You are on page 1of 24

Accepted Manuscript

A rapid and eco-friendly route to synthesize graphene-doped silica nanohybrids

Andrea Maio, Simonpietro Agnello, Reza Khatibi, Luigi Botta, Antonino Alessi, Aurora
Piazza, Gianpiero Buscarino, Alessio Mezzi, Giuseppe Pantaleo, Roberto Scaffaro

PII: S0925-8388(15)31928-9
DOI: 10.1016/j.jallcom.2015.12.137
Reference: JALCOM 36226

To appear in: Journal of Alloys and Compounds

Received Date: 23 November 2015


Revised Date: 16 December 2015
Accepted Date: 17 December 2015

Please cite this article as: A. Maio, S. Agnello, R. Khatibi, L. Botta, A. Alessi, A. Piazza, G. Buscarino,
A. Mezzi, G. Pantaleo, R. Scaffaro, A rapid and eco-friendly route to synthesize graphene-doped silica
nanohybrids, Journal of Alloys and Compounds (2016), doi: 10.1016/j.jallcom.2015.12.137.

This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to
our customers we are providing this early version of the manuscript. The manuscript will undergo
copyediting, typesetting, and review of the resulting proof before it is published in its final form. Please
note that during the production process errors may be discovered which could affect the content, and all
legal disclaimers that apply to the journal pertain.
ACCEPTED MANUSCRIPT
A RAPID AND ECO-FRIENDLY ROUTE TO SYNTHESIZE GRAPHENE-
DOPED SILICA NANOHYBRIDS.

Andrea Maio1, Simonpietro Agnello2, Reza Khatibi1, Luigi Botta1, Antonino Alessi2,3, Aurora
Piazza2,4,5, Gianpiero Buscarino2, Alessio Mezzi6, Giuseppe Pantaleo7, Roberto Scaffaro1*.

1 Department of Civil, Environmental, Aerospace, Materials Engineering, University of Palermo,


Viale delle Scienze, Ed. 6, 90128, Palermo (Italy)

PT
2 Department of Physics and Chemistry, University of Palermo, Via Archirafi 36, 90123 Palermo,
Italy
3 Laboratoire H. Curien, UMR CNRS 5516, Université de Lyon, 18 rue du Pr.Benoît Lauras 42000,

RI
Saint-Etienne, France
4 CNR-IMM, Strada VIII, 5, Zona Industriale, 95121, Catania, Italy
5 Department of Physics and Astronomy, Via Santa Sofia, 64, 95123 Catania, Italy

SC
6 ISMN, CNR, Palermo, CNR, Via Ugo La Malfa 153, 90146 Palermo, Italy
7 ISMN, CNR, Monterotondo Stazione, Roma, Italy

U
Abstract AN
In the present study, the possibility to synthesize graphene oxide (GO)-based nanohybrids with pure
and O2-doped silica nanoparticles by a rapid and easy hydrothermal process has been explored. The
nanohybrids were prepared by varying the type of silica nanoparticles (average diameter 7 nm or 40
M

nm) and the silica/GO weight ratio. All the materials were fully characterized by spectroscopic and
morphological techniques.
D

The experimental results revealed that it is possible to tune the characteristics of the obtained
TE

nanohybrids, such as morphology and amount of ester/ether linkages upon varying the preparation
parameters, together with the nanosilica’s typology and the silica to GO ratio. By Fischer
EP

esterification it was possible to achieve GO-silica nanohybrid lamellae to be then reduced into
nanostructured films by a hydrothermal process. These latter materials show a “lasagna-like”
C

structure in which it is possible to observe fully exfoliated (and partially reduced) GO lamellae
intercalated by silica nanoparticles agglomerates. The extension of silica layers, film morphology
AC

and structure, degree of functionalization, and thermal stability are strongly affected by the type of
silica. Furthermore, after the hydrothermal treatment, the nanohybrids were found to be insoluble in
water.

Keywords:

Graphene oxide, nanosilica, XPS, nanohybrids, Raman spectroscopy.

*
Corresponding author. E-mail: Roberto.scaffaro@unipa.it (Roberto Scaffaro)
ACCEPTED MANUSCRIPT
Introduction

The increasing interest in carbon based nanostructured materials like graphene [1-3], graphene
oxide [4-6] and carbon nanotubes is related to their potential applications in many fields of
materials science, medicine and environment [7-11].

Among the above materials, graphene, a flat monolayer of sp2-bonded carbon atoms, is very

PT
promising due to its unique characteristics such as high electronic conductivity, large specific
surface area, and high mechanical strength [1-3]. However, the poor dispersion of graphene in the

RI
most common solvents limits the preparation of nanohybrid materials, another task of particular
concern. A possible way to overcome this problem is the use of a precursor of graphene: the

SC
Graphene Oxide (GO). GO honeycomb lattice is composed of two equivalent sub-lattices of carbon
atoms bonded together with σ bonds and containing oxygen related moieties [12]. The presence of
both sp2-conjugated atoms and oxygen-containing functional groups provides to GO a strong

U
hydrophilicity and the possibility to further functionalize it with other molecules (i.e. π- π
AN
interactions, covalent attachment, etc.) and to reduce it during or after processing through chemical
or thermal reduction [13], evidencing interesting functionalization and tunable physical and
chemical properties [14]. Furthermore, since the GO is biocompatible and noncytotoxic, many
M

studies have been recently focused on the development of GO-based nanodevices for bioimaging,
DNA detection, drug delivery [17-21]. On the other side, due to their low cytotoxicity and large
D

specific surface, silica nanoparticles have been taken into account as promising material for
TE

biolabeling and drug loading/delivery [15]. In this context, silica nanoparticles opportunely
functionalized show interesting emission properties in the Near Infrared spectral range (about 1.3
EP

