You are on page 1of 10

j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

journal homepage: www.elsevier.com/locate/jmatprotec

Densification behaviour of nanocrystalline


hydroxyapatite bioceramics

S. Ramesh a,∗ , C.Y. Tan a , S.B. Bhaduri b , W.D. Teng c , I. Sopyan d


a Ceramics Technology Laboratory, University Tenaga Nasional, 43009 Kajang, Selangor, Malaysia
b School of Materials Science & Engineering, Clemson University, Clemson, SC 29634, USA
c Ceramics Technology Group, SIRIM Berhad, 40911 Shah Alam, Selangor, Malaysia
d Department of Manufacturing and Materials Engineering, Faculty of Engineering,

International Islamic University Malaysia, Malaysia

a r t i c l e i n f o a b s t r a c t

Article history: The sinterability of nanocrystalline hydroxyapatite (HA) particles by microwave sinter-
Received 6 June 2007 ing was compared with conventional pressureless sintering. The results revealed that
Received in revised form microwave heating was effective in producing a useful HA body in a very short sinter-
10 November 2007 ing cycle without disrupting the HA phase stability. The maximum hardness of 7.21 GPa
Accepted 6 December 2007 and 6.38 GPa was obtained for HA sintered at 1050 ◦ C by the conventional method and
1150 ◦ C by microwave sintering, respectively. The maximum fracture toughness measured
for the microwave-sintered and conventional-sintered HA was 1.45 MPam1/2 at 1050 ◦ C and
Keywords: 1.22 MPam1/2 at 1000 ◦ C, respectively. Although the relative density of microwave-sintered
Hydroxyapatite HA was slightly lower than the conventional-sintered HA throughout the sintering regime
Bioceramic employed, taking into account of the heating and soaking periods, the time taken by
Sintering microwave sintering to achieve a relative density of 96.5% was about 3% of the time con-
Mechanical properties sumed for samples sintered by the conventional heating. Microwave heating was found to be
an effective technique to produce a useful HA body for clinical applications without causing
grain coarsening.
© 2007 Elsevier B.V. All rights reserved.

1. Introduction studies have been carried out to improve the mechani-


cal properties of sintered HA (Suchanek and Yoshimura,
Hydroxyapatite, Ca10 (PO4 )6 (OH)2 (HA) material has been clin- 1998; Rodrı́guez-Lorenzo et al., 2002; Kokubo et al., 2003).
ically applied in many areas of dentistry and orthopaedics These include the development of new powder process-
because of its excellent osteoconductive and bioactive prop- ing/synthesis techniques and composition modification. The
erties which is due to its chemical similarity with the mineral ultimate goal is to identify the most appropriate synthesis
portion of hard tissues (Hench, 1998; Afshar et al., 2003; method and conditions to produce well-defined particle mor-
Suchanek and Yoshimura, 1998; Aoki et al., 1987; Metsger et phology (Suchanek and Yoshimura, 1998; Aoki et al., 1987;
al., 1982). Metsger et al., 1982; Gabriel Chu et al., 2002; Muralithran and
One of the major concern with HA is the low fracture Ramesh, 2000; Rodrı́guez-Lorenzo et al., 2002; Kokubo et al.,
toughness, i.e. <1 MPam1/2 of the sintered body (Gabriel Chu et 2003) that is sinter active at low temperatures to produce body
al., 2002; Muralithran and Ramesh, 2000). As a result various that exhibits high mechanical properties.


Corresponding author. Tel.: +60 3 89287282; fax: +60 3 89212116.
E-mail address: ramesh@uniten.edu.my (S. Ramesh).
0924-0136/$ – see front matter © 2007 Elsevier B.V. All rights reserved.
doi:10.1016/j.jmatprotec.2007.12.027
222 j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

