You are on page 1of 9

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

Nanoscale Dynamic Article Links < C

Cite this: Nanoscale, 2012, 4, 2937


www.rsc.org/nanoscale PAPER
Graphene oxide versus functionalized carbon nanotubes as a reinforcing agent
in a PMMA/HA bone cement
Gil Gonçalves,a Sandra M. A. Cruz,a A. Ramalho,b Jose Gracioa and Paula A. A. P. Marques*a
Received 8th February 2012, Accepted 23rd February 2012
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

DOI: 10.1039/c2nr30303e

Graphene oxide (GO) and functionalized carbon nanotubes (f-CNTs) (each in the concentration range
of 0.01–1.00 wt/wt%) were investigated as the reinforcing agent in a poly(methyl methacrylate)
(PMMA)/hydroxyapatite (HA) bone cement. Mixed results were obtained for the changes in the
Downloaded by University of Sussex on 14 January 2013

mechanical properties determined (storage modulus, bending strength, and elastic modulus) for the
reinforced cement relative to the unreinforced counterpart; that is, some property changes were
increased while others were decreased. We postulate that this outcome is a consequence of the fact that
each of the nanofillers hampered the polymerization process in the cement; specifically, the nanofiller
acts as a scavenger of the radicals produced during polymerization reaction due to the delocalized p-
bonds. Results obtained from the chemical structure and polymer chain size distribution determined,
respectively, by nuclear magnetic resonance and size exclusion chromatography analysis, on the
polymer extracted from the specimens support the postulated mechanism. Furthermore, in the case of
the 0.5 wt/wt% GO-reinforced cement, we showed that when the concentration of the radical species in
the PMMA bone cement was doubled, mechanical properties markedly improved (relative to the value
in the unreinforced cement), suggesting suppression of the aforementioned scavenger activity.

Introduction addressed in many studies.6,7,21,22 The results show that the


answer depends on many factors, such as the polymer, the
Single-layer graphene has received a lot of research attention method of preparation of the nanofiller, and the nanofiller
because of its remarkable thermal, electronic, and mechanical content. Also, improvement of the physicochemical properties of
properties; moreover, graphene can significantly improve the the nanocomposite depends on the distribution of nanofiller in
physical properties of many polymers at low loading.1–5 Gra- the polymer matrix, which, in turn, depends on the chemical
phene has been referred to as a credible alternative to carbon affinity between them.23
nanotubes (CNTs) for mechanical reinforcement of polymeric Since its discovery in 1960, PMMA bone cement remains one
matrixes.6,7 Graphene powder can be produced in a way scalable of the most widely used biomedical materials, being employed to
to mass production,2 which, in combination with low production anchor total joint replacements.24 One of the principal draw-
costs, makes graphene-powder-based composite materials backs of this cement is that it is not bioactive and, hence,
attractive for a variety of uses. In addition to their excellent provides no osseointegration with periprosthetic bone.17,25–27 To
physical and electrical properties and outstanding performance improve this situation, cements containing a bioactive agent,
in composite materials,8–12 CNTs have proven to be highly bio- usually, hydroxyapatite (HA) (which is both osseoconductive
compatible, a fact that has led them to be used for biomedical and osseoinductive28), have been prepared. To limit the potential
applications.13,14 In particular, the use of CNTs in biomaterials for adversely affecting the mechanical properties of the cement
applied to bone is also expected to act as scaffolds to promote when HA is employed, only low loadings (<10 vol./vol.%) have
and guide bone tissue regeneration as well as improve mechanical been used.29,30 On the other hand, studies reveal that, in order for
properties.13,15,16 One example is the use of CNTs as mechanical a bioactive cement to display osseoconductivity after setting, the
reinforcement of PMMA bone cements.17–20 bioactive filler loading should be >60 wt/wt% (35 vol./vol.%).31
The question as to which of these two nanofillers—graphene Thus, Singh et al.32 prepared a PMMA bone cement formulation
or CNTs—is the better reinforcing agent in polymeric nano- with 67 wt/wt% (42 vol./vol.%) of HA and 0.1 wt/wt% of CNTs
composites, in terms of mechanical properties, has been and showed that it was bioactive, possessed suitable mechanical
properties, and in vitro and in vivo tests showed accelerated cell
a
Nanotechnology Research Division, Center for Mechanical Technology & maturation.32,33
Automation, University of Aveiro, 3810-193 Aveiro, Portugal. E-mail:
paulam@ua.pt; Fax: +351234370953; Tel: +351234370830
The aforementioned advantages of graphene that make it
b
CEMUC, Department of Mechanical Engineering, University of Coimbra, more attractive than CNTs and the lack of a critical comparison
Portugal between these two nanofillers as the reinforcing agent in

