a3-F curves at 25 O C for eight surface combinations. The solid lines are based on fitting the advancing and receding branches to the JKR theory, eq 4, from which the indicated values of yA and yR were determined, in units of mJ/m2 or erg/cm2. (In performing these fits, both y and K were determined to

within about 10%as deduced from the x2 values.-") Typical experimental values were R = 1-2 cm, K = 2-7 X lotoN/m2. Typical advancing/ receding rates were 1 pm/s from one position to the next, with a wait of 1 min at each place before a was recorded. Thus, the measurement of one branch took about 10 min. At the end of each a'-F measurement the pull-off force, Fs, and the monolayer thickness were again measured. If the thickness was unchanged and the pull-off force stayed constant to within 5%, the monolayer was considered stable.

these and similar experiments are also given in Tables I and 11, where they are also compared with the (independently measured) values of ys determined from the pull-off forces. Before considering the adhesion-decohesion hysteresis in detail, it is useful to mention some basic aspects of yA and yR. First, we note from Figure 5 and Table I that all the aduancing surface energies yA lie between 18 and 29 mJ/m2. This is what one expects for the equilibrium values of y which, as noted at the end of the Introduction, are in the range 23-3 1 mJ/m2. The slightly smaller values sometimes obtained for the amorphous/amorphous combinations may reflect some roughness on the molecular scale that effectively decreases their initial (advancing) contact area which would act to reduce the advancing adhesion energy. Second, the values of yR obtained from the a3-F curves should be the same as the values of ys measured from the pull-off forces, F R . Two equations were used to determine ysfrom F R : the JKR equation, eq 7, and the DMT equation, eq 9. From the results shown in Tables I and I1 it is clear that only the JKR theory adequately describes the receding adhesion of all the monolayer surfaces (except for some of the extremely hysteretic and timedependent CTAB monolayers). In view of the above, we conclude that the JKR theory, both as regards its prediction for the pull-off force ( F = 3xRy) and area-load profiles (a3-F curves), is essentially correct, and that previous indication^^*-^^ that the DMT theory, which predicts higher pull-off forces given by F = 4xRy, is correct were fortuitous and apply only to those systems where adhesion hysteresis enhances the receding surface energy.

10742 The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

Chen et al.

TABLE I: Average Surface Energies (in mJ/m*) of Various Surfactant-Coated Mica Surfaces at 25 OC (Unless Stated Otherwise) As Deduced from Loading-Unloading Cycles, Eq 4, and Pull-Off Forces, Eq 7"

phase states of contacting monolayers monolayer combinations

advancing, y,, receding, yR pull-off, ys

crystalline/crystalline, crystalline/amorphous DMPE(42)/DMPE(42) DPPE/CTAB DPPE/DHDAA 28 f 1 28 f 1 28 f 1

amorphous/amorphous coverage C40 A2/chain DMPE(53-67)/ DMPE(53-67) DMPE(62)/DHDAA DMPE(62)/CTAB DHDAA/CTAB 21 f 3 28 f 4 21 f 3

Pull-off forces were measured after a contact time of 30-60

'Overage

liquid/liquid CTAB/CTAB (35 "C) CaABS/CaABS

'40 A2/chain

DHDAA/ CTAB

CTAB/CTAB

28 46

25

25 f 3

72 f 20 61 i 20

44 f 3

(15 "C, 25 "C) 46

37 f 1

s.

