ArticlePDF Available

Abstract and Figures

We have used composition depth profiling, based on nuclear-reaction analysis and secondary ion mass spectroscopy, to study segregation at the free surface of a partly miscible binary mixture consisting of random olefinic copolymers. The equilibrium surface excess data, analysed within a mean-field Cahn approach, point to a wetting transition. The surface phase diagram obtained was confirmed by the observed dynamics of the segregation from a coexistence composition: the monotonic and halted growth of the surface layer was observed at temperatures above and below the predicted wetting point, respectively.
Content may be subject to copyright.
Europhys. Lett.,50 (1), pp. 35–40 (2000)
EUROPHYSICS LETTERS 1 April 2000
Wetting transition in a binary polymer blend
J. Rysz 1, A. Bud kowski 1(),A.Bernasik
2,J.Klein
3,K.Kowalski
2,
J. Jedli´
nski2and L. J. Fetters4
1Smoluchowski Institute of Physics, Jagellonian University
Reymonta 4, 30-059 Krak´ow, Poland
2Surface Spectroscopy Laboratory, University of Mining and Metallurgy and
JCCASR Jagellonian University - Reymonta 23, 30-059 Krak´ow, Poland
3Department of Materials and Interfaces, Weizmann Institute of Science
76100 Rehovot, Israel
4Exxon Research and Engineering Company - Annandale, NJ 08801, USA
(received 15 November 1999; accepted 18 January 2000)
PAC S. 36.20.-r Macromolecules and polymer molecules.
PAC S. 68.10.-m Fluid surfaces and fluid-fluid interfaces.
PAC S. 68.45.Gd Wetting.
Abstract. We have used composition depth profiling, based on nuclear-reaction analysis
and secondary ion mass spectroscopy, to study segregation at the free surface of a partly
miscible binary mixture consisting of random olefinic copolymers. The equilibrium surface
excess data, analysed within a mean-field Cahn approach, point to a wetting transition. The
surface phase diagram obtained was confirmed by the observed dynamics of the segregation
from a coexistence composition: the monotonic and halted growth of the surface layer was
observed at temperatures above and below the predicted wetting point, respectively.
Wetting phenomena are classified primarily [1] in terms of the contact angle Θ, describing
the geometrical arrangement of two coexisting phases φ1and φ2at the surface. An alternative
approach (due to Cahn [2]) is provided by the picture of surface segregation described by
the profile φ(z) of blend composition φvs. distance zfrom the surface [2–5] (fig. 1a)): A
surface composition φs
2decaying smoothly to its bulk value φ1characterises partial
wetting. In the case of complete wetting a macroscopically thick layer of the second phase φ2
(say with thickness larger than the double width of the φ12interface 2w) resides at the
surface and excludes the bulk phase φ1from the surface. In a generic situation, complete
wetting occurs close to the critical temperature TC(critical-point wetting) and the transition
to partial wetting is observed [6,7] at TWfor a larger |TCT|value as predicted by the Cahn
model assuming short-ranged, temperature-independent surface potential fs[2,3, 5]. Different
scenarios, such as reversal wetting transition [8] or the sequence of two transitions [3, 9],
were revealed in recent experiments. The discontinuous (first-order) [6, 9] and continuous
(second-order or critical) [7–9] (generic) wetting transitions have been reported for small
() E-mail: ufbudkow@cyf-kr.edu.pl
c
EDP Sciences
36 EUROPHYSICS LETTERS
molecule systems, where observations could be convoluted by the presence of gravitational and
convective fields. These effects are negligible for thin films composed of polymer blends [10,11].
In addition, macromolecular length scales are large enough to facilitate the determination of
the composition profiles φ(z) with modern profiling techniques [12, 13]: the complete wetting
behaviour [10, 11, 14] and the reversal wetting transition [8] have been observed for various
polymer mixtures, mostly in spinodal demixing experiments [8,14]. The wetting transition in a
polymer blend, despite its relevance to modern technological applications [15], huge theoretical
attention [5] and very intensive research [10,11], so far eluded experimental observation.
