ArticlePDF Available

Effects of Fumed and Mesoporous Silica Nanoparticles on the Properties of Sylgard 184 Polydimethylsiloxane

Authors:

Abstract and Figures

The effects of silica nanoparticles on the properties of a commonly used Sylgard 184 polydimethylsiloxane (PDMS) in microfluidics were systemically studied. Two kinds of silica nanoparticles, A380 fumed silica nanoparticles and MCM-41 mesoporous silica nanoparticles, were individually doped into PDMS, and the properties of PDMS with these two different silica nanoparticles were separately tested and compared. The thermal and mechanical stabilities of PDMS were significantly enhanced, and the swelling characteristics were also improved by doping these two kinds of nanoparticles. However, the transparency of PDMS was decreased due to the light scattering by nanoparticles. By contrast, PDMS/MCM-41 nanocomposites showed a lower coefficient of thermal expansion (CTE) owing to the mesoporous structure of MCM-41 nanoparticles, while PDMS/A380 nanocomposites showed a larger elastic modulus and better transparency due to the smaller size of A380 nanoparticles. In addition, A380 and MCM-41 nanoparticles had the similar effects on the swelling characteristics of PDMS. The swelling ratio of PDMS in toluene was decreased to 0.68 when the concentration of nanoparticles was 10 wt %.
Content may be subject to copyright.
Micromachines 2015, 6, 855-864; doi:10.3390/mi6070855
micromachines
ISSN 2072-666X
www.mdpi.com/journal/micromachines
Article
Effects of Fumed and Mesoporous Silica Nanoparticles on the
Properties of Sylgard 184 Polydimethylsiloxane
Junshan Liu
1,2,
*, Guoge Zong
1
, Licheng He
1
, Yangyang Zhang
1
, Chong Liu
1
and Liding Wang
1
1
Key Laboratory for Micro/Nano Technology and System of Liaoning Province, Dalian University
of Technology, Dalian 116024, China; E-Mails: zongguoge@163.com (G.Z.);
helicheng868@163.com (L.H.); 15104504983@163.com (Y.Z); chongl@dlut.edu.cn (C.L.);
wangld@dlut.edu.cn (L.W.)
2
Key Laboratory for Precision and Non-traditional Machining Technology of Ministry of Education,
Dalian University of Technology, Dalian 116024, China
* Author to whom correspondence should be addressed; E-Mail: liujs@dlut.edu.cn;
Tel.: +86-411-8470-7713 (ext. 2171); Fax: +86-411-8470-7940.
Academic Editor: Hongrui Jiang
Received: 21 April 2015 / Accepted: 2 July 2015 / Published: 8 July 2015
Abstract: The effects of silica nanoparticles on the properties of a commonly used Sylgard
184 polydimethylsiloxane (PDMS) in microfluidics were systemically studied. Two kinds
of silica nanoparticles, A380 fumed silica nanoparticles and MCM-41 mesoporous silica
nanoparticles, were individually doped into PDMS, and the properties of PDMS with these
two different silica nanoparticles were separately tested and compared. The thermal and
mechanical stabilities of PDMS were significantly enhanced, and the swelling
characteristics were also improved by doping these two kinds of nanoparticles. However,
the transparency of PDMS was decreased due to the light scattering by nanoparticles. By
contrast, PDMS/MCM-41 nanocomposites showed a lower coefficient of thermal
expansion (CTE) owing to the mesoporous structure of MCM-41 nanoparticles, while
PDMS/A380 nanocomposites showed a larger elastic modulus and better transparency due
to the smaller size of A380 nanoparticles. In addition, A380 and MCM-41 nanoparticles
had the similar effects on the swelling characteristics of PDMS. The swelling ratio of PDMS
in toluene was decreased to 0.68 when the concentration of nanoparticles was 10 wt %.
Keywords: polydimethylsiloxane; PDMS; Sylgard 184; silica; nanoparticles
OPEN ACCESS
Micromachines 2015, 6 856
1. Introduction
Polydimethylsiloxane (PDMS) has been widely used in microfluidics due to its flexibility,
biocompatibility, transparency, low cost and ease of fabrication [1–7]. However, PDMS has a high
coefficient of thermal expansion (CTE) and a low elastic modulus, which makes its poor thermal and
mechanical stabilities and hinders many practical applications. For example, the metal microelectrode
has become one of the essential compositions of microfluidic chips. However, wrinkles or even cracks
were caused in metal films deposited on PDMS substrates owing to poor stabilities of PDMS [7–9]. In
addition to stabilities, swelling of PDMS is another important issue [10–12]. The swelling ratios of
PDMS are very large in common organic solvents, such as a swelling ratio of 1.15 in toluene [10].
Hence, it is difficult to perform experiments in organic media for PDMS microfluidic chips.
To improve the stabilities of PDMS, many kinds of nano materials, such as carbon nanotubes
(CNTs) [1315], silica nanoparticles [16–18], and graphene [19] have been doped into PDMS. Wu et al.
reported that the elastic modulus of PDMS was increased to 2.34 MPa by adding carbon nanotubes [13].
Camenzind et al. reported that the elastic modulus of PDMS was increased to 3.18 MPa by admixing
fumed silica nanoparticles [16]. Chen et al. reported a method for embedding CNT sheets in PDMS,
and the CTE of PDMS was reduced to 6 ppm/°C in the direction parallel to CNT alignment [15].
Yamauchis group has investigated thermal, mechanical, optical and swelling properties of PDMS/silica
nanocomposites [17,18]. The mechanical and thermal stabilities were effectively enhanced, and the
swelling characteristics of PDMS were also improved. However, in their studies, a hard PDMS
(X-32-3095, Shin-Etsu Chemical Co., Ltd, Tokyo, Japan), which does not require a curing agent, was used.
In this paper, we systematically studied the effects of silica nanoparticles on the properties of the
most commonly used PDMS in microfluidics, Sylgard 184 (Dow Corning Corporation, Midland, MI,
USA) [3,5,10,11,20]. The thermal and mechanical stabilities, swelling characteristics and transparency
were all examined. Two kinds of silica nanoparticles, fumed silica nanoparticles and mesoporous silica
nanoparticles were doped and compared.