This journal is ª The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 2937–2945 | 2937
View Article Online

a PMMA/high load HA bone cement were the motivation for the step, the nanofiller (GO or f-CNTs) was suspended in water and
present study, which had two primary purposes. The first was tip-sonicated for 30 min by using a probe-tip sonicator (Sonics
to critically compare graphene oxide (GO) and functionalized Model VC 130, 20 kHz, probe tip (6 mm diameter), amplitude
(f-CNTs) as the reinforcing agent PMMA/high-loading HA bone 80%, 2.2 kW). After that, the nanofiller, the cement powder, and
cement in terms of a large array of physical and mechanical the HA (particle size 2–3 mm; purity > 98%; Agoramat-Advanced
properties of the cement. The second was to propose a mecha- Materials, Portugal) were thoroughly blended. In the second
nism for the influence of these carbon nanostructures on the step, a freeze-granulation technique (Power Pro Freeze-granu-
properties determined. We selected graphene in its oxidized state, lator L5-2, Sweden) was used to increase the homogeneity of the
that is, GO, because, in previous studies, it was shown that the blended powder.32 The suspension was then dried by lyophili-
presence of oxygen functional groups at carbon nanostructure sation. After this final treatment, the nanocomposite was kept in
surfaces simultaneously facilitates their dispersion in the polymer a desiccator.
matrix and reduces their toxicity.32–35 The combination of Nine sets of solid component compositions were prepared:
nanofillers with in situ polymerization is considered an excellent PMMA/67 wt/wt% HA (unreinforced cement) herein referred to
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

method for producing truly synergetic composite materials with as PMMA/HA and PMMA/HA reinforced by 0.01, 0.1, 0.5 and
carbon structures.36 1 wt/wt% of GO or f-CNTs.

Experimental section 5. Preparation of PMMA/high load HA/nanofiller bone cement


specimens
Downloaded by University of Sussex on 14 January 2013

1. Materials
These specimens were fabricated by mixing a given solid
The PMMA bone cement used was a commercially available component with the liquid monomer of the PMMA bone cement
brand, CMW1 (DePuy, Warsaw, IN, USA). For its composition, under ambient conditions (temperature: 22  1  C; relative
see Table 1. humidity: $40%). Hand mixing was performed in a suitable,
clean, dry mixing ceramic bowl when the liquid component was
2. Synthesis of graphene oxide (GO) added to the solid component. The dough was mixed carefully to
The method used for the exfoliation of graphite was based on minimize the air entrapment, and then it was taken into gloved
that detailed in our previous work.37 Briefly, 2 grams of graphite hands and kneaded thoroughly according to the CMW1
(99.99%; Sigma-Aldrich) was dispersed into a flask containing 50 instruction leaflet. The dough was then poured into poly(tetra-
mL of concentrated H2SO4 and 7 g of KMnO4 and the solution fluorethylene) (Teflon) moulds designed according to each test,
was magnetically stirred for 2 h. After that, the solution was and pressed at 9.8 MPa using a uniaxial press for 10 min. The
treated with H2O2 until the gas evolution ceased and then the specimens were then stored under ambient laboratory conditions
resultant suspension was intensively washed, first with HCl for 24 h prior to testing.
solution (0.1 mol dm3) and then with distilled water by filtration
and centrifugation until a pH of 7 was achieved before drying by 6. Characterization of GO and f-CNTs
lyophilisation. The morphology of each of these nanofillers was determined
using a high-resolution field-emission scanning electron micro-
3. Functionalization of multiwalled carbon nanotubes scope (HRFE-SEM) (Hitachi SU-70; Japan) and a high resolu-
Multiwalled CNTs (purity > 95%; Nanocyl-3150; Belgium) were tion transmission electron microscope (HR-TEM) (JEOL 2200F;
chemically oxidised using a mixture of concentrated acids Japan) operated at 200 kV. The TEM specimens were prepared
H2SO4/HNO3 (3 : 1) for 24 h, with further details as given by Jia by depositing an aliquot of the solid component onto a carbon
et al.38 The resulting f-CNTs were intensively washed by centri- grid and then letting the solvent evaporate.
fugation with distilled water until the pH of their aqueous The thickness and size of GO sheets deposited on SiO2
dispersion was neutral. substrate were determined using an atomic force microscope
(AFM) (VEECO Multimode; USA).
The elemental composition and atomic concentration of
4. Preparation of a PMMA/high-load HA/nanofiller cement
each constituent in each of the nanofillers were determined using
powder mixture (solid component)
an X-ray photoelectron spectroscope (XPS) (VG Scientific
Three steps were used in the preparation of the cement powder ESCALAB 200 A; UK).
mixture (herein designed solid component) (Fig. 1). In the first
7. Determination of cement properties
Table 1 Composition of CMW1 PMMA bone cement
Dynamic mechanical analysis (DMA) (Tritec 2000; UK) tests
Bone Cement Powder: were performed on the specimens (20.0 mm  4.0 mm  2.2 mm)
Polymethyl methacrylate (wt/wt%) 88.85 in single cantilever bending mode, at a frequency of 1 Hz, an
Benzoyl peroxide (BPO) (wt/wt%) 2.05
Barium sulfate (wt/wt%) 9.10 amplitude of 0.010 mm and with a specimen heating rate of 2  C
Bone Cement Liquid: min1 from room temperature (22  C) to 150  C. The properties
Methyl methacrylate (wt/wt%) 98.5 determined were the storage modulus (E0 ), loss modulus (E00 )
N,N-Dimethyl-p-toluidine (DMT) (wt/wt%) #1.50
Hydroquinone/ppm 75
and, hence, damping factor (tan d), as a function of specimen
temperature (T). The temperature corresponding to the peak in

