TheEffectsofRigidPolystyreneParticlesontheGlassTransitionCharacteristicsandMechanicalWaveAttenuationofStyrene-ButadieneRubber ParticleSizeandAcousticMismatch

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

The Effects of Rigid Polystyrene Particles on the Glass Transition

Characteristics and Mechanical Wave Attenuation of Styrene-Butadiene


Rubber: Particle Size and Acoustic Mismatch

FARHAD FAGHIHI, NASER MOHAMMADI, MAJID HAGHGOO


Loghman Fundamental Research Group, Polymer Engineering Department, Amirkabir University of Technology,
P. O. Box 15875-4413, Tehran, Iran

Received 13 June 2009; revised 7 September 2009; accepted 8 September 2009


DOI: 10.1002/polb.21846
Published online in Wiley InterScience (www.interscience.wiley.com).

ABSTRACT: Glass transition characteristics and mechanical wave ticles filled systems led to a broad absorption peak for the former
attenuation of the neat and filled styrene-butadiene rubber (SBR) and appearance of a secondary absorption peak at lower frequen-
containing 10 wt % of rigid monosize polystyrene particles of var- cies for the latter. Intensity of this secondary peak was highest
ious diameters from several hundred microns down to several for the system containing PS nanoparticles. Finally, ultrasonic
tens of nanometers were investigated by dynamic mechanical attenuation enhanced by the PS particle size to wavelength ratio
thermal analysis, impedance tube, and ultrasonic spectroscopy. increase according to asca  (d/k)0.38 scaling law and declined by
The results showed the matrix damping capacity and the breadth replacing the dense particles with larger hollow PS particles.
of glass transition increase by reducing the size of rigid particles Comparison of the normalized attenuation of the PS particle filled
due to the matrix-particles interfacial area increase as the major SBR in various mechanical wave attenuation regimes implied
governing parameter. Matrix glass transition broadening toward low sensitivity to particle size in vibration, mild differentiation in
higher temperatures was attributed to the increased dynamic het- the sound, and finally severe differentiation in the ultrasound
erogeneity induced by fillers, whereas the damping capacity regimes. V C 2009 Wiley Periodicals, Inc. J Polym Sci Part B: Polym

increase was assigned to contribution of interfacial friction loss Phys 48: 82–88, 2010
mechanism. The proposed postulation was confirmed based on
the calculated temperature distribution of the relaxing matrix vol- KEYWORDS: acoustic; attenuation; glass transition; loss area;
ume fraction. Sound wave attenuation by the matrix and PS par- mechanical waves; nanoparticles

INTRODUCTION Absorption and attenuation of mechanical tion of local dynamics of filled elastomers by different
waves with polymers is of great importance in diverse appli- techniques, such as, nuclear magnetic resonance,4 dielectric
cations. The behavior of polymers in response to mechanical spectroscopy,5 and neutron scattering6 have revealed some
waves is dominated by their viscoelastic state, which evidences regarding the existence of an interfacial layer of
depends among other factors, to the frequency and the am- matrix chains with highly restricted segmental mobility.
plitude of incident wave. Various modes of molecular Based on NMR spectroscopy, Berriot et al.7 confirmed glass
motions, specifically local segmental one, provide the main transition gradient in the vicinity of dispersed particles.
mechanism for mechanical energy damping with polymers. Induction of a network of domains with slow dynamics
Local segmental dynamics has the maximum contribution in extending few nanometers from particles surfaces toward
energy dissipation when its characteristic time coincides the bulk was suggested due to polymer-particles interac-
with the vibration frequency.1 Therefore, polymers show tions.7 However, controversial results have been published
maximum mechanical wave energy dissipation in their on the extent to which this localized effect could influence
glass–rubber transition. Numerous attempts have been made the glass transition of the filled polymer.8 Tsagaropoulos and
to extend the frequency span of the glass transition and Eisenberg9 reported an additional peak in tan d of nanopar-
improve the damping capacity in this region by designing ticles filled vinyl polymers at temperatures higher than the
phase separated multicomponent polymeric systems with main glass transition peak of matrix. This new relaxation
modified molecular dynamics.2,3 Among different app- was attributed to the glass transition of the matrix chains
roaches, addition of rigid particles, specifically in nanosize with restricted mobility at particles surfaces. The appear-
scale, to soft elastomeric matrices seems promising via mod- ance of a secondary relaxation in tan d spectra of nanosilica
ifying local segmental dynamics in a desired way and acti- filled styrene-butadiene rubber (SBR) matrix has also been
vating additional wave attenuation mechanisms. Investiga- reported.10 The intensity of this relaxation was proposed to

