Anda di halaman 1dari 11

FULL PAPER

DOI: 10.1002/adem.201400505

Tuning the Electromechanical Properties of Silicones by


Crosslinking Agent**
By Adrian Bele, Maria Cazacu,* Carmen Racles, George Stiubianu, Dragos Ovezea
and Mircea Ignat

Five R-trialkoxysilanes, with R: CH3, C6H5, NH2(CH2)3, Cl(CH2)3, or NC(CH2)3 are used as
crosslinkers for two polydimethylsiloxane-a,v-diols with different molecular masses (35,000 and
125,000 g  mol 1) The crosslinking occurs by condensation at room temperature under the influence of
the environmental moisture and in presence of dibutyltindilaurate as a catalyst. After aging, the films
are characterized by mechanical testing, dielectric spectroscopy, and thermal analysis. Moisture
sorption capacity is evaluated by dynamic vapour sorption analysis, while the morphology of the
crosslinked films is observed by scanning electron microscopy in cryo-fracture. The mechanical
response to an applied electric field is measured. The results are discussed in correlation with polymeric
chain length and the nature of R from crosslinking agent emphasizing in principle an increasing of the
dielectric permittivity and actuation with the polarity of the organic group but a worsening of these
with increasing length of the polymer chain matrix.

1. Introduction
To date, silicone rubbers constitute, besides acrylics and
polyurethanes, the basis of most successful dielectric elastomer devices.[1] Their use in actuators is due to their good
performances and stable properties over a wide range of
temperatures.[2] Due to the high flexibility of the SiO bonds,
silicones possess good elastomer behavior,[1,3] which is useful

*[*] Dr. M. Cazacu, A. Bele, Dr. C. Racles, Dr. G. Stiubianu


Petru Poni Institute of Macromolecular Chemistry, Aleea Gr.
Ghica Voda 41A, Iasi 700487, Romania
E-mail: mcazacu@icmpp.ro
Dr. D. Ovezea, Dr. M. Ignat
National Institute for Research and Development in Electrical
Engineering ICPE-CA, 313 Splaiul Unirii, Bucharest 030138,
Romania
[**] Support from Romanian-Swiss Research Programme (RSRP)
NR: 10 RO-CH/RSRP/01.01.2013, is gratefully acknowledged. Authors also thank to dr. Dorina Opris from Empa,
Swiss Federal Laboratories for Materials Science and
Technology, Laboratory for Functional Polymers, D
ubendorf,
Switzerland, for dielectric measurements. (Supporting Information is available online from Wiley Interscience or from the
author.)
Paper dedicated to the 65th anniversary of Petru Poni
Institute of Macromolecular Chemistry of Romanian Academy, Iasi, Romania.
ADVANCED ENGINEERING MATERIALS 2015,
DOI: 10.1002/adem.201400505

for the electromechanical transduction properties. However,


silicones have the disadvantages of low values for dielectric
permittivity (e0 ), requiring increased voltages to obtain
reasonable actuation effects.
Different methods were used to increase the dielectric
permittivity (e0 ) of silicones. One approach is the chemical
attachment of polar groups, by synthesis of the copolymers
from monomers containing polar groups, by grafting of
polar groups to a preformed polymer, or by using a
crosslinker having a polar group.[36] Thus, dipoles as
N-allyl-N-methyl-p-nitroaniline[4] or cyanpropyl[5,7] were
attached to the polysiloxane chain by hydrosilylation
reaction. Significant increases of e0 were reported for the
high content of polar groups.[47] Unfortunately, with
increasing the amount of cyanopropyl groups the materials
became brittle.[7] Another approach which led to better
mechanical characteristics is based on the dispersion of
a filler of high e0 into the elastomeric matrix before
vulcanization.[811]
Three major curing reactions are frequently used to convert
silicone polymers from liquid into the elastomer state: free
radical (with peroxides), addition (with platinum complex),[12] and condensation.[13] The condensation of silanol
groups to form siloxane bonds is frequently used to crosslink
siloxanes in commercial applications. A typical system
contains silanol-terminated polydimethylsiloxane (PDMS),
crosslinker (a tri- or tetrafunctional molecule which produces
a three dimensional network) and catalyst. Depending on the