2938 | Nanoscale, 2012, 4, 2937–2945 This journal is ª The Royal Society of Chemistry 2012
View Article Online
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

Fig. 1 Schematic representation of the three steps used in preparing the solid component of the PMMA/HA/nanofiller bone cement. The morphology
Downloaded by University of Sussex on 14 January 2013

of the solid component is shown in the large SEM micrograph, while the morphology of the PMMA bead in the component, with nanofiller particles on
its surface, is shown in the inset.

the tan d vs. T plot is the glass transition temperature of the dissolved in deuterated chloroform and then the spectra were
cement (Tg). For each study group, 5 specimens were tested. acquired with a spectrometer (DRX 500, Bruker; Germany),
Four point bending tests were performed on the specimens operated at 300.13 MHz. For the SEC analysis, 5 mg of polymer
(25.0 mm  10.0 mm  3.3 mm) using a universal tensile tester fraction extracted from the specimen was dissolved in 1 mL of
machine (AG-X; 5 kN load cell; Japan) controlled with dimethylacetamide (DMAc) solution, at 20  C, for 30 min and
a commercially available computer software (Trapezium X; then filtered through a 0.3 mm filter, yielding a sample concen-
Shimadzu; Japan), at a crosshead displacement rate of 5 mm tration of 0.5% (5 mg mL1). The columns, injector system, and
min1. For each study group, 5 specimens were tested. SEM detector were each maintained at 70  C during the analysis. The
images were obtained from the fractured surfaces of the eluent (DMAc) was pumped at a flow rate of 0.9 mL min1. The
specimens. analytical columns were calibrated with polystyrene standards
The modulus of elasticity (E) was determined using the (Polymer Laboratories; UK) in the range 1.7–100.0 kDa. The
impulse excitation technique (ASTM C1259-96). The specimen injected volume was 100 mL. Average molecular weight (Mw),
(25.0 mm  10.0 mm  3.3 mm) was set in free flexural vibration number average molecular weight (Mn) and polydispersity
by a light mechanical impulse. A strain gage was used to assess (PDI ¼ Mw/Mn) were determined from these analysis using
the vibration. Fast Fourier transform was applied to analyse the Cirrus software.
vibration signals to obtain the frequency of the first flexural
vibrational mode. E was calculated using the expression: 8. Preparation and characterization of cements with different
! ! concentrations of radicalars
mft2 l3
E ¼ 0:9465 T1 $
d t3 Four varieties of the nanocomposite cements that contained
0.1 wt/wt% f-CNTs loading were prepared, the difference being
where m is the mass of the specimen, ft is the fundamental in the additional concentration of the radicalars, namely, BPO
frequency, d is the width of the specimen, l the length of the and DMT (RIC), that were mixed in the solid component and
specimen, t is the specimen thickness, and T1 is a correction liquid monomer during preparation of the cements (Table 2). For
factor that takes into account the finite dimensions of the spec- each cement, the bending strength and E were determined
imen. For the calculation of T1, a constant Poisson ratio of 0.3 (in each case, sample size ¼ 5).
was used. For each study group, 5 specimens were tested. Also, using RIC ¼ 2X, nanocomposite cements that contained
The maximum exothermic temperature reached during the 0.5 wt/wt% of either GO or f-CNTs were prepared. For each
polymerization stage (Tmax) was determined per ISO 5833, using cement, bending strength, E, and Tg were determined (in each
a Ni–Cr–Al k-type thermocouple and a commercially available case, sample size ¼ 5).
data acquisition system (PicoLog Data Acquisition Software; For each of the above described cements, the steps in the
Pico Tecknology Ltd., UK). For each study group, the tests were preparation and the methods of determination of the properties
run in triplicate. were the same as given in the previous sub-sections.
Preparation of specimens (1.5 g each) for the nuclear magnetic
resonance (NMR) and size exclusion chromatography (SEC)
9. Statistical analysis
analysis needed a preliminary Soxhlet extraction of the polymer
from the specimen. This was carried out for 8 h using 150 mL of In the case of each of the mechanical properties and Tg, the
dichloromethane. For the 1H NMR analysis, the specimen was results were evaluated for statistical significance using a one-way

This journal is ª The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 2937–2945 | 2939
View Article Online

Table 2 Additional quantities of BPO and DMT added to the solid


component (5 g) and the liquid component (1.8 mL), respectively

Radicalar agents

Concentration BPO (solid component) DMT (liquid component)

0 Initial Initial
1 40 mg (0.165 mmol) 26.25 mL (0.182 mmol)
2 80 mg (0.330 mmol) 52.5 mL (0.364 mmol)
3 120 mg (0.495 mmol) 78.75 mL (0.546 mmol)
10 400 mg (1.65 mmol) 262.5 mL (1.819 mmol)

(ANOVA) analysis of variance, with p < 0.05 being considered


Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

significant. Post-hoc tests were conducted using Duncan


methods (IBM SPSS Statistics; USA).