Correspondence to: N. Mohammadi (E-mail: mohamadi@aut.ac.ir)


C 2009 Wiley Periodicals, Inc.
Journal of Polymer Science: Part B: Polymer Physics, Vol. 48, 8288 (2010) V

82 INTERSCIENCE.WILEY.COM/JOURNAL/JPOLB
ARTICLE

be proportional to the filler volume fraction and polymer–fil- agent) were supplied by Merck Co. and used without further
ler interfacial strength. Rao and Pochan11 have quantitatively purification.
related the depression in the tan d peak of the matrix glass
Synthesis and Fractionation of Polystyrene Particles
transition to the number of immobilized chain segments
Narrow size distribution polystyrene nanoparticles, 80 nm in
in the interfacial region. Nevertheless, recent works12–14
diameter with particle size distribution (PSD) of 1.05, were
showed no indication of additional relaxation or perceivable
synthesized by emulsion polymerization.19 Monosize 1 lm
change in overall glass transition of filled polymers associ-
with PSD of 1.07 and larger PS particles were synthesized
ated with immobilized matrix layer around particles. The
by dispersion polymerization20 and suspension polymeriza-
decrease in the intensity of dielectric loss peak at a-relaxa-
tion, respectively. Due to broad size distribution of suspen-
tion and tan d peak at glass transition temperature was
sion polymerized PS and EPS samples, they were sieved to
attributed to the reduced polymer volume fraction and
narrow size distribution fractions with different average
increased storage modulus of filled polymer, respectively.
diameters and PSD of around 1.15.
Besides changing molecular dynamics of the matrix, the
Particle Size Characterization
presence of rigid particles may also activate additional me-
The size and size distribution of nanoparticles were deter-
chanical wave dissipation and attenuation mechanisms. It
mined using dynamic laser light scattering (SEMAtech,
has been demonstrated that dispersed rigid particles with
France) by a He-Ne laser beam with wave length of 632.8
high specific surface area cause significant increase in loss
nm. The average size of dispersion particles was determined
modulus of polymeric composites.15 This is due to energy
by scanning electron microscope (Cambridge, 630ST).
dissipation by frictional sliding at the polymer-particle inter-
face. The extent of contribution of this mechanism to overall Sample Preparation
energy dissipation is determined by several factors, such as, Neat and filled SBR sheets, 2 mm in thickness, were pre-
the oscillatory deformation amplitude, the strength of inter- pared by latex-film formation in a Teflon-coated pan at room
facial interactions, and the effective sliding area.15 Alterna- temperature for 1 week. To prepare the filled systems, poly-
tively, acoustic attenuation in filled polymers can be acti- styrene latex (15 wt % solid content) or dried polystyrene
vated due to wave scattering from rigid inclusions. This particles were mixed with the SBR latex before film forma-
additional mechanism becomes specifically important at ul- tion using magnetic stirrer at 300 rpm for 5 min.
trasonic frequencies. A theoretical model for ultrasonic wave
Techniques
attenuation in viscoelastic particulate composites was pro-
Dynamic mechanical behavior of the neat and filled SBR
posed by Beltzer and Brauner assuming independent scatter-
samples was studied in temperature range of 100 to 50  C
ing from particles.16 Biwa et al.17,18 utilized a differential
with heating rate of 5  C/min at fixed frequency of 1 Hz by
scheme to investigate ultrasonic attenuation in rigid particle
Dupont-DMA, model 983. The test was performed in cantile-
filled polymers. It was shown that the scattering loss
ver bending deformation mode on rectangular samples,
depends remarkably on the wavelength-to-particle size ratio.
2 mm in thickness, at strain amplitude around 0.1%.
When the wavelength is sufficiently large compared with the
Dynamic strain sweep experiments were performed at 1 Hz
particle radius, scattering loss is negligibly small and pres-
and two different temperatures, 20 and 23  C in the
ence of rigid particles decrease wave attenuation capacity by
dynamic strain range of around 0.01–1% before the main
decreasing matrix volume fraction. However, as the particle
experiments. The results clearly indicated that all samples
size to wavelength ratio grows toward unity, the contribution
are within linear viscoelastic region at 0.1% strain ampli-
of scattering loss rises significantly and improves overall ul-
tude. Normal incident sound absorption coefficient of the
trasonic attenuation of the composite.
samples was measured in the frequency range of 500–6300
In this article, vibration damping and acoustic attenuation Hz using B&K impedance tube according to ASTM-E1050.
behavior of SBR filled with polystyrene particles of various Ultrasound attenuation was measured by trough-transmis-
sizes are investigated. The influence of rigid particles on sion method21 at 3.8 MHz and room temperature. Two ultra-
glass transition characteristics and various mechanical sound transducers (Karl Deutsch, TS 10wB4) were mounted
energy dissipating mechanisms of the elastomeric matrix are, coaxially with 10 cm separation inside a water bath. The test
then, evaluated. specimen was placed halfway between two transducers. The
transmitted wave through the specimen was analyzed by an
oscilloscope.
EXPERIMENTAL
RESULTS AND DISCUSSION
Materials
Styrene-butadiene random copolymer (styrene content  23 Dynamic mechanical behavior of the neat and three filled
wt %) latex with solid content of 20 wt % was obtained SBR samples containing 10 wt % of monosize polystyrene
from Bandar Imam petrochemical complex (Iran). Expanda- particles with different sizes are shown in Figures 1–3. Addi-
ble polystyrene (EPS) was obtained from Tabriz petrochemi- tion of the glassy polystyrene particles increased the filled
cal complex (Iran). Styrene monomer, sodium lauryl sulfate systems storage modulus at glass transition and rubbery
and Triton X-100 (surfactants), potassium persulfate (initia- regions, Figure 1. The storage modulus increase is rather
tor), and tertiary dodecyl mercaptan (TDM, chain transfer similar for micron size particle filled samples. However,