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

wileyonlinelibrary.com

FULL PAPER

A. Bele et al./Electromechanical Tuning by Crosslinker


application, reinforcing fillers and other additives may be
added. By curing, flexible silicone elastomers with excellent
retention of their elastic properties over a wide temperature
range are formed. Tri- and tetrafunctional crosslinkers, such as
trialkoxysilanes (R0 Si(OR00 )3), are capable of vulcanization by
exposure to air moisture. The alkoxy functional groups are
hydrolyzed to form silanol and the corresponding organic
alcohol. The silanol can then condense with the polydimethylsiloxane-a,v diol and produce a crosslinked
network. Suitable catalysts for this chemistry are tin
carboxylates (i.e., dibutyltin dilaurate and dibutyltin
octoate) and there are commercially available kits,
which contain two-component systems. One part would
typically contain the silanol-terminated PDMS and the
other the crosslinker and catalyst. Both components are
mixed at the application stage.[1416] Among curing
technologies, the silane crosslinking shows some advantages that are a consequence of the specific structure of the
network, which is responsible for better thermomechanical
properties compared to those of the network formed by
peroxide crosslinking, for example.[17]
In this work, various functional trialkoxysilanes that
contain groups of different polarity attached to the silicone
were used as crosslinking agents for PDMS. Trialkoxysilanes
with methyl, phenyl, chloropropyl, aminopropyl, and
cyanpropyl groups were used in excess so that, along with
the crosslinking, they also allow in situ formation of
silsesquioxane (SSQ) domains. The influence of the functional group of the crosslinker on the properties that
determine the electromechanical actuation (mechanical,
dielectric, thermal properties, surface, as well as morphology) was investigated.
2. Experimental Section
2.1. Materials
The polydimethylsiloxane-a,v-diols, PDMSs, were synthesized according to the already described procedure: cationic
ring-opening polymerization of octamethylcyclotetrasiloxane
in the presence of a cation exchanger as catalyst.[18] Molecular
masses were estimated on the basis of gel permeation
chromatography (GPC) analysis, Mn values being 34,500
and 125,000 g  mol 1, respectively. Triethoxymethylsilane,
MTES (99%), (3-&chloropropyl)&trimethoxysilane, CPTMS
(97%), (3-&aminopropyl)triethoxysilane, APTES (98%),
(3-&cyanopropyl)triethoxysilane, CyPTES (98%), triethoxyphenylsilane, PhTES (98%), and dibutyltindilaurate,
DBTDL, (95%) were commercial compounds (Aldrich) and
were used as received.
2.2. Measurements
GPC measurements were made in DMF on a PL-EMD 950
ChromatographEvaporative Mass Detector using polystyrene standards. Stressstrain measurements were performed
on a TIRA test 2161 apparatus, Maschinenbau GmbH
Ravenstein, Germany on dumbbell-shaped cut samples with

http://www.aem-journal.com

dimensions of 50  8.5  4 mm. Measurements were run at an


extension rate of 20 mm  min 1, at room temperature. All
samples were measured three times and the averages of the
obtained values were taken into consideration. The acquired
data were processed with MatLab program. Thermogravimetric (TG) measurements were conducted on a STA 449 F1
Jupiter device (Netzsch, Germany). About 10 mg of each
sample was weighed and heated in alumina crucibles.
Nitrogen was purged as inert atmosphere at a flow rate of
50 mL  min 1. Samples were heated in the temperature range
from 30 C to 700 C at a heating rate of 10 C min 1.
Differential scanning calorimetry (DSC) measurements were
conducted with a DSC 200 F3 Maia (Netzsch, Germany).
About 10 mg of sample was heated in pressed and punched
aluminium crucibles at a heating rate of 10 C  min 1.
Nitrogen was used as inert atmosphere at a flow rate of
100 mL  min 1. The dielectric constant measurements, have
been performed by using equipment (HP 4284 A LCR meter)
and experimental procedure described previously.[3] The
measurements for dynamic vapour sorption and sorption
isotherms were performed with an IGAsorp Dynamic
Vapour Sorption apparatus with the following characteristics: minimum gas pressure, 2 bar; resolution of 0.1 mg for
100 mg and sample containers made out of stainless steel
micron size mesh. Before sorption measurements, the
samples were dried at 25 C in a flow of dry nitrogen
(250 mL  min 1) until the weight of the sample was in
equilibrium at a relative humidity (RH) less than 1%.
Scanning electron microscopy (SEM) images on film surfaces
were taken with an Electron Microscope type Quanta 200
operating at 30 kV with secondary and backscattering
electrons in low or high vacuum mode. The Energy
Dispersive X-Ray system (EDX) available on this equipment
was used for qualitative analysis and elemental mapping.
The electromechanical actuation measurements were made
using an AGILENT 5529 A System based on LASER
interference, with 10 nm resolution and setup to measure
linear displacement.
2.3. Crosslinking Reactions
To a solution of polydimethylsiloxane-a,v-diol (10 g) in
chloroform (5 ml), silane crosslinker (1.5 ml), and DBTDL
catalyst (0.1 ml) were added. The mixture was mechanically
stirred for about 2 min, and then poured on a 15  5 cm Teflon
plate. The samples were maintained in air at room temperature for 15 days (depending on the crosslinking rate) and
subsequently the formed films were easily peeled off from the
substrate. The films were then kept in the laboratory
environment (around 25 C and 28% humidity) about a
month for aging before characterization by different techniques (scanning electron microscopy, differential scanning
calorimetry, thermogravimetric analysis, dynamic vapours
sorption, dielectric, mechanical and electromechanical
testing).
The recipes used to prepare films with different crosslinking agents are presented in Table 1.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ADVANCED ENGINEERING MATERIALS 2015,