$

1.4

2 w .-6

1.3

.c Iu

30

1.2

2b

1.1

w

2

1.o

5

0

10

20

15

2.0 >

F Figure 6. Schematics of chain interdigitations (see text) occurring after

two surfaces have been brought into contact. (A) Both surfaces in the solid crystalline state-no interdigitation. (B) Both surfaces in the amorphous or fluid states-interdigitation occurs slowly for two amorp hous surfaces and rapidly for two fluid surfaces. If the surfaces are separated sufficiently quickly, the effective molecular areas that are being separated from each other will be greater than the "apparent" area, and the receding adhesion will be greater than the advancing adhesion. Note that various surface restructuring effects can occur, e.g., interdiffusion, reorientation, and exchange, and that these effects can occur at either constant or decreasing or increasing film volume (thickness). In all experiments described here no surfactant was forced out of the contact zone during the loading-unloading cycles, as was ascertained from the identical film thicknesses measured, to within f0.5 A, at the start and end of each cycle (this should not be. confused with the reversible change in film thicknesses during loading and unloading due to the finite elastic compressibility of the films40). (C) One surface solid crystalline, the other amorphous or fluid-no interdigitation. We now proceed with a more detailed account of our results, starting with the effect of monolayer phase state on adhesion. Figure 5 shows that there is no detectable hysteresis so long as at least one of the monolayers is in the solid crystalline state (cf. Figure 5A,B,D) and that maximum hysteresis occurs for the amorphous/amorphous combination, CTAB/CTAB (Figure 5F). This indicates that Y~ = yA for two crystalline surfaces or for a crystalline with an amorphous surface, but that Y~ > for two amorphous or fluid surfaces. This suggests that chain interdiffusion or some other rearrangement occurs after two amorphous or fluid surfaces are brought into contact which enhances their adhesion during separation. The observation that a crystalline and an amorphous surface do not exhibit hysteresis is consistent with this scenario since only one surface needs to be frozen to prevent interdigitation from occurring with the other. All this is illustrated schematically in Figure 6. With this scenario in mind, we present further results which provide more insights into the factors that favor hysteresis. These include the length of time two surfaces are in contact before they are separated, the applied load, the speed of loading and unloading, the surface coverage of the molecules, and other factors. Effects of Contact Time on tbe Adhesion. The adhesion energy as determined from the pull-off force generally increased with contact time for almost all the monolayers studied (Figure 7A). The effect was most pronounced for amorphous monolayers at

B

- CTAB

2

1.8-

L

u

-

15% 25OC 4'

35'C I

0.8'

I

'

I

1

1

I

I

I

I

-

Monolayer Surfaces

The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

5-

A

-

CTAB

T 4 -

-5

*

3-

Do'

2

SLOWLOOP

-

1-

:i/

10743

-

7

"*- 2

1

-20

/I -10

I

I

I

I

0 10 20 Applied Load, F (mN)

I

I

30

I

I 40

Figure 8. Effect of advancing/receding rates on a'-F curves. For the

"fast" loops each point was measured after 1 min equilibration time; for the "slow" loops a 5-min wait was allowed. Solid lines are theoretical fits to the JKR theory, eq 4. Dashed lines are simply drawn through the experimental points to serve as a guide to the eye. (A) Two CTAB covered surfaces at 25 OC exhibit a pronounced hysteresis. By fitting the data points to eq 4 the following sets of parameters were obtained: fast loop on loading (yA = 20 mJ/m2, K A = 2.6 X 1O'O N/m2) and unloading (yR= 50 mJ/m2, KR = 3.2 X 1O'O N/m2); slow loop on loading (yA = 23.7 mJ/m2, KA = 2.7 X lotoN/m2) and unloading (yR = 44.5 mJ/m2, KR = 3.2 X 1O'O N/m2). At the end of each cycle the surface energy on separation, ys, was determined via eq 7 to be 42 mJ/m2 (fast loop) and 53 mJ/m2 (slow loop), respectively. (B) One surface is covered with a solid crystallinemonolayer of DPPE, the other with an amorphous layer of DHDAA. There was no detectable hysteresis. A fit of the data to eq 4 gave for the fast loop (yA = y R = 29 mJ/m2, K A =: KR = 2.9 X 10'O N/m2), and for the slow loop (yA = YR = 27 mJ/m2, KA = KR = 2.75 X 10'' N/m2). The pull-off forces were 26.7 and 27.1 mJ/m2, respectively, which is the same to within 2%, as expected. unloading behavior of an adhesive junction. In contrast to two CTAB monolayers, we find no hysteresis when one of the surfaces is in the solid crystalline state, on both slow and fast cycling. This is shown in Figure 8B where the fast and slow loops correspond to the same system with the same adhesion energy y, but at two different K values, and where no hysteresis was observed in either case. In an earlier series of measurements of hysteretic a3-F curves for two adhering untreated mica surfaces exposed to vapor, Horn et al.25suggested that the hysteresis was due to plastic deformation in the glue (though no hysteresis was observed between nonadhering surfaces immersed in aqueous salt solution). The present results suggest that the hysteresis observed in those experiments was more likely due to the presence of a molecularly thin organic layer on the surfaces (which is known to adsorb onto mica from ambient air3'). This would also be consistent with one of the implications of our findings, viz., that molecular restructuring effects occurring within a surface layer as thin as a few Hngstroms are enough to yield significant changes in the surface energy and thus to large hysteresis effects. In the present experiments it was also ascertained that the type of glue used to support the mica sheets was not responsible for the observed hysteresis: no hysteresis was observed between solid crystalline monolayer surfaces supported by either the softer polymer glue (type 1004 of Tg 100 "C) or the harder glue (type 1007 of higher T J . Effects of Applied Load on B 3-FCwes. Figure 9A shows that increasing the applied load leads to an increased y R and an increased hysteresis in the n3-F curves of D H D A A monolayers. This is what one would expect if the degree of interdiffusion of the two monolayers is enhanced by forcing them together more-a reasonable supposition. To analyze the effect of the load on the surface energy further

