Nanomechanical Properties of Interfacially

Total Page:16

File Type:pdf, Size:1020Kb

Nanomechanical Properties of Interfacially

Surface Glass Transition ChemE 554/Overney

4.5 Surface Glass Transition and Transitions in Thin Films

Index

4.5 Surface Glass Transition and Transitions in Thin Films...... 1 4.5.1 Background...... 2 4.5.2 Interfacial confinement effects and film preparation history...... 3 4.5.3 Liquid-like surface models that address Tg depletion in thin films...... 5 Capillary wave induced surface melting...... 5 Near surface polymer chain sliding (sliding model)...... 5 4.5.4 Shear modulation scanning force microscopy (SM-SFM)...... 6 4.5.5 Mobile surface layer theories and preliminary SM-SFM results...... 8 4.5.6 SM-SFM transition measurements of ultrathin supported films...... 8 References...... 10

4.5.1 Background Extensive literature deals with the determination and the interpretation of the glass transition temperature, Tg, of homopolymer systems. Within bulk systems, many debates have meanwhile been settled. At free surfaces and in thin films however, because of constraints and size effects, the relaxation dynamics is still poorly understood. It has been recognized theoretically and experimentally that in thin homopolymer films the proximity of a free surface, substrate interactions, and stress induced anisotropy within the film, are responsible in an intricate way for shifts of the glass transition temperature with respect to the bulk.1-20 Glass transition temperatures have been reported to be depressed up to tens of degrees Celsius for films thinner than a few hundred angstroms.11,14,20 It has also be recognized that substrate effects, can cause an opposite 11,17 shift in Tg from the bulk to higher values. Models that have been developed over the last few years favor the idea of a liquid- like surface layer that is responsible for a Tg depletion in none-substrate-confined thin films.5,20,21 One of them, inspired by dewetting studies of thin films,22,23 in which morphological changes below the bulk Tg value were reported, proposes coupling of capillary waves with flow properties in thin films responsible for surface melting.21 In another liquid-like surface model (referred to it as the sliding model), the scenario of the glass transition is split into two; i.e., a standard bulk transition that is based on freezing- unfreezing of certain local degrees of freedom), and a near-surface transition that originates from polymer chain sliding motions. Depending on the temperature and thickness of the film, the model suggest either a sandwich structure of bulk and surface phases, or a single semifluid phase.5,24 The two liquid surface models are applied to different regimes of molecular weights: The capillary wave model21 that is based on capillary wave-induced dewetting deals with low molecular weight polymers melts. The sliding model 5,20 that is considering in parallel standard bulk transition and near-surface chain sliding motions is proposed to

1 Surface Glass Transition ChemE 554/Overney apply to polymers above 100k. Both models do conceptually not depend on the film thickness, i.e., they apply for ultrathin films (< 100 nm) and at surfaces of bulk films. Liquid-like surface theories have been partially supported by macroscopically averaging techniques, such as, for instance, ellipsometry11,25,26, Brillouin light scattering20,27, and thermal expansion spectroscopy28. The macroscopically determined Tg depletion is often related to the film thickness. However, more local probes used in dielectric measurements 29 and scanning force microscopy, SFM,12,17 led to results, where the glass transition temperature is much less affected by the film thickness. Contrary to the last statement about local probes are some recently published interpretations of SFM friction force measurements30. Possible misinterpretations are addressed below. After the development of a nanorheological technique that has shown to be quite successful in determining Tg values of polymer films over a wide range of molecular weights, one has recently investigated (a) the possibility of surface melting below the glass transition temperature, and (b) interfacial substrate confinement effects on apparent Tg values. Glass transition values obtained of monodisperse, homopolymer bulk material or at surfaces of >200 nm thick films by macroscopic techniques (i.e., differential scanning calorimetry (DSC) and electron spin resonance (ESR)), and by local techniques (i.e., shear modulation SFM), respectively, are found to correspond over a wide range of molecular weights, Figure 1. The finding that surface and bulk transition properties correspond agrees well with the low surface energy of these systems, and the van der Waals liquid-like description of polymer melts.