DOI: 10.1002/adem.201400505

A. Bele et al./Electromechanical Tuning by Crosslinker

Sample
R8M
R8Ph
R8C
R8A
R8Cy
R9M
R9Ph
R9C
R9A
R9Cy

[a]

Polymer code/Mn
[g/mol]

Crosslinking
agent

R8/34500

MTES
PhTES
CPTMS
APTES
CyPTES
MTES
PhTES
CPTMS
APTES
CyPTES

R9/125000

SSQ, wt.%
[mmol/100 g][a]
5.2
7.2
8.7
6.8
7.0
5.3
7.4
8.9
6.9
7.2

(78)
(56)
(67)
(61)
(59)
(79)
(57)
(69)
(63)
(60)

Aspect of the film


transparent, adhering to the substrate
transparent, easy removal from the support
transparent, easy removal from the support
opaque, very easy removal from the support
slightly opaque, easy removal from the support
transparent, adhering to the substrate
transparent, easy removal from the support
slightly opaque, easy removal from the support
opaque, very easy removal from the support
slightly opaque, easy removal from the support

Theoretical estimation of the silsesquioxane (SSQ) quantity that could form by full hydrolysis and condensation of the excess crosslinker.

3. Results and Discussion


Two polydimethylsiloxane-a,v-diols with Mn of 34,500 and
125,000 g  mol 1 were prepared and crosslinked by condensation (Scheme 1) using five organosilanes differing by the
nature of the organic group attached to silicon for example
methyl (M), 3-chloropropyl (C), 3-aminopropyl (A), 3cyanopropyl (Cy), or phenyl (Ph) (Table 1). In each
experiment, 15 wt.% crosslinking agent reported to the
polymer mass was used, this being in excess towards
the amount due to provide a full crosslinking. According
to the well-known mechanism is to assume that this excess
will be converted in silsesquioxane (SSQ) domains with
generic formula [RSiO1.5-x(OH)2x]n, which can be polymeric
networks or polyhedral clusters and could act as fillers for the
polymeric matrix.[19] Assuming that all the crosslinking
agents condensed completely, different amounts of SSQ will
result depending on the silane crosslinker used (Table 1), that
is, 5.28.9 wt% (or 5679 mmol/100 g). By considering
the samples containing polar groups, chloropropyl, aminopropyl, and cyanopropyl, these values range between 59
69 mmol/100 g within the two series. The highest values
result for the chloropropyl groups, R8C and R9C with 67 and
69 mmol/100 g, respectively that could influence some
properties of the material. Due to higher molecular mass of
R9, the amount of SSQ generated in situ (from the same
feeding amount) will be slightly higher than the R8 series for
the same cross-linker (by 0.10.2 wt% or 12 mmol/100 g).
These differences cannot influence very much the behaviour
of the films. The dilution of the initial mixture with chloroform
allows mixing the components at the molecular level, thus
creating the premise for a uniform distribution of the
hydrolysis-condensation products of the crosslinker excess
within polydimethylsiloxane matrix.[20]
The stress-strain curves are presented in Figure 1 and data
are centralized in Table 2. It can be observed that elongations
as large as 640 and over 900% were obtained in the case of
ADVANCED ENGINEERING MATERIALS ,
DOI: 10.1002/adem.201400505

samples crosslinked with CH3- substituted trialkoxysilane.