In this letter we present results of two types of studies on the surface segregation, free
of the spinodal decomposition effects, indicating the wetting transition at the free surface of
a model binary polymer blend. First, the equilibriu m properties of the segregation from the
one-phase region of phase diagram are analysed within the Cahn approach [2], pointing to the
transition [16]. This is confirmed by experiments focused on the dynamics of the segregation
Fig. 1 Segregation and wetting characteristics of the d75/h66 mixture. Different symbols (data
points) and line types (calculated from a mean-field model [2, 5]) correspond to various temperatures:
T=59C; and T=64C; T=79C; and---T=92C. a) Composition-depth profiles
φ(z) corresponding to the free surface (z= 0) partially (T=64
C, w=30.1 nm) and completely
(T=92
C, w=59.8 nm) wetted from the blend at coexistence. b) Temperature variation of
coexistence compositions φ1and φ2denoted by binodal [17] (dotted line). Bulk compositions φbulk
are marked for which the segregation from the one-phase region and from the coexistence composition
was studied. Solid bars represent the evaluated uncertainty of TC,TW,φ1and φ2. c) Variation of
the equilibrium surface excess zas a function of φbulk1. The inset shows a typical NRA profile
with zmarked by the shaded area. d) The Cahn construction [2] predicting TW=67±5C: The
trajectories 2κ(dφ/dz)vs. φ, marked by thick lines, correspond to the profiles of fig. 1a). The
“bare” surface energy derivative (dfs/dφ), derived from the zdata (panels c) and b)), determines
the starting points (at the surface composition φs) of the trajectories.
J. Rysz et al.:Wetting transition in a binary polymer blend 37
from the coexistence composition.
We used the mixture composed of two random poly(ethylene-ethylethylene) copolymers of
mean microstructure (C4H8)1x(C2H3(C2H5))x. Such blends, where the two components have
different ethylethylene fractions x1and x2, create an attractive model system, as bulk [17–19]
and surface [10, 11, 18, 20–22] interactions may be tailored by a suitable choice of x-values.
The molecular characteristics of the pair used in this study, d75 (x=0.75), which is partially
deuterated, and h66 (x=0.66), are as follows: for d75: degree of polymerisation N=
1625; statistical segment length a=0.64 nm; degree of deuteration = 0.40; glass transition
temperature Tg=46 C; for h66: N= 2030; a=0.68 nm; Tg=54 C. Both polymers
had polydispersity index <1.08. The d75/h66 mixture displays a phase equilibrium with
TC= 101 ±4C and the binodal (see fig. 1b)) described within the Flory-Huggins model by
the interaction parameter χ=(0.371/T 2.7×105)(1+0.212φ), dependent on the d75
volume fraction φ[17].
The segregation equilibrium studies were performed for monolayer samples of various over-
all d75 concentration (and thickness Dca. 500 nm), prepared by spin coating from a toluene
solution onto polished silicon wafers. The segregation dynamics observations were made for
bilayers of a pure h66 film (D= 220 ±50 nm) on top of a pure d75 layer (D= 170–800 nm):
the h66 film was spin coated onto freshly cleaved mica and then float-mounted onto the d75
layer. The samples were annealed in a vacuum oven (102Torr) at temperatures T(stable
to ±1C) for different times t, and stored at T<T
guntil required for the experiments. The
profiles φ(z) of the deuterated d75 blend component normal to the sample surface were deter-
mined either by nuclear reaction analysis (NRA) [12] or by secondary ion mass spectroscopy
(SIMS) [13] with a depth resolution in the range of 9 to 20 nm, as described earlier [12,13].
The segregation from the one-phase region of the phase diagram is described in terms of
the (integrated) surface excess zof d75, represented in the inset to fig. 1c) by the shaded
area. zwas measured from the profiles of the monolayer samples annealed at T=64and
92 Cfort>1 day. The times used were sufficient to reach the equilibrium, due to a large
molecular mobility of the polyolefines [19]. In contrast to the situation at the free surface, no
segregation was observed at the blend interface with the substrate [20]. The equilibrium z
values were determined for bulk compositions φbulk marked by diamonds in the phase diagram
of fig. 1b). Figure 1c) shows the corresponding segregation isotherms with zplotted as a
function of the normalised bulk composition φbulk1.