400 Figure 1. Glass transition of atactic bulk polystyrene

380 as function of the molecular weight determined independently by three different methods: 360 3

] - Differential scanning calorimetry (DSC),

K 340

[

g 31 T - Electron spin resonance (ESR) and

320 DSC (Claudy) ESR (Kumler) - Shear modulation scanning force microscopy shear modulation SFM 300 1 Fox Flory Fit (SM-SFM)

280 The solid line represents a Fox-Flory fit. 1x103 1x104 1x105 1x106 1x107 MOLECULAR WEIGHT [M ] n

However, the more complex the polymer system, i.e., the more anisotropic it is, the less correspondence is found between bulk, thin film and surface Tg measurements. Origins for deviations have to be carefully analyzed. Common assumptions have to be revisited and challenged. A common hypothesis is that thermal annealing of a spin coated homopolymer film, with a thickness exceeding the pinning regime of a few nanometers, is sufficient to relax the film back to the bulk phase. It is interesting to note that near surface Tg values determined with shear modulation SFM method correspond well to bulk Tg measurements as illustrated above in Figure 1. Hence, the annealing presumption of ultrathin films seems to be justified. Over the last few years however, various experimental findings have challenged common hypotheses such as the annealing presumption. It has, for instance, been found that the polymer rheological properties are modified over tens to hundreds of nanometers.1,17,32-34

2 Surface Glass Transition ChemE 554/Overney

4.5.2 Interfacial confinement effects and film preparation history While materials such as ceramics, metals, oxides, exhibit size limitations ('quantum well effects') only noticeable below 10 nm, it was found that in polymer systems, interfacial effects could be noticeable over distances of tens to hundreds of nanometers. Over the last few years, various groups reported bulk-deviating structural and dynamic properties for polymers at interfaces.15,34-39 For instance, increased molecular mobility was observed at the free surface for thick films.15 Reduced molecular mobility at the film surface of ultrathin films was reported based on forward recoil spectroscopy measurements.35 In secondary ion mass spectrometry (SIMS) and scanning force microscopy (SFM) studies of graft-copolymers, it was found that the degree of molecular ordering significantly affects dynamic processes at interfaces.34 Self-organization of graft and block-copolymers at surfaces and interfaces were found with transmission electron microscopy (TEM) and neutron reflectivity (NR).36-39 Application of mean-field theories to interfacially constrained and size-limited polymer systems failed to describe the rather unexpected mesoscale behavior observed experimentally. The extension of the interfacial boundary far into the bulk is unexpected because many amorphous polymer systems are theoretically well treated as van der Waals liquids with an interaction length on the order of the radius of gyration, i.e., the effective molecular size. At solid interfaces the radius of gyration is further compressed, like a pancake, and thus, any memory effects of the solid are expected to be even more reduced to a pinning regime of only 0.5 to 2 nm.40 Within the pinning regime, it is commonly accepted that the material is structurally altered and exotic properties (for instance, quantum-well effects) are expected. Outside the pinning regime, the polymer is expected to behave bulk-like. Experiments show however, that such scaling theories, i.e., mean-field theories, fail in describing the observed unique mesoscale properties because they do not consider effects that occur during the film coating process, e.g., rapid solvent evaporation. For instance, recent SFM experiments revealed that the spin coating process altered the structural properties of polyethylene-copropylene (PEP) at silicon interfaces due to anisotropic molecular diffusion that is caused by process-induced structural anisotropy.41 The polymer structure at the interface affects properties such as the shear mechanical properties, the entanglement strength, and dewetting instabilities and velocities, as illustrated in Figure 2. An extensive SFM analysis involving also density measurements by X-ray diffraction and self-diffusion measurements by neutron reflectivity (NR) revealed a three component system after spin coating: (i) An adjacent to the surface immobilized and fully disentangled sublayer (20 nm thick, Fig. 2), (ii) a partially disentangled intermediate layer (100-200 nm thick, Fig. 2), and (iii) beyond the intermediate layer the bulk polymer phase. he strained interfacial sublayer can be pictured as highly disentangled and laterally anisotropic system with a thickness on the order of the polymer's radius of gyration. NR reveals that the polymers adjacent to the surface immobilized sublayer can diffuse through the sublayer's pores.39 X-ray diffraction measurements exhibit a shrinkage of the pores if annealed which immobilizes the diffused polymers 'permanently'.32 Hence, a boundary layer that exceeds by orders of magnitudes the pinning regime is formed between the interfacial sublayer and the polymer bulk phase. Naturally one could expect that the glass transition temperature be affected within the constrained boundary regime. Indeed, it has been observed that substrate supported spin

3 Surface Glass Transition ChemE 554/Overney

o coated ultrathin films of polystyrene exhibit an increase in Tg (by 5 C at a thickness of 11,17 30 nm) compared to the bulk. Much larger changes in Tg by tens of degrees Celsius and in the opposite direction have been observed for substrate unsupported ultrathin homopolymer monodisperse PS films.11,14,20 These findings have let to theoretical models that go beyond the free volume theory and consider either coupling of capillary waves with flow properties, or near surface polymer chain sliding motions responsible for 5,20,21 the observed depletion in Tg in ultrathin films.