Results and discussion


Downloaded by University of Sussex on 14 January 2013

1. Characteristics of the carbon nanostructures


The as-synthesized GO consisted of thin flakes, with a well-
defined few-layer structure (Fig. 2a and b). Note the wrinkled
surface, which could play an important role in enhancing
Fig. 3 Topographic view of contact-mode AFM scan of GO deposited
mechanical interlocking and load transfer to the matrix in
on SiO2 glass and height profile through the line.
a GO-reinforced bone cement. The f-CNTs were 1.5 mm long,
with tube diameter ranging from 30 nm to 50 nm (Fig. 2c and
d), and hollow in structure (Fig. 2c). The plate size distribution in Table 3 Results of XPS analysis of the C1s content of the surfaces of
GO ranged from 200 nm to 3 mm (Fig. 3) and the mean GO and f-CNTs
thickness of GO nanosheets was 1.2 nm, corresponding to
GO f-CNT
nanosheets with two or three layers.39
Two clear trends are seen in the XPS results (Table 3). First, Binding Atomic conc. Binding Atomic conc.
the oxidation level of the GO surface (48.57%) is markedly higher energy/eV (at%) energy/eV (at%)
than that of f-CNTs surface (13.96%). This difference is consis- C–C 285 15.67 285.4 12.89
tent with the differences in the methods used to prepare these C–O 286.9 42.55 286.4 6.57
nanofillers. Second, while the GO surface is mainly functional- C¼O 288.2 4.14 287.4 3.51
ized with epoxy groups, the f-CNTs surface presents a higher C(O)O 289.2 1.88 288.6 3.88
amount of carboxylic groups.

2. Characteristics of the PMMA/HA/nanofiller cements


2.1 Mechanical properties. The results for the mechanical
properties (Table 4 and Fig. 4) showed that, relative to the
unreinforced cement, (1) neither of the filler enhanced E0 ; (2) each
nanofiller led to a broadening of the tan d plot (Fig. 4); (3) while
GO enhanced bending strength, f-CNTs did not; and (4) GO
markedly enhanced E but f-CNTs did not.
The trends in the mechanical properties are consistent with the
morphologies of the GO sheets and the f-CNTs. The GO sheets
display a high specific area and wrinkled surface, contributing to
a better integration of the filler in cement matrix and high matrix
adhesion/interlocking. On the other hand, the cylindrical shape
of the f-CNTs does not contribute to identical interlocking with
the micron sized elements present in the solid component
(PMMA beads and HA). The aforementioned trends in
mechanical properties are also consistent with the morphologies
of the surfaces of fractured bending test specimens (Fig. 5).
Specifically, fewer and smaller pores were seen in the GO-rein-
Fig. 2 HRFSEM and HRTEM images of graphene oxide respectively forced cements, regardless of the loading (for example, Fig. 5a
(a) and (b) and carboxyl functionalized MWCNTs respectively (c) and and b), whereas a porous structure was observed in f-CNTs
(d). reinforced cements, especially at high loadings (Fig. 5c and d). In

2940 | Nanoscale, 2012, 4, 2937–2945 This journal is ª The Royal Society of Chemistry 2012
View Article Online

Table 4 Mechanical properties of the PMMA/HA nanocomposites prepared with different loadings of GO and f-CNTs

Nanofiller loading (wt/wt%)

Nanofiller Properties 0 0.01 0.1 0.5 1


0
GO E /GPa 2.67  0.31 a
2.49  0.23a
2.26  0.14 a
2.20  0.37 a
2.36  0.19a
Tg/ C 118.0  0.8 118.1  2.3a 98.2  0.4 98.0  0.9 98  1.2
Bending strength/MPa 20.00  1.46 23.01  0.37 22.34  0.74 24.89  1.19 28.28  1.09
E/GPa 6.09  0.42 9.53  0.46 9.84  0.47 9.63  0.50 10.22  0.67
Tmax/ C 82.4  1.3 71.3  1.4 64.7  0.9 56.3  1.1 38.6  2.1
f-CNTs E0 /GPa 2.67  0.31 1.66  0.23 2.13  0.34 1.35  0.21 0.99  0.27
Tg/ C 118.0  0.8 95.3  1.1 91.2  1.5 69.9  1.3 68.0  1.5
Bending strength/MPa 20.00  1.46 14.22  0.63 15.11  0.90 12.57  0.63 11.81  1.69
E/GPa 6.09  0.42a 6.89  0.68a 8.47  0.38a 6.93  0.65a 6.80  0.68a
Tmax/ C 82.4  1.3 61.9  1.2 53.8  0.7 28  1.1 22.4  1.4
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

a
Indicates no statistically significant difference between control and reinforced cements (ANOVA at 0.05 level of significance); all the other values show
statistically significant difference between control and reinforced cement (ANOVA at 0.05 level of significance).
Downloaded by University of Sussex on 14 January 2013

Fig. 4 Summary of the tan delta results for PMMA/HA reinforced with different percentages of GO (a) and f-CNTs (b).

attributed to the release of the unreacted methyl methacrylate


(MMA) monomer (which is volatile) during the polymerization
of the cement. In other words, each of the nanofillers, but
particularly f-CNTs, captures the radicals produced during the
cement polymerization process, thereby hampering their normal
course.