To analyse the segregation data of fig. 1c) we follow the standard Cahn procedure [2],
reviewed recently [5] and used previously [18,20–22]. The excess free energy functional F[φ]
is represented as a sum of the bulk and “bare” surface fsfree energies, dependent on φ(z)and
its surface value φs, resp ectively [2, 23]:
F[φ]
kBT=
0
dz[∆f(φ)+κ(dφ/dz)2]+fs(φs).(1)
Here κis the φ-dependent coefficient [24] and f(φ) is the energy needed to create a unit
volume with composition φfrom a bulk region with φbulk [24].
To minimise F[φ] we analyse the Cahn construction (fig. 1d)), that is, the φ-dependence
of two quantities: the trajectory 2κ(dφ/dz) (equal to 2(κf)1/2and specified entirely by
bulk parameters [24]) and the derivative (dfs)/dφ) driving the segregation. While each
equilibrium profile φ(z) [25] is determined by its own trajectory, the starting point of the
latter (at φs) is given by the intersection of both, 2κ(dφ/dz) and (dfs/dφ), relations. The
reversal procedure [18, 20–22] allows us to determine the (dfs/dφ) relation, whenever the
profiles φ(z) with the φsvalues, or equivalently the segregation data, are known. Diamonds
38 EUROPHYSICS LETTERS
Fig. 2 Dynamics of surface enrichment from a coexistence phase φ1:T=59
C<T
W(w=
28.3 nm); T=79
C>T
W(w= 38.6 nm). SIMS comp osition-depth profiles φ(z) of the surface
layer (of thickness L) growing from the bulk phase φ1(of width d), a) after 12.6 days at 59 C, b) after
73.2 days at 59 C(z=19.2±1.5 nm), c) after 3.4 days at 79 C corresponding [19] to 18.8 days at
59 C. An additional film, rich in d75 (at z>200–300nm), acted as a material reservoir. d) Variation
of Lwith the parameter (t/d) reduced to 59 C [11]. Dashed lines are a guide to eye.
in fig. 1d) represent the (dfs/dφ)vs. φrelation derived from the experimental input of
fig. 1c). An additional point () corresponds to the final stage of the segregation dynamics
detected at 59 C (fig. 2b)), where the equilibrium surface excess is attained for the bulk
composition φ1. A non-linear shape of the (dfs/dφ) function was observed and examined
previously [18, 20–22] for the other dx1/hx2mixtures. Figure 1d) shows that the (dfs/dφ)
vs. φrelation is, for the values <4×103nm, hardly temperature dependent. This result
enables us to evaluate the wetting point TW=67±5C. Typical computed trajectories
(fig. 1d)) and profiles (fig. 1a)) correspond to partial (solid lines) and complete (dashed lines)
wetting at T=64and92
C, respectively. The Cahn analysis suggests the critical wetting
transition. This model also predicts a phenomenon prerequisite to the continuous transition
(an enrichment-depletion duality), which was observed recently in the d52/h66 blend [22].
To study the dynamics of the segregation from the coexistence composition we followed
the idea of Steiner et al. [10, 11] and measured the bilayer samples annealed at T=59
and 79 C for times from 5 h to 2.4 months. Such long annealing times (10 times longer
than used previously [10, 11]) were necessary to distinguish the partial from the complete
wetting behaviour. Typical profiles corresponding to two temperatures are shown in figs. 2a)-
b) and c), respectively. They reflect the structure of the sample after the transient initial
stage, characterised by interdiffusion leading to coexisting compositions, is completed. The
profiles φ(z) may be divided into three regions: i) the surface layer of thickness L; ii) the bulk
phase φ1of width d[26]; iii) the d75-rich layer acting as a material reservoir. At the late stage
of the annealing a different shape of the surface layer was observed at both temperatures:
J. Rysz et al.:Wetting transition in a binary polymer blend 39
While the profile characteristic of complete wetting was measured at T=79
C (fig. 2c)),
it was never observed at T=59
C for equivalent and even much longer annealing times.
