ArticlePDF Available

Effect of Low Surface Energy Chain Ends on the Glass Transition Temperature of Polymer Thin Films

Authors:

Abstract

Many polymer thin film applications are intimately related to the chemical functionality of the chain end groups. For example, use of hydrophilic end groups is a common strategy to end-graft hydrophobic polymer chains onto high-energy substrate surfaces; introduction of polar end groups to the polyfluoroalkyl ether improves its ability to protect a computer hard disk against wear; using high molecular weight end-functionalized chains as additives can improve the durability of polymer thin films against deweting; polymer surfaces that are modified by thermodynamically segregated low-energy chain ends can be exploited to make “self-healing” surfaces.4 Previous experiments showed that polymer surfaces enriched with segregated chain ends exhibit higher molecular mobility,5-7 which are consistent with theoretical predictions8 assuming the generally larger free volume associated with chain ends compared to chain segments. A local interfacial layer with higher molecular mobility has been attributed to be the cause of a global reduction in the Tg of polymer thin films observed recently. In this study, we investigate the Tg of thin films of PS with low surface energy fluoroalkylsilyl (Rf) termination on both ends (alpha,omega-PS-(Rf)2). Results show that the amount of reduction in the Tg found in thin films of alpha,omega-PS-(Rf)2 with decreasing film thickness is about twice of that found in thin films of sec-butyl-initiated-proton-terminated PS.
Effect of Low Surface Energy Chain Ends on
the Glass Transition Temperature of Polymer
Thin Films
Fengchao Xie, H. F. Zhang, Fuk Kay Lee,
Binyang Du, and Ophelia K. C. Tsui*
Department of Physics, Hong Kong University of Science
and Technology, Clear Water Bay, Kowloon, Hong Kong
Y. Yokoe, K. Tanaka, A. Takahara, and T. Kajiyama
Department of Applied Chemistry, Faculty of Engineering,
Kyushu University, 6-10-1 Hakozaki, Higashi-ku,
Fukuoka 812-8581, Japan
Tianbai He
State Key Laboratory of Polymer Physics and Chemistry,
Changchun Institute of Applied Chemistry, Chinese
Academy of Sciences, Changchun, Jilin 130022, China
Received September 27, 2001
Introduction. Many polymer thin film applications
are intimately related to the chemical functionality of
the chain end groups. For example, use of hydrophilic
end groups is a common strategy to end-graft hydro-
phobic polymer chains onto high-energy substrate sur-
faces;1introduction of polar end groups to the polyfluoro-
alkyl ether improves its ability to protect a computer
hard disk against wear;2using high molecular weight
end-functionalized chains as additives can improve the
durability of polymer thin films against deweting;3
polymer surfaces that are modified by thermodynami-
cally segregated low-energy chain ends can be exploited
to make “self-healing” surfaces.4Previous experiments
showed that polymer surfaces enriched with segregated
chain ends exhibit higher molecular mobility,5-7which
are consistent with theoretical predictions8assuming
the generally larger free volume associated with chain
ends compared to chain segments. A local interfacial
layer with higher molecular mobility has been attrib-
uted to be the cause of a global reduction in the Tgof
polymer thin films observed recently.9-14 In this study,
we investigate the Tgof thin films of PS with low surface
energy fluoroalkylsilyl (Rf) termination on both ends
(R,ω-PS-(Rf)2). Results show that the amount of reduc-
tion in the Tgfound in thin films of R,ω-PS-(Rf)2with
decreasing film thickness is about twice of that found
in thin films of sec-butyl-initiated-proton-terminated PS
(H-PS).
Experimental Section. The R,ω-PS-(Rf)2(Mn)
47.8K Da, Mw/Mn)1.18, TgDSC )376 K) was synthe-
sized by living anionic polymerization using potassium
naphthalene as a bifunctional anionic initiator and
(tridecafluoro-1,1,2,2-tetrahydrooctyl)dimethylchloro-
silane as a terminator. The H-PS, synthesized by anionic
polymerization using sec-butyllithium as the initiator,
had been purchased from Scientific Polymer Products
(abbreviated com. H-PS,15 Mn)530K Da, Mw/Mn)1.04,
TgDSC )382 K) and also self-made in our own laboratory
(abbreviated home H-PS, Mn)52K Da, Mw/Mn)1.05,
TgDSC )372 K). Thin films of the polymers with different
thicknesses (8-200 nm) were prepared by spin-coating
solutions of the polymers in toluene (0.3-4 wt %) onto
cleaned silicon substrates16 covered with a 1.2 nm
thick native oxide layer. Upon spin-coating, the R,ω-
PS-(Rf)2thin films were kept under vacuum at ambient
temperature for 3 days to remove the residual solvent.