2. Experimental Section
2.1. Materials
A380 fumed silica nanoparticles were purchased from Evonik Degussa (Essen, Germany), and
MCM-41 mesoporous silica nanoparticles were provided by Nanjing XFNANO Materials (Nanjing,
China). These two types of nanoparticles are abbreviated henceforward as A380 and MCM-41,
respectively. Specific characteristics of these two types of silica nanoparticles are listed in Table 1.
According to the density of silica (2.2 g/cm
3
) and the particle size and pore volume of silica
nanoparticles listed in Table 1, the mass of a single nanoparticle was approximately calculated. A
single A380 nanoparticle was about 4 × 10
19
g, and a single MCM-41 nanoparticle was about
1 × 10
14
g.
PDMS (Sylgard 184) consisted of a base and a curing agent was purchased from Dow
Corning Corporation.
Micromachines 2015, 6 857
Table 1. Specific characteristics of silica nanoparticles (SSA, specific surface area).
The Type
SSA (m
2
/g)
Particle size (nm)
Pore diameter (nm)
Pore volume (cm
3
/g)
A380
~380
~7
-
-
MCM-41
~800
~300
3.5–4
0.80.9
2.2. Fabrication of PDMS/Silica Nanocomposites
During the process of fabricating PDMS/silica nanocomposites, the agglomeration of nanoparticles
is inevitable. In order to obviate the agglomeration and achieve a homogeneous dispersion, magnetic
stirring and sonication were often used [13,19]. In addition, organic solvents were often used to help
the dispersion of nanoparticles, such as toluene [21,22] and xylene [13,19]. Our group once used
toluene to dilute PDMS for making a thin film [23], hence, here, toluene was also chosen as the dilute
solvent. In this study, PDMS/silica nanocomposites were fabricated by combining magnetic stirring,
sonication and the dilute solvent. The whole fabrication procedures were illustrated in Figure 1:
(a) Silica nanoparticles were mixed with toluene. The mixture was magnetically stirred for 1 h, and
ultrasonically oscillated for 3 h. (b) The nanoparticles in toluene were mixed with the PDMS base.
Similarly, the mixture was stirred for 1 h and oscillated for 3 h. (c) Toluene in the mixture was
thoroughly evaporated in chemical hood. (d) The PDMS base and curing agent were mixed at a weight
ratio of 10:1. The mixture was stirred uniformly, and then put into a vacuum oven to remove bubbles.
(e) The mixture was poured on a glass plate, and cured at 80 °C for 2 h. Then the PDMS/silica
nanocomposites were peeled off from the glass plate.
2.3. Characterization
Silica nanoparticles were observed by a transmission electron microscope (TEM) (Tecnai G2 F30,
FEI Company, Hillsboro, OR, USA), as shown in Figure 2. The thermal expansion property of
PDMS/silica nanocomposites was examined by a thermomechanical analyzer (Q400, TA Instruments,
New Castle, DE, USA). A specimen with a dimension of 8 mm × 5 mm × 5 mm was prepared. During
the test, the temperature was increased from room temperature to 180 °C at a rate of 5 °C/min. The
CTE of PDMS/silica nanocomposites was determined based on the curves from the analyzer. The
elastic modulus of PDMS/silica nanocomposites was measured by using a universal tensile testing
machine (5657, Instron Company, Norwood, MA, USA). The specimen was fabricated according to
American Society for Testing of Materials (ASTM) standards, and the rate of crosshead motion was
kept at 50 mm/min. To measure the swelling property of PDMS/silica nanocomposites, a specimen
with a dimension of 25 mm × 25 mm × 2 mm was put into toluene, and the weight of the specimen
was recorded every half an hour until the swelling of nanocomposites reached equilibrium. In
accordance with the change in weight caused by absorption, the swelling ratio was calculated [10]. The
transmittance of PDMS/silica nanocomposites was measured by a UV-Vis spectrophotometer
(Cary300, Agilent Technologies, Santa Clara, CA, USA), and the thickness of the specimen was
100 μm.
Micromachines 2015, 6 858
Figure 1. Fabrication procedures of PDMS/silica nanocomposites. (a) Dispersion of silica
nanoparticles in toluene under magnetic stirring and ultrasonic vibration; (b) mixing silica
nanoparticles with PDMS base under magnetic stirring and ultrasonic vibration;
(c) evaporation of toluene; (d) mixing PDMS base with curing agent and removing
bubbles; and (e) cast molding of PDMS/silica nanocomposites on a glass plate.
Figure 2. Transmission electron microscope (TEM) images of silica nanoparticles. (a) A380
fumed silica nanoparticles; (b) MCM-41 mesoporous silica nanoparticles.
3. Results and Discussion
It is worth noting that the Sylgard 184 PDMS base itself contains silica filler. However, these silica
filler are dimethylvinylated and trimethylated silica [24], and different from the silica nanoparticles
(A380 and MCM-41) used in this study. Moreover, during the curing process, the vinyl of the silica
Micromachines 2015, 6 859
filler and the silicon hydride groups of the curing agent would undergo a hydrosilylation reaction to
form a Si-C bond [24]. Therefore, it is believed that the silica nanoparticles used in this work would
not interact with the silica filler, but with the whole three-dimensional PDMS network.
3.1. The CTE of PDMS/Silica Nanocomposites
Strain-temperature curves of PDMS/silica nanocomposites were shown in Figure 3, and the CTE of
PDMS/silica nanocomposites was calculated from the slope of the curves. The CTE of pure PDMS
was 301 ppm/°C. The CTE of PDMS decreased with the silica concentration. The decrease of the
CTE was mainly attributed to the following two factors. On one hand, the CTE of silica is only
0.54 ppm/°C [25], which is almost four orders of magnitude lower than that of pure PDMS. Hence the
CTE of PDMS can be decreased by adding silica nanoparticles. On the other hand, covalent bonds
between silica nanoparticles and PDMS and hydrogen bonds between silica nanoparticles formed [16,26].
The bonds caused large interaction between PDMS and silica nanoparticles, and the interaction
restricted the heat deformation of PDMS.