Instead the profile typical of partial wetting was monitored persistently (figs. 2a) and b)).
The profiles resembling fig. 2c) were observed previously [10, 11] for the d88/h78 and
d66/h52 mixtures: It was established that the surface layer thickness Lgrows at the expense
of the material reservoir, while the thickness dof the bulk phase φ1remains unchanged. This
process is limited by the diffusion of the dx1blend component across the bulk phase and
therefore Ldepends on the parameter (t/d) rather than time t. The surface layer thickness L,
growing with (t/d), attained macroscopic dimensions (L/(2w)1.2 [10]) indicating a complete
wetting regime.
To examine the segregation dynamics data of this study we represent them in fig. 2d) as a
plot of Lvs. the parameter (t/d) reduced to 59 C [11]. The data set corresponding to 79 C()
reproduces the monotonic evolution of the surface layer to a macroscopic surface phase φ2(here
with L/(2w)=1.3) reported previously for critical-point wetting [10,11]. On the contrary,
the data points of T=59
C() after an initial increase level off (compare figs. 2a) and b))
at the value L=38±2 nm, smaller than the double interfacial width (2w=56.6 nm). This
observation clearly indicates a completed build-up of the surface enriched layer corresponding
to a partial wetting regime.
There are two conclusions. First, we have demonstrated for the first time the wetting
transition for a polymer blend: the dynamics of the segregation from the coexistence com-
position characteristic of complete and partial wetting was observed in the d75/h66 mixture
at T=79and59
C, respectively (fig. 2d)). Second, this observation is in accord with the
prediction, TW=67±5C, of the Cahn approach based on the surface excess data (fig. 1).
Previously this model has been used to describe other wetting-related phenomena, such as
the enrichment-depletion duality [22] or the extended critical-point wetting regime [21]. The
present results show that for polymer mixtures the Cahn theory can provide reasonable pre-
dictions of the surface phase diagram. Further experimental studies are needed to explain
how this diagram is altered by the long-ranged surface forces [3, 4, 7, 9].
∗∗∗
We thank Profs. K. Binder and U. Steiner for useful discussions. Partial support from
the Reserve for Individual Research of the Rector of Jagellonian University, the German-Israel
Foundation (GIF) and the Ministry of Science and Arts (Israel) is gratefully acknowledged.
REFERENCES
[1] Young T.,Philos. Trans. R. Soc. London,95 (1805) 65.
[2] Cahn J. W.,J. Chem. Phys.,66 (1977) 3667; Schmidt I. and Binder K.,J. Phys. (Paris),
46 (1985) 1631.
[3] de Gennes P.-G.,Rev. Mod. Phys.,57 (1985) 827.
[4] Dietrich S., in Phase Transitions and Critical Phenomena, edited by C. Domb and J. L.
Lebowitz, Vol . II (Academic, New York) 1988, pp. 1-218.
[5] Binder K.,Acta Polym.,46 (1995) 204; Adv. Polym. Sci.,138 (1999) 1 and references therein.
[6] Moldover M. R. and Schmidt J. W.,J. Chem. Phys.,79 (1983) 379; Bonn D., Kellay H.
and Wegdam G. H.,Phys. Rev. Lett.,69 (1992) 1975.
[7] Ragil K. et al.,Phys. Rev. Lett.,77 (1996) 1532.
[8] Genzer J. and Kramer E. J.,Phys. Rev. Lett.,78 (1997) 4946; Europhys. Lett.,44 (1998)
180.
40 EUROPHYSICS LETTERS
[9] Shahidzadeh N. et al.,Phys. Rev. Lett.,80 (1998) 3992.
[10] Steiner U., Klein J., Eiser E., Budkowski A. and Fetters L. J.,Science,258 (1992)
1126.
