You are on page 1of 4

Experimental Report 04/04/2008

Proposal: 6-04-237 Council: 10/2006

Title: Nanomechanics and Glass Transitionof Confined Polymers within Spherical Nanosized Particles
This proposal is a new proposal
Researh Area: Soft condensed matter
Main proposer: RHARBI Yahya

Experimental Team: RHARBI Yahya


YOUSFI Mohamed
Local Contact: LINDNER Peter
Samples: deuterated polystyrene, polybuthyl methacrylate and polymethyl methacrylate
Instrument Req. Days All. Days From To
D11 4 2 07/05/2007 09/05/2007
Abstract:
It is most likely that the polymer chain environment modifies their dynamics, their rheology and more particularly their glass
transition temperature (Tg). This idea is backed up by "Brillouin Light scattering", and ellipsometry results from both
supported and free standing ultrathin films. They suggest that the Tg of polystyrene is reduced below the bulk values with a
decrease in the film thickness (h) for h< 50 nm. For free standing film, where thickness h ~ 20 nm, the Tg is reduced by
about 60°C below the bulk value. However, the situation of a polymer confined within another polymer has not been
thoroughly investigated. This situation is particularly crucial for a good number of applications; polymer blends, copolymers,
composites… In this work we propose to study the influence of proximity of two polymers on their glass transition and their
dynamics. We developed a method which combines Nanomechanics with SANS to study the deformation modes of
polymers confined within nanosized domains . The work consists of analyzing the deformation modes of deuterated
nanosized polymer particles, dispersed individually within matrices of another polymer using SANS.
Experimental Report 6-04-237
Nanomechanics and Glass Transition of Confined Polymers
Within Spherical Nanosized Particles
Yahya Rharbi*, Mohamed Yousfi
Laboratoire de Rheologie, UJF/INPG/CNRS, PB 53, Domaine universitaire 38041, Grenoble, France
Local contact: Peter Lindner
Introduction. The glass transition (Tg) of nanoconfined polymers has been extensively studied throughout the last decade
[1-12]. The majority of this research has focused on thin film geometry and was carried out on supported and free-standing films
using various experimental techniques; Brillouin light scattering, [2] ellipsometry [3,5], dielectric relaxation [4], x-ray reflectivity
[10] and fluorescence [7,8]. Mechanical measurements on nano-bubble inflation have also been used to probe Tg in ultrathin
films [9]. These reports substantially agreed that the Tg of polystyrene (PS) decreases with decreasing film thickness for h < 50
nm. They demonstrate the relevance of polymer structure and polymer/substrate interactions in confined Tg. Some reports also
raised the question of the influence of thin film sample preparation (spin coating) on Tg reduction [13]. This issue could be settled
by studying samples prepared by other means.
MacKeena et al.[9] carried out mechanical tests using the “bubble inflation” technique on ultra-thin films and reported
spectacular results on the dependence of the rubbery complaisance on film thickness. Their results infer that film stiffness
increases with decreasing film thickness, which implies that thin films have a higher entanglement density than bulk. Bodiguel et
al.[14] carried out experiments on the contraction of ultrathin film under surface tension. They did not found any dependence of
the rubbery plateau on the film thickness. These results are in contradiction with theoretical predictions and with the results on the
reduction of glass transition with film thickness. The liquid-like surface layer model, which predicts a high surface mobility, also
predicts low surface entanglement and a low rubbery plateau.
Although polymers are often found nanoconfined in geometries such as spheres, little effort has been dedicated to the study of
their glass transition. This information could be useful in many industrial and environmental applications; blends, copolymers,
nanocomposites, colloids, coatings… One example is zero-VOC coating, which uses polymer nanoparticles in the film-making
process [15]. VOCs are used in coatings to lower the particle Tg, which permits the fabrication of crack-free films at room
temperature. If the Tg is reduced by decreasing the nanoparticle size, the use of VOCs could be avoided, which would have a
positive impact on the environment. Another example is in nano-blends, where hard polymer nanoparticles are used to reinforce
soft matrices. Reducing nanoparticle size would be counterproductive in this application if it causes the reduction of their Tg. It is
therefore of utmost importance to tune in to the study of nano-confined glass transition of polymers in spherical geometries. There
is an urgent need for developing novel experiments to study nanomechanical properties of nanoconfined polymers: glass
transition, relaxation, creep compliance experiments…
In this work we present a novel method to study the nanomechanical properties of nanoconfined polymer nanoparticles. We
prepared dispersions of uncrosslinked polystyrene and highly crossslinked PBMA nanoparticles. We made PBMA films
containing individual PS nanoparticles via evaporation of the dispersion mixture (Schema1). These films are made at
temperatures well below the glass transition of PS and above the Tg of crosslinked particles. In these conditions the crosslinked
PBMA particles deform under capillary pressure to fill the void between particles. The capillary pressure applied to the particles
is G = 35 γaw/R where γ is the surface tension of air/water and R is the radius of the particle (G ~ 30 MPas). The crosslinked
particles deform under these conditions because they have an elastic modulus in the order of 0.2 - 5 MPas. On the other hand, the
PS particles remain spherical because they are in the glassy state at this temperature (G ~GPas). The elastic energy is stored
within the crosslinked PBMA nanoparticles. Upon annealing these films above PS Tg, the PBMA nanoparticles regain their
original form and release the stored elastic energy which results in the deformation of neighboring PS nanoparticles (Schema 1).
We used SANS to follow the shape of the PS nanoparticle which allowed us to estimate the relaxation time and the compliance of
the nanoparticles.