The phenyl groups gave the second (series R8) or third (series
R9) best result, while the samples crosslinked with trialkoxysilanes having polar groups showed lower values for
ultimate strain. These data seem to support the assumption
that nonpolar substituents improve the elastic properties of
the material, due to their better compatibility with the matrix.
Between the polar crosslinkers, the highest values for tensile
strength were obtained for cyanopropyl groups in both series,
while the elongation at break was around 200%. The stress
values and the Young modulus for the higher Mn matrix are
significantly lower than for the polymer with lower Mn. This is
due to lower crosslinking density in R9 series with longer
chains. More specifically, modulus values of 0.0620.375 MPa
were calculated for R9 series towards 0.0980.948 MPa for
series R8 and strength values one order of magnitude lower in
series R9 compared to R8. On the other hand, increased
elongation for films based on higher Mn polymer would be
expected. Overall, this expectation was also confirmed as a
general trend; elongations between 141 and 918% for series R9
compared to 119640% in series R8 were obtained.
The R-trialkoxysilanes used assure the crosslinking reaction by the known hydrolytical sensitivity of Si O C bonds
while the R-functional group remaining in the network node
will confer certain behaviours or properties, for example low
crystallization temperature,[13] increasing in refractive index,[21] improved oil resistance,[22] hydrophilicity,[23] or on the
contrary nonpolar, hydrophobic effect,[24] biocidal activity, or
softening properties[23] in dependence on its nature. In our
study, the main desired effect of using crosslinkers with
organic polar groups was to improve the dielectric properties
of the crosslinked film. The presence of randomly dispersed
dipols as cyanopropyl groups was reported to increase the e0 of
silicones.[5,7] In order to compare the molar parameters of the
substituents (R) from the crosslinkers used in our study, some
theoretical calculations were done using well-established
methods[25] on pentamethyldisiloxanes with R substituents

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.aem-journal.com

FULL PAPER

Table 1. The cured polydimethylsiloxane films prepared by using different crosslinking agents at room temperature: the same amounts of PDMS (10 g), crosslinker (1.5 ml)
and DBTDL catalyst (0.1 ml) were used.

FULL PAPER

A. Bele et al./Electromechanical Tuning by Crosslinker

Scheme 1. The crosslinking reaction of polydimethylsiloxane-a,v-diol with trialkoxysilanes.

Fig. 1. Stress-strain curves for crosslinked films.

http://www.aem-journal.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ADVANCED ENGINEERING MATERIALS 2015,


DOI: 10.1002/adem.201400505

A. Bele et al./Electromechanical Tuning by Crosslinker

Sample
R8M
R8Ph
R8C
R8A
R8Cy
R9M
R9Ph
R9C
R9A
R9Cy

Youngs modulus
[MPa][a]

Tensile strength[b]
[MPa]

Elongation
at break [%]

Dielectric permittivity,
e0 (at 10 kHz)

Dielectric loss,
e00 (at 10 kHz)

0.098
0.680
0.723
0.378
0.948
0.062
0.087
0.375
0.155
0.320

0.32
1.20
0.63
0.28
2.14
0.17
0.08
0.17
0.11
0.20

640
361
180
119
247
918
388
141
470
176

2.5
2.8
3.1
3.2
3.7
2.8
2.9
3.1
3.1
3.3

0.0045
0.0034
0.0167
0.0441
0.0215
0.0037
0.0037
0.0121
0.0292
0.0124

[a]
Youngs modulus was calculated as a ratio between the stress and strain when the latter is 10%. Linear stress-strain dependence is considered on this region of the
curve.
[b]
Corresponding for the value of elongation at break.

model molecules (Table 3). As is known, dipole moment is a


factor that influences the value of the dielectric constant.
Within the series of substituents used for the crosslinker,
the dipole moment values increase in the following
order: methyl < phenyl < aminopropyl < chloropropyl < cyanopropyl (Table 3). As expected, the e0 values measured on the
silicone films follow this trend within both series of materials
(Figure 2, Table 2), except for the case of chloropropyl
substituent, which gave e0 values lower than expected. This
could be due to larger free volume induced by chloropropyl
group which, after phenyl group, is the bulkiest (Table 3). The
highest values for the e0 at 10 kHz, that is, 3.7 and 3.3, were
obtained in the case of the films crosslinked with excess of
cyanopropyl- and aminopropyltrialkoxsilane, folowed by
those with chloropropyl-, phenyl-, and methyltrialkoxysilane.
These values are higher as compared with those calculated by
Bicerano method[26,27] for model disiloxanes, but in the same
order as predicted from the theory. Very large differences
between the dielectric loss values (one order of magnitude)
were observed for the samples containing different crosslinkers. However, the loss values were lower than 4  10 2 for
all the samples.
With increasing molecular weight of the polymer matrix,
the crosslinking density thereof by the ends of the chain
decreases. This should lead to increased free volume thereby

decreasing e0 .[28] On the other hand, because we used large


excesses of thrialkoxysilanes, only part of them are consumed
as crosslinkers, mostly being converted in hydrolysis and
partial or total self-condensation products. They contain polar
OH groups which, besides already existing polar R groups,
could contribute to increasing the dielectric permittivity. The
balance between the effects of molecular weight increasing
and trialkoxysilane excess will define the final dielectric
behavior of the material. Even though the amount of SSQ is
slightly higher in R9 series, the presence of polar groups
didnt have the expected effect of increasing e0 . Between the
two series, higher e0 values were obtained in R9 series for
RM and Ph, and lower for Cy substituent, while for C it
didnt change. R9A had a continuous decrease with increasing
frequency, thus it is difficult to consider a correct value for
comparison. Probably this result might be explained by the
agglomeration of the SSQs formed into the nodes of the
networks, and thus a lesser dense distribution of the polar
groups within the matrix in the case of R9 series, with longer
chains. Assuming that all crosslinkers have similar conversion
(thus similar number of OH groups), the ones with higher
dipole moments would adopt aggregation patterns as to have
the polar groups far away from the PDMS (at the interior of
such aggregates), thus their polarization in electric field

Table 3. Calculated values of some physical parameters that influence dielectric constant.