2.2 Tg results. Besides the already mentioned broadening of


the tan d plot in the presence of both nanofillers (Fig. 4), f-CNTs
had a pronounced effect on the Tg values: Tg decreased
substantially with f-CNTs increase on the cement composition
(Fig. 4b). This is indicative of a decrease in the polymer chain size
and the formation of some oligomeric species.

2.3 Tmax results. The trend in the results, of a decrease of


Tmax with the presence of either GO or f-CNTs (Table 3), is the
Fig. 5 Morphologies of fracture surfaces of bending test specimens of same as reported by Ormsby et al.40 for PMMA-CNT bone
nanocomposite PMMA bone cements reinforced with 1 wt/wt% GO (a) cement. The decrease is more pronounced when f-CNTs were
and (b) and 1 wt/wt% of f-CNTs (c) and (d). used compared to when GO was used. These results may be
interpreted as evidence of the nanofiller acting as a radical
each type of carbon nanostructure-reinforced cement, pores scavenger, with f-CNTs having a more pronounced scavenging
occur in spite of the homogeneous distribution of the nanofiller effect. The higher oxygen density of GO (Table 2) and the
cement matrix (Fig. 5a and d). The origin of these pores may be presence of lattice defects (localized sp3 defects within the sp2

This journal is ª The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 2937–2945 | 2941
View Article Online

analysis of their relative peaks in the spectra showed that the


former and latter end groups predominate in the case of the
polymer grown in the presence of GO and f-CNTs,
respectively.

2.5 SEC results. These results (Table 5) confirm that GO had


a less detrimental effect on cement polymerization compared
to f-CNTs. Specifically, the organic fraction extracted from the
GO-reinforced specimen displayed only one main peak whereas
two main peaks were displayed in the case of f-CNT-reinforced
cement specimen (corresponding to two main-sized fractions).
These results suggest a more pronounced detrimental effect of
f-CNTs on the MMA polymerization, with the appearance of an
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

oligomeric fraction.

3. Influence of radicalar concentration on properties of


Fig. 6 Results of the 1H NMR analysis of PMMA with (a) GO and (b)
PMMA/HA/nanofiller bone cement
Downloaded by University of Sussex on 14 January 2013

f-CNTs.
When 0.1 wt/wt% f-CNTs were used as the reinforcing agent,
the highest value of each of the two mechanical properties
carbon network)6,41 lead to a lower density of p-bonds when determined for the cement (relative to both the unreinforced
compared with f-CNTs, thereby reducing GO’s scavenging cement and the reinforced cement when the liquid monomer in
action. CMW1 cement was used (that is, RIC ¼ 0X)) was obtained with
RIC ¼ 2X (Fig. 7).
2.4 1H NMR results. Two key trends are seen in these With RIC ¼ 2X, there were significant increases in both
results (Fig. 6). First, the tacticity is dependent on the nanofiller bending strength and E for cements in which GO was used as the
used; specifically, when the polymer was grown in the presence reinforcing agent (relative to the corresponding values for the
of GO, isotactic conformation increases significantly, whereas unreinforced cement), regardless of the loading of the nanofiller
heterotactic and syndiotactic conformations were dominant for (Table 6). Furthermore, for a cement reinforced with 0.5 wt/wt%
PMMA grown in the presence of f-CNTs. Further work is GO, Mw was 79 100 (PDI ¼ 1.38) compared to 47 700
needed to explain this finding. Second, there are two terminal
end groups, namely benzoyloxy and unsaturated, and careful
Table 6 Mechanical properties of PMMA/HA reinforced with different
loadings of GO and with RIC ¼ 2X
Table 5 Results of SEC analysis of the polymer fraction extracted from GO (wt/wt%) E0 /GPa Bending strength/MPa E/GPa
the nanocomposites prepared with 0.5 wt/wt% of GO and 0.5 wt/wt% of
f-CNT 0 2.67  0.31a 20.0  01.46 6.09  0.42
0.01 3.20  0.36a 21.17  0.37 12.57  0.46
Cement Peak 1 Peak 2 0.1 2.96  0.21a 31.45  0.74 11.52  0.52
0.5 3.15  0.43a 37.53  1.19 12.00  0.49
0.5GO Mw, 47 700 — 1.0 2.77  0.52a 29.50  1.90 12.24  0.67
Mn, 22 800 —
a
PDI, 2.09 — Indicates no statistically significant difference between control and
0.5CNT Mw, 40 400 Mw, 870 reinforced cements (ANOVA at the 0.05 level of significance); all the
Mn, 28 230 Mn, 570 other values show statistically significant difference between control
PDI, 1.43 PDI, 1.51 and reinforced cement (ANOVA at he 0.05 level of significance).

Fig. 7 Variation in bending strength (a) and modulus of elasticity (b) values obtained for the PMMA/HA matrix and for PMMA/HA reinforced with
0.1 wt/wt% of f-CNTs, with cements prepared using additional quantities of radicalar agents (RIC ¼ 0X, 1X, 2X, 3X and 10X).