One of the critical controlling parameter that requires green compacts were sintered in air atmosphere at a ramp
attention during the processing of hydroxyapatite is the selec- rate of 2 ◦ C/min and soaking time of 2 h was employed for each
tion of a suitable sintering method to obtain a solid, high firing cycle.
density HA body that is characterized by having fine-grained The microwave sintering was carried out using a
microstructure. The most commonly used sintering tech- microwave furnace (M.M.T., Knoxville, TN) provided with a
nique is the conventional pressureless sintering (CPS) in air variable power output magnetron source capable of operat-
atmosphere. This method of densification normally requires ing from 0 kW to 3 kW at 2.45 GHz. The cavity is large and
high temperature, slow heating rate and long holding time to overmoded, thus ensuring mixing of the microwave modes
produce >99% dense HA body with large-grained microstruc- and resulting in a homogeneous field distribution. The HA
ture and low mechanical properties (Metsger et al., 1982; compacts were sandwiched between two SiC susceptors and
Muralithran and Ramesh, 2000). surrounded with alumina fiber before sintering. In this pro-
In contrast, sintering by microwave heating has been cess, the sintering schedule for all samples was kept constant
reported to be a viable processing technique to produce a at duration of 30–35 min. The temperature of the samples was
dense HA body that possessed fine microstructure coupled measured via a pyrometer (Model M77, Mikron Corp.) and the
with improved mechanical properties (Agrawal et al., 1992; power of the microwave system adjusted accordingly to main-
Fang et al., 1994; Yang et al., 2002; Vijayan and Varma, 2002). tain the required temperature.
For instance Yang et al. (2002) prepared green compacts of All the sintered disc samples were polished to a 1 ␮m
stoichiometric HA synthesized by the chemical precipitation surface finish prior to testing. Phase analyses by X-ray diffrac-
method. The sintering behaviour of the synthesized HA was tion (XRD: Rigaku Geiger-Flex, Japan) of synthesized powder
compared by microwave heating at 1050–1100 ◦ C for 5–10 min and polished compacts were carried out at room temperature
in a 10 kW, 2.45 GHz microwave furnace and conventional using Cu K␣ radiation at a scan speed of 0.5◦ /min and a step
heating at 1100 ◦ C for 1 h. SEM investigation performed on scan of 0.02◦ . The microstructural evolution of the sintered
the fracture surface indicated that the microwave-sintered HA samples was performed by scanning electron microscopy
had very fine grain sizes, in the range of 0.3–0.5 ␮m, as com- (SEM: Philips XL30). The sintered grain size of polycrystalline
pared to 3–5 ␮m observed for the conventional-sintered HA. HA was determined from the SEM micrographs using the line
The authors inferred that the finer grain size and higher den- intercept method (Mendelson, 1969).
sities of the microwave-sintered HA observed in their work The bulk density of sintered samples was determined by
was due to the higher efficiency of microwave heating. This the Archimedes technique. The relative density was calcu-
can be attributed to the rapid and uniform generation of heat lated by taking the theoretical density of HA as 3.156 g cm−3 .
within the body instead of being transferred from outside the The microhardness (Hv ) and fracture toughness (KIc ) of the
body as in the conventional sintering (Sutton, 1989). Although samples were determined using the Vickers indentation
microwave sintering (MS) has great potential in fabricating method. The indentation load (<200 g) was applied and held
dense body, there are concerned regarding the stability of HA in place for 10 s. Five indentations were made for each sam-
phase (Fang et al., 1994). ple and the average value was taken. The indentation fracture
The objective of the present paper was to compare the toughness was determined from the equation derived by
effect of microwave sintering with that of conventional Niihara (Niihara, 1985):
method on the densification behaviour and mechanical prop-
 c −1.5
erties of synthesized nanocrystalline hydroxyapatite. 0.5
KIc = 0.203 (Hv )(a) (1)
a

2. Experimental details where Hv is the Vickers hardness, a the half diagonal of the
indentation, c the radial crack dimension measured from the
centre of the indent impression, i.e. c = L + a and L is the crack
In the present work, the nanocrystalline hydroxyapatite pow-
length.
der was prepared according to a wet chemical precipitation
method from aqueous medium involving calcium hydroxide
and orthophosphoric acid (Ramesh, 2004). The resulting pre- 3. Results and discussion
cipitate was filtered, washed, dried and then ground to a fine
powder. 3.1. HA powder characteristics
The specific surface area of the powder was measured
by the Brunauer–Emmett–Teller (BET) method. Nitrogen gas The properties of the synthesized HA powder is shown in
adsorption analysis was performed on a Coulter SA 3100 Table 1. The estimates of crystal size from the XRD peak broad-
Analyzer. The calcium-to-phosphorous (Ca/P) ratio of the ening based on Scherrer’s formula (Cullity and Stock, 2001)
synthesized powder was determined by inductively coupled for the HA powders taken at the most prominent peaks were
plasma-atomic emission spectrometry (ICP-AES) analysis.
The as-received HA powder was isostatic pressed at
200 MPa into rectangular and discs samples. These samples
Table 1 – Properties of synthesized HA powders
were then sintered by conventional method and microwave
Ca/P molar ratio 1.67 ± 0.02
technique over the temperature range of 1000–1300 ◦ C.
Colour White
The conventional pressureless sintering was performed
SBET 71.2 m2 /g
using a rapid heating box furnace (ModuTemp, Australia). The
j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230 223

Fig. 1 – Comparison of XRD patterns of synthesized HA with the standard JCPDS card for stoichiometric HA.

