Vous êtes sur la page 1sur 5

Journal of

Materials Chemistry

View Online / Journal Homepage / Table of Contents for this issue

Dynamic Article Links <

Cite this: J. Mater. Chem., 2012, 22, 6213

PAPER

www.rsc.org/materials

Downloaded by Donghua University on 30 March 2012


Published on 17 February 2012 on http://pubs.rsc.org | doi:10.1039/C2JM16111G

Deposition of amino-functionalized polyhedral oligomeric silsesquioxanes on


graphene oxide sheets immobilized onto an amino-silane modified silicon
surface
Luca Valentini,*a Silvia Bittolo Bon,a Orietta Monticellib and Jose M. Kennya
Received 23rd November 2011, Accepted 3rd February 2012
DOI: 10.1039/c2jm16111g
The reaction between amino-functionalized polyhedral oligomeric silsesquioxanes (POSS-NH2) and
graphene oxide (GO) sheets was used to graft POSS-NH2 onto a GO layer immobilized onto a layer of
(3-aminopropyl)triethoxysilane, self-assembled onto Si substrate. The chemical composition and
surface morphology as well as the surface properties of the prepared films were investigated by means of
Fourier transform infrared spectroscopy, water contact angle measurements and field emission electron
microscopy. Tribological measurements performed with a nanoindenter showed that such hydrophobic
trilayer film exhibited a reduced friction coefficient suitable for applications in lubricant coatings.

1. Introduction
Polyhedral oligomeric silsesquioxanes (POSS) are organicinorganic molecules, approximately 1 to 3 nm in diameter, with the
general formula (RSiO1.5)n where R is hydrogen or an organic
group, such as alkyl, aryl or any of their derivatives.13 The
incorporation of POSS molecules into polymeric materials leads
to significant improvements in the polymeric mechanical properties, thermal behavior and flame retardancy.46
Graphene oxide (GO) has created a lot of attention as it
provides a pathway to produce large quantities of graphene
sheets in solution at low cost.79 The easy processibility of GO
and its compatibility with various substrates, including plastics,
makes it an attractive candidate for the high yield manufacturing
of graphene based flexible electronics.
The structure of graphene oxide sheets consists of a basal plane
of the sheet, decorated with hydroxyl and epoxy functional
groups. Carbonyl groups are present as carboxylic acids along
the sheet edge but also as organic carbonyl defects within the
sheet.10,11
These functional groups provide a variety of surface-modification reactions, which can be used to develop functionalized
GO-based materials. For example graphene oxide is intrinsically
hydrophilic and its dispersion in water is easily achieved.1215
The chemical interaction between amino groups and functional groups of graphene oxide has also been extensively
debated in the literature.1619 For example, the treatment of GO
a

Dipartimento di Ingegneria Civile e Ambientale, Universit


a di Perugia,
Strada di Pentima 4, INSTM, UdR Perugia, 05100, Terni, Italy. E-mail:
mic@unipg.it; Tel: 0039-0744-492924
b
Dipartimento di Chimica e Chimica Industriale, Universit
a di Genova, Via
Dodecaneso 31, 16146 Genova, Italy
Electronic supplementary information (ESI) available. See DOI:
10.1039/c2jm16111g

This journal is The Royal Society of Chemistry 2012

with alkylamine can lead to the derivatization of both the edge


carboxyl and the surface epoxy functional groups, as suggested
by Bourlinos et al.,17 who proposed the covalent bonding of
amine on epoxy groups or by Matsuo et al.,18 who suggested the
hydrogen bonding between amines and hydroxyl groups on the
basal plane. Lerf et al.10 discussed the interaction of the amine
with carboxyl groups along the graphene oxide sheet edge.
The approach, involving a direct coupling of amino functionalized POSS with graphene oxide sheets to introduce amino
groups via amide formation, could lead to a novel nanostructured hybrid material that merges the properties of the two
starting nanomaterials. The approach presented in this work is
the grafting of an amino functionalized POSS onto a GO layer
immobilized onto a (3-aminopropyl)triethoxysilane modified Si
substrate. In this paper, we report on the realization of such
layered film with hydrophobic properties, along with a low
friction coefficient. To the best of our knowledge, the utilization
of POSS assembled onto GO sheets as thin-film lubricants has
not been investigated.