EFFECTS OF RIGID POLYSTYRENE PARTICLES, FAGHIHI, MOHAMMADI, AND HAGHGOO 83


JOURNAL OF POLYMER SCIENCE: PART B: POLYMER PHYSICS DOI 10.1002/POLB

FIGURE 1 Temperature dependence of the storage modulus


FIGURE 3 Temperature dependence of the tan d of the neat
of the neat SBR and monosize polystyrene particles filled
SBR and monosize polystyrene particles filled systems.
systems.

particle size decrease to 80 nm, markedly improved the rein- of matrix chains with hindered segmental relaxation.7 How-
forcing effect of the dispersed particles and more interest- ever, this local effect only becomes appreciable when the ma-
ingly, reduced the slope of the storage modulus–temperature trix-particle interfacial area grows sufficiently high with par-
variation at the matrix glass–rubber transition region. At ticle size decrease, Table 1. This could be the underlying
glass transition the loss modulus peak height decreased reason differentiating viscoelastic behavior of microparticle
from around 370 MPa for the neat SBR to about 280 MPa and nanoparticle filled systems. To further investigate the
for the filled samples, whereas the peak temperature varia- influence of rigid particles on matrix segmental relaxation,
tion did not follow any definite trend with particle size. The the relaxing matrix volume fraction distribution in the glass
influence of particle size decrease on segmental relaxation of transition region was derived from storage modulus data. It
the matrix is also evident in Figures 2 and 3 as the transi- is well established that increased dynamic heterogeneity in
tion region is extended toward higher temperatures by glass formers increases the breadth of a-relaxation22 and
including the nanoparticles. Its worth mentioning that the this, in turn, reduces the slope of storage modulus decrease
damping capacity variation may also be analyzed based on at Tg, as for instance is the case for interpenetrating polymer
tan d–temperature curves, Figure 3. Nonetheless, appreciable networks. Therefore, the slope of storage modulus can be
E0 increase by addition of rigid particles, specifically nano- considered as a measure of the breadth of a-relaxation.
particles, can reduce the intensity of tan d peak significantly. Moreover, the slope of storage modulus versus time curve is
Therefore, determination of energy damping capacity and the proportional to the intensity of relaxation at each relaxation
quantity of relaxing segments based on tan d curves can be time within the whole spectrum of glass transition.23 Assum-
misleading for the case of filled systems. In filled systems, ing that the normalized relaxation intensity is proportional
matrix-particle interaction may cause significant local dy- to the fraction of relaxing domains, and considering the
namics slowdown and lead to formation of interfacial layer time–temperature correspondence, we can calculate the vol-
ume fraction of matrix segments relaxing between T and T þ
DT from normalized storage modulus decrease, DE0 , at this
temperature interval24:

TABLE 1 Interfacial Characteristics and Loss Area of the Neat


and 10 wt % PS Particle Filled SBR

The Size of the PS Filler


Neat
Sample SBR 500 lm 250 lm 75 lm 1 lm 80 nm

Interfacial – 11 22 74 5560 69,500


area per
unit volume
(cm2/cm3)
Ligament – 461 230 69 1 0.07
thickness (lm)
FIGURE 2 Temperature dependence of the loss modulus of the
neat SBR and monosize polystyrene particles filled systems. Loss area (a.u.) 7150 – 6830 – 6860 8010