10.8 nm at (2 1 1) reflection corresponding to diffraction angle sintering were compared in terms of HA phase stability, rel-
2 = ∼31.9◦ and 31.9 nm at (0 0 2) reflection corresponding to ative density, Vickers hardness, fracture toughness and grain
diffraction angle 2 = ∼25.9◦ . Thus, the HA primary particles size measurement.
consisted of extremely fine precipitates. The XRD analysis of all the samples, regardless of sintering
X-ray diffraction analysis of the synthesized HA powder method, fired between 1000 ◦ C and 1300 ◦ C indicated no sec-
produced only peaks which corresponded to the standard ondary phases were present in the HA lattice. Further evidence
JCPDS for stoichiometric HA as shown in Fig. 1. Secondary of the high temperature phase stability of the material was
phases such as tricalcium phosphate (TCP), tetracalcium obtained by sintering at higher temperature of 1350 ◦ C. The
phosphate (TTCP) and calcium oxide (CaO) were not evident in phase analysis as shown in Fig. 3 for both the MS-HA and CPS-
the synthesized powder. This is consistent with the Ca/P ratio HA compacts showed no sign of HA decomposition, and the
obtained for the powder (Table 1). Additionally, the XRD pat- phases present in the lattice was in good agreement with the
tern of the synthesized powder was in good agreement with JCPDS card number 74-566 for stoichiometric HA. The results
that of enamel of a human tooth as shown in Fig. 2. obtained in this study are comparable to the results reported
by Muralithran and Ramesh (Muralithran and Ramesh, 2000)
3.2. HA phase stability for the effect of sintering temperature on the phase stability
of synthetic HA. These authors found that decomposition of
The sinterability of the HA powder under both sintering condi- HA started at 1400 ◦ C and sign of melting of the material was
tions, the microwave sintering and conventional pressureless observed at 1450 ◦ C.

Fig. 2 – A good agreement was observed between the XRD traces of the synthesized HA powder with that of a human tooth.
224 j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

Fig. 3 – Comparison of XRD patterns between microwave-sintered and conventional pressureless sintered HA at 1350 ◦ C. All
diffraction peaks belong to the HA phase.

In general, high temperature sintering of HA can lead to the sphere resulted in a slight dissociation of the HA phase only
partial thermal decomposition of HA into TCP and/or TTCP. after 8 h of holding at 1300 ◦ C, whereas the stability of the HA
The thermal decomposition is accompanied in two steps, phase was maintained in the structure after sintering for 3 h
i.e. dehydroxylation and decomposition (Wang et al., 1998; at 1450 ◦ C. This observation prompted them to believe that the
LeGeros and LeGeros, 1993). Wang and Chaki (1993) and Gu dissociation of HA occurred by a nucleation and growth mech-
et al. (2002) reported that the dehydroxylation phenomenon anism, involving an incubation period which decreased with
could be observed by comparing the XRD peaks position of increasing sintering temperature.
the sintered material with that of the standard JCPDS data for If an incubation period is one of the factor governing the
stoichiometric HA. decomposition of HA, then the rapid sintering by microwave
In the present work, XRD peak shifting was observed would be beneficial in suppressing the dissociation of the OH
mainly for the conventional pressureless sintered HA com- ions and thus preventing decomposition of the HA phase. This
pacts as typically shown in Fig. 4. The peak shifting for CPS-HA could explained the observation made in the present work
was found to vary significantly between 0.08◦ and 0.16◦ when since no decomposition as well as minimal degree of XRD peak
compared to less than 0.06◦ for MS-HA. shifting was observed upon microwave sintering.
The dehydroxylation phenomenon observed for the CPS-
HA is consistent with the findings of other workers (Wang and 3.3. Bulk density
Chaki, 1993; Gu et al., 2002; Kijima and Tsutsumi, 1979; Santos
et al., 1991; Zhou et al., 1993; Ruys et al., 1995). Although the The effects of CPS and MS on the densification of HA compacts
dehydroxylation phenomenon in HA is still unclear, a general are shown in Fig. 5. In general, both the bulk density variation
consensus that has been reached in the reported literatures of the HA samples sintered by conventional and microwave
is that dehydroxylation neither collapse the HA structure nor methods increases with increasing sintering temperature but
is responsible for the declined in mechanical properties when at different rates. The onset of densification, indicated by a
sintered at elevated temperatures. sharp increase in the sintered density, for both sintering meth-
In the present work, decomposition of HA was not observed ods was between 1000 ◦ C and 1100 ◦ C. Nevertheless, the bulk
in both the conventional-sintered and microwave-sintered density of the CPS-HA was higher than the MS-HA throughout
compacts. The processing route employed to synthesize the the sintering regime investigated.
HA crystallites could have resulted in the retention of absorbed The conventional-sintered material achieved a relative
water in the HA lattice. However, it is not clear if the loss of density of 96% when sintered at 1000 ◦ C and exhibited above
water during sintering plays a role in suppressing dehydroxy- 99% of theoretical density when the temperature increased
lation. to above 1100 ◦ C. In contrast, the microwave-sintered ceramic
Van Landuyt et al. (1995) studied the effects of sinter- exhibited a lower value of 90–91% when sintered at 1000 ◦ C
ing temperature (1200–1500 ◦ C) on the densification behaviour and achieved above 96% of theoretical value when sintered
of stoichiometric HA, and their XRD results showed a simi- at 1100 ◦ C. The maximum value of 98% theoretical density
lar observation to both the CPS-HA and MS-HA produced in was attained for compacts that were microwave-sintered at
this study for temperatures between 1000 ◦ C and 1350 ◦ C. The 1300 ◦ C. These results are in good agreement with that in the
authors found that conventional sintering under air atmo- reported literature for microwave-sintered HA (Agrawal et al.,
j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230 225