-

Chen et al.

10744 The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

0.1

!IO

0

10

F ("1

20

30

40

-5

APPLIED LOAD

0

5 F (")

10

15

Figure 10. a'-F curves for DMPE monolayers at different deposited

areas per molecule. There was no hysteresis only for monolayers in the solid crystalline state (molecular area 43 A*). Figure 9. (A) Effect of maximum applied load an the a'-F curves for

-

two DHDAA surfaces at 25 OC with an equilibration time of 1 min between points. At the start of loading-unloading cycles 1 and 2 3 the surfaces were brought into contact under zero load, which lead to spontaneous adhesive flattening. In cycle 1, the surfaces were not subjected to any external load before they were separated, and the receding branch yielded yR 15-20 mJ/m2 and ys = 27 mJ/m2. In cycle 2 3 the load was first increased to about 20 mN before the surfaces were receded. The measured values were yA = 21 mJ/m2, yR = 30 mJ/m2, ys = 27 mJ,"*. (B) Stress distribution of adhesive surfaces under a compressive load according JKR theory. The compressive stress is maximum at the center of the contact zone and decreases to negative infinity at the edges. Solid and dashed lines are as defined in Figure 8.

-

-

it is helpful to consider the pressure distribution within the contact areal 9,25

which is shown schematically in Figure 9B, where r is the distance from the center. According to eq 10, the local pressure is highest (compressive) a t the center of the contact area and decreases to negative (tensile) values toward the edges. As indicated in Figure 9A, once the maximum contact radius, amaxis reached (branch 2) and the load is decreased (branch 3), the shape of the hysteretic receding branch is not like that of Figure 3B, but rather it starts off being reversible and only begins to deviate from the advancing branch after the radius has fallen by about 10%. This point is close to where the two surfaces were previously subjected to compressive pressures. The results of Figure 9A (of which there are other examples) suggest that the nonuniform stress distribution within the contact zone leads to regions of different adhesiveness and that at any time the value of yR is probably different at different places within the contact area. The results also suggest that there is a higher degree of interdiffusion or entanglements, and hence of adhesive energy, toward the center of the contact zone. Concerning the effect of load on interdiffusion, one might ask whether by increasing the compressive pressure one does not also slow down the diffusion rates of the chains, as occurs with bulk elastomeric materials.39 If so, this should act to reduce, not increase, the interdiffusion (over the same time period) and thus lead to a reduced adhesion. From our results on the enhanced adhesion with load, it appears that increasing the pressure across an interface separating two molecularly thin films or layers enhances the rate of interdigitation (or at least some adhesion-enhancing structural change) even though it may slow down some of the molecular motions involved. Similar effects have been observed with polymer surfaces,3s though in bulk systems interdiffusion rates are usually decreased with increasing load or pressure.39