Before measurement, all samples were annealed at
or above 120 °C for 5-9 h inside a vacuum oven to
relax the polymer, whereupon they were allowed to
cool to room temperature under vacuum. Spectroscopic
ellipsometric measurements over wavelengths 350-1300
nm were carried out in a J.A. Woollam (Lincoln, NE)
variable-angle spectroscopic ellipsometer (VASE). A
home-built hot-stage with temperature controllable to
within (1 °C was incorporated to heat the sample in
situ. To obtain the temperature scans, both the wave-
length and incident angle of incoming light were fixed
at settings empirically found to optimize the VASE
signal;9the ellipsometric angles, δand ξ, which are
directly related to the film thickness and sample refrac-
tive index,9were recorded as a function of temperature,
T, at a constant heating rate of 2 K/min. The Tgof a
polymer film was determined as the temperature at
which slope of its the temperature dependence data
displays a discontinuity.
Results and Discussion. Shown in Figure 1 are the
measured results plotted as Tg(t)-Tg()vst, where t
is the film thickness and Tg() is the asymptotic Tgin
the limit of large t. The data exhibit a monotonic
reduction in Tg(t) with decreased t, in accordance with
previous measurements.9-14 A number of models have
been proposed to explain this behavior.9,11-14 While they
may differ by the physical process involved, all these
models are based on the idea that a nanometer scale
mobile (or lower Tg) region exists, whose influence on
the Tgof the film increases as the film thickness is
decreased, and end in the same phenomenological
expression: Tg(t))Tg
[1 -(ξ0/t)ν], where ξ0is the
length scale of the mobile region and νa constant. We
have model-fitted our data to this expression and found
excellent agreement (smooth lines). From Figure 1, one
may also notice that data points of the two H-PS fall
within each other’s experimental error bars (or 1σ
confidence interval) and hence are indistinguishable. On
the other hand, the Tgof the R,ω-PS-(Rf)2thin films are
noticeably more reduced (by 2 times, which amounts
to a sizable difference of 3.5σon average). According
to Table 1, the fitted value of νand ξ0are within the
range of values obtained previously.9,13 Comparing
between polymers, the fitted values of νvary only by
Figure 1. Tg(t)-Tg() vs thickness, t, for thin films of PS
with different Mnand chain ends. Smooth lines through the
data are fits to the phenomenological expression discussed in
the text.
1491Macromolecules 2002, 35, 1491-1492
10.1021/ma011689a CCC: $22.00 © 2002 American Chemical Society
Published on Web 01/24/2002
10%. However, the fitted value of ξ0for the R,ω-PS-
(Rf)2films is about 2 times larger than those of the H-PS
films, suggesting a doubling in the length scale of the
mobile region in the former. In the following, we will
try to understand these different results in terms of the
degree of surface segregation of chain ends and surface
molecular motion in these two kinds of PS.
In H-PS, numerous measurements showed that the
sec-butyl end groups preferentially segregated to the
polymer surface.5,17 The preferential segregation of
surface chain ends is attributable to the enhanced
surface molecular motion found in H-PS by scanning
viscoelasticity microscopy (SVM)5and lateral force
microscopy (LFM),6as a later finding7showed that no
enhanced surface molecular mobility was present in PS
with higher surface energy chain end groups, namely
R,ω-PS-(COOH)2and R,ω-PS-(NH2)2.
In the fluoroalkylsilyl-terminated PS, angle-depend-
ent X-ray photoelectron spectroscopy (ADXPS) and
neutron reflectivity revealed that the polymer surface
was 5 times more concentrated with the fluoroalkyl-
silyl end groups than the bulk,18 which is about 2.5
times bigger than that found in H-PS.4In PS terminated
with both kinds of end groups, i.e., R-H-PS-ω-(Rf), the
fluoroalkylsilyl end group had been found to dominate
at the free surface.17 LFM study also revealed that
surface relaxations began to be discernible at a much
lower temperature (140 K) in R,ω-PS-(Rf)2than in
H-PS (300 K). It is thus justifiable to presume that
the degree of segregation of chain ends to the free
surface and the surface molecular mobility are substan-
tially larger with the fluoroalkylsilyl termination than
with the sec-butyl-initiated proton termination.