µm) arising from the presence of interstitial molecular oxygen [16]. This aspect is relevant for
labeling applications and also interesting in perspective, considering the paramagnetic character of
O2. Particular consideration has recently been demonstrated for GO-silica composites because of
C

the potentialities for electrical applications, their chemical inertia and stability toward ions exposure
AC

[17-20]. The possibility to combine the extraordinary properties of GO and silica offers several
advantages for the realization of nanoprobes for biological applications and for biosensor too [21].
Many studies tackled this task but principally by using full chemical procedures involving silica
preparation through sol-gel technique and GO by synthesis methods. This route has the limitation to
hinder the functionalization of the silica component because of the need to grow the raw material
itself.

Aim of the present work is the rapid and easy preparation of GO-silica nanocomposites starting
from commercial silica nanoparticles, also opportunely functionalized, and laboratory synthetized
ACCEPTED MANUSCRIPT
GO, a tactic that could find application in diverse fields and could enable to obtain composites with
functionalized silica.

The strategy for the fabrication of GO-nanosilica sandwich films, illustrated in Figure 1, can be
schematized as follows:

(i) synthesis of GO by oxidizing graphite powder through a modified Hummer method.

PT
(ii) preparation of O2 loaded silica nanoparticles of average diameter 40 or 7 nm [22]
(iii) Covalent immobilization of O2 loaded silica nanoparticles onto GO lamellae via rapid

RI
esterification without using silica precursors, such as TEOS, etc.
(iv) hydrothermal stratification at 120 °C to produce water-resistant composites with

SC
enhanced thermal stability.

+ O2 (ii) (i)

U
SiO2 AN Graphene Oxide (GO)

Fischer
T, P esterification
Agnello S et al
Patent WO2011128855-A1; IT1399551-B
H3PO4
KMnO4
AEOX50 AE300 H2SO4
GO-doped silica nanohybrids
M

Specific area 50+15 300+30 (iii)


Graphite
Size nm 40 7 Hydrothermal
(iv) treatment
D

O2 loading 5 1018 1 1018


TE
EP

GO-doped silica nanostructured film


C

Figure 1. Schematics of the route for the preparation of GO-silica bulky films.
AC

The process is ecofriendly and rapid, as the steps (iii) and (iv) require less than 1 hour.

To the best of our knowledge, the easiest method to prepare GO-silica nanohybrids via direct
esterification in water requires 24 h under high pressure [23], the method here proposed allows
instead to obtain nanohybrids in less than 1 hour.
ACCEPTED MANUSCRIPT
2. Experimental Details

2.1 Materials

Neat graphite (grade Ma 399, 45 µm) was purchased by NGS Naturgraphit (Germany). Two
samples of fumed silica nanoparticles, namely AEOX50 and AE300 (Aerosil OX50 and Aerosil
300), which properties are summarized in Table 1 [24], were provided by Evonik Industries

PT
(Germany). Finally, H3PO4, H2SO4, KMnO4, HCl, H2O2, C2H5OH and diethyl-ether were purchased
by Sigma Aldrich. All the reactants were used as received without any further purification.

RI
Table 1 Main properties of the samples of silica nanoparticles used in the frame of this work [24]

SC
Specific surface Particle diameter O2 content after
Name Nickname loading
[m2/g] [nm]
[molecules/cm3]
AerosilOX50 AEOX50 50+15 40 (5 ± 1 ) x 1018
Aerosil300 AE300
U
300+30 7 (1.0 ± 0.2) x 1018
AN
2.2 Synthesis of GO
M

The GO was synthesized via chemical oxidation and exfoliation of natural graphite, according to
D

our previous work [7], by slightly modifying the method illustrated by Marcano et al. [25]. Several
parameters, such as reaction time, solid reactant to liquid solvent ratio (S/L) and stirring speed (i.e.
TE

melt viscosity) and concentration of HCl during the purification step, were changed in order to
maximize the content in –OH and –COOH. Finally, the sample prepared according to the operative
EP

set listed in Table 2 was selected for the silica-functionalization. More in details for GO
preparation, a solid mixture of neat graphite flakes (5 g, 1 wt equiv) and KMnO4 (30 g, 6 wt equiv)
C

was added to a 9:1 mixture of concentrated H2SO4/H3PO4, which volume was 600 mL (540:60 mL)
AC

and stirred at 64 rpm. The reaction was heated to 50 °C and the temperature was kept constant in
order to avoid the formation of explosive manganese compounds. The reaction time was 18 h. The
reaction was then cooled to room temperature and quenched with ice (600 mL) and 30% H2O2 (4.5
mL). The viscous solution achieved was diluted by adding distilled water (100 mL) and sonicated
for 30 minutes. The workup consisted of repeated washing and centrifuging steps. For each step, the
filtrate was centrifuged (12000 rpm for 1 h) and the remaining solid was then washed in succession
with 400 mL of water, 400 mL of 10% HCl, and 200 mL of ethanol (2×), and the supernatant was
decanted away. The solid obtained on the filter was vacuum-dried overnight at room temperature,
obtaining 10 g of product.
ACCEPTED MANUSCRIPT
Table 2. Main properties of the GO sample used in the frame of this work, and the starting graphite.

[-COOH] [-OH] Thickness Spacing Z pot


Sample C/O % sp2
at % at % nm nm mV

GO 1.1 12 10.6 33.3 0,8 0.884 -45


Graphite 24 75,5 - - n.c. 0.132 n.c.