2942 | Nanoscale, 2012, 4, 2937–2945 This journal is ª The Royal Society of Chemistry 2012
View Article Online
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E
Downloaded by University of Sussex on 14 January 2013

Fig. 8 Schematic representation of a possible mechanism of the radical polymerization of PMMA under the influence of carbon nanostructures.

(PDI ¼ 2.09) for a cement with the same GO loading but since a different initiator was used, namely one that is activated
synthesized using RIC ¼ 0X (Table 5). This indicates a less broad at room temperature, it is highly unlikely that the p-bonds of
range of molecular weights in these samples showing that the carbon nanostructures could establish covalent bonds with the
increase of RIC compensates for the radical scavenger activity of polymer chains. Instead, the p-bonds are more likely to act as
the nanofillers. scavengers of primary and macromolecular radicals, as discussed
above, because of the high electron affinity of carbon
nanostructures.44–46
4. Postulated mechanism
We suggest two approaches to counter the negative
Carbon nanostructures, in particular carbon nanotubes, are influence of a nanofiller on the polymerization of the cement
large arrays of conjugated double bonds; they are expected to and, hence, its mechanical properties. The first involves
have large electron donor–acceptor capability. This specific modification of the surface of the nanofiller with PMMA
property makes them remarkably reactive toward free radi- chains via atom transfer radical polymerization, as proposed by
cals.42,43 We, therefore, postulate that both GO and f-CNTs are us in our previous study.47 In that work, we showed that when
active participants in the polymerization of the nanocomposites this method was applied to GO-reinforced PMMA/HA, there
cement (Fig. 8). In this scheme, we distinguish between two was good integration of GO in the matrix and enhanced
different types of action assigned to carbon nanostructures that mechanical properties. One limitation of this approach,
can restrict the PMMA radicalar polymerization. The first is however, is that it is difficult to scale up. The second approach
polymerization retardation (partial loss of initiation activity). In involves increasing the concentration of the radicalar species
this case, the primary radicals formed after the split of BPO during the polymerization, thereby attempting to compensate
molecules, when they contact the aromatic surface of carbon for a subsequent loss due to a scavenging activity. Results
nanostructures, lose their unpaired electron, giving rise to obtained in the present study (Fig. 7 and Table 6) show the
non-reactive species. The second is polymerization inhibition feasibility of the second approach, particularly in the case of
(chain termination). Here, during the growth of macromolec- a GO-reinforced bone cement. These results are consistent with
ular radicals, an electron transference may also occur to the an improvement in the polymerization, as reflected in the
carbon aromatic surface, ending the polymer chain growth, increase of Mw.
resulting in the formation of small-sized macromolecules. These
two types of actions may occur simultaneously and/or
5. Future work
competitively.
Jia W. et al.38 were probably the first workers to propose the Future work will be conducted on the determination of other key
participation of CNTs in the MMA radical polymerization, properties of GO-reinforced PMMA bone cements prepared
pointing out that CNTs could be attacked by AIBN, thereby with additional radical initiators, such as curing time, setting
opening their p-bonds. These opened p-bonds of CNTs could time, fracture toughness, fatigue life, and fatigue crack propa-
link with the PMMA, resulting in obstruction of the growth of gation rate. In addition, since the in vitro toxicity of graphene
PMMA and, possibly, producing a C–C bond between the CNT towards human cells appears to be lower than that of CNTs and
and the PMMA. A strong interface between the CNT and the does not hamper the proliferation of human mesenchymal stem
PMMA was thus claimed, without convincing evidence, to be cells and their differentiation into bone cells,48–50 another future
produced through that bonding.38 This explanation has been put study will be to determine the in vitro and in vivo biocompatibility
forward by other workers.18,19,36,40 In the present work, however, of GO-reinforced PMMA bone cements.

This journal is ª The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 2937–2945 | 2943
View Article Online

Conclusions 10 M. Mazaheri, D. Mari, Z. R. Hesabi, R. Schaller and G. Fantozzi,


Compos. Sci. Technol., 2011, 71, 939–945.
In the present study, we compared the influence of two carbon 11 P.-C. Ma, N. A. Siddiqui, G. Marom and J.-K. Kim, Composites, Part
A, 2010, 41, 1345–1367.
nanostructures (GO and f-CNTs) when blended with the powder 12 Z. Spitalsky, D. Tasis, K. Papagelis and C. Galiotis, Prog. Polym.
of a PMMA/high-load HA bone cement on a large collection of Sci., 2010, 35, 357–401.
properties of the cements. In the case of GO-reinforced cements, 13 K. Sahithi, M. Swetha, K. Ramasamy, N. Srinivasan and
we found higher values of some of the mechanical properties N. Selvamurugan, Int. J. Biol. Macromol., 2010, 46, 281–283.
14 D. A. Gomez-Gualdr on, J. C. Burgos, J. Yu and P. B. Balbuena,
determined, compared to the corresponding values for the Nanoparticles in Translational Science and Medicine, Academic
unreinforced cement. This is attributed to GO’s high specific Press, 2011, vol. 104, pp. 175–245.
area, high degree of surface functionalization, and wrinkled 15 E. Hirata, M. Uo, H. Takita, T. Akasaka, F. Watari and
A. Yokoyama, Carbon, 2011, 49, 3284–3291.
surface, all of which lead to high adhesion/interlocking of the
16 T. R. Nayak, L. Jian, L. C. Phua, H. K. Ho, Y. P. Ren and
nanofiller with the cement matrix. In the case of the other G. Pastorin, ACS Nano, 2010, 4, 7717–7725.
mechanical properties determined, there is depreciation no 17 D. Renteria-Zamarron, D. A. Cortes-Hernandez, L. Bretado-Aragon
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E