These previous findings enable useful insights to the
understanding of data displayed in Figure 1. The
notably lower Tgfound in R,ω-PS-(Rf)2than in H-PS thin
films reveals that when the surface of a polymer is
sufficiently plasticized, the reduction in the Tgof the
polymer confined in thin films may be promoted. To
inspect whether this conclusion can be generalized to
any kind of chain end that causes a lowering in the Tg
of the surface, we examine data of the two H-PS
displayed in Figure 1. Clearly, the same degree of
depression in Tgprevails in thin films of both polymers
despite their notably different Mn. This is consistent
with previous experiments on the Tgof similar system,
namely uncapped H-PS supported on Si substrates,
where no Mndependence was discernible for Mn)12.4-
2900 kDa.9,19,20 Correlating these findings with results
of ref 6 showing that the surface molecular motions of
H-PS with Mn<30K Da are enhanced, one may
conclude that existence of a plasticized surface layer
does not always warrant a promotion in the depression
of Tgin supported polymer films.
In conclusion, we have measured the Tgof thin films
of PS with sec-butyl-initiated proton terminations and
those with the lower energy fluoroalkylsilyl termination.
In comparison, the latter exhibit a drastically higher
reduction in the Tgwith decreasing film thickness. The
present result evidences the first time that a surface
layer sufficiently plasticized by segregated chain ends
may have a global effect on the Tgof a polymer film.
Acknowledgment. This work was supported by
HKUST through the University Grants Council of Hong
Kong under Project DAG 98/99.SC24.
References and Notes
(1) Mansky, P.; Liu, Y.; Huang, E.; Russell, T. P.; Hawker, C.
Science 1997,275, 1458.
(2) Marina, R.; Granick, S. J. Phys. Chem. B 1999,103, 8711.
(3) Henn, G.; Bucknall, D. G.; Stamm, M.; Vanhoorne, P.;
Je´roˆme, R. Macromolecules 1996,29, 4305.
(4) Schaub, T.; Kellogg, G. J.; Mayes, A. M.; Kulasekere, R.;
Ankner, J. F.; Kaiser, H. Macromolecules 1996,29, 3982.
(5) Tanaka, K.; Taura, A.; Ge, S.; Takahara, A.; Kajiyama, T.
Macromolecules 1996,29, 3040.
(6) Kajiyama, T.; Tanaka, K.; Takahara, A. Macromolecules
1997,30, 280.
(7) Tanaka, K.; Jiang, X.; Nakamura, K.; Takahara, A.; Kaji-
yama, T.; Ishizone, T.; Hirao, A.; Nakahama, S. Macromol-
ecules 1998,31, 5148.
(8) Mayes, A. M. Macromolecules 1994,27, 3114.
(9) Keddie, J. L.; Jones, R. A. L.; Cory, R. A. Faraday Discuss.
1994,98, 219.
(10) Forrest, J. A.; Dalnoki-Veress, K.; Stevens, J. R.; Dutcher,
J. R. Phys. Rev. Lett. 1996,77, 2002.
(11) Dalnoki-Veress, K.; Forrest, J. A.; de Gennes, P. G. J. Phys.
IV 2000,10, Pr7-221.
(12) Forrest, J. A.; Mattsson, J. Phys. Rev. E 2000,61, R53.
Mattsson, J.; Forrest, J. A.; Bo¨ rjesson, L. Phys. Rev. E 2000,
62, 5187.
(13) Fukao, K.; Miyamoto, Y. Phys. Rev. E 2000,61, 1743.
(14) Tsui, O. K. C.; Russell, T. P.; Hawker, C. J. Macromolecules
2001,34, 5535.
(15) The synthesis of the com. H-PS has been confirmed by the
supplier to be anionic polymerization using sec-butyllithium
as the initiator.
(16) Wang, X. P.; Xiao, X.; Tsui, O. K. C. Macromolecules 2001,
34, 4180.
(17) Elman, J. F.; Johs, B. D.; Long, T. E.; Koberstein, J. T.
Macromolecules 1994,27, 5341.
(18) Tanaka, K.; Yokoe, Y.; Takahara, A.; Kajiyama, T. Manu-
script in preparation.
(19) DeMaggio, G. B.; Frieze, W. E.; Gidley, D. W. Phys. Rev.
Lett. 1997,78, 1524.
(20) Tsui, O. K. C.; Zhang, H. F. Macromolecules, in press.