Substituent

Dipole moment[a]

e0 (298 K)[b]

Amorphous molar volume


at 298 K [cm3/mol][b]

Van der Waals Volume


at 298 K [cm3/mol][b]

CH3
C6H5
(CH2)3-NH2
(CH2)3-Cl
(CH2)3-CN

0.22
1.30
2.44
2.86
2.92

2.40
2.51
2.70
2.39
2.72

159.82
208.56
149.20
163.99
151.39

87.27
123.59
82.15
88.89
80.60

[a]
[b]

Estimated by using HyperChem(TM),[25]


Calculated according to Bicerano method.[26,27]

ADVANCED ENGINEERING MATERIALS ,


DOI: 10.1002/adem.201400505

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.aem-journal.com

FULL PAPER

Table 2. The main results of the mechanical and dielectric tests.

FULL PAPER

A. Bele et al./Electromechanical Tuning by Crosslinker

Fig. 2. Dielectric spectroscopy results for the crosslinked films.

would be hindered. This could explain the lower values of e0


in the case of R9Cy.
For the electromechanical actuation measurements by
using the AGILENT 5529 A System, the samples were placed
in a sandwich on a reference surface between two electrodes
made of 35 mm Cu on FR4 (PCB board). The sandwich was
pressed by the weight of the retro-reflector and the upper
electrode. The experimental setup diagram is shown in
Figure 3a. This setup eliminates the influence of variable force
contact probes. The thickness of the samples was measured
using an electronic comparator of 10 mm resolution while
being placed between the electrodes, in order to minimize the
deformation of the samples. The actual thickness was
calculated by subtracting the thickness of the electrodes.
The measurements were made in a 20 C room and at least
30 min were given to each sample to achieve mechanical and
thermal stabilisation at 0 V polarisation voltage before
beginning the testing. The measuring process consisted in
recording the position of the retro-reflector for 20 s at 0 V and
then establishing a step voltage of 50 500 V for 100 s,
followed by another 120 s of 0 V polarisation (electrodes were
shorted), giving the polymer enough time to respond for a

http://www.aem-journal.com

typical electromechanical application. The actual response


time of the polymers is much larger but since mechanical
positioning must be achieved in limited time, displacement
values after the specified time were considered a good
compromise, offering a practical assay of the performance.
The response of R8Cy sample at a 500 V/100 s polarisation
step is given in Figure 3b for exemplification purposes. The
500 V was the limit of the stabilized voltage supply and thus,
larger displacements are expected at higher electric voltages.
Measurements were taken at 50 V, 100 V, 150 V, 200 V, 250 V,
300 V, 350 V, 400 V, 450 V, and 500 V unless electrical breakdown occurred. The voltages were measured with a Meterman 37XR digital multimeter. The actual displacement values
were calculated as difference between the reference position
(first 20 s of data) and the position at the last 10 s of the voltage
step (110 s to 120 s of data), where the averages of each data
population were considered. As seen in Figure 3b, the
thickness of the sample decreases when voltage is applied
suggesting electrostrictive behavior. All measured samples
behaved in the same manner. The response of each actual
sample is presented in Figure 1S9S. For each sample, an
interpolation curve was traced using a 2nd order polynomial

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ADVANCED ENGINEERING MATERIALS 2015,


DOI: 10.1002/adem.201400505

A. Bele et al./Electromechanical Tuning by Crosslinker

FULL PAPER

Fig. 3. Experimental setup diagram for electromechanical actuation measurements a, and illustrative response of R8Cy sample at 500 V applied voltage b.