84 INTERSCIENCE.WILEY.COM/JOURNAL/JPOLB
ARTICLE

ing through the number of relaxing chains decrease. How-


ever, friction at the matrix-particle interface may provide
additional energy dissipating mechanism and increase the
area under the loss modulus curve of filled samples.15 The
contribution of the new mechanism becomes important
when the matrix-particle interfacial area grows sufficiently
high. Therefore, the loss peak area increase of samples con-
taining nanometer size polystyrene particles with high sur-
face area is observed due to the extra contribution of the
interfacial friction, whereas the ineffective large particles
reduced the loss peak area, Table 1. It should be mentioned
that particles synthesized by different methods may have dif-
ferent surface chemistry and roughness. However, consider-
ing possible influence of the particles surface chemistry and
roughness on viscoelastic behavior of filled samples besides
FIGURE 4 Distribution of relaxed matrix volume fraction of the the effect of particle size, can make the analysis far more
neat SBR and filled systems. Each curve is normalized to its complicated which is not in the scope of this article.
peak intensity. Viscoelastic behavior of the matrix phase in the acoustic
frequency range was estimated by applying WLF time–
temperature superposition principle to the DMTA results,
!
0
ETþDT  ET0 eq 2:
Durelaxed ¼ (1)
Er0  Eg0
C1 ðT  T0 Þ
log aT ¼  (2)
C2 þ ðT  T0 Þ
where Eg0 and Er0are the glass and rubber storage modulus,
respectively. The normalized relaxation curves of the neat The constants C1 ¼ 17.44 and C2 ¼ 51.6 were selected from
and filled SBR samples were plotted in Figure 4 (the relaxa- literature26,27 and confirmed by the time–temperature super-
tion spectrum of the sample containing 250 lm particles position studies on the SBR.28 Matrix glass transitions were
was omitted for the sake of clarity). Matrix-particle contact estimated to be 60  C and 38  C for the lower (500 Hz)
area increase clearly shifted the segmental relaxation of cer- and upper (3.8 MHz) acoustic frequencies, respectively.
tain fraction of matrix chains to higher temperatures, con- Therefore, the matrix phase was anticipated to behave rub-
firming local dynamics slowdown in the vicinity of particles. bery under conditions applied in both sound absorption and
Therefore, the breadth of matrix softening dispersion ultrasound attenuation measurements. The sound absorption
increase in the nanoparticle filled system could be attributed coefficient of the neat SBR and three filled samples in fre-
to the particle induced dynamics heterogeneity in the elasto- quency range of 500–6300 Hz at room temperature are pre-
meric matrix phase. It is to be noted that filled rubbers can sented in Figure 5. The sound absorption capacities of all
show strain-induced nonlinearity in their modulus and samples were rather low due to their poor impedance match
undergo pseudo-glass transition under oscillatory strain due with air. Nevertheless, the sound absorption spectra revealed
to dejaming process of filler network.25 This process is influ-
some interesting features. The neat SBR showed a single
enced by various factors, such as, dynamic strain amplitude,
broad absorption peak at high frequencies, whereas in all
filler volume fraction, and temperature.25 In our samples,