MA011689A
Table 1. Comparison between Fitted Parameters for
Tg(t)-Tg()vstData of Figure 1 by Using the
Phenomenological Expression Discussed in the Text
sample ξ0(nm) νTg() (K)
R,ω-PS-(Rf)21.46 (0.23 1.37 (0.1 375 (0.9
com. H-PS (Mn)530K) 0.86 (0.12 1.35 (0.06 373.5 (0.2
home H-PS (Mn)52K) 0.72 (0.06 1.22 (0.03 372.2 (0.3
1492 Communications to the Editor Macromolecules, Vol. 35, No. 5, 2002
... [81][82][83][84] A fundamental underpinning of Tg-confinement effects in polymers relates to the role of interfaces and how they modify cooperative segmental mobility. 1,16,21,[26][27][28][29][30]33,43,44,50,52 For example, free-surface effects, 1,2,16,17,[20][21][22][23][26][27][28][29][30][32][33][34][35][36][38][39][40][41][42][43]45,[47][48][49][50][51][52][53][54][61][62][63][64][65]67,[71][72][73][74] i.e., effects originating at the polymer-air interface, are understood to accommodate cooperative segmental mobility at a lower temperature than in the bulk state, [21][22][23][30][31][32][33][34][35][36]38,43,44,49,50,52,63,66,67,69 resulting in a reduced local Tg that can impact the overall average Tg in a nanoconfined film. Many examples of Tg reductions in thin polymer films that lack attractive polymer-substrate interactions at the substrate interface have been documented, 1,2,16,17,[19][20][21][22][23][26][27][28][29][30][31][32][33][34][35][36][37][38][39][40][41][42]45,48,[52][53][54][55][56][57][58][59][61][62][63][64][65]67,68,[70][71][72][73] with reports of Tg reductions as large as 53 ○ C in supported films 22 and ∼80 ○ C in freely standing films. ...
... [81][82][83][84] A fundamental underpinning of Tg-confinement effects in polymers relates to the role of interfaces and how they modify cooperative segmental mobility. 1,16,21,[26][27][28][29][30]33,43,44,50,52 For example, free-surface effects, 1,2,16,17,[20][21][22][23][26][27][28][29][30][32][33][34][35][36][38][39][40][41][42][43]45,[47][48][49][50][51][52][53][54][61][62][63][64][65]67,[71][72][73][74] i.e., effects originating at the polymer-air interface, are understood to accommodate cooperative segmental mobility at a lower temperature than in the bulk state, [21][22][23][30][31][32][33][34][35][36]38,43,44,49,50,52,63,66,67,69 resulting in a reduced local Tg that can impact the overall average Tg in a nanoconfined film. Many examples of Tg reductions in thin polymer films that lack attractive polymer-substrate interactions at the substrate interface have been documented, 1,2,16,17,[19][20][21][22][23][26][27][28][29][30][31][32][33][34][35][36][37][38][39][40][41][42]45,48,[52][53][54][55][56][57][58][59][61][62][63][64][65]67,68,[70][71][72][73] with reports of Tg reductions as large as 53 ○ C in supported films 22 and ∼80 ○ C in freely standing films. ...
... 1,16,21,[26][27][28][29][30]33,43,44,50,52 For example, free-surface effects, 1,2,16,17,[20][21][22][23][26][27][28][29][30][32][33][34][35][36][38][39][40][41][42][43]45,[47][48][49][50][51][52][53][54][61][62][63][64][65]67,[71][72][73][74] i.e., effects originating at the polymer-air interface, are understood to accommodate cooperative segmental mobility at a lower temperature than in the bulk state, [21][22][23][30][31][32][33][34][35][36]38,43,44,49,50,52,63,66,67,69 resulting in a reduced local Tg that can impact the overall average Tg in a nanoconfined film. Many examples of Tg reductions in thin polymer films that lack attractive polymer-substrate interactions at the substrate interface have been documented, 1,2,16,17,[19][20][21][22][23][26][27][28][29][30][31][32][33][34][35][36][37][38][39][40][41][42]45,48,[52][53][54][55][56][57][58][59][61][62][63][64][65]67,68,[70][71][72][73] with reports of Tg reductions as large as 53 ○ C in supported films 22 and ∼80 ○ C in freely standing films. 20 These effects have sometimes been ascribed to increased free volume at and near the free surface 30,49,50 and in other cases simply to a greater thermal expansivity with nanoconfinement. ...