D 1.0 Liquid 0.8 c Bulk Regime Frictiondecreasesbecauseof pinningforces! 0.6 < 200 nm Normalized Lateral Force 0.4 Thin Film 0.2 Regime 0.0 Pinning Regime 0 50 100 150 200 250 300 350 400 PEP Film Thickness D[nm] Si Figure 2(a): (Left) SFM rheological lateral force measurements on thermally annealed PEP reveal a Figure 2(b): Lateral (friction) force vs. load SFM comparable qualitative film thickness behavior-as the experiments provide fundamental insight into the dewetting velocities in Fig. 4(b). A decrease in the origin for changes of the rheological properties measured lateral force is interpreted as an increase in in spin coated thin PEP films. The transition the shear mechanical properties (moduli) of the point Px (x = thickness of polymer film), which interfacially confined polymer film. corresponds to the kink in the friction vs. loading (Right) The range over which the confinement is curve is a measure of the entanglement strength recognizable (Thin Film Regime, 100-200 nm,) of the polymer (PEP). The bulk value is reached exceeds by orders of magnitude the surface pinning for films thicker than 230 nm. Films thinner than regime (< 5 nm). Above the thin film regime, the 230 nm are partially disentangled due to the spin polymer behaves bulk-like. coating process, and thus, the transition point occurs earlier. Within a 20 nm boundary regime the film is entirely disentangled (gel-like).

0.8 0.7 0.6

0.5 Plot Title nm/s 0.4 0.3 0.2 0.1 0.0 0 100 200 300 400 500 Thickness of PEP [nm]

Figure 2(c): (left a+b) Illustrative sketch of the Figure 2(d): PS/PEP dewetting velocities dewetting process measured by SFM. (right) 5050 measured by optical microscopy reveals a m recorded dewetting pattern (topography (left) and decrease in velocity for PEP films thinner than friction (right)) of PS/PEP. (top) Large and deep 100-200 nm. The triangle indicates that the dewetting holes for thick (400 nm) PEP films. history of the sample preparation is very (bottom) Small and interfacially bound dewetting important ( PEP spin coated on silicon wafer, holes for thin (4 nm) PEP films.33,42 PEP floated on polyvinylpyridine (PVP)).34

4 Surface Glass Transition ChemE 554/Overney

4.5.3 Liquid-like surface models that address Tg depletion in thin films

Capillary wave induced surface melting In this model, capillary wave coupling on the free surface to the bulk flow of the polymer is responsible for the formation of a melt in the vicinity of the surface. Herminghaus' capillary model21 involves coupling between interfacial eigenmodes of the viscoelastic polymer melt with capillary waves. This model is considering only low molecular weight polymers. (Note: The polymer-chain-sliding model (discussed below) applies for high molecular weights exceeding 100k). The capillary model is not restricted in polymer film thickness, i.e., applies to the free surface. It suggests a continuum scenario responsible for the glass transition, which is largely independent of the molecular weight. The theory considers strain relaxation fluctuations (density fluctuations are neglected) and its viscoelastic eigenmode spectrum. The central assumption is that the physical cause for the melting or freezing of the film (or surface layer) are memory effects in the polymer material (i.e. using as a theoretical treatment: a convolution integral with memory kernel). It is found that only high frequency modes contribute appreciably to the reduction of the glass transition temperature. The model predicts that the highest eigenvalue modes are those of the molten layer alone, which depend inversely on the thickness of this layer. The mean square amplitude of the strain fluctuation of these high modes, a measurable quantity that can be obtained from rheological spectrum analysis, should be appreciable only down to a critical thickness of the melt. The model predicted - thickness of the molten layer, d, scales as d  (Tgb-T) with a critical exponent =1. Recently developed other multilayer models predict =0.5.20 Herminghaus' model consequently proposes for ultrathin films that the transition temperature value measured over the entire film is continuously changing from the time surface-melting occurs until the thickness of the molten layer corresponds to the thickness of the film. Note however, in thick films, i.e., in films in which the thickness of the film exceeds any molten surface layer thickness below the bulk Tgb value, two transition values should be obtainable with a surface sensitive tool.