Effect of Coverage. In a fully condensed ("frozen" or "crystalline") monolayer a single hydrocarbon chain has a cross-sectional area of 20 hi2. In our experiments we varied the chain area between 20 and 60 A*. A summary of our results is given in Table 11. These show that, for high-density coverage, where the area per chain is in the range 20-40 A2, the hysteresis is small, but that it increases dramatically for chain areas in the range 40-60 A2. This phenomenon will now be described in more detail. Figure 10A shows the absence of hysteresis for two DMPE monolayers with an area of 21.5 A2 per chain (yR-yA= 4-8 mJ/m2). Increasing the area to 33.5 A2 per chain introduces some hysteresis, but this remains small. Note that, while these chains are probably in the solid-amorphous state?2 it is still not easy for them to interdigitate fully until their areas exceed twice the fully dense area. Other monolayers where the area per chain is less than 40 AZalso show relatively small hysteresis effects (see Table I1 and Figure 5 ) . Whenever the hysteresis was large (yR-yA> 15 mJ/m2) it appears that the chains in one or both of the monolayers had a low coverage density, with chain areas in excess of 40 A2. This is illustrated in Figure 5 and Table I1 for the DHDAA/CTAB and CTAB/CTAB systems, where the CTAB chain area was 60 A* per chain and where opposite chains could now interdigitate fully. These monolayers provide the highest hysteresis observed, with cases where yRwas more than twice the value of yA,indicating that the true or "real"' molecular area was more than double the initial (or "apparent") area. Effects of Phase State on Adhesion. The results so far indicate that the more amorphous or disordered the molecules at a surface, the greater is their propensity to interdigitate with other surfaces, and the greater will be the receding adhesion energy. One might therefore expect two liquid layers to have the highest values for yR. But this is not the case: in addition to the coverage, the phase state, or chain mobility, is also important. For amorphous monolayers, their chains slowly interdigitate once the two surfaces are in contact, and the adhesion energy is therefore higher on separation unless the separation is done much more slowly than it takes the chains to reverse their path and disentangle. Thus, when the temperature is raised to above the temperature where the monolayer is in a fluid state the equilibrium chain configuration is now achieved within a small fraction of a second, so that on separation the chains can disentangle rapidly enough that the effective molecular area being separated is similar to the one on approach. The absence of a time effect with CTAB monolayers in the near-fluid state was already noted in Figure 7B. Figure 11 shows how the loading-unloading cycle of CTAB changes with temperature. The hysteresis appears to decrease at the highest temperature, but the effect is within our experimental error (see Table I1 for details). On the other hand, with the fluid CaABS mon-

The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

Adhesion on Monolayer Surfaces

Surfaces exposed to inert dry air

10745

Surfawse~posedto raturnled hydrocarbanvapor

A

- 3

0 *- 2 *T-

O l

-20

-10

0

10

20

30

APPLIED LOAD

Figure 11. a'-F curves for CTAB monolayers at different temperatures. The corresponding values of yR,and ys are given in Table 11. Solid and dashed lines are as defined in Figure 8.