matter which of the two nanofillers was used. We postulate that and W. Ortega-Lara, Mater. Des., 2009, 30, 3318–3324.
18 R. Ormsby, T. Mcnally, C. Mitchell, P. Halley, D. Martin,
this is a consequence of the fact that the carbon nanostructure T. Nicholson and N. Dunne, Carbon, 2011, 49, 2893–2904.
acts as a scavenger of the radicals that are produced during the 19 R. Ormsby, T. Mcnally, C. Mitchell and N. Dunne, J. Mater. Sci.:
polymerization process, causing the inhibition and retardation of Mater. Med., 2010, 21, 2287–2292.
the growth of polymer chains. This phenomenon is more 20 B. Marrs, R. Andrews, T. Rantell and D. Pienkowski, J. Biomed.
Downloaded by University of Sussex on 14 January 2013

Mater. Res., Part A, 2006, 77, 269–276.


pronounced when f-CNTs are used compared to when GO is 21 T. Chatterjee, K. Yurekli, V. G. Hadjiev and R. Krishnamoorti, Adv.
used because of the former’s higher density of p-bond areas. Funct. Mater., 2005, 15, 1832.
Based on this explanation, we proposed that one approach that 22 J. Du, L. Zhao, Y. Zeng, L. Zhang, F. Li, P. Liu and C. Liu, Carbon,
2011, 49, 1094–1100.
may be taken to counter the negative impact of a carbon nano- 23 J. M. Wernik and S. A. Meguid, Appl. Mech. Rev., 2010, 63,
structure on mechanical properties of the cement is to substan- 050801.
tially increase the concentration of the radicalar initiators (RICs) 24 M. R. S. Chandler, N. D. Watkins, A. Briscoe and A. M. New, Proc.
present in the cement during polymerization. We demonstrated Inst. Mech. Eng., Part H, 2006, 220, 321–331.
25 T. W. Bauer and J. Schils, Skeletal Radiol., 1999, 28, 483–497.
the feasibility of this approach for both GO- and f-CNT-rein- 26 R. L. Barrack, J. Arthroplasty, 2000, 15, 1036–1050.
forced cements. Taken as a whole, the present results should 27 P. E. Sinnett-Jones, M. Browne, A. J. Moffat, J. R. T. Jeffers,
contribute to the development of novel PMMA-based nano- N. Saffari, J. Y. Buffiere and I. Sinclair, J. Biomed. Mater. Res.,
Part A, 2009, 89, 1088–1097.
composite bone cements.
28 P. K. Stephenson, M. A. R. Freeman, P. A. Revell, J. Germain,
M. Tuke and C. J. Pirie, J. Arthroplasty, 1991, 6, 51–58.
Acknowledgements 29 M. J. Dalby, L. Di Silvio, E. J. Harper and W. Bonfield, Biomaterials,
2001, 22, 1739–1747.
This work was co-financed by International Iberian Nanotech- 30 M. J. Dalby, L. Di Silvio, E. J. Harper and W. Bonfield, Biomaterials,
2002, 23, 569–576.
nology Laboratory (INL)—Foundation of Science and Tech- 31 T. Tsukeoka, M. Suzuki, C. Ohtsuki, A. Sugino, Y. Tsuneizumi,
nology (FCT—Portugal—FCOMP-01-0124-FEDER-008439). J. Miyagi, K. Kuramoto and H. Moriya, Biomaterials, 2006, 27,
P. Marques thanks the Ciencia 2007 Program and G. Gonclaves 3897–3903.
thanks INL for a PhD grant. We thank Prof. Dmitry Evtyugin 32 M. K. Singh, T. Shokuhfar, J. J. D. A. Gracio, A. C. M. De Sousa,
J. M. D. F. Fereira, H. Garmestani and S. Ahzi, Adv. Funct.
(CICECO, University of Aveiro) for help with SEC analysis and Mater., 2008, 18, 694–700.
valuable discussions. We especially thank Prof. Alessandro 33 M. K. Singh, J. Gracio, P. Leduc, P. P. Goncalves,
Gandini for valuable and fascinating discussions regarding P. A. A. P. Marques, G. Goncalves, F. Marques, V. S. Silva,
F. Capela, E. Silva, J. Reis, J. Potes and A. Sousa, Nanoscale, 2010,
polymer chemistry. 2, 2855–2863.
34 S. K. Smart, A. I. Cassady, G. Q. Lu and D. J. Martin, Carbon, 2006,
References 44, 1034–1047.
35 H. Hu, Y. Ni, V. Montana, R. C. Haddon and V. Parpura, Nano
1 K. S. Novoselov, A. K. Geim, S. V. Morozov, D. Jiang, Y. Zhang, Lett., 2004, 4, 507–511.
S. V. Dubonos, I. V. Grigorieva and A. A. Firsov, Science, 2004, 36 C. Velasco-Santos, A. L. Martınez-Hernandez, F. T. Fisher, R. Ruoff
306, 666–669. and V. M. Casta~ no, Chem. Mater., 2003, 15, 4470–4475.
2 S. Stankovich, D. A. Dikin, G. H. B. Dommett, K. M. Kohlhaas, 37 G. Goncalves, P. A. A. P. Marques, C. M. Granadeiro,
E. J. Zimney, E. A. Stach, R. D. Piner, S. T. Nguyen and H. I. S. Nogueira, M. K. Singh and J. Gracio, Chem. Mater., 2009,
R. S. Ruoff, Nature, 2006, 442, 282–286. 21, 4796–4802.
3 B. G. Choi, J. Hong, Y. C. Park, D. H. Jung, W. H. Hong, 38 Z. Jia, Z. Wang, C. Xu, J. Liang, B. Wei, D. Wu and S. Zhu, Mater.
P. T. Hammond and H. Park, ACS Nano, 2011, 5, 5167–5174. Sci. Eng., A, 1999, 271, 395–400.
4 Y. Li, Polymer, 2011, 52, 2310–2318. 39 H. C. Schniepp, J.-L. Li, M. J. Mcallister, H. Sai, M. Herrera-Alonso,
5 L. L. Tian, P. Anilkumar, L. Cao, C. Y. Kong, M. J. Meziani, D. H. Adamson, R. K. Prud’homme, R. Car, D. A. Saville and
H. J. Qian, L. M. Veca, T. J. Thorne, K. N. Tackett, T. Edwards I. A. Aksay, J. Phys. Chem. B, 2006, 110, 8535–8539.
and Y. P. Sun, ACS Nano, 2011, 5, 3052–3058. 40 R. Ormsby, T. Mcnally, C. Mitchell and N. Dunne, J. Mech. Behav.
6 T. Kuilla, S. Bhadra, D. Yao, N. H. Kim, S. Bose and J. H. Lee, Prog. Biomed. Mater., 2010, 3, 136–145.
Polym. Sci., 2010, 35, 1350–1375. 41 C. Soldano, A. Mahmood and E. Dujardin, Carbon, 2010, 48, 2127–
7 M. A. Rafiee, J. Rafiee, Z. Wang, H. Song, Z.-Z. Yu and N. Koratkar, 2150.
ACS Nano, 2009, 3, 3884–3890. 42 A. Galano, Nanoscale, 2010, 2, 373–380.
8 M. K. Yeh, T.-H. Hsieh and N.-H. Tai, Mater. Sci. Eng., A, 2008, 43 M. Francisco-Marquez, A. Galano and A. MartıNez, J. Phys. Chem.
483–484, 289–292. C, 2010, 114, 6363–6370.
9 Y. Jia, K. Peng, X.-L. Gong and Z. Zhang, Int. J. Plast., 2011, 27, 44 P. J. Krusic, E. Wasserman, P. N. Keizer, J. R. Morton and
1239–1251. K. F. Preston, Science, 1991, 254, 1183–1185.