2. Experimental
Graphene oxide was purchased from Cheaptubes (GO; thickness
0.71.2 nm estimated by AFM, see supporting information).
Aminopropyl heptaisobutyl POSS (referred to as POSS-NH2
from now on) was purchased from Hybrid Plastics (USA) as
crystalline powder and used as received.
GO water solution (1 mg ml 1) was prepared and sonicated
(750 W, 60% amplitude) for 1 h to yield a yellow suspension.
After this mixing process, the solution was transferred to a vial
and it was centrifuged for 30 min at 9000 rpm. The supernatant
of the dispersion was carefully extracted and separated from the
residual visible at the bottom of the vial.
J. Mater. Chem., 2012, 22, 62136217 | 6213

Downloaded by Donghua University on 30 March 2012


Published on 17 February 2012 on http://pubs.rsc.org | doi:10.1039/C2JM16111G

View Online

Commercially available silicon with an electrical sheet resistivity of 14 Ohms/square has been used as substrate. The Si
surface was cleaned with ethanol and acetone, rinsed with water,
dried under nitrogen and taken inside a dry Ar glovebox.
Afterwards, the Si surface was treated in oxygen plasma for
10 min. The plasma was generated by a radio frequency power
supply (13.56 MHz) and carried out at room temperature with
the gas pressure fixed at 4.5  10 2 Torr. The power employed
was fixed at 20 W (substrate bias 300 V). The plasma-treated Si
surface was then immersed into a 1% (3-aminopropyl)triethoxysilane (APTES, purchased from Aldrich) solution in toluene for 1
h at room temperature. Following this treatment, the APTES
modified Si substrate was rinsed in toluene and dried in nitrogen.
The characterization of the APTES self-assembled monolayer
has already been reported in our previous study.20 The thickness
of the prepared films was about 58 nm as estimated by atomic
force microscopy.20
Subsequently, the APTES modified Si substrate was kept in
the prepared GO aqueous solution and rinsed at 80  C for 1 h.
This sample, referred to hereafter as APTES-GO, was immersed
into a POSS-NH2 solution of hexane at room temperature for
1 h. The prepared sample was washed with hexane to remove the
excess of POSS-NH2. After that, this sample, referred to as
APTES-GPOSS, was dried in a nitrogen stream.
Infrared spectroscopy in the 5004000 cm 1 range, was used to
confirm the chemical composition. The surface morphology of
the films was investigated by optical and field emission scanning
electron microscopy (FE-SEM).
The wettability of these films was investigated by contact
angle measurements. Contact angles were measured with a FTA
1000 Series instrument equipped with a CCD camera. Deionized water droplets were dropped carefully onto the surfaces
and the contact angle was monitored statically as a function of
time.
Tribological properties were investigated through instrumented indentation tests by using a Micro-Materials nanoindenter (Nanotest 600). A Berkovich tip was used, providing
a load of 0.1 mN. The friction coefficient versus scanning length
was recorded for each test. Five valid tests were performed on
each sample. Si samples coated with different films were mounted
onto the flat base and the friction coefficients were recorded
automatically.

3. Results and discussion


IR spectroscopy was used to verify the functionalization of the
oxygen plasma treated Si surface by the amino-silane functional
groups. The CH2 peaks were used rather than the NH2 peak to
confirm the film growth because of the overlap signal of NH2 at
32003500 cm 1 with the H2O peak, which is present in the
background. The positions of the two CH2 peaks in the spectrum
at 2854 and 2927 cm 1, corresponding to the symmetric and
asymmetric CH2 stretch respectively, are in agreement with
literature values2022 and confirm the presence of APTES on the
Si surface (see supporting information).
For the neat Si surface, the equilibrium contact angle was 51 ,
while that for the oxygen plasma treated one was about 41
(Fig. 1ab). The amino-silane modified surface showed an
increase of the contact angle value (Fig. 1c).
6214 | J. Mater. Chem., 2012, 22, 62136217

Fig. 1 Contact angle images of a) Si substrate, b) oxygen plasma treated


Si substrate, c) APTES modified Si substrate, d) APTES-GO sample and
e) APTES-GPOSS sample. The contact angle values are reported on the
right-hand side of the images.