olayers, the hysteresis was significantly less than with the amorphous CTAB monolayers even though the potential for entanglements is greater with the randomly branched CaABS surfactant. Effect of Organic Vapors on Adhesion. We found that the adhesion and the hysteresis were strongly affected by the presence of organic molecules in the vapors. Thus, when two surfaces covered with monolayers are exposed to a saturated or nearly saturated vapor of n-decane or n-dcdecane, as soon as they come into adhesive contact capillary condensation occurs and bulk liquid alkane condenses around the contact zone. Also, the thickness of the monolayers increases due to penetration of alkanes into the hydrophobic region of the mon01ayer.I~ Under such conditions, the two monolayer surfaces are now actually interacting across a film of liquid hydrocarbon, pressed together by the Laplace pressure of the hydrocarbon-vapor meniscus." The condensed liquid meniscus has two effects on the adhesion. First, the direct adhesion between the two contacting monolayers across the liquid hydrocarbon phase, ysL, is drastically reduced-falling from about 25 to 1 mJ/m2.28 This is due to the reduced van der Waals force across a hydrocarbon phase and the increased repulsion arising from hydrocarbon penetration into the monolayers. Even if interdigitation between the apposing monolayers would take place, it would increase the adhesive force only slightly, since ysL is so low. Second, the adhesion between the two surfaces is now determined almost entirely by the capillary forces arising from the Laplace pressure of the hydrocarbon liquid, which gives an additional term to the adhesion force of Fs = 47rRyL cos OR, where yLis the surface tension of the condensing liquid. For CTAB in dodecane vapor, it was found that OR = 10-13O. In other words, in the presence of a large capillary condensed liquid, since yL>> ysLand cos OR = 1, the total adhesion force is effectively given by" Fs = ~ R ( 4 7 ,COS O R + 3ysL) i= 47rRy~ (11) Thus, in the presence of condensing organic vapors we expect (i) that the adhesion energy be given by eq 11, which is the same as the DMT equation, eq 9, with the operative y being that of the condensed liquid and not that of the original solid surfaces, and (ii) that since the major adhesion force now involves a liq-

Figure 12. Disappearanceof adhesion hysteresis on exposure to various organic vapors. The adhesion energies ys as measured by the pull-off forces were calculated to be 21 mJ/m2 for CTAB, 18 mJ/m2 for DHDAA, and 21 mJ/m2 for CaABS.

uid-vapor rather than a solid-vapor interface, the hysteresis should be much reduced. Experiments were done in saturated vapors of decane and dodecane, for which yL a t 25 "C are 23.4 and 25 mJ/m2, re~pectively~~ (hexane was found to dissolve some of the monolayers, leading to phase separation). From eq 11 we would therefore expect yL = FS/4uR = 24 mJ/m2. Figure 12 shows that in the presence of organic vapors the hysteresis disappears in all cases, including with CTAB monolayers which were the most hysteretic in inert vapor. The measured adhesion energies, calculated according to eq 11, were in the range 18-21 mJ/m2-all slightly low (though if the JKR equation is used we would obtain yL = Fs/4?rR = 24-28 mJ/m2 which is more reasonable; however, the JKR theory should not apply in cases where there is a large meniscus bridging the two surfaces,34 as was the case here). Contact Angle Measurements. Contact angles of water on monolayer-coated surfaces were measured as described in the Experimental Section. Whereas all the data so far shown have been for fully dried monolayers exposed to an inert atmosphere, it is clear that as soon as one starts measuring liquid contact angles on a monolayer it must become wet as the liquid penetrates into it. This is especially true for water which has long been known to penetrate into adsorbed monolayers on solid surfaces and drastically change both their structure and dynamics. Indeed, as long ago as 1938, Langmuir12 proposed that contact angle hysteresis of water on monolayer surfaces was due to the flipping of surfactant molecules on becoming exposed to water, which rendered the surfaces hydrophilic thus reducing the receding angle. We have made many measurements on the adhesion and wetting properties of monolayers exposed to atmospheres of different vapors at different vapor pressures. These results will be reported in detail elsewhere. Here we briefly summarize those aspects of our findings that bear a relation to the adhesion hysteresis of these surfaces. As shown in Table 111, the advancing contact angles of all but the CaABS surfaces were about BA = 95O-1 loo, as expected for water on hydrocarbon surfaces. Receding contact angles OR were generally less, and clearly depended on (i) the advancing and ~

~~

~

~~

(33) Landolt-Bomstein Zahlenwerte und Funktionen, Band 11, 3 Teil, 6 Auflage; Springer Verlag: Berlin, 1956. (34) Fogden, A.; White, L. R. J. Colloid lnterface Sci. 1990, 183, 414. ( 3 5 ) Kausch, H. H.; Nguyen, T.Q. Acta Polym. 1988, 39, 5 5 .