[11] Steiner U. and Klein J.,Phys. Rev. Lett.,77 (1996) 2526; Mater. Res. Soc. Symp. Proc.,464
(1997) 121.
[12] Kerle T. et al.,Acta Polymer.,48 (1997) 548.
[13] Bernasik A. et al., in ECASIA ’97, edited by I. Olefjord, L. Nyborg and D. Bryggs (John
Wiley & Sons, Chichester) 1997, pp. 775-778; Schwarz S. A. et al.,Mol. Phys.,76 (1992) 937.
[14] Bruder F. and Brenn R.,Phys. Rev. Lett.,69 (1992) 624; Krausch G. et al.,Macromolecules,
26 (1993) 5566; Straub W. et al.,Europhys. Lett.,29 (1995) 353.
[15] Mayes A. M. and Kumar S. K.,MRS Bulletin, January issue (1997) 43; Service R. F.,
Science,278 (1997) 383; B¨
oltau M. et al.,Nature,391 (1998) 877.
[16] Such a conclusion, valid for the short-ranged surface potential fs, might not be true if the long-
range interfacial interactions are also present [11]. To check this conclusion we have performed
the dynamics experiments.
[17] Scheffold F. et al.,J. Chem. Phys.,104 (1996) 8786.
[18] Budkowski A.,Adv. Polym. Sci.,148 (1999) 1; Budkowski A. et al.,J. Polym. Sci. Polym.
Phys. Ed.,36 (1998) 2691.
[19] Losch A. et al.,J. Polym. Sci. Polym. Phys. Ed.,33 (1995) 1821.
[20] Scheffold F. et al.,J. Chem. Phys.,104 (1996) 8795.
[21] Budkowski A., Scheffold F., Klein J. and Fetters L. J.,J. Chem. Phys.,106 (1997) 719.
[22] Budkowski A., Rysz J., Scheffold F. and Klein J.,Europhys. Lett.,43 (1998) 404; Bud-
kowski A. et al.,Vacuum,54 (1999) 273.
[23] Equation (1) corresponds to the long-wavelength limit: R<1, where R=a(N/6)1/2(dφ/dz)max .
R=0.1atT=64C for φbulk =φ1,N=Nh66 and a=ah66.
[24] κ= ((1 φ)a2
d75 +φa2
h66)/(36φ(1 φ)), f(φ)=f(φ)f(φbulk )(φφbulk)(∂f /∂φ)φbulk,
f(φ)=φln φ/Nd75 +(1φ)ln(1φ)/Nh66 +χφ(1 φ).
[25] The replacement of the short- [18, 20–22] by long- [11] ranged fsterm in eq. (1) leads to no
experimentally detectable changes in the equilibrium profile φ(z) corresponding to the one-
phase region of the phase diagram; JonesR.A.L.,Phys. Rev. E,47 (1993) 1437; Genzer J.,
Faldi A., Oslanec R. and Composto R. J.,Macromolecules,20 (1996) 5438.
[26] A depletion region of the bulk phase φ1, adjacent to the surface layer, with local concentration
φdφ1was also detected [11].
Article
Full-text available
Within self-consistent field theory and Monte Carlo simulations the phase behavior of a symmetrical binary AB polymer blend confined into a thin film is studied. The film surfaces interact with the monomers via short ranged potentials. One surface attracts the A component and the corresponding semi-infinite system exhibits a first order wetting transition. The surface interaction of the opposite surface is varied as to study the crossover from capillary condensation for symmetric surface fields to interface localization/delocalization transition for antisymmetric surface fields. In the former case the phase diagram has a single critical point close to the bulk critical point. In the latter case the phase diagram exhibits two critical points which correspond to the prewetting critical points of the semi-infinite system. Only below a triple point there is a single two-phase coexistence region. The crossover between these qualitatively different limiting behaviors occurs gradually, however, the critical temperature and the critical composition exhibit a non-monotonic dependence on the surface field. The dependence of the phase behavior for antisymmetric boundaries is studied as a function of the film thickness and the strength of the surface interactions. Upon reducing the film thickness or decreasing the strength of the surface interactions we can change the order of the interface localization/delocalization transition from first to second. The role of fluctuations is explored via Monte Carlo simulations of a coarse grained lattice model. Close to the (prewetting) critical points we observe 2D Ising critical behavior. Also, there is a rich crossover behavior between Ising critical, tricritical and mean field behavior. At lower temperatures capillary waves of the AB interface lead to a pronounced dependence of the effective interface potential on the lateral system size.