The morphology of the aqueous GO dispersions deposited


onto neat Si and APTES modified Si substrate was evaluated
using optical microscopy. The optical image shows that GO,
deposited onto Si substrate, consists of isolated platelets
(Fig. 2a). What is observed for the APTES-GO sample was the
formation of an extended area with a paper-like structure
(Fig. 2b) which was found to be more hydrophilic than the
APTES modified Si surface (Fig. 1d) due to the surface hydroxyl
groups, which are expected to serve as the active points to induce
the subsequent anchoring of POSS-NH2 molecules, as confirmed
by the FTIR characterization reported below.
The structural morphology of the APTES-GO sample reported in Fig. 2b can be schematically rationalized as reported in
Scheme 1. The Si substrate is fully covered by APTES and GO
can be grafted onto the APTES layer due to the interaction
between NH2 (head groups of APTES) and the COOH groups
of the underlayer GO sheet. Consequently, APTES-GO appears
continuous on the APTES modified Si substrate (Fig. 2b).
The most characteristic features in the FTIR spectrum of the
GO sample (Fig. 3a) are the absorption bands corresponding to
the C]O carbonyl stretching at 1720 cm 1, and sp2-hybridized
C]C in-plane stretching at 1630 cm 1. A broad and intense
signal between 3000 and 3700 cm 1 can be assigned to hydroxyls
with contributions from COOH and H2O (i.e. COH).23
The IR spectrum of the APTES-GO sample (Fig. 3b) shows
the disappearance of the signal between 3000 and 3700 cm 1
observed for the GO sample and the presence of new bands
at 1573 and 1223 cm 1, corresponding to NH in-plane and CN
This journal is The Royal Society of Chemistry 2012

Downloaded by Donghua University on 30 March 2012


Published on 17 February 2012 on http://pubs.rsc.org | doi:10.1039/C2JM16111G

View Online

Fig. 2 Optical micrographs (450 mm  600 mm) of (a) neat SiGO and
(b) APTES-GO samples, respectively.

Fig. 3 FTIR spectra of a) GO, b) APTES-GO and c) POSS-NH2 and


APTES-GPOSS samples, respectively.

Scheme 1 Proposed schematic view for the construction of APTESGPOSS trilayer film.

bond stretching, respectively. This further confirms the presence


of the amide functional group when GO is deposited onto the
APTES modified Si surface.24
This journal is The Royal Society of Chemistry 2012

Fig. 3c shows the FTIR spectrum of POSS-NH2. The wide


band at 3350 cm 1 is attributed to the characteristic absorption
of the NH2 groups. The strong double peaks at 2930 and
2873 cm 1 correspond to the CH stretching of the CH2 groups
in the organic corner groups of the cage structure. The absorption band at 1120 cm 1 is the characteristic vibration of an
SiOSi bond. The absorption peak at 1030 cm 1 is attributed to
the special characteristic vibration of the silsesquioxane cage
SiOSi framework.25 The peak at 760 cm 1 is the bending
vibration of the SiC bond in SiCH2.25
The IR spectrum of the APTES-GPOSS sample (Fig. 3c),
shows the disappearance of the band at 1720 cm 1 observed for
the APTES-GO sample and the corresponding appearance of
a band with lower frequency (1661 cm 1), assigned to the amide
carbonyl stretch. Similar results for amide-functionalized carbon
nanostructures (i.e. single-walled carbon nanotubes) have been
reported previously in the literature.24,26
The reaction between the GO carboxylic groups and the
POSS-NH2 amino groups led to hydrophobic behavior of the
APTES-GPOSS sample (Fig. 1e). Considering the structure of
our amino functionalized POSS molecules, it is expected that
J. Mater. Chem., 2012, 22, 62136217 | 6215