Article
Nanoconfined poly(4-methylstyrene) [P(4-MS)] films exhibit reductions in glass transition temperature (Tg) relative to bulk Tg (Tg,bulk). Ellipsometry reveals that 15-nm-thick P(4-MS) films supported on silicon exhibit Tg − Tg,bulk = − 15 °C. P(4-MS) films also exhibit fragility-confinement effects; fragility decreases ∼60% in going from bulk to a 20-nm-thick film. Previous research found that incorporating 2–6 mol % 2-ethylhexyl acrylate (EHA) comonomer in styrene-based random copolymers eliminates Tg- and fragility-confinement effects in polystyrene. Here, we demonstrate that incorporating 3 mol % EHA in a 4-MS-based random copolymer, 97/3 P(4-MS/EHA), eliminates the Tg- and fragility-confinement effects. The invariance of fragility with nanoconfinement of 97/3 P(4-MS/EHA) films, hypothesized to originate from the interdigitation of ethylhexyl groups, indicates that the presence of EHA prevents the free surface from perturbing chain packing and the cooperative mobility associated with Tg. This method of eliminating confinement effects is advantageous as it relies on the simplest of polymerization methods and neat copolymer only slightly altered in composition from homopolymer. We also investigated whether we could eliminate the Tg-confinement effect with low levels of 2-ethylhexyl methacrylate (EHMA) in 4-MS-based or styrene-based copolymers. Although EHMA is structurally nearly identical to EHA, 4-MS-based and styrene-based copolymers incorporating 4 mol % EHMA exhibit Tg-confinement effects similar to P(4-MS) and polystyrene. These results support the special character of EHA in eliminating confinement effects originating at free surfaces.
... Kajiyama and co-workers proposed that the enrichment of chain-end groups led to the presence of excess free volume on/near the surface of the film [ 72-74 , 246 ], the surface layer was sufficiently plasticized by these segregated end groups, and thus induced a T g reduction on/near the surface [ 4 9 , 74 , 197 , 220 , 228 , 243 , 26 8 ]. They also found that PS end groups with lower surface energy exhibited a significantly larger T g reduction due to more enhanced surface mobility [218] . However, Bliznyuk and coworkers claimed that the entanglements, especially interchain entanglements, rather than segregation of the end groups, play a critical role in the T g depression at the surface [247] . ...
Article
The glass transition temperature (Tg) of polymer thin films, one of the most important parameters to characterize relaxation behavior of polymer chains in a thin film, has been extensively investigated. It was found that the relaxation behavior of polymer chains in a polymer thin film at different locations, e.g., near a free surface, near a buried interface, or inside a given position of the film bulk, can vary greatly. Such results are significant for both fundamental understanding of molecular behavior and practical use of polymer thin films in many important applications but have not been systematically summarized and reviewed. In this review, we look at two extensively studied and commonly used polymer thin films, i.e., poly(methyl methacrylate) (PMMA) and polystyrene (PS), as examples to systematically examine the effects of the free surface, buried interface (substrate), and geometrical confinement (encapsulated between two substrates) on the surface, interfacial, and bulk relaxation behavior of polymer thin films. Especially, the universal effects of “hard” and “soft” confinements and adsorbed nanolayer near a substrate on Tg are reviewed in detail. In addition, other important polymers such as poly(2-vinylpyridine) and the effects of other variables such as annealing are also discussed. This review elucidates variables which influence the relaxation behavior of polymer chains on the free surface, at buried interfaces, or in bulk region of polymer thin films, providing in-depth understandings on such behavior. The systematic knowledge reported here will help to guide future design of polymer thin films with desired properties used in a variety of applications, ranging from multilayer polymer packaging materials, polymer sheets, plastic microelectronics, to batteries and solar cells, and beyond.
Article
Full-text available
This review provides a comprehensive overview on the effects of plasticization on microporous polymer membranes, as well as strategies to mitigate this phenomenon for gas separation applications.
Article
Surface glass transition temperature (Tgsurf) and transition width (W) within 1 nm of the surface were measured by monitoring a qualitative change in the contact angle or density of end-groups (by Time-of-Flight Secondary Ion Mass Spectrometry) with temperature. Polystyrene (PS) films with various thicknesses (h) and molecular weights were studied. For unannealed PS supported on oxide-coated silicon or poly (dimethyl siloxane) with h > ~60 nm, Tgsurf approached a plateau value of ~25 K below the bulk Tg; below 60 nm, Tgsurf decreased with decreasing h. Separately, W exhibited a stepwise increase when h was decreased below the radius of gyration of the polymer. Upon thermal pre-annealing or deposition on a PS brush or adsorbed layer, the films ceased to exhibit Tgsurf reductions or stepwise change in W. We discuss how an out-of-equilibrium density profile with a deficit near the substrate and de Gennes’ sliding mode may explain these observations.