Fabrication of films above Tg of crosslinked PBMA


Annealing of films above Tg of PS
and below Tg of PS
Schema1: description of the experimental procedure

1
Experimental: Polybutylmethacrylate (PBMA) particles (56 nm) and deuterated polystyrene (dPS) particles were prepared
using emulsion polymerization, at 80°C for PBMA and 80°C for dPS. The polymer concentrations were 10 wt % for the PBMA
and 2 wt % for dPS. The PBMA particles were crosslinked at 10% using ethylene glycol dimethylacrylate (EGDMA) during
polymerization. The dPS particle size was controlled between 20 and 200 nm by the amount of surfactant SDS in the reaction.
The molecular weight measured using GPC was found to be around 400 kg/mol – 600 kg/mol for all the particles investigated
here. Particle diameters (D) were measured using dynamic light scattering (Malvern 5000). The bulk glass transition measured by
DSC (Perkin-Elmer, DSC7) yielded Tg =100°C. The surfactant and free ions were removed from the dispersions using a mixture
of anionic and cationic exchange resins (Dowex, Aldrich). The nano-blends were prepared by mixing dPS and PBMA dispersions
to make dPS concentrations of 1 wt % of the solid PBMA. Solid films were obtained after water evaporation at 45 °C.
Results and discussion. The SANS spectra of films containing up to 1 wt % dPS infer that the dPS nanoparticles are
indubitably individually dispersed within the PBMA matrices (Figure.1a). The scattering intensity I(q) is the product of the form
factor P(q) and the structure factor S(q) of the dPS particles. The structure factor equals 1 and I(q) = P(q). The I(q) intensities
were compared to the P(q) of poly-dispersed hard spheres and the diameter and poly-dispersity of the particles were calculated
from the best fit. These diameters are similar to those from quasi elastic light scattering.
The spectra of films made out of low crosslinked PBMA and PBA matrices remain the same upon annealing. On the other
hand annealing films of high crosslinked PBMA yield important changes in the SANS spectra. Since the measured intensity
contains just the form factor of the PS nanoparticles (low particle concentration and non-aggregated PS particles), the change in
the spectra is a signature of the change in the form factor (Pq) of PS nanoparticles. This infers that the volume of the particles
remains the same and their shape changes from spherical to a more elongated form. The pair correlation functions calculated from
the Inverse Fourier Transform of the scattering spectra is in agreement with elongated particles of the same volume. These spectra
are less likely to be due to the fusion between PS nanoparticles. If it were the case, the spectra would exhibit an increase in the
intensity for small q. The spectra are more in agreement with deformed PS particles in the manner described in Schema 1. As the
crosslinking ratio of PBMA increases from 1 to 10%, the magnitude of the deformation increases (figure 1).

Before
Uncrosslinked 104
104 annealing
PBMA
Intensity (u.a.)

Before After
103
annealing annealing
103
After
Annealing
102
102 crosslinked PBMA
at 10%

0,004 0,01
0,004 0,01
q (Å-1) q (Å-1)