Near surface polymer chain sliding (sliding model)

Experimentally observed extraordinary depleted Tg values for ultrathin freely suspended polystyrene films10,43 have recently been interpreted by de Gennes via a two "melting- scenario" 5: (a) a standard bulk transition, related to the freezing-unfreezing of certain local degrees of freedom, and (b) sliding motions of each polymer chain along its own path. In the bulk it is assumed that chain sliding is hindered by the end points. At or near the free surface, however, a thin fluid skin allows the chains to slide. The skin is estimated to be less than a nanometer thick. The two transition processes are illustrated in Figure 3.4 B C Figure 3: Model of polymer chain in a thin film. Two contributions arise from a segment formed by loop AB and a bridge BC. For films thinner than the coil size, the dominant process maybe the collective motion of a loop which does not A involve the chain ends.4

5 Surface Glass Transition ChemE 554/Overney

The bulk Tg motions are based on short-range rearrangements. These types of motions are contributions of bridge formations shown in the polymer chain touching points A and B in Figure 3. The other type of motion is based on polymer sliding, where a chain advances along a path via mobile kinks and the free volume required for sliding is less than the bulk cooperative motion. These types of motions are contributions of “loop” formations shown in the polymer chain touching points B and C in Figure 3. A loop at the surface should slide easily, however, it is believed that sliding is hindered in the bulk because chain ends would have to invade new territory, thus requiring more free volume. The two transition model lead to a Tg that depends on the distance from the free surface (h-h*), described by  M  T  T *  bln w  h  h*  g g   *   (1)  MW 

where Tg, Mw, and h refer to the glass transition temperature, molecular weight, and film * * * thickness respectively. The parameters b, Tg , Mw , and h are fitting parameters from Tg vs. h plots characteristic of the observed deviation from bulk Tg values. The model predicts bulk behavior at the surface for very high molecular weights (> several 1,000k), which is experimentally confirmed in ultrathin films.5,11 Although the mobile surface model has been motivated by ellipsometric ultrathin film studies, which lead to discussions about experimentally observed apparent transition values and theoretically predicted two scenario transitions, it is not restricted to thin films only.44 It could also be applied to free polymer surfaces of thick films. An experimental test of the model demands a surface sensitive tool such as the shear modulation SFM.

4.5.4 Shear modulation scanning force microscopy (SM-SFM) The working principle of the shear modulation scanning force microscopy, SM-SFM, method is sketched in Figure 4.17 The technique is well suited for any surface rheological study involving thermally activated transitions or relaxations. Over the last two years, the method has shown to be a highly accurate method for determining near surface glass transition temperatures of thin polymer films. The method involves a nanometer sharp SFM cantilever tip that is brought into contact with the sample surface as sketched in Figure 4. A constant load of a few nanonewton is applied, and the probing tip is laterally modulated (with a nanometer amplitude that guarantees no relative probe-sample slippage) while the temperature is stepwise increased by 0.1oC. At each temperature step the system is idle until thermal equilibrium is obtained before any viscoelastic responses are recorded. The recorded response amplitude, which is a measure of the contact 17 stiffness, is then plotted versus the temperature. The Tg value is determined from the "kink" in the response curve as documented in Figure 4(a). The figure also illustrates the high accuracy of the method. It allows Tg studies of various parameters such as, for instance, the molecular weight dependence as shown in Figure 4(b), (see also Fig. 1 above). It is important to note that the SM-SFM method is a non-scanning method. The reason is briefly describe here: To obtain high accuracy in Tg measurements, it is essential, not to induce by other means than temperature, changes in the contact area. This is to avoid system-driven artifacts in the contact stiffness, kc. To be precise, kc(AL, G*), i.e. the resistance of the contact to deform, is dependent on (a) the laterally projected

6 Surface Glass Transition ChemE 554/Overney

contact area, AL, (e.g., the side wall of an indentation dip), and (b) the relative shear properties of the two materials, G*. Thus, any local plastic deformation, for instance, the generation of a deformation wave (Schallamach wave)45 that travels ahead of a scanning SFM tip can change kc. Any plastic deformation is intrinsically rate and load dependent. Thus, it is not astonishing that scanning methods such as the lateral (friction) force microscopy revealed scanning velocity dependent apparent transition values for Tg, Fig. 5.8