10746 The Journal of Physical Chemistry, Vol. 95, No. 26, 1991

Chen et al.

TABLE 111 Advancing and Receding Contact Angles of Water on Silenated and Venous Surfactant-Coated Mica Surfaces at 25 OC“

phase state of monolayer O,, deg

e, deg

after 10 s after 2 min

AW, mJ/m2 after 10 s after 2 min

silane (chemisorbed)

crystalline DMPE (42 A2)

DMPE (62 A*)

amorphous DHDAA

CTAB

CaABS

112

113.5

111

106.5

95

34

105 105

53 15

53 13

12 13

56 13

26 20

8 8

72 98

69 96

43 91

47

5 8

76

“The stopping time between advancing and receding was either 10 s or 2 min as shown. Advancing and receding rates were kept roughly the same in each case at about 0 . 0 1 4 1 mm/s. The results were reproducible in the sense that after each droplet had receded and OR was measured, the subsequent advancing contact angles 8 , were always above 95’ (besides CaABS). receding rates, and (ii) the “stopping” or “resting“ time during which the liquid droplet was allowed to rest on the surface after an advance before it was made to recede. The longer the stopping time and the slower the receding rate, the lower was OR. In addition, values of OR were generally lower at the start of receding (static OR) than during motion (kinetic na& + observation that is reminiscent of, and probably related to, the phenomenon of static and kinetic friction that gives rise to stick-slip motion. In trying to quantitatively analyze the results of contact angle hysteresis it appears that the best criterion to use is not the hysteresis in the angle itself, but rather that of the adhesion energy that gives rise to it. Accordingly, we define the hysteresis as [see eq 21 AW = y(C0S OR - COS 0,) (12)

stiction forces are much stronger for amorphous than for solid or fluid monolayers. Our results have another bearing on friction, concerning the question of “real” versus “apparent” areas of contact.’ With some of the monolayers the value of yR was often more than twice the initial value of yA (which corresponded more closely to the thermodynamic value expected for molecularly smooth contacts). This indicates that during contact the molecular area rapidly grew to more than double the initial or apparent area-a possibility that can only arise with relatively soft, easily deformable surfaces. This is in contrast to the more common case, encountered in adhesion and friction of hard solid surfaces,l where the real area of contact between two solid surfaces is generally less than the apparent area, due to isolated asperity contacts.

Table 111 shows measured values for e,.!, and OR, and the corresponding values obtained for A W using the above equation. It appears that the hysteresis is low only when the surface groups are totally immobilized (silane surfaces) or highly free to reorient (CaABS). In all other cases,the rate of molecular rearrangements, flip-flop, etc.I3 is slower than the advancing/receding rate and so the hysteresis is high. The low hysteresis measured on the fluid monolayers of CaABS is most interesting because of the initially low advancing contact angle. It is likely that the initial (presumed) hydrophobic surface can equilibrate with the advancing water phase so fast (involving flip-flop of moleculed3) that the advancing angle is low, and not much different from the receding angle since the system is always close to equilibrium. Relation between Adhesion, Wetting, and Friction. Advancing menisci often move forward abruptly (“stickslip wetting”), and contact areas also often decrease abruptly on receding (“stick-slip dewetting”). Both these effects are reminiscent of stick-slip friction which normally occurs between solid surfaces.l Stick-slip wetting was found to occur particularly for water on the more solidlike monolayers of DMPE, but never on the fluid CaABS monolayers. Similarly, stick-slip adhesion, both on loading but especially on unloading, was found to occur with the monolayers, but was always eliminated when organic vapors were introduced (see Figure 12). All three phenomena may be related through the basic mechanisms shown in Figures 1 and 2, where instabilities must occur whenever the elastic restoring forces are weak, Le., under conditions where the backing material approximates more to a weak elastic solid. Furthermore, the increased adhesion or sticking with contact time arising from monolayer interdigitation may also be related to the increased friction between two solid surfaces the longer they have been allowed to remain in contact, and also with the common phenomenon of “stiction”-the very high initial friction preceding steady-state sliding. Merkel and recently found that the lateral diffusion coefficient of surfactants in monolayers in contact with each other decreased as their ability to interdigitate increased. Our preliminary results on the boundary friction of sliding monolayers also indicate that