Fig. 5 – Comparison between the relative density of MS-HA


and CPS-HA as a function of sintering temperature.

sive of heating, soaking and cooling to room temperature) as a


function of sintering temperature by both sintering methods
to achieve the respective densities.
Taking into consideration of the attained densities based
on the total time of sintering as shown in Table 2, the benefi-
cial effect of microwave heating in promoting densification of
HA can be realized. For example, a relative density of 96% is
achievable by CPS at 1000 ◦ C whereas a higher temperature of
1100 ◦ C was required by microwave sintering to achieve a simi-
lar relative density. However, this small difference of sintering
temperature is offset by the very short sintering time required,
i.e. microwave sintering took a fraction of less than 3% of the
total time taken by the conventional pressureless sintering to
achieve 96% relative density when sintered at 1000 ◦ C.
The densification of ceramics in a solid state sintering is
Fig. 4 – XRD signatures of HA phase taken at the three most
a process of pore elimination, accomplished by a diffusion
prominent diffraction angles for compacts fired at
process through the transfer of matter from the particle vol-
1000–1300 ◦ C. Peak shifting was evident for CPS-HA.
ume or from the grain boundary between particles (Kingery
et al., 1976). Thus, in the present work, it can be suggested
that grain boundary diffusion between the HA particles was
1992; Fang et al., 1994; Yang et al., 2002; Vijayan and Varma, substantially enhanced during the microwave sintering.
2002). Sutton (1989) and Rahaman (1995) explained that volumet-
Although the bulk density of the microwave-sintered HA ric heating in the microwave sintering is a consequence of a
is lower than the conventional pressureless sintered material, series of losses. In general, when microwaves penetrate and
the benefit and effectiveness of microwave sintering can be propagate through a dielectric material, the internal fields
viewed in the context of the time taken to achieve the respec- generated within the affected volume induce translational
tive bulk density shown in Fig. 5. This advantage is shown in motions of the free and bound charges and, rotate charge
Table 2 which compares the total sintering time taken (inclu- complexes such as dipoles. The resistance of these induced

Table 2 – Total sintering time taken to achieve the respective sintered density based on the two different sintering
techniques
Sintering temperature (◦ C) CPS MS

Relative density (%) Time taken (min) Relative density (%) Time taken (min)

1000 96.8 1090 90.8 29.4


1050 98.4 1140 95.2 30.4
1100 99.0 1190 96.5 31.4
1150 99.0 1240 96.7 32.4
1200 99.0 1290 96.6 33.4
1250 99.1 1340 97.0 34.4
1300 99.0 1390 98.0 35.4
226 j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

Fig. 6 – SEM images of (a) CPS-HA and (b) MS-HA surfaces fired at 1000 ◦ C.

motions due to inertial, elastic and frictional forces causes However, as the sintering temperature was increased
losses and in turn attenuates the electric field. As a con- beyond 1150 ◦ C, all the microwave-sintered samples exhibited
sequence of these losses, volumetric heating results. These very little porosity as shown in Fig. 7d–f. The SEM results cor-
principles should be applicable to the microwave sintering of related well with the measured density (Fig. 5), in particular
HA. In particular, the hydroxyl in the HA structure could have for the conventional pressureless sintered HA which showed
played a role in causing losses and thus promoting densifica- that almost full densification was accomplished when com-
tion of HA in a short sintering time. pacts sintered >1050 ◦ C. Increasing the sintering temperature
beyond 1050 ◦ C resulted in grain growth and the CPS-HA exhib-
3.4. Microstructural evolution and grain size ited larger grain sizes than the microwave-sintered HA for the
same sintering temperatures (Fig. 8).
The microstructural evolution of the conventional pressure- In general, the finer grain size clearly indicates that the
less sintering and microwave sintering HA at 1000 ◦ C is shown sintering regime employed for microwave processing did not
in Fig. 6. promote extensive grain growth even when sintered at high
It can be noted that the CPS-HA (Fig. 6a) exhibited an almost temperature of 1350 ◦ C as shown in Fig. 9.
dense microstructure that correspond to a relative density of The variation in average grain sizes with sintering tem-
96.8% if compare to the MS-HA (Fig. 6b) which had remnant perature is presented in Fig. 8. HA samples sintered by the
porosity (∼9%) remaining in the structure after sintering at conventional and microwave methods exhibited a similar
1000 ◦ C. grain size trend with increasing temperature from 1000 ◦ C to