Shear Response x L 380 (a) (b) 370 K

360 / Surface

g 350 Bulk T 340 Cantilever (1) 3 k 330 3 4 5 6 7 (2) 7 k 10 10 10 10 10 Tip Molecular weight (3) 9.5 k xs (4) 65 k Sample (5) 6,500 k

. 342 K u

Shear Displacement . a

/

e

d (1) 355 K u t

x i mod l p (2) m

A 360 K Figure 4: Schematic representation of the SFM based SM- (3) SFM method. The sample is sinusoidally modulated (xmod) 374 K relative to the probing cantilever tip. In response the contact (4) 374 K and the cantilever are deformed by xs and xL, respectively. (5) (a) Tg values correspond to dominating kinks in shear response amplitude vs. temperature curves, as here illustrated 300 320 340 360 380 400 Temperature / K on polystyrene for a wide range of molecular weights.

By placing the SFM tip stationary at constant load onto the polymer surface, contact area changes occur only due to temperature induced changes in the rheological properties of the material. Consequently the experimental observable in the SM-SFM method, kc, is changing only due to changes in the polymer material properties.

1 m/s 120 Figure 5: The absolute temperature for friction 5 m/s

] 20 m/s measurements is ill-defined due to sliding N n [ 100 velocity dependent changes in the contact area, E C

R i.e., the contact stiffness. The scan distance was

O F

L 80 5 m and the load 15 nN. At high speed, no A

R

E transition is observed. An apparent transition T 372K A

L 60 corresponding to Tg can be observed at very low 378K speed. At intermediate scan speeds, the apparent 40 8 360 365 370 375 380 385 transition is higher than Tg. TEMPERATURE [K] Hence, the "kinks", observed in Figure 4(a), are true measures of the transition property

4.5.5 Mobile surface layer theories and preliminary SM-SFM results

In SM-SFM experiments on high molecular weight polystyrene films (230 nm thick, Mw = 6.5M) a subtle change in the SM-SFM curve was found at about 25 oC below the bulk Tg value, Fig. 6(a). A more careful analysis using adhesion force SFM, Fig. 6(b),

7 Surface Glass Transition ChemE 554/Overney

confirmed that at Ta = 350 K the PS surface changed, i.e., softened, forming a larger contact. with the SFM tip. It is important to note that the SM-SFM measurements were contacted at lower loads (i.e., smaller penetration depth) than the prior experiments. Considering DeGennes model, the here discussed finding is not unexpected, as the model predicts Ta to be distance dependent from the surface.

0.030

30 Tgb = 374 K 28 ) N

n

( 26 Ta(h)  350 K T = 354K N 0.025

O 24 I S

E 22 H

D

A 20

0.020 18 340 360 380 400 300 320 340 360 380 400 420 TEMPERATURE (K) TEMPERATURE [K]

Fig. 6(a): Shear rheological response measurements Fig. 6(b): SFM adhesion pull-off force vs.

(SM-SFM, applied load:10 nN) on PS, 6.5M (Mw), temperature measurements (maximum applied 230 nm thickness. The graph shows a load load 10 nN). The transition value of 354 K dependent (i.e., initial penetration depth, h, corresponds to the apparent surface transition dependent) apparent surface transition Ta(h) of 350 value determined by SM-SFM. K. The bulk glass transition temperature, Tgb, is 374 K.

4.5.6 SM-SFM transition measurements of ultrathin supported films In thin film studies one has to pay particular attention to the film preparation technique employed. As discussed above, spin coated films in the vicinity to the substrate can exhibit quite complex strain structures that can impact the glass transition properties. This is illustrated in Figure 7. The plot in Figure 7 compiles distance-dependent Tg values in a single plot for spin coated polystyrene films. As expected from previous rheological studies on thin supported films (as discussed above), the glass transition for ultrathin homopolymer films deviates from the bulk Tg value for films that are thinner than about 100-200 nm. Astonishingly, two regimes were obtained: On one hand, the apparent Tg value increased by an average of 4 oC within 30-100 nm. On the other hand, the apparent o Tg value dropped by about 8 C compared to the bulk at a film thickness of about 15 nm. While an increase in Tg could be expected due to interfacial confinement effects, the finding of a reduction in Tg in the boundary regime to the substrate is rather unexpected. The Tg-thickness dependence can be interpreted with the multilayer layer model (sublayer and transition layer) in spin coated films. The fully disentangled layer that is directly adjacent to the silicon substrate exhibits a morphology in which Tg occurs earlier, while the intermediate, partially disentangled layer (sandwiched between the sublayer and the bulk) is constrained, and hence exhibits higher Tg values.