Discussion Our studies show that the adhesion of two molecularly smooth and chemically homogeneous surfaces can be hysteretic due to the finite elasticity of solid surfaces, the finite times of molecular rearrangements at interfaces, and the inevitably finite times of real loading/unloading and wetting/dewetting processes. Hysteresis increases with (i) the ability of molecular groups at surfaces to reorient and interdiffuse across the contact interface which is often determined by the phase state of the surface molecules, (ii) the time two surfaces remain in contact and the externally applied load during this time, and (iii) the rate of approach and separation (or peeling) of surfaces. Hysteresis effects are inherently energy-dissipative and thus indicative of the absence of true equilibrium. Our results suggest that one may distinguish between mechanical and chemical contributions toward hysteresis. The difference between the two was nicely exemplified in the following: When surface groups are in the solid crystalline state they are effectively immobilized (frozen). No interdiffusion occurs on contact and the adhesion does not increase with contact time. The loading/unloading process is elastic and nonhysteretic, and any irreversibility arises mainly from the mechanical nonequilibrium associated with approach and separation (Figure 2). By reducing the surface density (coverage) or changing the temperature the surface groups become disordered and freer to move and reorient. In the case of adsorbed monolayers this leads to some sort of amorphous or glassy state. With increasing disorder of hydrocarbon chains the receding adhesion energy increases above the value for solid crystalline surfaces, and the loading-unloading cycles now become very hysteretic. The hysteresis is now mainly due to chemical nonequilibrium effects associated with approach and separation (Figures 1 and 6B). As discussed in the Introduction, chemical nonequilibrium occurs when the surface groups on separation are differently positioned or oriented from those on approach. One consequence of this is that larger adhesion “pull-off“ forces would generally be measured than expected from equilibrium surface energy values. Our results indicate that the equilibrium JKR theory, both as regards its prediction for the pull-off force ( F = 3aRy) and load-area profiles, is essentially correct and that previous indic a t i o n ~ ~ ’that - ~ ~the equilibrium pull-off force is higher (and given

(36) Merkel, R.;Sackmann, E.; Evans, E. J . Phys. Fr. 1989, 15, 1535. (37) Bevington, P.R.Error Analysis and Data Reduction for the Physical Sciences; McGraw-Hill: New York, 1967.

Adhesion on Monolayer Surfaces by F = 4 ~ R y were ) fortuitously correct for those systems where adhesion hysteresis enhances the receding surface energy. In these as well as the present series of experiments the surface groups have all been mainly - C H 3 and -CH2 groups, which reorient and interdiffuse on contact so that the “chemistry” of the surfaces are different on separation. In another publication we intend to describe hysteresis effects between monolayer surfaces in the presence of water and other polar groups where additional chemical rearrangements occur involving the replacement of nonpolar groups by polar groups. At even higher temperatures or for totally fluid chains, the molecular groups at the surfaces are now able to relax very rapidly, and the loading/unloading cycles again become less hysteretic. Hysteresis is also reduced in the presence of condensable (and wetting) hydrocarbon vapors because of the fluidization of all the interfaces. Penetration of the alkanes into the chain regions of the monolayers also causes the monolayers to swell and become more fluidlike, which also lowers the relaxation times and thus speeds up the time to reach equilibrium. These results highlight the qualitative differences between chain entanglements and kinetic effects. For totally fluid chains, even though entanglements are high, one may expect reversible adhesion, with the system at equilibrium at all stages of a loading-unloading cycle. With amorphous, glassy, or gellike layers at low or intermediate coverage, their relaxation time is long and the hysteresis is high. As we approach the other extreme of totally frozen, high-density monolayers, no interdigitation is possible (at least over the times scales of the experiments) and the measured adhesion is again reversible. Note, however, that the frozen state of the monolayers may not be in the true thermodynamic equilibrium state when these monolayers are in contact. Contact angle measurements indicate that there are some correlations between adhesion hysteresis and contact angle hysteresis (of water on the same surfaces). Thus, in both cases the hysteresis was low for monolayers in the fluid state, and high for solid-amorphous layers. On the other hand, with crystalline monolayers of DMPE there was no adhesion hysteresis but a high contact angle hysteresis (though for chemisorbed sianated surfaces the contact angle hysteresis was again very low). It may be that water has a sufficiently strong effect that it can overturn even the more strongly physisorbed layers (but not chemisorbed layers), whereas no such strong forces existed during the adhesion measurements which were done in dry conditions. To support this view, we mention some publishedi3 and unpublished results on the adhesion of DMPE layers in water vapor which show that ys increases both with relative humidity and decreasing coverageexactly as would be expected. A natural question that always arises in contact angle and wetting studies is, “which is the thermodynamically correct contact angle, the advancing or receding one?” Our results suggest that