Figure 3. Strain-temperature curves of PDMS with different concentrations of (a) A380
fumed silica nanoparticles and (b) MCM-41 mesoporous silica nanoparticles; (c) CTEs of
PDMS/A380 and PDMS/MCM-41 nanocomposites. Data were shown as mean ± standard
deviation (SD). #, p < 0.05, compared with control of PDMS/A380 nanocomposites;
*, p < 0.05, compared with control of PDMS/MCM-41 nanocomposites (one-way analysis
of variance, Tukeys post hoc test).
Compared with A380, MCM-41 reduced the CTE of PDMS more efficiently. At a concentration of
10 wt %, the CTE of PDMS/MCM-41 nanocomposites was 272 ppm/°C, while that of PDMS/A380
nanocomposites was 280 ppm/°C. The width of the PDMS chain is about 0.7 nm [27], and the pore
diameter of MCM-41 is 3.54 nm (Table 1). A portion of PDMS chains could be easily penetrated into the
pores of MCM-41 due to capillary force [28]. According to the research by Yamauchis group [17,18], the
thermal expansion of these PDMS chains confined inside the pores of MCM-41 was restricted by the
framework of MCM-41. These confined PDMS chains made few contributions to the CTE of PDMS.
Therefore, PDMS/MCM-41 nanocomposites showed a lower CTE at the same concentration of
silica nanoparticles.
Moreover, it was observed that the viscosity of the mixture of PDMS and A380 was apparently
larger than that of the mixture of PDMS and MCM-41. At the same concentration of nanoparticles, due
to a smaller size, the number of A380 is about 4 orders of magnitude larger than that of MCM-41, and
Micromachines 2015, 6 860
the distance among A380 is also smaller. Therefore, we believed that the interaction among A380
could be stronger, and the relative motion among A380 could be more difficult, so the viscosity of the
mixture of PDMS and A380 was larger. It became very difficult to mix PDMS base and A380 when
the concentration of A380 was over 10 wt %. While the concentration of MCM-41 was easily
increased up to 20 wt %, and, correspondingly, the CTE of PDMS was decreased to 241 ppm/°C.
3.2. The Elastic Modulus of PDMS/Silica Nanocomposites
The elastic modulus (E) of PDMS/silica nanocomposites was calculated based on the stress-strain
region below 40%. Stress-strain curves of PDMS are shown in Figure 4a,b, and effects of
concentrations of nanoparticles on the modulus are shown in Figure 4c. The elastic modulus of pure
PDMS was 1.38 MPa. It can be seen that silica nanoparticles significantly improved the elastic
modulus of PDMS. The modulus of PDMS with 20 wt % MCM-41 was 9.44 MPa, which is almost
seven times of that of pure PDMS. The increase of the elastic modulus mainly resulted from two
aspects. First, the elastic modulus of silica is at least four orders of magnitude larger than that of
PDMS, and, therefore, the modulus of PDMS increased with the amount of nanoparticles. Second,
hydrogen bonds between silica nanoparticles and covalent bonds between silica and PDMS increased
the resistance to deformation [29].
In addition, the elastic modulus of PDMS/A380 nanocomposites was apparently larger than that of
PDMS/MCM-41 nanocomposites at the same concentration. At a concentration of 10 wt %, the elastic
modulus of PDMS/A380 nanocomposites was 7.83 MPa, while that of PDMS/MCM-41
nanocomposites was only 5.44 MPa. As stated above, the interaction among A380 and between A380
and PDMS could be stronger due to the smaller size of A380, and the stronger interaction made a
larger elastic modulus.
Figure 4. Stress-strain curves of PDMS with different concentrations of (a) A380 fumed
silica nanoparticles and (b) MCM-41 mesoporous silica nanoparticles; (c) Elastic moduli
of PDMS/A380 and PDMS/MCM-41 nanocomposites. Data were shown as mean ± standard
deviation (SD). #, p < 0.05, compared with control of PDMS/A380 nanocomposites;
*, p < 0.05, compared with control of PDMS/MCM-41 nanocomposites (one-way analysis
of variance, Tukeys post hoc test).
3.3. Swelling of PDMS/Silica Nanocomposites
The swelling property of PDMS/silica nanocomposites in toluene was examined. The swelling ratio
(R) was calculated as (WW
0
)/W
0
where W
0
and W denoted weights of the PDMS before and after
Micromachines 2015, 6 861
absorbing toluene. As shown in Figure 5, at the beginning, the swelling ratio of PDMS increased
quickly, and reached the maximum after about 2 h. Then the ratio kept nearly constant with time. The
swelling ratio of pure PDMS was 1.142. The ratio remarkably decreased with the silica concentration,
and the ratio was even down to 0.486 when the concentration of MCM-41 was 20 wt %, which is
almost 60% lower than that of pure PDMS. The silica nanoparticles increased the crosslinking density
of PDMS [26]. According to the Flory-Rehner rubber swelling theory [30], the increase of crosslinking
density decreased the swelling ratio of PDMS.
Compared with MCM-41, the reduction of the swelling ratio caused by A380 was expected to be
more effective because the crosslinking density of PDMS/A380 nanocomposites was denser due to its
smaller size [16]. However, as shown in Figure 5, the swelling ratios of PDMS/A380 and
PDMS/MCM-41 nanocomposites were very close at the same concentration. For example, at the
concentration of 10 wt %, the swelling ratio of PDMS/A380 nanocomposites was 0.681, and that of
PDMS/MCM-41 nanocomposites as 0.686. As mentioned above, a portion of PDMS was confined
inside the nanopores in MCM-41. We consider that the absorbing ability of this portion of PDMS was
suppressed by the frameworks of MCM-41, which led to an average smaller swelling ratio of the
whole PDMS.
Figure 5. Time-dependent swelling ratios of PDMS with different concentrations of (a) A380
fumed silica nanoparticles and (b) MCM-41 mesoporous silica nanoparticles in toluene.
3.4. Transparency of PDMS/Silica Nanocomposites
The effect of silica nanoparticles on the transparency of PDMS was studied. The ultravioletvisible
spectroscopy (UVVis) spectra of PDMS/silica nanocomposites are shown in Figure 6. The pure
PDMS had excellent transparency, and the transmittance was above 95% in the range from 350 to
800 nm. The silica nanoparticles can cause significant light scattering, and this will decrease the
transmittance of PDMS. As shown in Figure 6, the transmittance of PDMS decreased with the
concentration of nanoparticles. Moreover, the smaller the wavelength was, the lower the transmittance
of PDMS/silica nanocomposites was.