Article
Versatility of solution-processing strategy based on the simultaneous rather than additive deposition of different functional molecules is discussed. It is shown that spin-cast polymer blends result in films with domains that could form elements with complementary functions of (i) solar cells, (ii) electronic circuitries, and (iii) test plates for protein micro-arrays: Alternating layers, rich in electron-donating polyfluorene and electron-accepting fullerene derivative, result in optimized solar power conversion. Surface patterns, made by soft lithography, align conductive paths of conjugated poly(3-alkylthiophene) in dielectric polystyrene. Proteins, preserving their biologically activity, are adsorbed to hydrophobic domains of polystyrene in hydrophilic matrix of poly(ethylene oxide). The authors report the research progress on structure formation in three polymer blend families, resulting in films with complementary elements for electronics and biotechnology. Blend film structures are determined with secondary ion mass spectrometry, atomic force microscopy, and fluorescence microscopy. In addition, the authors present recent results on (i) structure formation in fullerene derivative/poly(3-alkylthiophene) blends intended for solar cells, (ii) 3-dimensional SIMS imaging of conductive paths of poly(3-alkylthiophene) in dielectric polystyrene, (iii) test plates for multiprotein micro-arrays fabricated with blend films of hydrophobic polystyrene and thermoresponsive poly(N-isopropylacrylamide). (C) 2012 Wiley Periodicals, Inc. J Appl Polym Sci, 2012
Article
A polymer vs solvent diagram of film structures, formed in polystyrene (PS) blends (1:1 w:w PS/PT) with poly(3-alkylthiophenes) PT [regioregular R-P3DDT, R-P3HT and regiorandom P3BT, P3DDT] spin-coated onto oxidized silicon surfaces from various common solvents [p-xylene, toluene, chloroform, chlorobenzene, cyclohexanone] is presented. The structures were determined with microscopic techniques (atomic, AFM and lateral, LFM, force microscopy, fluorescent microscopy FM) and dynamic secondary ion mass spectrometry (dSIMS). The diagram, arranged according to the solubility parameter of the PTs and the solvents, exhibits three main structural classes: dewetting, lamellar, and lateral (quasi-2-dim) morphology. Decrease in PT solubility parameter δPT inhibits dewetting of polymer films. It induces also a transition from lamellar to lateral film structure. Increase in solvent solubility parameter δS has similar effects. Such behavior is related to the stability of transient homogeneous films and multilayers in the course of spin-casting. The role of δPT and δS is elucidated based on the stability analysis performed in terms of spreading coefficient (dependent on δPT) and effective interfacial tension of solvent-rich polymer phase (dependent on δS).
Article
In this study the morphology of spin-casted films of polymers blended with [6,6]-phenyl C61-butyric acid methyl ester (PCBM) has been studied. It was found that the lateral structure formation in the films is favored by rapid solvent evaporation and strong polymer−PCBM repulsion. The formation of homogeneous films is favored by slow evaporation and weak polymer−PCBM repulsion. The effect of solvent evaporation rate is the opposite of what is found for spin-casting polymer−polymer blends. The results can be explained by the kinetics of phase separation and the phase behavior involving limited solubility and crystallization of PCBM.