Downloaded by Donghua University on 30 March 2012


Published on 17 February 2012 on http://pubs.rsc.org | doi:10.1039/C2JM16111G

View Online

the modification of the GO surface with alkyl chain molecules


can switch the surface properties from hydrophilic to
hydrophobic.
The FE-SEM characterization performed on the APTES-GO
(Fig. 4a) and APTES-GPOSS (Fig. 4b) samples demonstrates
that, after the grafting of POSS-NH2, the graphene sheets
roughness increases, with the presence of crystals. Corrugation
effects have been previously observed for chemically converted
graphene produced by reduction of graphene oxide.27,28 Simulations suggest that different chemical addends and their
arrangements may result in significant wrinkling and even
bending of graphene sheets.29 The removal of the oxygen-containing functional groups from the basal plane of graphene flakes
upon POSS-NH2 grafting could also contribute to the change in
its corrugation.30
The APTES-GPOSS trilayer film is thus expected to have
different tribological properties with respect to the APTES
modified Si substrate and APTES-GO sample, respectively. In
this regard it is known that silane based self-assembled monolayers on an Si substrate can be used as lubricant coatings for
nano/microelectromechanical systems.31,32 In particular, it has

Fig. 4 FE-SEM images for a) APTES-GO and b) APTES-GPOSS


samples.

6216 | J. Mater. Chem., 2012, 22, 62136217

been demonstrated that such self-assembled monolayers are


hydrophobic and can reduce friction.
Fig. 5 presents the curves of friction coefficient versus distance
for the APTES modified Si substrate, APTES-GO and APTESGPOSS samples, respectively. Compared with APTES modified
Si (Fig. 5a), APTES-GO (Fig. 5b) exhibits a lower friction
coefficient, as expected for a graphite-based lubricant that tends
to glide during the test.
Once the POSS-NH2 was grafted onto the GO outlayer,
a lower friction coefficient was obtained (Fig. 5c). From these
results, it can be summarized that the low friction is due to the
joint effect of GO and POSS molecules. Accordingly to what is
reported in Fig. 4b, we can reasonably retain the possibility that
the lower friction was due to the fact that the probe has less
surface contact due to the increased surface roughness. Further
investigations are in progress in this regard in our lab.

Fig. 5 Friction coefficient for: (a) APTES modified Si, (b) APTES-GO
and (c) APTES-GPOSS samples, respectively. The average friction
coefficient is reported above the corresponding curve.

This journal is The Royal Society of Chemistry 2012

View Online

4. Conclusions

Downloaded by Donghua University on 30 March 2012


Published on 17 February 2012 on http://pubs.rsc.org | doi:10.1039/C2JM16111G

Amino functionalized polyhedral oligomeric silsesquioxanes


were grafted onto GO sheets self-assembled onto an APTES
modified silicon surface. The morphology, the wettability and the
tribological properties were investigated. Results indicate that
such a trilayer film exhibits hydrophobicity with a better friction
coefficient compared with the Si substrate and APTES-GO
samples. This approach can be used for the preparation of
hydrophobic and lubricant coatings based on graphene hybrid
materials.

Acknowledgements
We are grateful to the Italian Ministry of Education and
University through the 2008 PRIN project (Grant No.
20089B75ML_001).
Dedication: In memoriam of Daniela Tabuani.

References
1 P. G. J. Harrison, J. Organomet. Chem., 1997, 542, 141183.
2 R. H. Baney, M. Itoh, A. Sakakibara and T. Suzuki, Chem. Rev.,
1995, 95, 14091430.
3 M. G. Voronkov and V. I. Lavrentyev, Top. Curr. Chem., 1982, 102,
199236.
4 A. Fina, O. Monticelli and G. Camino, J. Mater. Chem., 2010, 20,
9297.
5 O. Monticelli, A. Fina, A. Ullah and P. Waghmare, Macromolecules,
2009, 42, 6614.
6 O. Monticelli, A. Fina, E. S. Cozza, M. Prato and V. Bruzzo, J.
Mater. Chem., 2011, 21, 18049.
7 S. Park and R. S. Ruoff, Nat. Nanotechnol., 2009, 4, 217.
8 S. Park, J. An, I. Jung, R. D. Piner, S. Jin An, X. Li, A. Velamakanni
and R. S. Ruoff, Nano Lett., 2009, 9, 1593.
9 S. Stankovich, Carbon, 2007, 45, 1558.
10 A. Lerf, H. He, M. Forster and J. Klinowski, J. Phys. Chem. B, 1998,
102, 4477.