that intersects the origin. Polynomial formulas and R2 values


are shown in each respective Figure. It should be noticed that
R8C (Figure 3S) appears to have an inflexion point between
500 V/mm and 600 V/mm which suggests a possible saturation. For this reason, the interpolating polynomial was
calculated before reaching this point. It should also be noted
that the response of R9Ph is intensely affected by measurement errors below 300 V/mm. An interpolation curve was
calculated anyway (Figure 6S). Only nine samples were
measured, the sample R8A being compromised when
unpacked. Figure 4 presents the response of the samples
vs. the applied electrical field reported to the film thickness
while the actuation measurements data are centralized in
Table 4.
For a better comparison, the displacement values at
625 V/mm (which is the lowest voltage intensity that was
ADVANCED ENGINEERING MATERIALS ,
DOI: 10.1002/adem.201400505

applied on the thickest film, i.e., sample R8Cy) were also


provided in Table 4 and plotted in Figure 10S. By examining
the data from Table 2 and 4 and Figure 4 and 10 S, it can be
noticed that reasonable displacements were obtained in the
case of the samples containing chloro, amino, and cyano
groups when the dielectric constant of the films was >3,
but a significant influence has the modulus value. In
series R9, the sample R9Cy having slightly lower modulus
(0.320 MPa) as compared with R9C (0.375 MPa) and also
higher e0 value results in a higher electromechanical
displacement, 8,037 nm/mm compared to 2,591 nm/mm. In
series R8, larger differences in both dielectric and mechanical
properties were registered for samples R8C and R8Cy:
dielectric constants of 3.1 and 3.7 and modulus 0.723 and
0.948, respectively. Sample R8C proved to be more active in
electro-mechanical test having the highest displacement,

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.aem-journal.com

FULL PAPER

A. Bele et al./Electromechanical Tuning by Crosslinker

Fig. 4. The comparative values of displacement as a function of electric field expressed as V/mm thickness for the films based on PDMS crosslinked with different agents. (From
clarity as well as aesthetic reasons, the displacement axis of the chart was broken at 6000 nm while the displacement of the sample R8C reached 22311.10 nm, value that is specified
in Table 4 and taken into account in the further discussions.)

30,226 nm/mm at an applied voltage of 625 V/mm. Thus, it


seems that lower modulus is the most important factor that
governed the actuation behavior in this case. When we
compare the R8 and R9 series, the displacements obtained for
the higher Mn polymer were lower. Judging after the much
lower Young moduli in series R9, we would expect higher
response, since the decrease in dielectric permittivity was
insignificant, except for cyanopropyl substituent. In fact, the
displacement for R9Cy was much higher than for R8Cy, in
spite of lower e0 . We suppose that the distribution of polar
groups within the sample might also influence the actuation
behavior. Assuming that most of the formed SSQ structures
are agglomerated in the networks nodes where the trialkoxysilanes react with the PDMS chain ends, it results that in the
case of the R8 series (with shorter chains) the crosslinked
structures and thus the polar groups are more densely

distributed within the material. In R9 series where chains are


much longer, the number of the crosslinking points is lower
and, given the same amount of crosslinker used as in R8 series,
more of it is transformed into SSQ, which agglomerate into
fewer nodes. Thus, although R9 series films have a slightly
higher content of polar particles, larger clusters form, which
dont have reinforcement effect (as suggested by the low
Young modulus). On the other hand, the incompatibility of
the polar SSQs and the nonpolar silicone matrix might lead to
lower dielectric constant, as discussed before. This hypothesis
could be supported by the EDX mapping results, obtained in
the same conditions on randomly chosen regions of samples
containing heteroelements (Cl and N). In all cases, the density
of the zones containing polar groups is higher in the case of
the samples in Series R8 as compared with series R9
homologues (Figure 11S).

Table 4. The centralized data related to the actuation measurement.

Sample
R8M
R8Ph
R8C
R8A
R8Cy
R9M
R9Ph
R9C
R9A
R9Cy

Film thickness
[mm]

Applied tension
[V/mm]

Total displacement
[nm]

Displacement,
at 500 V [nm/mm]

Displacement,
at 632 V [nm/mm]

0.57
0.70
0.60
n.d.
0.80
0.57
0.68
0.43
0.54
0.68

877.19
714.29
847.46
n.d.
632.91
877.19
735.29
1162.79
925.93
735.29

1005.63
955.82
22311.10
n.d.
2394.24
973.03
347.06
5458.04
1858.53
4657.68

1764
1365
37185
n.d.
2993
1707
510
12693
3442
6850

945
1051
30226
n.d.
2993
763
451
2591
1558
8037

http://www.aem-journal.com

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ADVANCED ENGINEERING MATERIALS 2015,


DOI: 10.1002/adem.201400505

A. Bele et al./Electromechanical Tuning by Crosslinker

FULL PAPER

Fig. 5. SEM images on cryo-fractured section: a- R8M; b- R9M; c R8Ph; d R9Ph; e R8C; f R9C; g R8A; h R9A; i R8Cy; j R9Cy.