Article
We have examined the effect of deuterium labeling on surface interactions in mixtures of random olefinic copolymers [C4H8]1−x[C2H3(C2H5)]x. Based on surface segregation data we have determined a surface energy difference χs between pure blend constituents. In each binary mixture components have different fractions x1, x2 of the group C2H3(C2H5), and one component is labeled by deuterium (dx) while the other is hydrogenous (hx). The mixtures are grouped in four pairs of structurally identical blends with swapped labeled constituent (dx1/hx2, hx1/dx2). For each pair the surface energy parameter χs increases when the component with higher fraction x is deuterated, i.e., χs(dx1/hx2) > χs(hx1/dx2) for x1 > x2. A similar pattern has been found previously for the bulk interaction parameter χ. This is explained by the solubility parameter formalism aided by the lattice theory relating the surface excess to missing-neighbor effect. χs has also an additional contribution, insensitive to deuterium swapping effect, and related to entropically driven surface enrichment in a more stiff blend component with a lower fraction x. Both enthalpic and entropic contributions to χs seem to depend on the extent of chemical mismatch between blend components. © 1998 John Wiley & Sons, Inc. J Polym Sci B: Polym Phys 36: 2691–2702, 1998
Article
Surface segregation in thin films of binary liquid mixtures consisting of random olefine copolymers was studied by nuclear reaction analysis over a wide temperature and composition range. A divergence of the surface excess Γ was indicated as the binodal of each mixture was approached from the one-phase region, even at temperatures 100 °C below the critical point Tc, and interpreted as the advent of complete wetting behavior. A consistent description of the adsorption isotherms in terms of a mean field approach assuming a short-ranged surface potential fs is feasible, but requires an unexpected temperature dependence of fs. This dependence causes the wetting transition temperature to be located lower than expected on the basis of present models.
Article
It is shown that in any two-phase mixture of fluids near their critical point, contact angles against any third phase become zero in that one of the critical phases completely wets the third phase and excludes contact with the other critical phase. A surface layer of the wetting phase continues to exist under a range of conditions when this phase is no longer stable as a bulk. At some temperature below the critical, this perfect wetting terminates in what is described as a first-order transition of the surface. This surface first-order transition may exhibit its own critical point. The theory is qualitatively in agreement with observations.
Article
According to a common viewpoint the surface of the binary polymer mixture A/B could be enriched only in one component, say A, regardless of the value of its bulk volume fraction φbulk. From recent theoretical analyses and Monte Carlo simulations one can expect however that some mixtures can exhibit surface segregation in the minority blend component, i.e. enrichment in the component A for φbulk50%. Using composition-depth profiling techniques, we have observed both types of the segregation with only one- or both-blend component(s) preferred at the surface. These results were obtained for model mixtures composed of partly deuterated (dx) and protonated (hx) random olefinic copolymers of the structure –[C4H8]1-x [C2H3(C2H5)]x–. A simple mean-field model is presented to explain both situations.
Article
It has already been asserted, by Mr. Monge and others, that the phenomena of capillary tubes are referable to the cohesive attraction of the superficial particles only of the fluids em­ployed, and that the surfaces must consequently be formed into curves of the nature of lintearias, which are supposed to be the results of a uniform tension of a surface, resisting the pressure of a fluid, either uniform, or varying according to a given law. Segner, who appears to have been the first that maintained a similar opinion, has shown in what manner the principle may be deduced from the doctrine of attraction, but his demonstration is complicated, and not perfectly satisfactory; and in applying the law to the forms of drops, he has neglected to consider the very material effects of the double curvature, which is evidently the cause of the want of a perfect coinci­dence of some of his experiments with his theory. Since the time of Segner, little has been done in investigating accurately and in detail the various consequences of the principle. It will perhaps be most agreeable to the experimental phi­losopher, although less consistent with the strict course of logical argument, to proceed in the first place to the comparison of this theory with the phenomena, and to inquire afterwards for its foundation in the ultimate properties of matter. But it is necessary to premise one observation, which appears to be new, and which is equally consistent with theory and with experiment; that is, that for each combination of a solid and a fluid, there is an appropriate angle of contact between the surfaces of the fluid, exposed to the air, and to the solid. This angle, for glass and water, and in all cases where a solid is perfectly wetted by a fluid, is evanescent: for glass and mer­cury, it is about 140°, in common temperatures, and when the mercury is moderately clean.