Furthermore, PDMS/A380 nanocomposites showed much better transparency than PDMS/MCM-41
nanocomposites at the same concentration. At the concentration of 10 wt %, the transmittance of
PDMS/A380 nanocomposites was 81% at 450 nm, while that of PDMS/MCM-41 nanocomposites was
67% at 450 nm. The transmittance of PDMS/silica nanocomposites is mainly related to the size and
Micromachines 2015, 6 862
amount of nanoparticles [31]. Hence, when the concentration of nanoparticles is the same, the
transmittance is dependent on the size of nanoparticles. As shown in Table 1, the size of A380 is much
smaller than MCM-41. Therefore, the scattering loss caused by A380 is smaller, and a better
transparency of PDMS/A380 nanocomposites can be obtained.
Figure 6. Ultravioletvisible spectroscopy (UVVis) spectra of PDMS with different
concentrations of (a) A380 fumed silica nanoparticles and (b) MCM-41 mesoporous
silica nanoparticles.
4. Conclusions
The effects of two kinds of silica nanoparticles on the CTE, elastic modulus, swelling property and
transparency of Sylgard 184 PDMS were investigated and compared. By doping silica nanoparticles,
the stabilities and swelling characteristics of PDMS were remarkably improved, while the
transmittance was decreased. Compared with A380, the reduction of CTE of PDMS caused by MCM-41
was more effective due to its mesoporous structure. Moreover, a higher concentration of MCM-41
could be doped owing to its larger size, and a CTE of 241 ppm/°C was obtained at a concentration of
20 wt %. Owing to the smaller size of A380, the PDMS/A380 nanocomposites showed larger elastic
moduli and better transparency than PDMS/MCM-41 nanocomposites. A380 and MCM-41
nanoparticles had the similar effects on the swelling property of PDMS. In sum, by doping different
types of silica nanoparticles, Sylgard 184 PDMS with different properties can be achieved and applied
for a variety of applications.
Acknowledgments
This work was supported by the National Natural Science Foundation of China (51475080, 51321004),
the National High-tech R&D Program of China (2012AA040406), the Fundamental Research Funds for
the Central Universities (DUT14QY21), Dalian Foundation of Science and Technology (2012J21DW002),
and Funds of Key Laboratory of Liaoning Education Department (LZ2014005).
Author Contributions
Junshan Liu, Chong Liu and Liding Wang conceived and designed the experiments; Guoge Zong,
Licheng He and Yangyang Zhang performed the experiments; Junshan Liu and Guoge Zong analyzed
the data; Junshan Liu and Guoge Zong wrote the paper.
Micromachines 2015, 6 863
Conflicts of Interest
The authors declare no conflict of interest.
References
1. Xia, Y.N.; Whitesides, G.M. Soft lithography. Angew. Chem. Int. Ed. 1998, 37, 550–575.
2. McDonald, J.C.; Whitesides, G.M. Poly(dimethylsiloxane) as a material for fabricating microfluidic
devices. Accounts Chem. Res. 2002, 35, 491–499.
3. Bonk, S.M.; Oldorf, P.; Peters, R.; Baumann, W.; Gimsa, J. Fast prototyping of sensorized cell culture
chips and microfluidic systems with ultrashort laser pulses. Micromachines 2015, 6, 364–374.
4. Pinto, E.; Faustino, V.; Rodrigues, R.O.; Pinho, D.; Garcia, V.; Miranda, J.M.; Lima, R. A rapid
and low-cost nonlithographic method to fabricate biomedical microdevices for blood flow
analysis. Micromachines 2015, 6, 121–135.
5. Yu, Z.; Amirouche, F. An electromagnetically-actuated all-PDMS valveless micropump for drug
delivery. Micromachines 2011, 2, 345–355.
6. Wu, W.Y.; Zhong, X.; Wang, W.; Miao, Q.A.; Zhu, J.J. Flexible PDMS-based three-electrode
sensor. Electrochem. Commun. 2010, 12, 1600–1604.
7. Guo, L.; DeWeerth, S.P. An effective lift-off method for patterning high-density gold interconnects
on an elastomeric substrate. Small 2010, 6, 2847–2852.
8. Graudejus, O.; Goerrn, P.; Wagner, S. Controlling the morphology of gold films on
poly(dimethylsiloxane). ACS Appl. Mater. Interfaces 2010, 2, 1927–1933.
9. Bowden, N.; Brittain, S.; Evans, A.G.; Hutchinson, J.W.; Whitesides, G.M. Spontaneous
formation of ordered structures in thin films of metals supported on an elastomeric polymer.
Nature 1998, 393, 146–149.
10. Kim, B.Y.; Hong, L.Y.; Chung, Y.M.; Kim, D.P.; Lee, C.S. Solvent-resistant PDMS microfluidic
devices with hybrid inorganic/organic polymer coatings. Adv. Funct. Mater. 2009, 19, 3796–3803.
11. Lee, J.N.; Park, C.; Whitesides, G.M. Solvent compatibility of poly(dimethylsiloxane)-based
microfluidic devices. Anal. Chem. 2003, 75, 6544–6554.
12. Koh, K.S.; Chin, J.; Chia, J.; Chiang, C.L. Quantitative studies on PDMS-PDMS interface
bonding with piranha solution and its swelling effect. Micromachines 2012, 3, 427–441.
13. Wu, C.L.; Lin, H.C.; Hsu, J.S.; Yip, M.C.; Fang, W. Static and dynamic mechanical properties of
polydimethylsiloxane/carbon nanotube nanocomposites. Thin Solid Films 2009, 517, 4895–4901.
14. Vast, L.; Carpentier, L.; Lallemand, F.; Colomer, J.F.; van Tendeloo, G.; Fonseca, A.; Delhalle, J.
Multiwalled carbon nanotubes functionalized with 7-octenyltrichlorosilane and n-octyltrichlorosilane:
Dispersion in Sylgard
®
184 silicone and Young’s modulus. J. Mater. Sci. 2009, 44, 3476–3482.