Fig. 7 – Microstructural evolution of CPS-HA (a–c) and MS-HA (d–f) when sintered at 1150 ◦ C (a, d), 1250 ◦ C (b, e) and 1300 ◦ C
(c, f).
j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230 227

Fig. 8 – Average grain size variation with sintering Fig. 10 – Effect of sintering temperature on the fracture
temperature for CPS-HA and MS-HA. toughness of HA sintered by the conventional method
(CPS-HA) and microwave technique (MS-HA).

1200 ◦ C, i.e. the grain sizes of both the CPS-HA and MS-HA
increased slowly with sintering temperature. It is interesting al., 1998; Vaz et al., 1999; Duan et al., 2002). In general, it is per-
to note that in this temperature regime, the rates of grain ceived that the pores provide a mechanical interlock leading to
growth of both conventional-sintered and microwave-sintered firmer fixation of the material (Suchanek and Yoshimura, 1998;
HA were almost similar. However, as the temperature was Flautre et al., 2001). In addition, bone tissue grows well into the
increased further, above 1200 ◦ C, the rate of grain growth for pores and thus increasing the strength of the HA implant (Hing
the conventional pressureless sintered HA increased rapidly et al., 1997; Lu et al., 1999; Rivera-Munoz et al., 2001).
as compared to the microwave-sintered ceramic. The grain
size of the CPS-HA increased by a factor of >5 from 1.54 ␮m 3.5. Fracture toughness and Vickers hardness
(1200 ◦ C) to 8.78 ␮m (1300 ◦ C) whereas the grain size of MS-HA
increased by a factor of about 2.4 from 0.86 ␮m to 2.08 ␮m when The effect of sintering temperature on the fracture tough-
the temperature increased by 100 ◦ C from 1200 ◦ C. ness (KIc ) of synthesized HA sintered by the conventional
A similar observation was noted for the HA sintered at and microwave technique is shown in Fig. 10. The results
1350 ◦ C as shown in Fig. 9. The measured grain sizes for the show that the fracture toughness of the HA sintered by
ceramics sintered at 1350 ◦ C were 12.4 ␮m for the CPS-HA the conventional method decreases almost linearly with
and 2.77 ␮m for the MS-HA, respectively. These results clearly increasing sintering temperature, i.e. from a maximum
demonstrate another useful advantage of microwave sintering value of 1.22 ± 0.08 MPam1/2 at 1000 ◦ C to a minimum of
in preventing grain growth when sintered at high tempera- 0.91 ± 0.05 MPam1/2 for sintering at 1300 ◦ C. On the other hand,
tures. This is important, since material with smaller grain the average toughness of microwave-sintered HA was found
size would normally result in enhanced mechanical proper- initially to increase with density and grain size from an
ties when compared to the same material having a larger grain average value of 1.24 MPam1/2 to 1.45 MPam1/2 and thereafter
size. decrease almost linearly with further increase in grain size,
Although a fully dense body was not attained in the present i.e. down to 0.96–1 MPam1/2 when sintered at 1300 ◦ C despite
work, a useful HA body that has the potential for clinical having high relative density of 98%.
implant has been produced by microwave sintering in the The minimum fracture toughness obtained for both the
shortest possible sintering time. The low degree of porosity, i.e. conventional-sintered and microwave-sintered HA was in
<10%, in the microwave-sintered HA compacts in the present agreement with the values reported in the literature for syn-
work would be an added benefit when implanted in living thetic HA samples sintered up to 1300 ◦ C (Muralithran and
body. According to the literature, HA ceramics in porous form Ramesh, 2000; Van Landuyt et al., 1995; De With et al., 1981). In
(porosity up to 60%) has been widely used successfully as bone general, the maximum KIc values for most HA reported in the
substitutes (Akao et al., 1981; Hench, 1991; Liu, 1997; Yuan et literature varied between 0.96 MPam1/2 and 1 MPam1/2 . Thus,

Fig. 9 – SEM micrographs of HA sintered at 1350 ◦ C by (a) conventional pressureless sintering and (b) microwave sintering.
228 j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

Fig. 11 – Vickers hardness of HA sintered at various Fig. 12 – Variation in hardness with sintered density for
temperatures by the conventional and microwave methods. MS-HA and CPS-HA compacts.