PT
2.3 Preparation of GO-silica nanohybrids

RI
For the preparation of GO-silica nanohybrids with AE300 (GO-AE300) and AEOX50 (GO-
AEOX50), 20 mg of GO and 20 mg of silica nanoparticles, previously dispersed in water (100 mL)

SC
with 2% HCOOH and sonicated at 50 °C for 1 hr, were transferred to a Teflon-coated crystallizer,
where the dispersion was magnetic-stirred at 120 °C. Of course, as long as the solvent was present,

U
the temperature remained constant at 100 °C, after about 25 minutes the total evaporation of water
AN
occurred and the gel-like slurry was maintained at 120 °C for 35 minutes. Finally, a brown film was
peeled off. This temperature was chosen basing on preliminary tests in which the esterification
reactions were monitored as a function of the temperature with the other parameters fixed. The
M

nanohybrids obtained are summarized in Table 3. A different GO/nanosilica weight ratio of 10:1
was investigated for comparison utilizing both nanosilica types AE300 and AEOX50, but fixing all
D

the other process steps. Other preparations of nanohybrids at different temperature of 30°C and
TE

100°C were considered also to evaluate the effects of processing temperature in the final
nanocomposite material.
EP

Table 3. The four experimental conditions adopted for the preparation of GO-silica nanohybrids

Experimental runs Typology of silica GO/silica ratio


C

A AE300 1:1
B AEOX50 1:1
AC

C AE300 10:1
D AEOX50 10:1

2.4 Characterization

The samples were fully characterized by spectroscopy analyses (FT-IR, XPS, Raman), X-ray
Diffraction (XRD) and morphological analysis Atomic Force Microscopy (AFM) and Scanning
Electron Microscopy (SEM). Immersion tests were carried out to the final nanohybrids as well.
ACCEPTED MANUSCRIPT
More in details, the FT-IR/ATR analysis was carried out by using a Perkin-Elmer FT-IR/NIR
Spectrum 400 spectrophotometer. The spectra were recorded in the range 4000-400 cm-1.

For the Micro-Raman analysis a Bruker SENTERRA instrument, with diode laser excitation at 532
nm and spectral resolution 9-15 cm-1, has been used. The laser power was fixed at 0.2 mW to avoid
sample modifications; furthermore, measurements in three different sample positions have been
repeated for each treatment to check the homogeneity of the sample. Analysis of the spectra has

PT
been done after subtraction of a linear baseline in the range 1200-1800 cm-1, for the D and G bands
[26], and another linear baseline in the range 2500-3500 cm-1, for the 2D region [26]. The spectra

RI
have been successively normalized to the G band and to the central feature of the 2D region at
~2960 cm-1, respectively.

SC
The XPS experiments have been performed by using an ESCALAB MkII (VG Scientific)
spectrometer, equipped with a standard Al Kα excitation source and a 5-channeltron detection

U
system for spectroscopic analysis. The spectra were collected at 20 eV constant pass energy, and the
AN
analyzed area of the sample was about 5 mm in diameter. Spectroscopic data have been processed
by the Avantage v.5 software. More details are reported elsewhere [27].
The morphology of nanohybrids was observed by scanning electron microscopy (ESEM FEI
M

QUANTA 200). The samples were cryo-fractured in liquid nitrogen to investigate the morphology
of both longitudinal and cross section.
D
TE

Tapping mode amplitude modulation AFM measurements were performed by a Multimode V


(Veeco Metrology) scanning probe microscope. The instrument was equipped with a conventional
piezoscanner (maximum xy range ≈ 14 µm, and maximum z range ≈ 3.6 µm) and a four-segment
EP

photodetector for cantilever deflection monitoring. PointProbe®Plus Silicon-SPM-probes were


used with Al backside reflex coating, resonance frequency ≈300 KHz and tip apical diameter ≈10
C

nm. All the scans were executed at room temperature and in N2 atmosphere. The inert gas ambient,
AC

obtained by continuous purging the sample chamber with N2 gas (flux≈2 l/min), was necessary to
reduce the probability of adhesion of the nanocomposite to the tip, by efficiently reducing the
capillary forces acting between them.

The microstructure was investigated by XRD using an Empyrean PANalytical II diffractometer


with a Cu Kα radiation source (λ = 1.5406 Å).

Thermogravimetric analysis (TGA) was carried out by means of a Mettler Toledo TGA/DSC
instrument, all the experimental runs were performed under N2 flow (100 ml/min) and two different
ACCEPTED MANUSCRIPT
temperature ramps: a slow ramp of 1°C/min from room temperature up to 800 °C and a faster ramp
of 10°C/min within the range 800-1400 °C.

3. Results and discussion

GO was functionalized and partially reduced by fumed silica via Fischer esterification and
etherification reactions in water. This route is based on the reactivity of functional groups of GO,
such as hydroxyl, phenols and carboxylic acid with silanol (Si-OH) and siloxane (Si-O-Si) groups

PT
on silica nanoparticles surface to form Si-O-C and Si-COO-C, as illustrated in Figure 2.

RI
U SC
AN
Figure 2. Pictorial representation of a possible mechanism of esterification and etherification
reactions giving origin to stable GO-silica nanocomposite.
M

The possible reactions in aqueous environment are several, according to previous studies [23].
Silica nanoparticles surface could react with H2O to enrich itself with silanol and this latter can
D

react either with carboxyl or alcoholic moieties of the GO via esterification and etherification,
respectively as reported in the following reaction equations (1-3):
TE

SiO2 + 2H2O Si(OH)4 (1)


EP

Si-OH + HOOC-G G-COO-Si + H2O (2)

Si-OH + HO-G Si-O-G+ H2O (3)


C

where, in eq.(2) and (3), G stays for graphene oxide network and Si stays for the silica one. As the
pH of GO and silica in water is quite acidic, it is presumable that even the epoxy moieties,
AC

abundant in the GO, are extremely reactive and could turn into glycols.