15. Chen, L.; Liu, C.; Liu, K.; Meng, C.; Hu, C.; Wang, J.; Fan, S. High-performance, low-voltage,
and easy-operable bending actuator based on aligned carbon nanotube/polymer composites. ACS
Nano 2011, 5, 1588–1593.
16. Camenzind, A.; Schweizer, T.; Sztucki, M.; Pratsinis, S.E. Structure & strength of silica-PDMS
nanocomposites. Polymer 2010, 51, 1796–1804.
Micromachines 2015, 6 864
17. Suzuki, N.; Kiba, S.; Kamachi, Y.; Miyamoto, N.; Yamauchi, Y. Unusual reinforcement of
silicone rubber compounds containing mesoporous silica particles as inorganic fillers. Phys. Chem.
Chem. Phys. 2012, 14, 3400–3407.
18. Suzuki, N.; Kamachi, Y.; Takai, K.; Kiba, S.; Sakka, Y.; Miyamoto, N.; Yamauchi, Y. Effective
use of mesoporous silica filler: Comparative study on thermal stability and transparency of
silicone rubbers loaded with various kinds of silica particles. Eur. J. Inorg. Chem. 2014, 2014,
2773–2778.
19. Ozbas, B.; ONeill, C.D.; Register, R.A. Multifunctional elastomer nanocomposites with
functionalized graphene single sheets. J. Polym. Sci. B Polym. Phys. 2012, 50, 910–916.
20. Johnston, I.D.; McCluskey, D.K.; Tan, C.K.L.; Tracey, M.C. Mechanical characterization of bulk
Sylgard 184 for microfluidics and microengineering. J. Micromech. Microeng. 2014, 24, 035017.
21. Liu, C.X.; Choi, J.W. Patterning conductive PDMS nanocomposite in an elastomer using
microcontact printing. J. Micromech. Microeng. 2009, 19, 085019.
22. Kong, K.T.S.; Mariatti, M.; Rashid, A.A.; Busfield, J.J.C. Enhanced conductivity behavior of
polydimethylsiloxane (PDMS) hybrid composites containing exfoliated graphite nanoplatelets and
carbon nanotubes. Compos. Part B Eng. 2014, 58, 457–462.
23. Liu, J.; Xu, F.; Wang, S.; Chen, Z.; Pan, J.; Ma, X.; Jiao, X.; Xu, Z.; Liu, C.; Wang, L.
A polydimethylsiloxane electrophoresis microchip with a thickness controllable insulating layer for
capacitatively coupled contactless conductivity detection. Electrochem. Commun. 2012, 25, 147–150.
24. Lee, J.N.; Jiang, X.; Ryan, D.; Whitesides, G.M. Compatibility of mammalian cells on surfaces of
poly(dimethylsiloxane). Langmuir 2004, 20, 11684–11691.
25. Duwe, S.; Arlt, C.; Aranda, S. A detailed thermal analysis of nanocomposites filled with SiO
2
,
AlN or boehmite at varied contents and a review of selected rules of mixture. Compos. Sci. Technol.
2012, 72, 1324–1330.
26. Kaneko, M.L.Q.A.; Yoshida, I.V.P. Effect of natural and organically modified montmorillonite clays
on the properties of polydimethylsiloxane rubber. J. Appl. Polym. Sci. 2008, 108, 2587–2596.
27. Horn, R.G.; Hirz, S.J.; Hadziioannou, G.; Frank, C.W.; Catala, J.M. A reevaluation of forces
measured across thin polymer films: Nonequilibrium and pinning effects. J. Chem. Phys. 1989, 90,
6767–6774.
28. Suzuki, N.; Kiba, S.; Yamauchi, Y. Bimodal filler system consisting of mesoporous silica
particles and silica nanoparticles toward efficient suppression of thermal expansion in silica/epoxy
composites. J. Mater. Chem. 2011, 21, 14941–14947.
29. Shim, S.E.; Isayev, A.I. Rheology and structure of precipitated silica and poly(dimethyl siloxane)
system. Rheol. Acta 2004, 43, 127–136.
30. Flory, P.J.; Rehner, J., Jr. Statistical mechanics of cross-linked polymer networks II. Swelling.
J. Chem. Phys. 1943, 11, 521–526.
31. Novak, B.M. Hybrid nanocomposite materials-between inorganic glasses and organic polymers.
Adv. Mater. 1993, 5, 422–433.
© 2015 by the authors; licensee MDPI, Basel, Switzerland. This article is an open access article
distributed under the terms and conditions of the Creative Commons Attribution license
(http://creativecommons.org/licenses/by/4.0/).
... The silica serves as a reinforcing agent that improves the mechanical properties of the PDMS matrix due to its inherently weak intermolecular force [21]. Reinforcing silicone rubber with mesoporous silica nanoparticles significantly improved its mechanical and thermal stability [22][23][24][25] as compared to a nonporous silica reinforcement. The mesoporous silica reinforcement was more effective at reducing the CTE values because its robust silica framework suppressed the thermal expansion of the silicone rubber confined in the mesopores. ...
... Thus, the addition of TiO 2 /MCM induces a higher thermal stability of the material. The new reactive mesoporous adsorbent, which has a low CTE of silica-ceramic materials, introduced its higher dimensional stability when above room temperature (45-70 • C) to the final modified PDMS [25]. According to the literature [23,24], some PDMS chains may penetrate the pores of MCM-41 due to capillary forces. ...
... As the loading increased from 2.5 to 15 wt%, the transmittance of the PDMS composites decreased to be in the range of 27%-54%, as can be viewed due to the higher absorbance in the visible range. The TiO2/MCM particles can cause light scattering, which decreases the transmittance of PDMS gradually as the particle concentration increases [25]. In the UV range, pristine PDMS has a negligible absorbance between 250-400 nm. ...