This journal is The Royal Society of Chemistry 2012

11 H. Y. He, J. Klinowski, M. Forster and A. Lerf, Chem. Phys. Lett.,


1998, 287, 53.
12 T. Szabo, E. Tombacz, E. Illes and I. Dekany, Carbon, 2006, 44,
537.
13 G. I. Titelman, V. Gelman, S. Bron, R. L. Khalfin, Y. Cohen and
H. Bianco-Peled, Carbon, 2005, 43, 641.
14 M. Hirata, T. Gotou, S. Horiuchi, M. Fujiwara and M. Ohba,
Carbon, 2004, 42, 2929.
15 S. Stankovich, R. D. Piner, S. T. Nguyen and R. S. Ruoff, Carbon,
2006, 44, 3342.
16 Y. Matsuo, T. Miyabe, T. Fukutsuka and Y. Sugie, Carbon, 2007, 45,
1005.
17 A. B. Bourlinos, D. Gournis, D. Petridis, T. Szabo, A. Szeri and
I. Dekany, Langmuir, 2003, 19, 6050.
18 Y. Matsuo, K. Watanabe, T. Fukutsuka and Y. Sugie, Carbon, 2003,
41, 1545.
19 O. C. Compton, D. A. Dikin, K. W. Putz, L. C. Brinson and
S. T. Nguyen, Adv. Mater., 2010, 22, 892.
20 L. Valentini, M. Cardinali and J. M. Kenny, Carbon, 2010, 48, 861.
21 D. G. Kurth, Langmuir, 1993, 9, 2965.
22 K. Nakanishi, P. Solomon, Infrared absorption spectroscopy, San
Francisco: Holden-Day, 1977.
23 M. Acik, G. Lee, C. Mattevi, A. Pirkle, R. M. Wallace,
M. Chhowalla, K. Cho and Y. Chabal, J. Phys. Chem. C, 2011,
115, 19761.
24 D. B. Mawhinney, V. Naumenko, A. Kuznetsova, J. T. Yates, J. Liu
and R. E. Smalley, J. Am. Chem. Soc., 2000, 122, 2383.
25 F. J. Feher, D. A. Newman and J. F. Walzer, J. Am. Chem. Soc., 1989,
111, 1741.
26 J. B. Lambert, H. F. Shurvell, D. A. Lightner, R. G. Cook, Organic
structural spectroscopy, Prentice Hall: Upper Saddle River, NJ, 1998.
27 A. Lerf, H. He, M. Forster and J. Klinowski, J. Phys. Chem. B, 1998,
102, 4477.
28 S. Wang, Y. Zhang, N. Abidi and L. Cabrales, Langmuir, 2009, 25,
11078.
29 H. C. Schniepp, J. L. Li, M. J. McAllister, H. Sai, M. Herrera-Alonso
and D. H. Adamson, J. Phys. Chem. B, 2006, 110, 8535.
30 J. T. Paci, T. Belytschko and G. C. Schatz, J. Phys. Chem. C, 2007,
111, 18099.
31 S. L. Ren, S. R. Yang, Y. P. Zhao, J. F. Zhou, T. Xu and W. M. Liu,
Tribol. Lett., 2002, 13, 233.
32 I. H. Sung, J. C. Yang, D. E. Kim and B. S. Shin, Wear, 2003, 255,
808.

J. Mater. Chem., 2012, 22, 62136217 | 6217

Vous aimerez peut-être aussi