To investigate the thermal processing effects, FTIR and Raman measurements were carried out on
GO-silica nanocomposites prepared with hydrothermal casting at different temperatures of 30°,
100° and 120°C.
The results of GO-AE300 nanocomposite preparation routes are reported in Figure 3 and Figure 4.
Of particular interest are the spectral ranges 1400-700 cm-1 and 2000-1400 cm-1 since they are
related to C-O and C=C as well as to Si-O-Si and Si-O-C vibrations [28].
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 3. Deconvoluted spectral range (1900-1500 cm-1) in the FTIR spectra of GO-AE300 (run C)
nanocomposites prepared at 30 °C (a), 100 °C (b), 120 °C (c).
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M

Figure 4. Deconvoluted spectral range (1400-700 cm-1) in the FTIR/ATR spectra of GO-AE300 (run C)
nanocomposites prepared at 30 °C, 100 °C and 120 °C.
D

In details, in literature the features at about 1100, 1050, 1180, 1220 cm-1 were assigned to Si-O-Si,
TE

C-O-Si, C-O and C-O-C but the main contributes are indistinguishable each other [28-29].

The temperature was selected by studying the system GO-AE300 (10:1) (run C). In order to detect
EP

the presence of ethers and/or ester after the functionalization, the spectra related to GO-AE300
nanocomposites prepared at three different temperatures were qualitatively fitted by means of
C

Lorentzian or Gaussian functions within the spectral range 2000-1350 cm-1 and 1400-700 cm-1,
respectively. The fitted 2000-1350 cm-1 region of the spectra recorded for GO-AE300 (run C)
AC

prepared at 30 °C, 100 °C and 120 °C are reported in Figure 3 a-c. This spectral range is
characterized by four bands: one centered at about 1625-1630 cm-1, assigned to sp2 C=C bond of
GO, one located at about 1730 cm-1, assigned to C=O bond of carboxyl, one band located around
1800 cm-1 presumably attributable to ketones and/or lactones and the broad band centered at about
1410 cm-1 which is assigned to alcoholic moieties. The nanohybrids prepared at 30 °C (a) and 100
°C (b) retain essentially the same peak locations, although the bands are clearly different in
intensity and width. In the sample treated at 120 °C, instead, the carbonyl band exhibits a
remarkable shift from 1731 to 1743 cm-1, thus indicating the conversion of carboxyls into ester
ACCEPTED MANUSCRIPT
bonds. Furthermore, it is possible to identify at least one more band, located at about 1450 cm-1 and
a band centered at about 1388 cm-1. Likely, this latter feature could be assigned to the –C=O-O-Si
silicic ester, as asserted by other authors [23].

Similarly, in the 1400-700 cm-1 spectral range, some changes are detected when the process
temperature increases. More in particular, as shown in Fig. 4, at 30 °C it is possible to detect
different peaks, centered at about 900, 980, 1070, 1100 and 1225 cm-1 that could be representative

PT
of C-O-C bending of epoxy, Si-OH, alcoxy vibration stretching, asymmetric Si-O-Si stretching, v
(C-O-C) of epoxy and/or esters. These assignments are not unique due to the overlap of GO and

RI
silica spectral features [30-32]. At 100 °C substantially the same features are still detectable but the
reaction of epoxy moieties is confirmed by the disappearance of the band at 900 cm-1. When the

SC
reaction temperature is set to 120 °C, the spectrum exhibits clear differences, as new features appear
(see red-filled area under the curves labeled as ω1, ω2 and ω3). Those centered at about 1170 and

U
1300 cm-1 could be assigned to Si-O-C and Si-O-OC bonds, whereas the feature located at 720 cm-1
could be ascribed to C=C-H, as a consequence of a partial recovery of aromatic domains of GO or
AN
to the formation of five-membered-ring lactols due to thermal treatment. Finally, it is worth noting
(see blue-filled area under curves in fig. 4) that all samples show peaks centered at 1220 and 850-
M

900 cm-1, respectively assigned to esters moieties and epoxy groups. By analyzing their relative
intensity ratio, it is possible to detect that at 120 °C the main contribute is due to ester groups.
D

Nevertheless, even if this region is highly difficult to analyze, the changes observed in the spectrum
TE

certify that the chemical structure of the nanocomposites prepared at 120 °C underwent some
modification with respect to those prepared at 100 °C and 30 °C.
EP

To further investigate the dependence of the nanocomposites properties on the preparation


temperature, the Raman spectra have been registered.
C

Figure 5a reports micro-Raman plot in the range 1200-1800 cm-1, referred to the first order (one
AC

phonon) Raman region. The D-band, located at the K point in the Brillouin zone, is centered at 1350
cm-1 and is ascribed to the sp3 diamond-like carbons, whereas the G band (at point Γ), usually
located at around 1600 cm-1, refers to the sp2 hybridized carbons.