filled systems, additional sound absorption peak appeared at
possible occurrence of filler network breakdown under
lower frequencies. The intensity of low frequency peak
applied dynamic strain should increase the steepness of stor-
increased with the PS particle size decrease down to 80 nm.
age modulus, |dE/dT|, at matrix glass transition. However,
The absorption peak at high frequency, which is more pro-
the storage modulus steepness of micron and nanosize PS
nounced in the neat sample, can be attributed to the acoustic
particle filled samples did not change or even decreased in
energy dissipation by the molecular motions of elastomeric
comparison with the neat matrix. Therefore, we postulated
matrix. The low frequency peak, appeared only in the filled
that the network structure of rigid PS particles, if there
samples, might be due to the particles vibration under the
existed any, remained intact during matrix glass transition
influence of sound wave.29,30 Vibrational motions of particles
due to low temperature and strain amplitude (0.1%) and
provide additional energy dissipation mechanism by inducing
the storage modulus decrease was mainly originated from
shear deformation in the matrix ligament between adjacent
matrix chains segmental relaxation. Damping capacity of me-
particles. This mechanism is magnified by the particles num-
chanical energy in both neat and filled systems under
ber density increase and the matrix ligament thickness
dynamic deformation at temperatures below the glass transi-
decrease, Table 1.
tion of PS particles, 100  C, was calculated from the area
under the loss modulus curves (loss area), Figure 2, and The ultrasound attenuation of the neat and filled SBR con-
tabulated in Table 1. Replacing a part of matrix phase with taining 10 wt % of dense and hollow PS particles with dif-
rigid particles may decline the glass transition overall damp- ferent sizes is presented in Figure 6. Despite decreased

EFFECTS OF RIGID POLYSTYRENE PARTICLES, FAGHIHI, MOHAMMADI, AND HAGHGOO 85


JOURNAL OF POLYMER SCIENCE: PART B: POLYMER PHYSICS DOI 10.1002/POLB

FIGURE 5 Normal sound absorption coefficients of the neat SBR (a) and SBR filled with 80 nm (b), 250 lm (c) and 500 lm (d)
monosize polystyrene particles.

matrix viscoelastic attenuation in the filled systems, their loss contribution. To evaluate the influence of this factor quan-
attenuation coefficient rose monotonically with the dense PS titatively the scattering attenuation was correlated with the
particle size increase. This could be assigned to additional particle size to wavelength ratio, Figure 7. Scattering attenua-
energy dissipation due to wave scattering from the dispersed tion (asca) for each sample was derived by subtracting the ma-
particles. The contribution of wave scattering mechanism is trix attenuation (am) from the composite attenuation (am):
most significantly influenced by the particle size to wavelength
ratio (d/k) as well as the acoustic mismatch between the dis- asca ¼ a  /m am (3)
persed particles and the matrix.17,18 Considering the similarity
of acoustic mismatch for all filled samples, the particle size to where /m is the matrix volume fraction. The ultrasound
wavelength ratio is the key controlling factor of the scattering wavelength inside the filled systems was calculated 421 lm