Article
Polysiloxanes exhibit practical properties not observed in carbon or carbon-oxygen backbone polymers. Linear ionic polysiloxanes containing an amine hydrochloric salt exhibit a humidity-induced self-assembly (HiSA) behavior. HiSA leads to nanoscale structuring by the uptake of absorbed water around the ionic aggregates of the polymer, promoting the hydrophilic-hydrophobic phase separation of the polymer. The resulting polymer forms a self-assembled lamellar nanostructure with a period of ca. 1.6 nm. Herein, we investigate the factors affecting the HiSA behavior. Introducing a dimethylsiloxane unit into the ionic polysiloxane achieves a more ordered lamellar structure and improves the nanostructure stability at higher humidities. The flexibility of the polysiloxane backbone, formation of ionic aggregates, and hydrophilic-hydrophobic phase separation affect the HiSA behavior. This nanoscale assembly can be characterized as a lyotropic liquid crystal formed with the assistance of ion aggregation instead of the hydrophobic effect in aqueous systems of ordinary surfactants.
Article
A fast and simple bottom-up methodology is reported to allow, thanks to a spontaneous phenomenon, the grafting of various materials on polymer substrates. In 60 s, the inert surface of polystyrene material is converted into a functional platform onto which organic molecules and metal particles can be anchored. A polystyrene (PS) matrix is blended with a small amount of a diblock copolymer additive engineered to both segregate at the material surface and be able to involve chemical reaction. Indeed, poly(penta fluorostyrene) (PSF) contains surface active fluorine atoms and the nucleophuge para-fluorine atom that allow nucleophilic substitution with thiols. Films of blend of PS matrix and PS-b-PSF additive are submitted to a microwave annealing to accelerate the spontaneous surface segregation of fluorine low surface energy atoms. XPS was performed to find that the optimal annealing condition was a treatment at 30 W during 60 s. With this procedure, a blend of 5 w% in additive can lead to a surface composed of 13 at% of fluorine atoms corresponding to 55 w% of diblock. The pentafluorostyrene units decorating the surface were used as anchors to immobilize gold nanoparticles via the para-fluoro-thiol click reaction.
Article
The glass transition of supported polystyrene (PS) and poly(2-vinylpyridine) (P2VP) thin films in the vicinity of the substrate interface was studied by using a nanoplasmonic sensing (NPS) method. This "nanocalorimetric" approach utilizes localized surface plasmon resonance from two-dimensional arrangements of sensor-nanoparticles deposited on SiO2 coated glass substrates. The NPS results demonstrated the existence of a high glass transition temperature (Tg, high) along with the bulk glass transition temperature (Tg, bulk ≈ 100 ºC for PS and P2VP) within the thin films: Tg, high ≈ 160 ºC for PS and Tg, high ≈ 200 ºC for P2VP. To understand the origin of the Tg, high, we also studied the thermal transitions of lone polymer chains strongly adsorbed onto the substrate surface using solvent rinsing. Interestingly, the NPS data indicated that the Tg, high is attributed to the adsorbed polymer chains. To provide a better understanding of the mechanism of the Tg, high, molecular dynamics simulations were performed on a PS film on hydrophobic and hydrophilic substrates. The simulation results illuminated the presence of a higher density region closest to the substrate surface regardless of the magnitude of the polymer-solid interactions. We postulate that the highly packed chain conformation reduces the free volume at the substrate interface, resulting in the Tg, high. Moreover, the simulation results revealed that the deviation of the Tg, high from the bulk Tg, bulk becomes larger as the polymer-substrate interaction increases, which is in line with the experimental findings.
Article
Polymer chain ends play an important role in the glassy dynamics of polymeric materials. In this study, a combination of single molecule defocus fluorescence microscopy and well-controlled atom transfer radical polymerization was used to investigate site-dependent segmental mobility of poly(n-butyl methacrylate). As the temperature increased, the rotation of fluorophores, which were selectively labelled in chain end and chain middle, was gradually activated. The power spectra of rotation trajectories, the distribution of angular displacement as well as the population of rotating fluorophores demonstrated that the local dynamics was more activated at the chain ends than the middles, showing the unique contribution of the chain end to the dynamics of the system.