the high values of 1.22 MPam1/2 and 1.45 MPam1/2 obtained in The results clearly show the effect of sintered density
the present work for CPS-HA and MS-HA, respectively were on the hardness of HA ceramics, and that the hardness of
encouraging. It is believed that this improvement in fracture CPS-HA was greater than MS-HA for lower sintering tempera-
toughness in the present HA could be attributed to both the tures, where it achieves higher sintered densities. The results
improved properties of the synthesized HA powder as well as obtained in this study however were not in agreement with the
the improved sinterability of these powders with limited grain general trend of hardness increases with increasing sintering
growth during microwave sintering. temperature as reported by other workers (Van Landuyt et al.,
The effect of sintering temperature on the Vickers hard- 1995; Gibson et al., 2001; Lu et al., 1998; Yeong et al., 2001).
ness of HA sintered by both the conventional method and In the present study, it was found that the hardness of
microwave technique is shown in Fig. 11. The hardness of both both HA decreased with increasing sintering temperature
HA increases with sintering temperature and peaked at a cer- beyond the point at which maximum hardness was obtained
tain value at a particular temperature and then decreases with (i.e. >1050 ◦ C for conventional pressureless sintered HA and
further increase in the temperature up to 1300 ◦ C. For sintering >1150 ◦ C for microwave-sintered HA as shown in Fig. 12),
below 1150 ◦ C, the hardness of HA sintered by the conventional although the densities of these compacts above this tem-
pressureless sintering method was higher than the equivalent perature range were still high, i.e. 99% for the CPS-HA and
microwave-sintered material. However, sintering at 1150 ◦ C above 96% for MS-HA. This decreased in hardness observed for
and above, both the CPS-HA and MS-HA exhibited a similar both the CPS-HA and MS-HA when sintered above 1050 ◦ C and
trend with almost identical hardness values being attained at 1150 ◦ C, respectively could not be due to bulk density effect
the same temperature. This is an advantage since it is possible but rather can be interpreted as due to a grain size effect as
to attain comparable hardness by microwave heating the HA shown in Fig. 13.
ceramic in minutes rather than in hours as in the conventional As can be noted from Fig. 13, the hardness of both the
pressureless sintering method. CPS-HA and MS-HA increases with grain size and reached
The lowest hardness value was measured for HA compacts a maximum value at a certain grain size limit (dc ) before
sintered at 1000 ◦ C in both the conventional and microwave decreasing further with increasing grain size resulting from
sintering methods. The conventional-sintered HA exhibits sintering at higher temperatures. The results obtained in the
a large increase in hardness from 5.72 ± 0.21 GPa at 1000 ◦ C
to a maximum value of 7.21 ± 0.17 GPa at 1050 ◦ C. Further
sintering, however was detrimental as the hardness of the
CPS-HA decreased almost linearly with increasing temper-
ature. On the other hand, the hardness of MS-HA start to
increase sharply between 1000 ◦ C and 1050 ◦ C and then grad-
ually between 1050 ◦ C and 1150 ◦ C to attain a maximum value
of 6.38 ± 0.20 GPa at 1150 ◦ C as shown in Fig. 11.
Based on the density results in Fig. 5, the CPS-HA exhib-
ited a rapid increase in relative density from 96.8% at 1000 ◦ C
to 98.4% and 99% at 1050 ◦ C and 1100 ◦ C, respectively. Simi-
larly, the relative density of the MS-HA increased from 90.8%
at 1000 ◦ C to 95.2% and 96.7% at 1050 ◦ C and 1150 ◦ C, respec-
tively. This increase in density in both the CPS-HA and MS-HA
is clearly reflected in the improved hardness with increasing
sintering upto 1050 ◦ C and 1150 ◦ C, respectively as shown in Fig. 13 – Vickers hardness dependence on the grain size of
Fig. 12. hydroxyapatite.
j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230 229