Figure 5b reports the spectral range within 2500-3500 cm-1 (second order Raman scattering) where
the main features are the 2D mode (2690 cm-1), that is a second-order overtone of a different in-
plane vibration, and the bands resulting by a combination of scattering peaks, i.e. D+G, centered at
2940 cm-1, and 2D’ (or D+D’), located at 3150 cm-1.
ACCEPTED MANUSCRIPT
As one can see, for the GO-AE300 10:1 (run C) small differences are found depending on the
process temperature up to 100 °C of preparation. At variance, a comparison with the GO-AE300 1:1
(run A) prepared at 120°C shows that both the D, G bands and the 2D spectral features are
modified. In particular, the preparation at 120°C induces a shrinking of both D and G bands.
Furthermore, at the 120°C preparation temperature, the 2D component at 2680 cm-1 clearly
increases and the 2D’ component at about 3150 cm-1 decreases with respect to the amplitude of the

PT
D+G component at 2950 cm-1 [26, 33]. The observed features are in good agreement with the
FTIR/ATR data suggesting that the 120°C treatment induces modifications of the GO, as previously

RI
guessed in a dedicated study [20], and will be further investigated.

U SC
AN
M
D

Figure 5. Raman spectra of D, G bands and the 2D feature of GO and nanosilica composites as a
TE

function of processing temperature for GO-AE300.

The GO nanocomposites with different silica nanoparticles are compared in Figure 6. This
EP

comparison includes also nanocomposites with silica nanoparticles previously functionalized with
interstitial O2. It can be observed that the employed preparation at 120 °C induces spectral
modifications both in the D and G bands and also in the 2D region. Furthermore, the smaller
C

nanoparticles (AE300) are responsible for the larger changes with respect to the GO starting batch
AC

(see further the XPS results). It is also worth to note that the functionalization of silica nanoparticles
with O2, compare GO-AEOX50 with GO-AEOX50 TTO2, does not affect the final Raman spectra
suggesting that essentially the same structural properties of the nanocomposite are achieved.
ACCEPTED MANUSCRIPT

PT
RI
Figure 6. Raman spectra of the raw graphene oxide and of the different nanocomposites including

SC
that with nanosilica functionalized with interstitial oxygen molecules; D, G bands (a) and 2D region
(b).

U
In Figure 7 the comparison of samples prepared at different process temperatures is shown. It is
AN
observed that the nanocomposites at 80°C processing are not modified with respect to the starting
GO batch, showing that functionalization of nanoparticles can be maintained since the processing
temperature is low enough to avoid outgassing of entrapped O2 [34].
M
D
TE
C EP
AC

Figure 7. Raman spectra of the raw graphene oxide and of the AEOX50 nanocomposites with
interstitial oxygen molecules at two processing temperatures; D, G bands (a) and 2D region (b).

XPS analysis was carried out to achieve further information about the covalent immobilization of
silica onto GO. Figure 8 shows C1s, O1s and Si2p spectra of GO-AE300. The obtained results
revealed that, after the combination of silica NPs with GO, the presence of C, O and Si are present
can be detected on both GO-AE300 and GO-AEOX50. Moreover, the esterification reaction
ACCEPTED MANUSCRIPT
occurred preferentially on the GO-AE300 samples, as indicated by the chemical shift registered in
the Si2p signal and assigned to Si – O – C bond (binding energy BE = 102.3 eV) [35].

PT
RI
U SC
AN
M
D

Figure 8. C 1s, O 1s and Si 2p spectra of GO-AE300 nanohybrids (a-c)


TE

The C 1s signal of GO-AE300 was also characterized by the presence of a component localized at
BE = 289.9 eV, assigned to the carbonate group, due to the residual of the esterification reaction.
EP

Finally, combing the results of the peak fitting analysis of C1s and the relative increase of the D
Raman peak amplitude with respect to that of the GO (compare fig. 6-7), it is possible to assert that
C

in GO-AE300 sample a partial reduction from GO to rGO occurred.


AC

The GO-AE300 and GO-AEOX50 1:1 nanohybrids were analyzed by AFM images, as shown in
Figure 9. In these samples the nanosilica/GO ratio is the largest investigated and it can be
considered representative of the maximum expected effect. It is clear that the mechanism of the
formation of hybrid structures is deeply different. In fact, in the case of AE300, by comparing
different sample zones, as shown in fig.9a and fig.9b, the silica immobilization is discontinuous and
extended regions of naked GO appear. At variance, for the AEOX50 (fig.9c) a remarkable tendency
to agglomerate and to build up a sort of bulky paper covering the GO lamellae is predominant and
uniform in different regions of the sample.
ACCEPTED MANUSCRIPT

PT
RI
Figure 9. AFM micrographs of 1:1 GO-AE300 (a, b) and 1:1 GO-AEOX50 (c) nanocomposites,
evidencing the presence (a, c) and absence (b) of silica nanoparticles (scale bar=1 µm).

SC
This feature, after the hydrothermal treatment, results into two different conformations, showed by

U
SEM micrographs reported in Fig.10 a-f. Even if both the types of silica led to the formation of a
lasagna-like structure with silica nanoparticles intercalated between GO lamellae (see panels a-d), it
AN
appears clear that AEOX50 has a more bulky uniform distribution onto the planes of GO with
respect to AE300 (see panels e-f).
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
M
D
TE
C EP
AC

Figure 10. SEM micrographs of the cross sections at different magnitude of 1:1 GO-AE300 (a-b),
1:1 GO-AEOX50 (c-d) and surface area of 1:1 GO-AE300 (e) and 1:1 GO-AEOX50 (f)

XRD patterns of nanohybrids, reported in Figure 11, put into evidence that the interparticle spacing
of both AE300 and AEOX50 decrease after hydrothermal casting, thus confirming the densification
ACCEPTED MANUSCRIPT
and stratification phenomena occurring during this treatment, whereas GO peak disappears
(nanohybrid A) or shifts to lower 2θ angles, thus suggesting that after the functionalization the GO
could remain expanded by intercalation of silica nanoparticles between lamellae or fully exfoliated
when GO/silica ratio is 1.