Article
Many studies have established a major effect of nanoscale confinement on the glass transition temperature (Tg) of polystyrene (PS), most commonly in thin films with one or two free surfaces. Here, we characterize smaller yet significant intrinsic size effects (in the absence of free surfaces or significant attractive polymer-substrate interactions) on the Tg and fragility of PS. Melt infiltration of various molecular weights (MWs) of PS into anodic aluminum oxide (AAO) templates is used to create nanorods supported on AAO with rod diameter (d) ranging from 24 to 210 nm. The Tg (both as Tg,onset and fictive temperature) and fragility values are characterized by differential scanning calorimetry. No intrinsic size effect is observed for 30 kg/mol PS in template-supported nanorods with d = 24 nm. However, effects on Tg are present for PS nanorods with Mn and Mw ≥ ∼175 kg/mol, with effects increasing in magnitude with increasing MW. For example, in 24-nm-diameter template-supported nanorods, Tg, rod − Tg, bulk = −2.0 to −2.5 °C for PS with Mn = 175 kg/mol and Mw = 182 kg/mol, and Tg, rod − Tg, bulk = ∼−8 °C for PS with Mn = 929 kg/mol and Mw = 1420 kg/mol. In general, reductions in Tg occur when d ≤ ∼2Rg, where Rg is the bulk polymer radius of gyration. Thus, intrinsic size effects are significant when the rod diameter is smaller than the diameter (2Rg) associated with the spherical volume pervaded by coils in bulk. We hypothesize that the Tg reduction occurs when chain segment packing frustration is sufficiently perturbed by confinement in the nanorods. This explanation is supported by observed reductions in fragility with the increasing extent of confinement. We also explain why these small intrinsic size effects do not contradict reports that the Tg-confinement effect in supported PS films with one free surface exhibits little or no MW dependence.
Article
Full-text available
Lifetime analysis of positronium annihilating in nanometer voids is used to study the thermal expansion behavior of thin, Si-supported polystyrene films near the glass transition temperature Tg. A reduction in void volume expansion is correlated with a reduction in the apparent Tg as film thickness decreases. Our results can be fitted using a three-layer model incorporating a 50 Å constrained layer at the Si interface and a 20 Å surface region with reduced Tg.
Article
Full-text available
Previous studies had demonstrated that the Tg of polymer thin films is strongly dependent on the interactions, γs, between the polymer and the underlying substrate. We present a study of the glass transition temperature, Tg, in thin films of polystyrene, PS, as a function of film thickness and as a function of γs by measuring the change in the thermal expansion using X-ray reflectivity. The Tg for PS on native silicon oxide was found to decrease with decreasing film thickness, consistent with results by others. Using random copolymers of styrene and methyl methacrylate anchored to the substrate, γs could be varied by changing the styrene fraction, f. With a constant PS film thickness of 330 Å, the Tg was depressed by 20 °C as f was decreased from 1 to 0.75. An analysis analogous to the Gibbs−Thompson model indicated that the surface energy was not a suitable parameter to use to describe the effects of interfacial interactions on the Tg of polymer thin films. An associated local fractional change in the segment density at the substrate interface is instead proposed to describe the changes in Tg observed with different γs. Only a small change (<1.4%) in the density at the interface is required to produce a 20 °C depression in Tg found in this experiment.
Article
Full-text available
In this paper we present a new analysis of data on the reduced glass transitions observed in thin freely-standing films. The empirical analysis presented is highly suggestive of the existence of a mechanism of mobility in thin freely-standing films that is inhibited in the bulk and distinct from the usual cooperative motion near the glass transition temperature T g . A tentative mechanism for T g reductions in thin freely-standing films is discussed.
Article
A simple technique for precisely controlling the interfacial energies and wetting behavior of polymers in contact with solid surfaces is described. End-functionalized statistical random copolymers of styrene and methylmethacrylate were synthesized, with the styrene fraction f varying from 0 to 1, and were end-grafted onto silicon substrates to create random copolymer brushes about 5 nanometers thick. For f < 0.7, polystyrene (PS) films (20 nanometers thick) rapidly dewet from the brushes when heated well above the glass transition temperature. The contact angle of the resulting polymer droplets increased monotonically with decreasing f. Similar behavior was observed for poly(methylmethacrylate) (PMMA) films but with an opposite dependence on f. The interfacial energies of the random copolymer brushes with PS and PMMA were equal when f was about 0.6. Thus, precise control of the relative surface affinities of PS and PMMA was possible, demonstrating a way to manipulate polymer-surface interactions.