2944 | Nanoscale, 2012, 4, 2937–2945 This journal is ª The Royal Society of Chemistry 2012
View Article Online

45 P. C. P. Watts, P. K. Fearon, W. K. Hsu, N. C. Billingham, 48 Z. M. Markovic, L. M. Harhaji-Trajkovic, B. M. Todorovic-


H. W. Kroto and D. R. M. Walton, J. Mater. Chem., 2003, 13, Markovic, D. P. Kepic, K. M. Arsikin, S. P. Jovanovic,
491–495. A. C. Pantovic, M. D. Dramicanin and V. S. Trajkovic,
46 I. Fenoglio, M. Tomatis, D. Lison, J. Muller, A. Fonseca, J. B. Nagy Biomaterials, 2011, 32, 1121–1129.
and B. Fubini, Free Radical Biol. Med., 2006, 40, 1227– 49 Y. Zhang, S. F. Ali, E. Dervishi, Y. Xu, Z. Li, D. Casciano and
1233. A. S. Biris, ACS Nano, 2010, 4, 3181–3186.
47 G. Goncalves, P. Marques, A. Barros-Timmons, I. Bdkin, 50 T. R. Nayak, H. Andersen, V. S. Makam, C. Khaw, S. Bae, X. F. Xu,
M. K. Singh, N. Emami and J. Gracio, J. Mater. Chem., 2010, 20, P. L. R. Ee, J. H. Ahn, B. H. Hong, G. Pastorin and B. Ozyilmaz,
9927–9934. ACS Nano, 2011, 5, 4670–4678.
Published on 29 February 2012 on http://pubs.rsc.org | doi:10.1039/C2NR30303E
Downloaded by University of Sussex on 14 January 2013

This journal is ª The Royal Society of Chemistry 2012 Nanoscale, 2012, 4, 2937–2945 | 2945

You might also like