present work suggest that below some critical grain size of dc Duan, Y., Wang, C., Cheng, J., Zhang, X., 2002. Bone-like apatite
the hardness is governed by bulk density. In contrast, above formation in intramuscularly implanted calcium phosphate
dc the bulk density is not the controlling parameter but rather ceramics in different kinds of animal. J. Mater. Sci. Lett. 21,
775–778.
grain growth. This is evident from Fig. 13, where the dc for the
Fang, Y., Agrawal, D.K., Roy, D.M., Roy, R., 1994. Microwave
conventional-sintered (CPS-HA) and microwave-sintered (MS- sintering of hydroxyapatite ceramics. J. Mater. Res. 9, 180–
HA) hydroxyapatite are ∼0.5 ␮m and ∼0.2 ␮m, respectively. 187.
Flautre, B., Descamps, M., Delecourt, C., Blary, M.C., Hardouin, P.,
2001. Porous HA ceramic for bone replacement: role of the
4. Conclusions pores and interconnections–experimental study in the
rabbits. J. Mater. Sci.: Mater. Med. 12, 679–
The sinterability of nanocrystalline HA by microwave sintering 682.
was studied and compared with that of conventional pressure- Gabriel Chu, T.-M., Orton, D.G., Hollister, S.J., Feinberg, S.E.,
Halloran, J.W., 2002. Mechanical and in vivo performance of
less sintering method. The XRD patterns of the microwave
hydroxyapatite implants with controlled architectures.
sintering and conventional pressureless sintering HA com- Biomaterials 23, 1283–1293.
pacts that were fired between 1000 ◦ C and 1300 ◦ C, respectively Gibson, I.R., Ke, S., Best, S.M., Bonfield, W., 2001. Effect of powder
corresponded to that of stoichiometric HA. Decomposition characteristics on the sinterability of hydroxyapatite powders.
of HA phase was not observed throughout the sintering J. Mater. Sci.: Mater. Med. 12, 163–171.
regime employed. However, substantial grain coarsening was Gu, Y.W., Loh, N.H., Khor, K.A., Tor, S.B., Cheang, P., 2002. Spark
plasma sintering of hydroxyapatite powders. Biomaterials 23,
observed in the conventional-sintered HA when compared to
37–43.
compacts sintered via microwave method, particularly for sin-
Hench, L.L., 1991. Bioceramics: from concept to clinical. J. Am.
tering above 1200 ◦ C. Ceram. Soc. 74, 1487–1510.
The fracture toughness values obtained for CPS-HA was Hench, L.L., 1998. Biomaterials: a forecast for the future.
lower compared to MS-HA, with maximum average value Biomaterials 19, 1419–1423.
of 1.45 MPam1/2 being attained for microwave compacts at Hing, K.A., Best, S.M., Tanner, K.E., Revell, P.A., Bonfield, W., 1997.
1050 ◦ C. The maximum average hardness of 7.21 GPa and Biomechanical assessment of bone ingrowth in porous
hydroxyapatite. J. Mater. Sci.: Mater. Med. 8, 731–736.
6.38 GPa was measured for CPS-HA and MS-HA sintered at
Kijima, T., Tsutsumi, M., 1979. Preparation and thermal properties
1050 ◦ C and 1150 ◦ C, respectively. of dense polycrystalline HA. J. Am. Ceram. Soc. 62, 455–460.
Regardless of sintering methods, there is a definite correla- Kingery, W.D., Bowen, H.K., Uhlmann, D.R., 1976. Introduction to
tion between hardness and grain size in the present HA. The Ceramics, second ed. John Wiley & Sons, Inc., Singapore, p.
hardness of HA would start to decrease at a certain critical 448.
grain size limit despite exhibiting high bulk density. Kokubo, T., Kim, H.-M., Kawashita, M., 2003. Novel bioactive
materials with different mechanical properties. Biomaterials
24, 2161–2175.
Acknowledgements LeGeros, R.Z., LeGeros, J.P., 1993. Dense hydroxyapatite. In:
Hench, L.L., Wilson, J. (Eds.), An Introduction to Bioceramics.
World Scientific, Singapore, p. 139 (Chapter 9).
This work was supported by the Ministry of Science, Technol- Liu, D.-M., 1997. Fabrication of hydroxyapatite ceramic with
ogy and Innovation of Malaysia (Grant No. 03-02-03-SF0073). controlled porosity. J. Mater. Sci.: Mater. Med. 8, 227–232.
The authors would like to thank Dr. M.G. Kutty and J. Jok- Lu, H., Qu, Z., Zhou, Y., 1998. Preparation and mechanical
isaari for technical support, Clemson University for microwave properties of dense polycrystalline hydroxyapatite through
facility, SIRIM and TNBR for providing support with testing freeze-drying. J. Mater. Sci.: Mater. Med. 9, 583–587.
Lu, J.X., Flautre, B., Anselme, K., Hardouin, P., Gallur, A.,
equipment.
Descamps, M., Thierry, B., 1999. Role of interconnections in
porous bioceramics on bone recolonization in vitro and in
references vivo. J. Mater. Sci.: Mater. Med. 10, 111–120.
Mendelson, M.I., 1969. Average grain size in polycrystalline
ceramics. J. Am. Ceram. Soc. 52, 443–446.
Metsger, D.S., Driskell, T.D., Paulsrud, J.R., 1982. Tricalcium
Afshar, A., Ghorbani, M., Ehsani, N., Saeri, M.R., Sorrell, C.C., 2003. phosphate ceramic—a resorbable bone implant: review and
Some important factors in the wet precipitation process of current status. J. Am. Dent. Assoc. 105, 1035–1038.
hydroxyapatite. Mater. Des. 24, 197–202. Muralithran, G., Ramesh, S., 2000. The effects of sintering
Agrawal, D.K., Fang, Y., Roy, D.M., Roy, R., 1992. Fabrication of temperature on the properties of hydroxyapatite. Ceram.
hydroxyapatite ceramics by microwave processing. Mater. Inter. 26, 221–230.
Res. Soc. Symp. Proc. 269, 231–236. Niihara, K., 1985. Indentation microfracture of ceramics—its
Akao, M., Aoki, H., Kato, K., 1981. Mechanical properties of application and problems. Ceram. Jap. 20, 12–18.
sintered hyroxyapatite for prosthetic applications. J. Mater. Rahaman, M.N., 1995. Densification process variables and
Sci. 16, 809–812. densification practice. In: Ceramic Processing and Sintering.
Aoki, H., Akao, M., Shin, Y., Tsuji, T., Togawa, T., 1987. Sintered Marcel Dekker, Inc., New York, pp. 725–734.
hydroxylapatite for percutaneous devices and its clinical Ramesh, S., A method for manufacturing hydroxyapatite
application. Med. Prog. Tech. 12, 213. bioceramic, Malaysia Patent No. PI. 20043325 (2004).
Cullity, B.D., Stock, S.R., 2001. Elements of X-Ray Diffraction, third Rivera-Munoz, E., Diaz, J.R., Rodriguez, J.R., Brostow, W., Castano,
ed. Prentice Hall Inc., New Jersey, pp. 167–170. V.M., 2001. Hydroxyapatite spheres with controlled porosity
De With, W., Van Dijk, J.A., Hattu, H.N., Prijs, K., 1981. Preparation, for eye ball prosthesis: processing and characterization. J.
microstructure and mechanical properties of dense Mater. Sci.: Mater. Med. 12, 305–311.
polycrystalline hydroxyapatite. J. Mater. Sci. 16, 1592–1598.
230 j o u r n a l o f m a t e r i a l s p r o c e s s i n g t e c h n o l o g y 2 0 6 ( 2 0 0 8 ) 221–230