Article
Forced modulation scanning force microscopic (SFM) and lateral force microscopic (LFM) measurements of the monodisperse polystyrene (PS) films were carried out at 293 K in order to reveal surface molecular motion. Surface dynamic storage modulus, E‘, and surface loss tangent, tan δ, of the monodisperse PS films were evaluated on the basis of forced modulation SFM measurement. It was revealed that the magnitudes of surface E‘ and surface tan δ were lower and higher than those for its bulk state, respectively, in the case of the number-average molecular weight (Mn) lower than 26.6k. Based on forced modulation SFM measurements, the surface of the PS film with Mn lower than 26.6k was in a glass−rubber transition state even at 293 K, in spite of that the bulk Tg was far above 293 K. LFM measurements for the PS films revealed that the magnitude of lateral force was dependent on the scanning rate of the cantilever tip in the case of Mn lower than 40.4k. The scanning rate dependence of lateral force appeared in the case that the surface of the PS film was in a glass−rubber transition state. LFM results agreed well with forced modulation SFM ones if the scanning rate of the cantilever tip for LFM measurement was converted to the measuring frequency for forced modulation SFM measurement. The active thermal molecular motion on the polymeric surface was explained by the excess free volume induced due to the surface localization of chain end groups. The surface enrichment of chain end groups was confirmed by dynamic secondary ion mass spectroscopic measurement.
Article
The molecullar weight dependence of the surface viscoelasticity function for the monodisperse PS (polystyrene) films has been investigated on the basis of SVM (scanning viscoelasticity microscope) measurement. In the case of Mn < 26.6K, the film surface is found to be in a glass-rubber transition state or a rubbery state, even at 293K. A depression of the surface Tg compared with that of bulk samples is explained in the basis of the surface localization of chain end groups. SIMS measurements revealed that chain end groups are enriched in the depth range from the outermost surface to Rg.
Article
We have explored the use of chain end segregation as a means of controlling the properties of a polymer surface. Thin film blends of homopolystyrene (PS) and PS synthesized with low-energy oligotetrafluoroethylene chain ends (PS-TFE) were studied using neutron reflectivity. The fraction of PS-TFE that localizes near the surface was found to increase as a function of its concentration in the blend. Contact angle measurements indicate corresponding reductions in the surface tension due to the surface localization of the TFE chain ends. For a 10% blend of 6000 mol wt PS-TFE in 3 × 105 mol wt PS, the surface coverage of fluorocarbon ends was found to be >20%. A free energy model of the blends gives good qualitative agreement with the experimental results.
Article
Many applications of polymer thin films are determined by the molecular structure and chain mobility at the polymer-air surface. In ultrathin polymer films where the film thickness is comparable to the size of the macromolecule, chain connectivity, for example, may mediate effects of the substrate to the polymer-air surface, giving rise to suppression of the surface segmental mobility. In a previous study, we showed that atomic force microscopic adhesion measurements (AFMAM) could be used to quantify the dynamics in the upper 10 nm surface layer of polymer thin films near the glass-to-rubber transition. In this study, we use AFMAM to investigate the surface dynamics of ultrathin films of poly(tert-butyl acrylate) (PtBuA), to look for any systematic variation in the surface chain mobility as the film thickness, d is varied from 1Rg to 13Rg. Suppression in the tip−sample adhesion was observed when d became < 20 nm ( 2Rg), suggesting that the extent of the substrate effect is of the order of the size of a macromolecule in the present system.
Article
The distribution of polymer terminal groups at surfaces and interfaces is assessed by neutron reflectometry (NR) experiments on end-functional polystyrenes. Mono-terminated polystyrenes (PS) are synthesized anionically to include a short perdeuteriostyrene sequence adjacent to the end groups for the purpose of selective contrast labeling of the end groups for NR. The location of deuterium serves as a marker to indicate the location of the adjacent end group. Three cases of end group surface segregation are examined: a ''neutral'' control specimen prepared by proton termination, a ''repulsive'' end group system terminated with high surface energy carboxylic acid end groups, and an ''attractive'' end group system containing low surface energy fluorocarbon chain ends. All three systems exhibit damped oscillatory end group concentration depth profiles at both the air and substrate interfaces. The periods of these oscillations correspond approximately to the polymer chain dimensions. The surface structure of the ''control'' sample is dominated by the sec-butyl initiator fragment located at one end of the chain. This end group has a lower surface energy than that of the PS backbone and segregates preferentially to both the air and substrate interfaces. In the fluorosilane-terminated material, the low energy fluorinated end groups are depleted from the substrate interface but are found in excess at the air interface. In the carboxy-terminated material, the high energy carboxyl end group segregates preferentially to the silicon oxide overlayer on the substrate and is depleted at the air surface. X-ray photoelectron spectroscopy (XPS) is utilized to provide complementary characterization of the atomic surface compositions for the three systems.