The Journal of Physical Chemistry, Vol. 95, No. 26, 1991 10747

in general neither is, since both the advancing and receding processes are occurring through nonequilibrium molecular mechanisms. Indeed, if we consider the more general case of an advancing or receding interface separating two immiscible liquids in contact with a surface, it becomes immediately obvious that how one assigns the advancing and receding angles is purely a matter of convenience. The solid-amorphous monolayers provided the highest hysteresis observed, with cases where YR was more than twice the value of yA,indicating that the molecular area was more than double the initial area. This is in contrast to the more common case, encountered in adhesion and friction phenomena,’ where the real area of contact between two solid surfaces is generally much less than the apparent area, due to asperity contacts. Chemical hysteresis appears to be far stronger than mechanical hysteresis, and even with organic molecules chemical hysteresis can give rise to changes in Wof 100 d / m 2 arising from molecular and segmental rearrangements occurring at the last 1-2 nm below a surface. By contrast, surface roughness and elastic instability effects, even when these occur over much larger scales, will rarely give rise to such large energy dissipative effects. In a similar study to ours Chaudhury and Whitesides3*found hysteresis in the adhesion (a3+ curves) between functionalized PDMS surfaces but not between unmodified (pure) PDMS. This is consistent with the notion that molecular reconfigurations and exchange can occur in the former system but not in the latter (or less so). In that study, too, surface roughness appears to have played an insignificant role. Our findings thus question the traditional view that hysteresis is due to the inherent or static roughness or chemical heterogeneity of surfaces and focuses more on the dynamics of these effects. Furthermore, the results suggest that many energy-dissipating processes at interfaces are determined by structural and chemical changes occurring at the molecular (or even Angstrom) level. Acknowledgment. We thank Patty McGuiggan for synthesizing the DHDAA by ion exchange and Michelle Gee for purifying the CaABS. Y.L.C. thanks Exxon Chemical Co. and IBM for a graduate fellowship sponsored by SUR Research Award No. 800612 for the work on the self-assembled monolayers. We thank the DOE for funding the work on the Langmuir-Blodgett deposited monolayers and for the development of the experimental techniques used in this study under grant DE-FG03-91ER45331. Registry NO. CTAB, 57-09-0; DHDAA, 71326-37-9; DMPE, 99807-2; DPPE, 923-61-5; OTS, 112-04-9. (38) Chaudhury, M. K.; Whitesides, G . M. Langmuir 1991, 7, 1013. (39) Kausch, H. H.; Tirrell, M. Annu. Rev. Mater. Sci. 1989, 19, 341. Yuan, B. L.; Wool, R. P. Polym. Eng. Sci. 1990, 30 (22), 1454-1464. (40) Chen, Y. L.; Helm, C. A.; Israelachvili, J. N. Langmuir, in press.