PT
RI
U SC
AN
M
D
TE

Figure 11. XRD patterns of nanohybrids (and starting materials as control)


EP

After the hydrothermal treatment, the nanohybrids become totally insoluble in water, as evidenced
C

by the immersion tests carried out onto A and B nanohybrids. Fig.12 depicts A and B nanohybrids
after 200 days of immersion in a water bath. The XPS and AFM analysis (here not reported)
AC

performed onto the samples showed no change in morphology and chemical composition, whereas
the FTIR analysis performed onto the water used for the tests did not detect any presence of silica
nor GO proving the strong stability of the nanohybrids.
ACCEPTED MANUSCRIPT

PT
RI
U SC
AN
Figure 12. Immersion tests of GO-AE300 and GO-AEOX50 nanohybrid films after 200 days.
M

Figure 13.a reports TGA results of starting materials and nanohybrids. As one can see, both samples
of silica are stable within the range investigated, and GO shows the characteristic weight losses
D

between 150 and 300 °C, ascribed to CO, CO2, and steam release from the most labile functional
TE

groups. The nanohybrids showed an enhanced thermal stability if compared with GO and this
feature makes these nanohybrids particularly suitable for high temperature applications [37]. The
presence of silica did not have much effect on the initial decomposition temperature but increased
EP

the char residues. In Figure 13.b the final char residue is plotted as a function of silica weight
fraction and compared to the trend expected by linearly combining the char residue of GO (12%)
C

and silica (99.4%) at 1400 °C. In our opinion, silica nanoparticles (NP) has two antagonistic effects
AC

on the thermal stability of the nanohybrids:

(i) the formation of a silica layer covering the surface of lamellae avoids the thermal
decomposition of GO;
(ii) O2 molecules entrapped in the NP could be released at high temperatures, thus
promoting the further degradation of graphene oxide.

With respect to AE300, AEOX50 induces systematically a weaker thermal stability. This can be
explained by the larger tendency to stratify into gel layer, (i), but conversely the amount of oxygen
molecules could also be relevant, (ii), since that of AEOX50 is typically five-fold that of AE300
ACCEPTED MANUSCRIPT
[36]. At a lower content (i.e. 0.09), the presence of silica provides better thermal stability (compare
A,B nanohybrids with C,D in fig. 13.a), since the benefits produced by the flame-retardant
mechanism of silica prevail on the degradative phenomena caused by the presence of O2 molecules.
The samples containing AE300 showed less mass loss than those containing AEOX50 and this
result could be likely explained by considering the different O2 content. When silica weight fraction
is 0.5, the amount of oxygen molecules dramatically increases, and (ii) counterbalances or nullifies

PT
(i), depending on the O2-doping level of the nanoparticles.

RI
U SC
AN
M
D
TE
C EP
AC

Figure 13. TGA traces (a), residual char percentage plotted as a function of type and weight
fraction of silica (b), DTG traces (c).

Since no thermal degradation was observed in the silica nanoparticles, we could gather that the
oxygen-promoted degradation involves only GO. Indeed, by analyzing the first derivative of mass
loss (DTG), reported in Fig.13.c, several peaks are detectable above 500 °C, and this issue is more
evident in the samples containing a higher content of GO.
ACCEPTED MANUSCRIPT
4. Conclusion and perspectives

In conclusion, a new and easy method to prepare graphene-silica nanocomposites is presented.


During the reactions in water, the GO is converted to rGO and the silica, also containing
paramagnetic oxygen molecules, can be anchored onto the edges of GO lamellae. When the
nanosilica particles and GO are allowed to stay at 120 °C for less than one hour, they stratify into a
bulky paper, consisting of lasagna-like structured layers of reduced GO and doped silica. The

PT
possibility to achieve nanohybrid films by low temperature preparation routes has been shown and
could extend the application fields of these materials as, for example, by the presence of trapped

RI
oxygen molecules inside the silica nanoparticles for its successive release or probing. The
nanohybrids films achieved feature a larger thermal resistance than pure GO and high water

SC
stability. These aspects are particularly promising for a large variety of applications, ranging from
(opto)electronics to energy, while the direct combination with nanoparticles is potentially suitable

U
for the development of nanoprobes and biosensors. Studies on further properties of these materials
are currently in progress.
AN
M

Acknowledgments

Partial financial support by the FAE-PO FESR SICILIA 2007/2013 4.1.1.1, by HIPPOCRATES -
D

PON02_00355 and by the FFR 2012/2013 project of the University of Palermo is acknowledged.
TE

References
EP

[1] K.S. Novoselov, D. Jiang, F. Schedin, T.J. Booth, V.V. Khotkevich, S.V. Morozov, et al.
Two-dimensional atomic crystals. Proc Natl Acad Sci U S A, 102 (2005) 10451–3.
C

doi:10.1073/pnas.0502848102.
AC

[2] S. Wang, R. Wang, X. Liu, X. Wang, D. Zhang, Y. Guo, et al., Optical Spectroscopy
Investigation of the Structural and Electrical Evolution of Controllably Oxidized Graphene
by a Solution Method, J. Phys. Chem. C. 116 (2012) 10702–10707. doi:10.1021/jp212184n.

[3] S. Stankovich, D.A. Dikin, G.H.B. Dommett, K.M. Kohlhaas, E.J. Zimney, E.A. Stach, et al.,
Graphene-based composite materials., Nature. 442 (2006) 282–6. doi:10.1038/nature04969.