Article
Full-text available
In this study, a reactive adsorbent filler was integrated into a polymeric matrix as a novel reactive protective barrier without undermining its mechanical, thermal, and chemical properties. For this purpose, newly synthesized TiO2/MCM/polydimethylsiloxane (PDMS) composites were prepared, and their various properties were thoroughly studied. The filler, TiO2/MCM, is based on a (45 wt%) TiO2 nanoparticle catalyst inside the pores of ordered mesoporous silica, MCM-41, which combines a high adsorption capacity and catalytic capability. This study shows that the incorporation of TiO2/MCM significantly enhances the composite’s Young’s modulus in terms of tensile strength, as an optimal measurement of 1.6 MPa was obtained, compared with that of 0.8 MPa of pristine PDMS. The composites also showed a higher thermal stability, a reduction in the coefficient of thermal expansion (from 290 to 110 ppm/°C), a 25% reduction in the change in the normalized specific heat capacity, and an increase in the thermal degradation temperatures. The chemical stability in organic environments was improved, as toluene swelling decreased by 40% and the contact angle increased by ~15°. The enhanced properties of the novel synthesized TiO2/MCM/PDMS composite can be used in various applications where a high adsorption capacity and catalytic/photocatalytic activity are required, such as in protective equipment, microfluidic applications, and chemical sensor devices.
... To fabricate the BCZT/MCNTs/PDMS composites, the potential agglomeration of BCZT and MCNTs particles was mitigated through magnetic stirring and sonication according to prior research [28,29]. In addition, organic solvents such as toluene [30][31][32] and xylene [28,29] can often be used to enhance nanoparticle dispersion. In our study, we utilized toluene to dilute the viscosity of PDMS from BCZT/MCNTs particles via magnetic stirring before casting a thick film. ...
... Silica particles are chosen as the filler material due to their excellent compatibility with CNTs and PDMS for a uniform dispersion and stable network formation. 30 Furthermore, their wide availability and cost-effectiveness make them a practical option for large-scale applications, and their biocompatibility renders the composite along with PDMS suitable for wearable applications, such as body-heat harvesting. Since both PDMS and silica particles are electrically insulating, all the carrier transport occurs through CNT networks. ...
Article
Carbon nanotube (CNT) networks embedded in a polymer matrix have been extensively studied as a flexible thermoelectric transport medium over the recent years. However, their power factor has been largely limited by the relatively inefficient tunneling transport at junctions between CNTs and the low-density conducting channels throughout the networks. This work demonstrates that significant power factor enhancements can be achieved by adding electrically insulating microscale particles in three-dimensional CNT networks embedded in the polymer matrix. When silica particles of a few μm diameters were co-embedded in single-walled CNT (SWCNT)-polydimethylsiloxane (PDMS) composites, both the electrical conductivity and the Seebeck coefficient were simultaneously enhanced, thereby boosting the power factor by more than a factor of six. We found that the silica microparticles excluded a large volume of the composite from the access of CNTs and caused CNT networks to form around them with the polymer as a binder, resulting in improved network connectivity and alignment of CNTs. Our theoretical calculations based on junction tunneling transport for three-dimensional CNT networks show that the significant power factor enhancement can be attributed to the enhanced tunneling with reduced junction distance between CNTs. Additional power factor enhancement by a factor of three was achieved by sample compression, which further reduced the mean junction distance to enhance tunneling but also reduced the geometric factor at the same time, limiting the enhancement of electrical conductivity.
... Parylene HT exhibits low initial autofluorescence with higher UV light stability, and its transmittance is also superior to that of other polymer materials 47 . Regarding the transmittance of Parylene HT, values greater than 90% can be measured between the biologically relevant wavelength range of 400-980 nm, which is significantly higher than that of polyimide (PI) 10 and polyethylene terephthalate (PET) 73 and similar to that of SU-8 substrates 74 and polydimethylsiloxane (PDMS) 75 . The lower autofluorescence of Parylene HT films is associated with the presence and strength of the C-F bond. ...
Article
Full-text available
The aim of this review is to present a comprehensive overview of the feasibility of using transparent neural interfaces in multimodal in vivo experiments on the central nervous system. Multimodal electrophysiological and neuroimaging approaches hold great potential for revealing the anatomical and functional connectivity of neuronal ensembles in the intact brain. Multimodal approaches are less time-consuming and require fewer experimental animals as researchers obtain denser, complex data during the combined experiments. Creating devices that provide high-resolution, artifact-free neural recordings while facilitating the interrogation or stimulation of underlying anatomical features is currently one of the greatest challenges in the field of neuroengineering. There are numerous articles highlighting the trade-offs between the design and development of transparent neural interfaces; however, a comprehensive overview of the efforts in material science and technology has not been reported. Our present work fills this gap in knowledge by introducing the latest micro- and nanoengineered solutions for fabricating substrate and conductive components. Here, the limitations and improvements in electrical, optical, and mechanical properties, the stability and longevity of the integrated features, and biocompatibility during in vivo use are discussed.
... The CTE (coefficient of thermal expansion) of pure PDMS and UV adhesive is 300 ppm/°C and 73ppm/°C, respectively. Doped graphene PDMS has been shown to have a lower CTE because of the negative CTE of graphene at room temperature [39]. The CTE of graphene-doped PDMS is about 220 ppm/°C [40]. ...
Article
Full-text available
Fiber optic Fabry-Perot Interferometer benefits many requirements in pressure sensing. The variation of pressure sensitivity of the polymer-diaphragm-based Fabry-Perot pressure sensor with temperature is studied by investigating the thermal effect of the cavity air and the diaphragm separately. FP cavity vacuum treatment and multi-curvature diaphragm simulation and experimental studies are conducted. Experimental results show that the sensor pressure sensitivity decreases with increasing temperature by 0.46nm/(kPa·°C). The diaphragm’s thermal effect is the leading cause of temperature-induced inconsistency in pressure sensitivity, accounting for 0.43nm/(kPa·°C).
Article
Full-text available
Boron (specifically ¹⁰B) is the element of choice to shield thermal neutrons due to its large (n, α) cross‐section; however, very few polymer composites containing high boron concentrations are available. This study aimed to determine the maximum possible amount of boron that could be introduced into a polymer matrix. Diverse manufacturing techniques, ranging from additive manufacturing to compression molding, were employed to fabricate inks and filaments for 3D printing, foams, and flexible pads. Composites using siloxanes, poly(lactic acid), and acrylonitrile butadiene styrene containing up to 80 wt% boron were sucessufully fabricated. The addition of known plasticizers (polyethylene glycol) and reinforcing agents (carbon nanofibers and fumed silica) helped to overcome fabrication problems such as clogging of the printing nozzle or crumbling of compression molded parts. In addition, the thermal‐mechanical properties of these novel boron composites were determined and shown to vary according to boron concentration, presence of additives, and fabrication techniques utilized.