Article
Dynamic secondary ion mass spectrometry (SIMS) has recently been employed to obtain high resolution depth profiles in polymer blend thin films and is now regarded as a key probe of surface and interfacial segregation in these systems. Segregation phenomena strongly impact blend properties such as adhesion, friction and weatherability. The strengths and limitations of the SIMS polymer profiling technique are described and contrasted with the complementary techniques of forward recoil elastic scattering (FRES) and neutron reflectivity (NR). The procedures developed for sample preparation and data acquisition are discussed. Experimental results for the effect of incident O2+ energy and angle on depth resolution and sputtering rate in polystyrene (PS) are presented. Ongoing SIMS studies of model blend systems are described: Segregation from dPS (deuterated)/PS blends to vacuum and Si interfaces is examined as a function of the molecular weight of the blend components and preparation of the Si substrate, revealing the importance of long range interactions. Similar behaviour in an acrylonitrile blend is demonstrated. The surface segregation profiles in the immiscible blend PBrS (polybromostyrene)/PS are discussed for samples annealed in the one and two phase regions. The conformation of carboxy terminated PS and dPS chains grafted to the Si oxide interface in a melt is studied as a function of grafting density, temperature, and matrix molecular weight. Diffusion of homopolymer dPS in a crosslinked PS matrix is examined and the observed diffusion coefficients are in good agreement with calculated values using rubber elasticity theory. Interdiffusion of PS/PS bilayer samples annealed above the glass transition temperature is studied. Trapped oxygen at the original bilayer interface is observed, indicating UV crosslinking of the individual film surfaces.
Article
We investigate a wetting reversal transition in thin films of two-phase mixtures of poly(ethylene-propylene) (PEP) and its deuterated analog (dPEP) on substrates covered by self-assembled monolayers (SAM) whose surface energy, gammaSAM, is tuned by varying the SAM composition. As gammaSAM increases from 21 to 24 mJ/m2, a transition from a dPEP/PEP/dPEP/SAM to a dPEP/PEP/SAM structure occurs at increasing TC-T, where TC and T are the critical and transition temperatures, respectively. The dependence of T on gammaSAM is predicted by a simple model from surface and interfacial energies of PEP/dPEP.
Article
We report an ellipsometry study of the wetting of hexane on water. By adding salt to the water, we are able to tune the Hamaker constant of this system. This allows us to demonstrate, for the first time, that two rather than one wetting transitions can exist in a single system. Upon increasing the temperature, a discontinuous (first-order) transition from a microscopic film to a mesoscopic film occurs, followed by a continuous (critical) wetting transition that leads to a thick adsorbed film. The latter is due to the Hamaker constant which changes sign with temperature. The first-order transition temperature changes by the same amount as the critical wetting temperature upon changing the Hamaker constant.
Article
Using nuclear reaction analysis, we have measured the enrichment by one of the components at the surface of a binary mixture of random olefinic copolymers, with components of monomer structure E1−x1EEx1 and E1−x2EEx2. Here E and EE are the linear ethylene and branched ethylethylene groups (C4H8) and [C2H3(C2H5)], respectively, and x represents the fraction of the EE group randomly distributed on the chains. We examined 12 different couples covering a range x=0.38–0.97. The mixtures, whose thermodynamic behavior was established in our earlier paper, were cast in the form of films on both a silicon and on a gold‐covered silicon surface, and were investigated in the one‐phase region of the binodal in the vicinity of the critical temperature. We find that it is always the more flexible component—the one with a shorter statistical step length, corresponding to the higher ethylethylene fraction (higher x)—that is enriched at the polymer/air surface. Within our resolution neither component is enriched at the polymer/solid interface. These results show clearly that enthalpic rather than entropic factors dominate the surface potential driving the surface enrichment. For two of the mixtures we determined the excess of the surface‐preferred species as a function of mixture composition along an isotherm in the one‐phase region of the binodal. A consistent description of our data in terms of a mean‐field model is provided by including in the surface potential a term in the mixture composition gradient at the polymer surface. © 1996 American Institute of Physics.