[4] D. Chen, H. Feng, J. Li. Graphene oxide: preparation, functionalization, and electrochemical
applications. Chem Rev, 112 (2012) 6027–53. doi:10.1021/cr300115g.

[5] D.A. Dikin, S. Stankovich, E.J. Zimney, R.D. Piner, G.H.B. Dommett, G. Evmenenko, et al.
Preparation and characterization of graphene oxide paper. Nature, 448 (2007) 457–60.
doi:10.1038/nature06016.
ACCEPTED MANUSCRIPT
[6] F. Perrozzi, S. Prezioso, M. Donarelli, F. Bisti, P. De Marco, S. Santucci, et al., Use of
Optical Contrast To Estimate the Degree of Reduction of Graphene Oxide, J. Phys. Chem. C.
117 (2013) 620–625. doi:10.1021/jp3069738.

[7] A. Maio, R. Fucarino, R. Khatibi, S. Rosselli, M. Bruno, R. Scaffaro, A novel approach to


prevent graphene oxide re-aggregation during the melt compounding with polymers. Compos
Sci Technol. 119 (2015) 131–7. doi: 10.1016/j.compscitech.2015.10.006.

[8] A. Maio, L. Botta, A.C. Tito, L. Pellegrino, M. Daghetta, R. Scaffaro, Statistical Study of the

PT
Influence of CNTs Purification and Plasma Functionalization on the Properties of
Polycarbonate-CNTs Nanocomposites, Plasma Process. Polym. 11 (2014) 664–677.
doi:10.1002/ppap.201400008.

RI
[9] R. Scaffaro, A. Maio, A.C. Tito, High performance PA6/CNTs nanohybrid fibers prepared in
the melt, Compos. Sci. Technol. 72 (2012) 1918–1923.

SC
doi:10.1016/j.compscitech.2012.08.010.

[10] R. Scaffaro, A. Maio, Enhancing the mechanical performance of polymer based


nanocomposites by plasma-modification of nanoparticles, Polym. Test. 31 (2012) 889–894.

U
doi:10.1016/j.polymertesting.2012.06.006.
AN
[11] R. Scaffaro, A. Maio, S. Agnello, A. Glisenti, Plasma Functionalization of Multiwalled
Carbon Nanotubes and Their Use in the Preparation of Nylon 6-Based Nanohybrids, Plasma
Process. Polym. 9 (2012) 503–512. doi:10.1002/ppap.201100140.
M

[12] Y. Zhu, S. Murali, W. Cai, X. Li, J.W. Suk, J.R. Potts, et al., Graphene and graphene oxide:
synthesis, properties, and applications., Adv. Mater. 22 (2010) 3906–24.
D

doi:10.1002/adma.201001068.
TE

[13] V. Georgakilas, M. Otyepka, A.B. Bourlinos, V. Chandra, N. Kim, K.C. Kemp, et al.,
Functionalization of graphene: covalent and non-covalent approaches, derivatives and
applications., Chem. Rev. 112 (2012) 6156–214. doi:10.1021/cr3000412.
EP

[14] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of graphene oxide., Chem.
Soc. Rev. 39 (2010) 228–40. doi:10.1039/b917103g.
C

[15] A. Ambrosone, M.R. Scotto di Vettimo, M.A. Malvindi, R. Modi, O. Levy, M. Valentina,
P.P. Pompa, C. Tortiglione, A. Tino, Impact of Amorphous SiO2 Nanoparticles on a Living
AC

Organism: Morphological, Behavioral, and Molecular Biology Implications, Front. Bioeng.


Biotechnol. 2 (2014) 1–12. doi: 10.3389/fbioe.2014.00037.

[16] S. Agnello, M. Cannas, L. Vaccaro, G. Vaccaro, F.M. Gelardi, M. Leone, et al., Near-
Infrared Emission of O 2 Embedded in Amorphous SiO 2 Nanoparticles, J. Phys. Chem. C.
115 (2011) 12831–12835. doi:10.1021/jp2035554.

[17] S. Watcharotone, D.A. Dikin, S. Stankovich, R. Piner, I. Jung, G.H.B. Dommett, et al.,
Graphene-silica composite thin films as transparent conductors., Nano Lett. 7 (2007) 1888–
92. doi:10.1021/nl070477+.

[18] W.L. Zhang, H.J. Choi, Silica-graphene oxide hybrid composite particles and their
electroresponsive characteristics., Langmuir. 28 (2012) 7055–62. doi:10.1021/la3009283.
ACCEPTED MANUSCRIPT
[19] L. Kou, C. Gao, Making silica nanoparticle-covered graphene oxide nanohybrids as general
building blocks for large-area superhydrophilic coatings., Nanoscale. 3 (2011) 519–28.
doi:10.1039/c0nr00609b.

[20] S. Agnello, A. Piazza, A. Alessi, A. Maio, R. Scaffaro, G. Buscarino, et al., Graphene oxide
and Fumed silica graphene oxide nanocomposites modification by thermal treatments, in:
NANOCON 2013 - Conf. Proceedings, 5th Int. Conf., TANGER Ltd., 2013: pp. 56–61.
http://www.scopus.com/inward/record.url?eid=2-s2.0-84929398033&partnerID=tZOtx3y1.

PT
[21] S. Abraham, V. Ciobota, S. Srivastava, S.K. Srivastava, R.K. Singh, J. Dellith, et al.,
Mesoporous silica particle embedded functional graphene oxide as an efficient platform for
urea biosensing, Anal. Methods. 6 (2014) 6711. doi:10.1039/C4AY01303D.