Article
Full-text available
Polydimethylsiloxane (PDMS)‐based membranes containing amine‐functionalized and unfunctionalized silica particles were fabricated via direct ink writing for CO2/N2 gas separation. The printability of the inks was evaluated by rheological measurements, while spectroscopy, microscopy, thermal measurements, and mechanical testing were employed to characterize the printed membranes. The surface morphology of the membranes revealed the absence of voids, demonstrating their suitability for gas separation. The printed membranes also exhibited desirable thermal and mechanical properties (i.e., thermal degradation temperature of 518 °C and tensile strength as high as 1.178 MPa with 529% elongation). The PDMS‐based membranes generally displayed high permeability for CO2 but slightly low selectivity for the CO2/N2 gas pair. The best‐combined permeability‐selectivity performance of 8794 barrer and selectivity of 11.64 was demonstrated by the printed PDMS membrane containing no SiO2 fillers. The inclusion of unfunctionalized SiO2 particles generally increased the membrane's gas permeability but compromised the selectivity. In contrast, membranes with amine‐functionalized silica showed improved selectivity compared to membranes containing unfunctionalized silica. Overall, the performance and characteristics offered by the PDMS/silica composite membranes demonstrated the potential of 3D printing as an economical and sustainable fabrication approach to developing materials for carbon capture applications.
Article
Full-text available
Mesoporous silica nanoparticles (MSN) characterized by large surface area, pore volume, tunable chemistry, and biocompatibility have been widely studied in nanomedicine as imaging and therapeutic carriers. Most of these studies focused on spherical particles. In contrast, mesoporous silica rods (MSR) that are more challenging to prepare have been less investigated in terms of toxicity, cellular uptake, or biodistribution. Interestingly, previous studies showed that silica rods penetrate fibrous tissues or mucus layers more efficiently than their spherical counterparts. Recently, we reported the synthesis of MSR with distinct aspect ratios and validated their use in multiple imaging modalities by loading the pores with maghemite nanocrystals and functionalizing the silica surface with green and red fluorophores. Herein, based on an initial hypothesis of high liver accumulation of the MSR and a future vision that they could be used for early diagnosis or therapy in fibrotic liver diseases; the cytotoxicity and cellular uptake of MSR were assessed in zebrafish liver (ZFL) cells and the in vivo safety and biodistribution was investigated via fluorescence molecular imaging (FMI) and magnetic resonance imaging (MRI) employing zebrafish larvae and rodents. The selection of these animal models was prompted by the well-established fatty diet protocols inducing fibrotic liver in zebrafish or rodents that serve to investigate highly prevalent liver conditions such as non-alcoholic fatty liver disease (NAFLD). Our study demonstrated that magnetic MSR do not cause cytotoxicity in ZFL cells regardless of the rods' length and surface charge (for concentrations up to 50 mg ml −1 , 6 h) and that MSR are taken up by the ZFL cells in large amounts despite their length of ∼1 mm. In zebrafish larvae, it was observed that they could be safely exposed to high MSR concentrations (up to 1 mg ml −1 for 96 h) and that the rods pass through the liver without causing toxicity. The high accumulation of MSR in rodents' livers at short post-injection times (20% of the administered dose) was confirmed by both FMI and MRI, highlighting the utility of the MSR for liver imaging by both techniques. Our results could open new avenues for the use of rod-shaped silica particles in the diagnosis of pathological liver conditions.
Article
Full-text available
Microfluidic devices are electrical/mechanical systems that offer the ability to work with minimal sample volumes, short reactions times, and have the possibility to perform massive parallel operations. An important application of microfluidics is blood rheology in microdevices, which has played a key role in recent developments of lab-on-chip devices for blood sampling and analysis. The most popular and traditional method to fabricate these types of devices is the polydimethylsiloxane (PDMS) soft lithography technique, which requires molds, usually produced by photolithography. Although the research results are extremely encouraging, the high costs and time involved in the production of molds by photolithography is currently slowing down the development cycle of these types of devices. Here we present a simple, rapid, and low-cost nonlithographic technique to create microfluidic systems for biomedical applications. The results demonstrate the ability of the proposed method to perform cell free layer (CFL) measurements and the formation of microbubbles in continuous blood flow.
Article
Full-text available
We developed a confined microfluidic cell culture system with a bottom plate made of a microscopic slide with planar platinum sensors for the measurement of acidification, oxygen consumption, and cell adhesion. The slides were commercial slides with indium tin oxide (ITO) plating or were prepared from platinum sputtering (100 nm) onto a 10-nm titanium adhesion layer. Direct processing of the sensor structures (approximately three minutes per chip) by an ultrashort pulse laser facilitated the production of the prototypes. pH-sensitive areas were produced by the sputtering of 60-nm Si3N4 through a simple mask made from a circuit board material. The system body and polydimethylsiloxane (PDMS) molding forms for the microfluidic structures were manufactured by micromilling using a printed circuit board (PCB) milling machine for circuit boards. The microfluidic structure was finally imprinted in PDMS. Our approach avoided the use of photolithographic techniques and enabled fast and cost-efficient prototyping of the systems. Alternatively, the direct production of metallic, ceramic or polymeric molding tools was tested. The use of ultrashort pulse lasers improved the precision of the structures and avoided any contact of the final structures with toxic chemicals and possible adverse effects for the cell culture in lab-on-a-chip systems.