FIGURE 6 The influence of particles size and morphology on FIGURE 7 Scattering attenuation versus particle size to wave-
ultrasound attenuation of filled systems. length ratio of the rigid polystyrene particle filled systems.

86 INTERSCIENCE.WILEY.COM/JOURNAL/JPOLB
ARTICLE

not affected by the rigid particles, acoustic attenuation of the


filled systems was significantly modified in comparison with
the neat SBR. Additional peak appeared in the sound absorp-
tion spectra of the filled systems at frequencies lower than
the matrix absorption. The intensity of the secondary peak
was much more pronounced in the nanofilled sample. By
increasing the applied wave frequency to the ultrasonic
range, inverse trend was found in the variation of normal-
ized attenuation with PS particle size. It was attributed to
the contribution of the scattering loss mechanism which fol-
lowed the scaling law asca  (d/k)0.38 in the range of studied
particle size to wavelength ratio. Furthermore, ultrasound
attenuation coefficient was decreased in the filled samples
containing hollow PS particles, due to their lower acoustic
FIGURE 8 Normalized attenuation of the neat SBR and PS par- mismatch with the matrix phase.
ticle filled systems in different regimes of mechanical waves.

(k ¼ cf1) assuming the longitudinal phase velocity to be


unaffected by the presence of particles. To avoid mathemati- REFERENCES AND NOTES
cal complexities in calculation of scattering cross-section,16,17 1 Sound and Vibration Damping with Polymers, ACS Symposium Se-
a simple power equation was fitted to the experimental data, ries 424; Corsaro, R. D.; Sperling, L. H., Eds.; American Chemical Soci-
Figure 7. Least square fitting procedure yielded the scaling ety: Washington, DC, 1990; p 5.
law asca  (d/k)0.38 in the studied range. This relation clearly 2 Amiri Omaraee, I.; Mohammadi, N. Polym Eng Sci 2002, 42,
indicates the increasing contribution of scattering loss mech- 2328–2335.
anism as the particle size to wavelength ratio grows to unity.
3 Gomez Ribelles, J. L.; Monleon Pradas, M.; Meseguer, J. M.; Torre-
It is interesting to note that the variation rate of asca is
grosa, C. J Non-Cryst Solids 2002, 35, 731–737.
higher for smaller values of d/k and tends to level off as d/k
approaches unity and grows higher. The above trend did not 4 ten Brinke, J. W.; Litvinov, V. M.; Wijnhoven, J. E. G. J.; Noordermeer,
hold for EPS filled systems as they showed lower attenuation J. W. M. Macromolecules 2002, 35, 10026–10037.
despite greater particle size. The lower sound attenuation 5 Fragiadakis, D.; Pissis, P.; Bokobza, L. Polymer 2005, 46,
might be originated from lower density of the EPS particles 6001–6008.
(q  0.92 g/cm3), which reduced the acoustic mismatch and 6 Gagliardi, S.; Arrighi, V.; Ferguson, R.; Telling, M. T. F. Phys B:
resultant scattering loss. Condens Matter 2001, 301, 110–114.
Normalized energy dissipation of monosize polystyrene 7 Berriot, J.; Montes, H.; Lequeux, F.; Long, D.; Sotta, P. Macromole-
particle filled systems in different regimes of mechanical cules 2002, 35, 9756–9762.
waves is summarized in Figure 8. In the vibration regime, 8 Alcoutlabi, M.; McKenna, G. B. J Phys: Condens Matter 2005, 17,
addition of PS particles with different sizes did not improve R461–R524.
the energy dissipation capacity so much. Taking the samples
9 Tsagaropoulos, G.; Eisenberg, A. Macromolecules 1995, 28,
to the acoustics domain differentiated the filled samples in
6067–6077.
two distinct directions. The sound absorption capacity was
magnified by inclusion of nanosize rigid particles. However, 10 Arrighi, V.; McEwena, I. J.; Qiana, H.; Serrano Prietob, M. B. Poly-
ultrasound attenuation capacity was improved by increasing mer 2003, 44, 6259–6266.
particle size to the micron range. The observed trends of 11 Rao, Y.; Pochan, J. M. Macromolecules 2007, 40, 290–296.
attenuation capacity are originated from different major 12 Robertson, C. G.; Lin, C. J.; Rackaitis, M.; Roland, C. M. Macromole-
energy dissipation mechanisms effective in various regimes cules 2008, 41, 2727–2731.
of mechanical waves.
13 Bogoslovov, R. B.; Roland, C. M.; Ellis, A. R.; Randall, A. M.; Robert-
son, C. G. Macromolecules 2008, 41, 1289–1296.
CONCLUSIONS 14 Wang, X.; Hall, J. E.; Warren, S.; Krom, J.; Magistrelli, M.; Rackaitis,
It was observed that the matrix glass transition characteris- M.; Bohm, G. A. G. Macromolecules 2007, 40, 499–508.
tics in rigid particles filled samples were dominated by ma- 15 Suhr, J.; Koratkar, N. A. J Mater Sci 2008, 43, 4370–4382.
trix-particles interfacial properties. Addition of rigid particles
16 Brauner, N.; Beltzer, A. I. Ultrasonics 1988, 26, 328–334.
with high surface area imparted additional dynamics hetero-
geneity to the matrix and consequently extended the matrix 17 Biwa, S.; Idekoba, S.; Ohno, N. Mech Mater 2002, 34, 671–682.
glass transition toward higher temperatures. Although the 18 Biwa, S.; Watanabe, Y.; Motogi, S.; Ohno, N. Ultrasonics 2004, 43,
matrix viscoelastic state at applied acoustic frequencies was 5–12.