FIG. 1. SANS spectra of 87 nm polystyrene nanoparticles (1 wt %) in PBMA matrices before annealing (line) and after 1 h
annealing at 140°C (◊). Left) uncrosslined PBMA matrix, , right) 10% crosslinked PBMA matrices.
Before annealing, the scattering intensities fit well to the form factor of spherical particles with diameters similar to those
measured via quasi elastic light scattering. Upon annealing the intensity deviates from the P(q) of spherical particles. We
compared the gyration radius ( Rg) of the particles before and after annealing in order to monitor the deformation of PS particle.
The Rg was estimated using the Guinier plot for small particles. Because of the limited q window it was not possible to access the
Rg of large particles. The deformation ratio of the particles was estimated using λ(t) = Rg(t) / Rg(t=0) and the strain ε(t) = [Rg(t)-
Rg(0)]/ Rg(0). λ(t) increases with time during annealing. This increase is activated with temperature, the larger the temperature
the higher the increase rate of λ(t). The increase rate of λ(t) is used to estimate the relaxation time of the nanoconfined
nanoparticles.
This experiment is equivalent to a creep compliance test.: The crosslinked PBMA particles store the elastic stress during the
film formation process and apply it to the PS nanoparticles during annealing process. One of the obstacles to PS deformation is
the Laplace pressure due to surface tension σST = 2 γ/R, where γ is the surface tension between PS and PBMA, and R the radius of
the PS particles. σST is in the order of 0.08 MPs for 30 nm PS particles. The complaisance J could be estimated using J = ε(t)/σ,
where ε(t) is deformation strain and σ the applied stress. The elastic modulus of the crosslinked PBMA particles is in the order of
2 MPas and therefore σ is at the most equal to 2 MPas. It is most likely that the applied stress σ is less than the elastic modulus of
the crosslinking particles because part of this stress is dissipated via other means in the film formation process (film bending,
cracks …).
In Figure 2, the deformation rate measured at 6000s is plotted vs. the annealing temperature. The apparent compliance is
found to level off between 80 and 100 °C and then increases rapidly with temperature. If the applied stress is the same for all
particles investigated here the plateau between 80 and 100 °C is likely to be proportional to the compliance rubbery plateau of
2
polystyrene. This plateau increases with increasing PS particle size, which infers that the rubbery elastic plateau decreases with
decreasing particle size. This result is in agreement with the large body of results on the reduction of Tg with decreasing film
thickness.[1-8] It also agrees with simulations on the dynamics of confined polymers.[11] On the other hand it is in total
contradiction with the bubble inflation experiment from MacKeena et al.[9] who reported an unusual stiffening of ultrathin films
and an increase in the elastic rubbery plateau. The major difference between the two systems is; a) sample preparation b)
environment confinement and c) geometry.
a) Thin films are made via spin coating which is believed to result in important residual stress, which is not the case of
nanoparticles, which are made via emulsion polymerization directly from monomers.
b) Thin films are in contact with air, which results in high surface tension γ = 35 mN/m and therefore air might act as a
repulsive barrier (solid barrier). On the other hand, PS nanoparticles in contact with PBMA generate γ = 0.72 mN/m.
c) Chains expelled from the surface in thin film geometry will be forced to stretch in 2D within the film. Chains
expelled from the surface in nanoparticles are forced to keep their spherical geometry.

1,8
2,2

dPS-30nm
1,6 140 °C
130 °C
1,8
120 °C
1,4 110 °C
100 °C
dPS-56nm
80 °C 1,4
1,2

1 1
0 100 200 300 100 150
Annealing Time (min) Annealing Temperature (°C)

FIG. 1. (left) Rg(t)/Rg(t=0) vs. annealing time for different temperatures for 36 nm deuterated PS particles in 10%
crosslinked PBMA matrices. The films were annealed at temperatures ranging from 80 to 140 °C. Rg(t) is calculated from the
Guinier approximation at small q values of the SANS spectra. (Right) Rg(6000 s)/Rg(t=0) vs. annealing temperature for dPS
particles 30 nm and 56 nm in 10% crosslinked PBMA matrices. The films were annealed at temperatures ranging from 80 to 140
°C. Rg(6000 s)/Rg(t=0) is taken from Figure 2 for an annealing time of 6000s.

Conclusions and perspectives: We have developed a novel method for measuring the nanomechanical properties of
nanoconfined polymers in nanoparticles. We are able to measure an apparent compliance for particles as low as 30nm for large
spectra of time and temperature. These results infer that small particles exhibit lower entanglement densities than larger particles
and lower rubbery plateau. Further experiments are needed to investigate the compliance of large particles and to confirm these
results on other matrices with different crosslinking ratios.
References.
[1] G. Reiter, Europhys. Lett. 23, 579 (1993).
[2] J.A. Forrest, K. Dalnoki-Veress, J.R. Stevens and J.R. Dutcher, Phys. Rev. Lett. 77, 2002 (1996).
[3] J.L. Keddie, R.A.L Jones and R.A.Cory, Europhys. Lett. 27, 59 (1994).
[4] K. Fukao and Y. Miyamoto, Phys. Rev. E. 61, 1743 (2000).
[5] S. Kawana and R.A.L. Jones, Phys. Rev. E 63, 021501 (2001).
[6] G.B. DeMaggio et al. Phys. Rev. Lett. 78, 1524 (1997).
[7] C.J. Ellison, S.D. Kim, D.B. Hall and J.M. Torkelson, Eur. Phys. J.E 8, 155 (2002).
[8] C.J. Ellison and J.M. Torkelson, Nat. Mater. 2, 695 (2003).
[9] P.A. O’Connell, G.B. McKenna, Eur., Phys., J. E. 20, 143 (2006).
[10] J.H. van Zanten, W.E. Wallace and W.L. Wu, Phys. Rev. E 53, R2053 (1996).
[11] F. Varnik, J. Baschnagel and K. Binder, Phys. Rev. E 65, 021507 (2002).
[12] R.A. Riggleman, K. Yoshimoto, J.F. Douglas and J.J. dePablo, Phys. Rev. Lett. 97, 045502 (2006).
[13] M. Alcoutlabi, and G.B. McKeena, J. Phys.: Condens. Matter 17, 461 (2005).
[14] H Bodiguel . Fretigny. Phys. Rev. Lett. 97, 266105 (2006)
[15] M. S. Tirumkudulu and W. B. Russel, Langmuir 21, 4938 (2005).

You might also like