RI
[22] S. Agnello, R. Boscaino, M. Cannas, F.M. Gelardi, M. Leone, V. Militello, Silica-Based NIR
Nano-Emitters for Applications in Vivo and Process for Production Thereof, patent

SC
WO2011128855-A1; IT1399551-B, 2011.

[23] J.Q. Dalagan, E.P. Enriquez, One-step synthesis of mesoporous silica-graphene composites
by simultaneous hydrothermal coupling and reduction of graphene oxide. Bull Mater Sci. 37

U
(2014) 589–95.
AN
[24] J.Q. Dalagan, E.P. Enriquez, L.-J. Li, Simultaneous functionalization and reduction of
graphene oxide with diatom silica, J. Mater. Sci. 48 (2013) 3415–3421. doi:10.1007/s10853-
012-7128-1.
M

[25] S. Agnello, D. Di Francesca, A. Alessi, G. Iovino, M. Cannas, S. Girard, et al., Interstitial O2


distribution in amorphous SiO2 nanoparticles determined by Raman and photoluminescence
D

spectroscopy, J. Appl. Phys. 114 (2013) 104305. doi:10.1063/1.4820940.


TE

[26] D.C. Marcano, D. V Kosynkin, J.M. Berlin, A. Sinitskii, Z. Sun, A. Slesarev, et al., Improved
synthesis of graphene oxide., ACS Nano. 4 (2010) 4806–14. doi:10.1021/nn1006368.

[27] S. Kaciulis, A. Mezzi, P. Calvani, D.M. Trucchi, Electron spectroscopy of the main
EP

allotropes of carbon, Surf. Interface Anal. 46 (2014) 966–969. doi:10.1002/sia.5382.

[28] T. Oh, Correlation between potential barrier and FTIR spectra in SiOC film with the C-O
bond of sp3 structure, Bull. Korean Chem. Soc. 30 (2009) 467–470.
C

http://www.scopus.com/inward/record.url?eid=2-s2.0-84894897078&partnerID=tZOtx3y1.
AC

[29] A.A.R. de Oliveira, V.S. Gomide, M. de Fátima Leite, H.S. Mansur, M. de Magalhäes
Pereira, Effect of polyvinyl alcohol content and after synthesis neutralization on structure,
mechanical properties and cytotoxicity of sol-gel derived hybrid foams, Mater. Res. 12
(2009) 239–244. http://www.scopus.com/inward/record.url?eid=2-s2.0-
68649116302&partnerID=tZOtx3y1

[30] E.F. Vansant, P. Van der Voort, K.C. Vrancken, Characterization and Chemical Modification
of the Silica Surface, Studies in Surface Science and Catalysis, first ed., Elsevier,
Amsterdam, 1995

[31] F.L. Galeener, Band limits and the vibrational spectra of tetrahedral glasses, Phys. Rev. B. 19
(1979) 4292–4297. doi:10.1103/PhysRevB.19.4292.
ACCEPTED MANUSCRIPT
[32] R. Scaffaro, L. Botta, G. Lo Re, R. Bertani, R. Milani, A. Sassi, Surface modification of
poly(ethylene-co-acrylic acid) with amino-functionalized silica nanoparticles, J. Mater.
Chem. 21 (2011) 3849. doi:10.1039/c0jm03310c.

[33] X. Díez-Betriu, S. Álvarez-García, C. Botas, P. Álvarez, J. Sánchez-Marcos, C. Prieto, et al.,


Raman spectroscopy for the study of reduction mechanisms and optimization of conductivity
in graphene oxide thin films, J. Mater. Chem. C. 1 (2013) 6905. doi:10.1039/c3tc31124d.

[34] G. Iovino, S. Agnello, F.M. Gelardi, R. Boscaino, O 2 Diffusion in Amorphous SiO 2

PT
Nanoparticles Probed by Outgassing, J. Phys. Chem. C. 116 (2012) 11351–11356.
doi:10.1021/jp3006734.

RI
[35] Y.L. Khung, S.H. Ngalim, L. Meda, D. Narducci, Preferential formation of Si-O-C over Si-C
linkage upon thermal grafting on hydrogen-terminated silicon (111)., Chemistry. 20 (2014)
15151–8. doi:10.1002/chem.201403014.

SC
[36] A. Alessi, G. Iovino, G. Buscarino, S. Agnello, F.M. Gelardi, Entrapping of O 2 Molecules
in Nanostructured Silica Probed by Photoluminescence, J. Phys. Chem. C. 117 (2013) 2616–
2622. doi:10.1021/jp310314t.

[37]
U
A. Maio, R. Fucarino, R. Khatibi, L. Botta, S. Rosselli, M. Bruno, et al., Graphene oxide-
AN
silica nanohybrids as fillers for PA6 based nanocomposites, in: Times of polymers (ToP) and
composites 2014 – Proceedings, 7th Int. Conf. on Times of Polymers (TOP) and Composites,
vol. 1599, AIP Publishing, 2014, May, pp. 438-441.
M
D
TE
C EP
AC
ACCEPTED MANUSCRIPT

Highlights

• A fast and simple route to achieve graphene-silica nanohybrids is presented


• It does not require any toxic solvent or long time consuming treatment.
• It allows selectively doping the silica prior to reaction with graphene.

PT
O2 molecules were entrapped in the silica as non-toxic luminescence agents.
• Nanostructured films provide strong thermal stability and insolubility in water

RI
U SC
AN
M
D
TE
C EP
AC

You might also like