Rodrı́guez-Lorenzo, L.M., Vallet-Regı́, M., Ferreira, J.M.F., Ginebra, Vaz, L., Lopes, A.B., Almeida, M., 1999. Porosity control of
M.P., Aparicio, C., Planell, J., 2002. A hydroxyapatite ceramic hydroxyapatite implants. J. Mater. Sci.: Mater. Med. 10,
bodies with tailored mechanical properties for different 239–242.
applications. J. Biomed. Mater. Res. 60, 159–166. Vijayan, S., Varma, H., 2002. Microwave sintering of nanosized
Ruys, A.J., Wei, M., Sorrell, C.C., Dickson, M.R., Brandwood, A., hydroxyapatite powder compacts. Mater. Lett. 56, 827–831.
Milthorpe, B.K., 1995. Sintering effect on the strength of Wang, P.E., Chaki, T.K., 1993. Sintering behaviour and mechanical
hydroxyapatite. Biomaterials 16, 409–415. properties of hydroxyapatite and dicalcium phosphate. J.
Santos, J.D., Morrey, S., Hastings, G.W., Monteiro, F.J., 1991. The Mater. Sci: Mater. Med. 4, 150–158.
production and characterisation of a hydroxyapatite ceramic Wang, C.K., Ju, C.P., Chern Lin, J.H., 1998. Effect of doped bioactive
material. In: Bonfield, W., Hastings, G.W., Tanner, K.E. (Eds.), glass on structure and properties of sintered hydroxyapatite.
Proceedings of the 4th International Symposium on Ceramics Mater. Chem. Phys. 53, 138–149.
in Medicine. vol. 4. Butterworth-Heinemann, London, pp. Yang, Y., Ong, J.L., Tian, J., 2002. Rapid sintering of hydroxyapatite
71–78. by microwave processing. J. Mater. Sci. Lett. 21, 67–69.
Suchanek, W., Yoshimura, M., 1998. Processing and properties of Yeong, K.C.B., Wang, J., Ng, S.C., 2001. Mechanochemical
hydroxyapatite-based biomaterials for use as hard tissue synthesis of nanocrystalline hydroxyapatite from CaO and
replacement implants. J. Mater. Res. 13, 94–117. CaHPO4 . Biomaterials 22, 2705–2712.
Sutton, W.H., 1989. Microwave processing of ceramic materials. Yuan, H., Yang, Z., Li, Y., Zhang, X., DeBruijn, J.D., DeGroot, K.,
Am. Ceram. Soc. Bull. 68, 376–386. 1998. Osteoinduction by calcium phosphate biomaterials. J.
Van Landuyt, P., Li, F., Keustermans, J.P., Streydio, J.M., Delannay, Mater. Sci.: Mater. Med. 9, 723–726.
F., Munting, E., 1995. The influence of high sintering Zhou, J., Zhang, X., Chen, J., Zeng, S., DeGroot, K., 1993. High
temperatures on the mechanical properties of temperature characteristics of synthetic hydroxyapatite. J.
hydroxylapatite. J. Mater. Sci.: Mater. Med. 6, 8–13. Mater. Sci: Mater. Med. 4, 83–85.

You might also like