SEM images (Figure 5) taken on cryo-fractured films


revealed very similar aspect for the pairs of films with the
same crosslinker. The samples generally showed a layer
structuration, probably as a result of crosslinking dynamics
(evaporation of the solvent and crosslinking kinetics). In detail
ADVANCED ENGINEERING MATERIALS ,
DOI: 10.1002/adem.201400505

images, micrometer particles can be observed in all samples,


embedded into the matrix polymer. In R8Cy and R9Cy, well
defined particles stand out of the matrix. In R9A large
inclusions of a few microns were observed, which were
probably responsible for the dielectric behavior.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.aem-journal.com

FULL PAPER

A. Bele et al./Electromechanical Tuning by Crosslinker


For further applications in microelectromechanic devices
(MEMS domain), the elastomer should ensure a stable
operation. This implies, among others, the stability of the
properties of interest with moisture and temperature.[29] To
verify this, the moisture sorption-desorption isotherms were
registered in dynamic regime for the crosslinked composite
films (Figure 12S). In general, the sorption values are low but,
as expected, the highest values were registered for samples
containing the most hydrophilic and polar groups. Thus
the samples with amino groups which, as is known are
hydrophilic, show sorption values of 1.86 (R8A) and 1.56 wt.%
(R9A), while the samples crosslinked with methyltrialkoxysilane (nonpolar and hydrophobic) show 0.17 (R8M) and
0.21 wt.% (R9M) sorption values (Table 1S).
All samples were analyzed from the point of view of their
thermostability in inert atmosphere, the main parameters of
the thermograms (Figure 13S) being centralized in Table 2S. In
general, the samples decompose in two or more steps with the
main mass losses around 300 C, except the sample containing
chlopropyl group where the main decomposition occurs
around 500 C. The fire retardant effect of the chlorine might
cause this behavior; the residue is also higher in this case. In all
cases the residue amounts are much lower as would be
expected if all silicon converts in SiO2. But this happens in
small extent only, likely due to formation of volatile
byproducts by breaking siloxanic chain during the heating
process. In DSC curves (Figures 14S, 15S, all transitions
characteristic for long polydimethylsiloxane chains are
visible: glass transition around 122 C, crystallization around
75 C, and melting around 40 C with differences of a few
degrees between the samples. For each substituent pair
samples, the crystallization and melting temperatures registered for series R9 were slightly higher than those for series R8
films, generally by 24 C. The glass transition temperatures
for series R9 are generally lower than those registered for
series R8 (Table 2S). These differences are due to higher
molecular weight in series R9 and to higher amount of filler
particles. In certain cases, a polymorphic melting was
registered, probably due to different size and degree of
organization in the crystalline domains formed on cooling.
4. Conclusions
Two polydimethylsiloxane-a,v-diols with different molecular masses (35,000 and 125,000 g  mol 1) were crosslinked
with trialkoxysilanes containing organic groups with different
polarities attached to the silicon: methyl, phenyl, chloropropyl, aminopropyl, and cyanopropyl. While the presence of
the polar groups leads in principal to the raising of the
dielectric permittivity (from 2.5 in case of the methyltrialkoxysilane to 3.7 when cyano-derivative was used as a
crosslinker), the increasing of the chain length of the
polymeric matrix has somewhat a contrary action, lowering
it. The long chains also lead to lowering of the modulus of 1.6
up to 7.7 times depending on the nature of crosslinking agent,
the latter recording in case of the samples containing phenyl

10

http://www.aem-journal.com

groups. The actuation behavior depends in a complex manner


on these parameters. The best results were obtained for the
lower Mn PDMS crosslinked with chloropropyltrialkoxysilane
and for the higher Mn PDMS crosslinked with cyanopropyl
derivative: 30,226 and 8,037nm  mm 1 at a normalized
voltage of 625 V/mm. All samples showed high thermostability, up to 300 and even 500 C in the case of the samples
containing chlorine. The moisture sorption values are slightly
increased in the samples containing polar groups but they
remained at a low level (the highest values were measured for
the sample containing amino groups, 1.86 and 1.56 wt.%).
Received: November 6, 2014
Final Version: December 19, 2014