EFFECTS OF RIGID POLYSTYRENE PARTICLES, FAGHIHI, MOHAMMADI, AND HAGHGOO 87


JOURNAL OF POLYMER SCIENCE: PART B: POLYMER PHYSICS DOI 10.1002/POLB

19 Mohammadi, N.; Yoo, J. N.; Klein, A.; Sperling, L. H. J Polym Sci 25 Wang, X.; Robertson, C. G. Phys Rev E: Stat Phys Plasmas 2005,
Part B: Polym Phys 1992, 30, 1311–1319. 72, 031406-1-031406-9.
20 Song, J. S.; Winnik, M. A. Macromolecules 2005, 38, 26 Youssef, M. H. Polym Test 2003, 22, 235–242.
8300–8309.
27 Oprisoni, C. A.; Alshuth, T.; Schuster, R. H. Presented at the
21 Lee, K. I.; Roh, H. S.; Yoon, S. W. J Acoust Soc Am 2003, 115, Forschungsprojekte-Präsentationstag der DKG; Fulda, Germany,
2933–2938. 2008.
22 Sillescu, H. J Non-Cryst Solids 1999, 243, 81–108. 28 Faghihi, F.; Mohammadi, N., unpublished work.
23 Ferry J. D. Viscoelastic Properties of Polymers; Wiley: New York, 29 Zhou, H.; Li, B.; Huang, G. J Appl Polym Sci 2006, 101,
1980; pp 60–67. 2675–2679.
24 Matsuoka, S.; Quan, X. Macromolecules 1990, 24, 2770–2779. 30 Hung, Y.; Hong, L. J Appl Polym Sci 2006, 102, 1202–1212.

88 INTERSCIENCE.WILEY.COM/JOURNAL/JPOLB

You might also like