8 Surface Glass Transition ChemE 554/Overney

FILM THICKNESS, d ( nm ) 0 50 100 150 200 250 300 105 12.0 kDa PS FOX-FLORY (BULK) Figure 7: Glass transition of PS vs. film 100 thickness determined from SM-SFM plots, for

) various molecular weight PS (MW). The films

C had been spin coated directly on silicon and o

o

( 95 thermally annealed over 4 hours at 130 C.

g The near surface measured Tg value is bulk-like

T for t > 200 nm. Within a "thin film transition 90 region" of about 20 to 200 nm (compare with o Fig. 2), the Tg value is increased by up to 5 C from the bulk value. A significant decrease in 85 Tg is observed for 15 nm thick films. Note: No bulk deviating Tg values were found for low interaction substrate surfaces, such as S

U PVP or OTS coated silicon surfaces, after

B temperature annealing.17 S T R A T E

SL INTERMEDIATE REGIME BULK

References

1 C. Buenviaje, F. Dinelli, and R. M. Overney, in ACS Symposium Series "Interfacital Properties on the Submicron Scale", 781 edited by J. Frommer and R. M. Overney, (Oxford University Press, Oxford, 2000), pp. 76-92. 2 H. P. Chen, D. Katsis, J. C. Mastrangelo, S. H. Chem, S. D. Jacobs, and P. J. Hood, Advanced Materials 12, 1283-1286 (2000). 3 P. Claudy, J. M. Letoffe, Y. Camberlain, and J. P. Pascault, Polymer Bulletin 9, 208- 215 (1983). 4 K. Dalnoki-Veress, J. A. Forrest, P. G. d. Gennes, and J. R. Dutcher, J. Phys. IV 10, 221-226 (2000). 5 P.-G. de Gennes, C.R. Acad. Sci. IV, 1-8 (2000). 6 G. B. DeMaggio, W. E. Frieze, D. W. Gidley, M. Zhu, H. A. Hristov, and A. F. Yee, Phys. Rev. Lett. 78, 1524-1527 (1997). 7 E. A. DiMarzio and A. J. M. Yang, J. Res. Natl. Inst. Stand. Technol. 102, 135-157 (1997).

9 Surface Glass Transition ChemE 554/Overney

8 F. Dinelli, C. Buenviaje, and R. M. Overney, Journal of Chemical Physics 113, 2043- 2048 (2000). 9 A. Eisenberg, ACS Symposium Series "Physical Properties of Polymers", Vol. 2nd edition, edited by J. E. M. e. al. (ACS, Washington DC, 1993). 10 J. A. Forrest, K. Dalnoki-Veress, J. R. Stevens, and J. R. Dutcher, Phys. Rev. E 56, 5705-5716 (1997). 11 J. A. Forrest and J. Mattsson, Physical Review E 61, R53-R61 (2000). 12 S. Ge, Y. Pu, W. Zhang, M. Rafailovich, J. Sokolov, C. Buenviaje, R. Buckmaster, and R. M. Overney, Physical Review Letters 85, 2340-2343 (2000). 13 T. Kajiyama, K. Tanaka, and A. Takahara, Macromolecules 28, 3482-3484 (1995). 14 J. L. Keddie, R. A. L. Jones, and R. A. Cory, Europhys. Lett. 27, 59-64 (1994). 15 Y. Liu, T. P. Russell, M. G. Samant, J. Stohr, H. R. Brown, A. CossyFavre, and J. Diaz, Macromolecules 30, 7768-7771 (1997). 16 G. F. Meyers, B. M. DeKoven, and J. T. Seitz, Langmuir 8, 2330-5 (1992). 17 R. M. Overney, C. Buenviaje, R. Luginbuehl, and F. Dinelli, J. Thermal Anal. and Cal. 59, 205-225 (2000). 18 D. Schaefer, H. W. Spiess, U. W. Suter, and W. W. Fleming, Macromolecules 23, 3431-3439 (1990). 19 K. C. Tseng, N. J. Turro, and C. J. Durning, Physical Review E 61, 1800-1811 (2000). 20 J. H. Kim, J. Jang, and W. C. Zin, Langmuir 17, 2703-2710 (2001). 21 S. Herminghaus, Eur. Phys. J. E in press. 22 W. J. Orts, J. H. Van Zanten, W. L. Wu, and S. K. Satija, Phys. Rev. Lett. 71, 867 (1993). 23 G. Reiter, Macromolecules 27, 3046 (1994). 24 P. G. de Gennes, Eur. Phys. J. E 2, 201 (2000). 25 S. Kawana and R. A. L. Jones, Phys. Rev. E 63, 21501 (2001). 26 K. Dalnoki-Veress, J. A. Forrest, C. Murray, C. Gigault, and J. R. Dutcher, Phys. Rev. E 63, 31801 (2001). 27 J. Mattsson, J. A. Forrest, and L. Boerjesson, Phys. Rev. E 62, 5187 (2000). 28 K. Fukao and Y. Miyamoto, Phys. Rev. E 64, 11803 (2001). 29 K. Fukao and Y. Miyamoto, Phys. Rev. E 61, 1743 (2000). 30 T. Kajiyama, K. Tanaka, and A. Takahara, Macromolecules 32, 4474 (1999). 31 P. L. Kumler, S. E. Keinath, and R. F. Boyer, J. Macromol Sci -Phys B13, 631-646 (1977). 32 C. Buenviaje, S. Ge, M. Rafailovich, J. Sokolov, J. M. Drake, and R. M. Overney, Langmuir 15, 6446-6450 (1999). 33 R. M. Overney, D. P. Leta, L. J. Fetters, Y. Liu, M. H. Rafailovich, and J. Sokolov, J. Vac. Sci. Technol. B 14, 1276-1279 (1996).