[1] F. Carpi, G. Gallone, F. Galantini, D. De Rossi,


Enhancing the dielectric permittivity of elastomers
(pp 5168) in Dielectric elestomers as electromechanical
transducers, (Eds: F. Carpi, D. De Rossi, R. Kornbluh,
R. Pelrine, P. Sommer-Larsen), Elsevier, Amsterdam,
2008, Ch. 5.
[2] R. Kornbluh, R. Pelrine, High-performance acrylic and
silicone elastomers in Dielectric elestomers as electromechanical transducers, (Eds: F. Carpi, D. De Rossi,
R. Kornbluh, R. Pelrine, P. Sommer-Larsen), Elsevier,
Amsterdam 2008, Ch. 4.
[3] D. M. Opris, M. Molberg, C. Walder, Y. S. Ko, B. Fischer,
F. A. N
uesch, Adv. Funct. Mater. 2011, 21, 3531.
[4] B. Kussmaul, S. Risse, M. Wegener, G. Kofod, H. Kr
uger,
Smart. Mater. Struct. 2012, 21, 064005.
[5] S. Risse, B. Kussmaul, H. Kr
uger, G. Kofod, Adv. Funct.
Mater. 2012, 22, 3958.
[6] F. B. Madsen, I. Dimitrov, A. Daugaard, S. Hvilsted,
A. L. Skov, Polym. Chem. 2013, 4, 1700.
[7] C. Racles, M. Cazacu, B. Fischer, D. M. Opris, Smart
Mater. Struct. 2013, 22, 104004.
[8] D. Khastgir, K. Adachi, Polymer 2000, 41, 6403.
[9] Z.-H. Lin, Y. Yang, J. M. Wu, Y. Liu, F. Zhang, Z. L. Wang,
J. Phys. Chem. Lett. 2012, 3, 3599.
[10] H. Zhao, D.-R. Wang, J.-W. Zha, J. Zhao, Z.-M. Dang,
J. Mater. Chem. A 2013, 1, 140.
[11] M. Cazacu, M. Ignat, C. Racles, M. Cristea, V. Musteata,
D. Ovezea, D. Lipcinski, J. Compos. Mater. 2014, 48, 1533.
[12] Y. T. Looi, R. Ramli, M. B. H. Othman, A. Zulkifli, Adv.
Mat. Res. 2011, 295297 2393.
[13] J. Heiner, B. Stenberg, M. Persson, Polym. Test. 2003, 22,
253.
[14] W. Noll, in Chemistry and Technology of Silicone, Academic
Press, Orlando 1968.
[15] S. J. Clarson, J. A. Semlyen, in Siloxane Polymers, P T R
Prentice Hall, New Jersey 1993.
[16] A. Colas, Silicones: Preparation, Properties and Performance; Dow Corning: Life Sciences. http://www.
dowcorning.com/content/publishedlit/013077.pdf.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

ADVANCED ENGINEERING MATERIALS 2015,


DOI: 10.1002/adem.201400505

A. Bele et al./Electromechanical Tuning by Crosslinker

ADVANCED ENGINEERING MATERIALS ,


DOI: 10.1002/adem.201400505

[23]
[24]
[25]
[26]
[27]
[28]
[29]

b) C. J. Waschinski, J. C. Tiller, ;1;Biomacromolecules


2005, 6, 235.
M. W. Skinner, C. Qian, S. Grigoras, D. J. Halloran,
B. L. Zimmerman, Textile Res. J. 1999, 69, 935.
A Guide to Silane Solutions, Dow Corning 2005, www.
dowcorning.com/silanes.
HyperChem(TM) Professional 7.51, Hypercube, Inc.,
1115 NW 4th Street, Gainesville, Florida 32601, US.
J. Bicerano, in Prediction of polymer properties, Marcel
Dekker, Inc., New York 2002.
J. Bicerano, J. Macromol. Sci. C Polym. Rev. 1996, 36, 161.
R. Bakule, A. Havranek, J. Polym. Sci. : Polym. Sympos.
1975, 53, 347.
E. A. Chigorina, T. M. Chigorina, A. A. Arutyunyants,
M. V. Bestaev, Polym. Sci. D 2010, 3, 228.

2015 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim

http://www.aem-journal.com

11

FULL PAPER

[17] J. J. Chru ciel, E. Le niak, in Modification of Thermoplastics with Reactive Silanes and Siloxanes in Thermoplastic Elastomers (Eds: A. Z. El-Sonbati), InTech, Rijeka
2012.
[18] M. Cazacu, M. Marcu, Macromol. Rep. A 1995, 32, 1019.
[19] a) H. Mori, Int. J. Pol. Sci. 2012, 173624, 17 pp. b) P. P.
Pescarmona, T. Maschmeyer, Aust. J. Chem. 2001, 54,
583.
[20] Q. Lu, M. E. Mullins, MRS Proceedings 2012, 1400,
mrsf11-1400-s06-02.
[21] L. Y. Tyng, M. R. Ramli, M. B. H. Othman, R. Ramli,
Z. Arifin, M. Ishak, Z. Ahmad, Polym. Int. 2012, 62,
382.
[22] a)G. Sauvet, W. Fortuniak, K. Kazmierski, J. Chojnowski, J. Polym. Sci: Part A: Polym. Chem. 2003, 41, 2939.

Anda mungkin juga menyukai