Article
Full-text available
Polydimethylsiloxane (PDMS) elastomers are extensively used for soft lithographic replication of microstructures in microfluidic and micro-engineering applications. Elastomeric microstructures are commonly required to fulfil an explicit mechanical role and accordingly their mechanical properties can critically affect device performance. The mechanical properties of elastomers are known to vary with both curing and operational temperatures. However, even for the elastomer most commonly employed in microfluidic applications, Sylgard 184, only a very limited range of data exists regarding the variation in mechanical properties of bulk PDMS with curing temperature. We report an investigation of the variation in the mechanical properties of bulk Sylgard 184 with curing temperature, over the range 25 °C to 200 °C. PDMS samples for tensile and compressive testing were fabricated according to ASTM standards. Data obtained indicates variation in mechanical properties due to curing temperature for Young's modulus of 1.32–2.97 MPa, ultimate tensile strength of 3.51–7.65 MPa, compressive modulus of 117.8–186.9 MPa and ultimate compressive strength of 28.4–51.7 GPa in a range up to 40% strain and hardness of 44–54 ShA.
Article
Elastomeric stamps and molds provide a great opportunity to eliminate some of the disadvantages of photolithograpy, which is currently the leading technology for fabricating small structures. In the case of "soft lithography" there is no need for complex laboratory facilities and high-energy radiation. Therefore, this process is simple, inexpensive, and accessible even to molecular chemists. The current state of development in this promising area of research is presented here.
Article
Elastomeric stamps and molds provide a great opportunity to eliminate some of the disadvantages of photolithograpy, which is currently the leading technology for fabricating small structures. In the case of “soft lithography” there is no need for complex laboratory facilities and high-energy radiation. Therefore, this process is simple, inexpensive, and accessible even to molecular chemists. The current state of development in this promising area of research is presented here.
Article
Polydimethylsiloxane (PDMS) hybrid composites consisting of exfoliated graphite nanoplatelets (xGnPs) and multiwalled carbon nanotubes functionalized with hydroxyl groups (MWCNTs-OH) were fabricated, and the effects of the xGnP/MWCNT-OH ratio on the thermal, electrical, and mechanical properties of polydimethylsiloxane (PDMS) hybrid composites were investigated. With the total filler content fixed at 4 wt%, a hybrid composite consisting of 75% x GnP/25% MWCNT-OH showed the highest thermal conductivity (0.392 W/m K) and electrical conductivity (1.24 x 10(-3) S/m), which significantly exceeded the values shown by either of the respective single filler composites. The increased thermal and electrical conductivity found when both fillers are used in combination is attributed to the synergistic effect between the fillers that forms an interconnected hybrid network. In contrast, the various different combinations of the fillers only showed a modest effect on the mechanical behavior, thermal stability, and thermal expansion of the PDMS composite.
Article
Here we propose a bimodal inorganic filler system consisting of mesoporous silica particles and fine silica nanoparticles to fabricate silica/epoxy composites with low thermal expansion property. Two types of inorganic fillers are mechanically mixed with epoxy resin firstly. Then, by adding a curing agent, silica/epoxy composites are prepared. From TEM observation, it is proved that both mesoporous silica particles and silica nanoparticles are well dispersed in the composites. Through the theoretical calculation, more than 90 vol% of the mesopores is estimated to be filled with the epoxy polymers. Various types of the composites are prepared by changing the amounts of doped silica nanoparticles. (The amounts of doped mesoporous silica particles are constant.) Their thermal expansion behaviours are systematically investigated by using thermal mechanical analysis (TMA). With the increase of the amounts of silica nanoparticles, the glass transition temperatures (Tg) are gradually shifted to the higher temperature range and the coefficient of linear thermal expansion (CTE) values are decreased. Thus, by utilizing the bimodal inorganic filler system, we can efficiently reduce thermal expansion in the silica/epoxy composites.
Article
Four types of silica fillers (nonporous silica particles, mesoporous silica particles, silica gels with micropores, and fumed silica particles) have been used as inorganic filler materials for silicone rubber. To determine their effectiveness as filler materials, a systematic study on the thermal strength and transparency of the obtained silica/silicone composites was conducted. The mesoporous silica/silicone composite shows the lowest thermal expansion and higher transparency than the other composites. The thermal strength and transparency of silica gel/silicone composite is quite similar to those of nonporous silica/silicone composite. Fumed silica/silicone composite is highly transparent (indeed, its transparency is the same as pristine silicone), but the thermal strength is at the same level as nonporous silica/silicone composite. These results clarify that mesoporous silica particles are the most promising filler materials for silicone rubber.
Article
We demonstrate the use of functionalized graphene sheets (FGSs) as multifunctional nanofillers to improve mechanical properties, lower gas permeability, and impart electrical conductivity for several distinct elastomers. FGS consists mainly of single sheets of crumbled graphene containing oxygen functional groups and is produced by the thermal exfoliation of oxidized graphite (GO). The present investigation includes composites of FGS and three elastomers: natural rubber (NR), styrene–butadiene rubber, and polydimethylsiloxane (PDMS). All of these elastomers show similar and significant improvements in mechanical properties with FGS, indicating that the mechanism of property improvement is inherent to the FGS and not simply a function of chemical crosslinking. The decrease in gas permeability is attributed to the high aspect ratio of the FGS sheets. This creates a tortuous path mechanism of gas diffusion; fitting the permeability data to the Nielsen model yields an aspect ratio of ∼1000 for the FGS. Electrical conductivity is demonstrated at FGS loadings as low as 0.08% in PDMS and reaches 0.3 S/m at 4 wt % loading in NR. This combination of functionalities imparted by FGS is shown to result from its high aspect ratio and carbon-based structure. © 2012 Wiley Periodicals, Inc. J Polym Sci Part B: Polym Phys, 2012
Article
A polydimethylsiloxane (PDMS) electrophoresis microchip with a thickness-controllable insulating layer for capacitatively coupled contactless conductivity detection is presented. A PDMS film is spin-coated on a glass slide with Pt microelectrodes to be the insulating layer, and then permanently bonded with a PDMS channel plate to make the whole microchip. The thickness of PDMS films can be precisely controlled down to submicrometers. With a microchip with a 0.6 μm-thick PDMS insulating layer, a superior limit of detection (LOD) of 0.07 μM for Na+ was obtained. For comparison, another two microchips with the same design but different insulating layer thicknesses (15 μm and 50 μm) were tested, and LODs were 1 μM and 3 μM respectively, which are almost two orders of magnitude higher. Moreover, the microchip detector also exhibited excellent linearity, reproducibilities and separation efficiencies.