10 Surface Glass Transition ChemE 554/Overney

34 R. M. Overney, L. Guo, H. Totsuka, M. Rafailovich, J. Sokolov, and S. A. Schwarz, in Dynamics in Small Confining Systems IV, 464, edited by J. M. Drake, J. Klafter, and R. Kopelman (Material Research Society, 1997), pp. 133-144. 35 B. Frank, A. P. Gast, T. P. Russel, H. R. Brown, and C. Hawker, Macromolecules 29, 6531-6534 (1996). 36 M. Rabeony, D. G. Pfeiffer, S. K. Behal, M. Disko, W. D. Dozier, P. Thiyagarajan, and M. Y. Lin, J. Che. Soc. Faraday Trans. 91, 2855-61 (1995). 37 P. F. Green, T. M. Christensen, T. P. Russel, and J. J. Jérôme, J. Chem. Phys. 92, 1478 (1990). 38 T. P. Russel, A. Menelle, S. H. Anastasiadis, S. K. Satija, and C. F. Majkrzak, Macromolecules 24, 6263 (1991). 39 X. Zheng, M. H. Rafailovich, J. Sokolov, Y. Strzhemechny, S. A. Schwarz, B. B. Sauer, and M. Rubinstein, Phys. Rev. Lett. 79, 241-244 (1997). 40 M. Brogley, S. Bistac, and S. J., Macromol. Theor. Simul. 7, 65-68 (1998). 41 C. Buenviaje, S. Ge, M. Rafailovich, J. Sokolov, J. M. Drake, and R. M. Overney, Langmuir 15, 6446-6450 (1999). 42 E. Meyer, R. Overney, K. Dransfeld, and T. Gyalog, Nanoscience: Friction and Rheology on the Nanometer Scale (World Scientific Publishing Co. Ltd., Singapore, 1998). 43 J. A. Forrest, K. Dalnoski-Veress, J. R. Stevens, and J. R. Dutcher, Phys. Rev. Lett. 77, 2002 (1996). 44 P. G. de Gennes, (2000). 45 A. Schallamach, Wear 17, 301-312 (1971). 46 L. E. Govaert, H. G. H. van Melick, and H. E. H. Meijer, Polymer 42, 1271 (2001). 47 R. M. Overney, E. Meyer, J. Frommer, H.-J. Guentherodt, M. Fujihira, H. Takano, and Y. Gotoh, Langmuir 10, 1281-1286 (1994). 48 H. W. Spiess, Annu. Rev. Mater. Sci. 21, 131-158 (1991). 49 C. A. Fyfe, Solid State NMR For Chemists (C.F.C. Press, Guelph, Ontario, 1983). 50 J. D. Ferry, Viscoelastic properties of polymers (Wiley, New York, 1980).

11

Recommended publications