Dielectric and Electrical Conductivity Properties of Multi-Stage Spark Plasma PDF

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Available online at www.sciencedirect.

com

Journal of the European Ceramic Society 33 (2013) 3445–3453

Dielectric and electrical conductivity properties of multi-stage spark plasma


sintered HA–CaTiO3 composites and comparison with conventionally
sintered materials
Ashutosh K. Dubey 1 , P.K. Mallik 1 , S. Kundu, B. Basu ∗
Materials Research Centre, Indian Institute of Science, Bangalore 560012, India
Received 13 March 2013; received in revised form 5 July 2013; accepted 13 July 2013
Available online 8 August 2013

Abstract
One of the different issues limiting the wider application of monolithic hydroxyapatite (HA) as an ideal bone replacement material is the lack
of reasonably good electrical transport properties. The comprehensive electrical property characterization to evaluate the efficacy of processing
parameters in achieving the desired combination of electroactive properties is considered as an important aspect in the development of HA-based
bioactive material. In this perspective, the present work reports the temperature (RT-200 ◦ C) and frequency (100 Hz–1 MHz) dependent dielectric
properties and AC conductivity for a range of HA–CaTiO3 (HA–CT) composites, densified using both conventional pressureless sintering in air as
well as spark plasma sintering in vacuum. Importantly, the AC conductivity of spark plasma sintered ceramics [∼upto 10−5 ( cm)−1 ] are found to
be considerably higher than the corresponding pressureless sintered ceramics [∼upto 10−8 ( cm)−1 ]. Overall, the results indicate the processing
route dependent functional properties of HA–CaTiO3 composites as well as related advantages of spark plasma sintering route.
© 2013 Elsevier Ltd. All rights reserved.

Keywords: Hydroxyapatite; Calcium titanate; Spark plasma sintering; Dielectric properties

1. Introduction as CaTiO3 and BaTiO3 attracted major scientific attention in


the context of the development of electroactive biomaterials.
Hydroxyapatite (HA) is well known as a highly biocompat- Among these, CaTiO3 (CT) is reported as a good substrate for
ible and bioactive ceramic that promotes bone growth in vivo.1 apatite growth with the potential to enhance osseointegration
However, poor electrical and mechanical properties limit the and osteoconduction.12,13
application of monolithic HA as an implant material for Despite significant bioactivity of HA, the development of
orthopaedic applications. The addition of ceramic/metallic rein- HA-based electrically active materials can provide the improved
forcement (mullite, Ti) has been studied extensively to improve functional response of the prosthetic implants. Also, reasonable
the mechanical properties of HA.2–6 It is known that the natu- substrate conductivity facilitates better cell-to-cell signalling
ral bone is a piezoelectric material in vivo that generates the processes involved in various cell fate processes (prolifera-
electric potential upon the application of mechanical stress, tion, differentiation etc.). In this context, our research group
which helps in bone metabolism and growth.7,8 To this end, recently reported the substrate conductivity dependent mod-
it has been demonstrated that the electrical charges on bio- ulation of myoblast proliferation and differentiation on spark
material surfaces can enhance the osteobonding.9,10 The new plasma sintered HA–CaTiO3 composites in vitro.14 The present
bone formation has been found to be significantly improved on work focuses on the evaluation of dielectric and electrical prop-
negatively charged surfaces.11 Because of high polarizability erties for a range of HA–CaTiO3 composites, synthesized using
and reasonable biocompatibility, few of the perovskites, such innovative multi stage spark plasma sintering (SPS) as well as
conventional pressureless sintering routes.
The main advantage of using SPS technique over conven-
∗ Corresponding author. Tel.: +91 80 2293 3256.
tional sintering route is to activate the different densification
E-mail address: bikram@mrc.iisc.ernet.in (B. Basu). mechanisms by enhancing the mass transport and diffu-
1 These authors contributed equally to this work. sion, which considerably reduces the processing time and

0955-2219/$ – see front matter © 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.jeurceramsoc.2013.07.012
3446 A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453

Table 1
Multi-stage spark plasma sintering conditions and relative density of HA–CT
composites. All the powder compacts are held at each temperature for 5 min.
The pressure applied during sintering was 50 MPa, except for HA (30 MPa).
Composition Sample Sintering temperature (◦ C) Sinter density
designation (% ρth )

Pure HA HA 850 ◦ C/950 ◦ C/1100 ◦ C 99.4


Pure CT CT 950 ◦ C/1100 ◦ C/1200 ◦ C 98.2
HA-20 wt% HA20CT 950 ◦ C/1100 ◦ C/1200 ◦ C 99.3
CaTiO3
HA-40 wt% HA40CT 950 ◦ C/1100 ◦ C/1200 ◦ C 99.0
CaTiO3
HA-60 wt% HA60CT 950 ◦ C/1100 ◦ C/1200 ◦ C 99.3
CaTiO3
HA-80 wt% HA80CT 950 ◦ C/1100 ◦ C/1200 ◦ C 98.6
CaTiO3

Fig. 1. The heating profile of the HA–CaTiO3 powder compact during spark
plasma sintering. at 1400 ◦ C for 4 h in air with heating and cooling rates of
5 ◦ C/min, while pure HA was sintered in air at 1200 ◦ C for
temperature.15 The multistage SPS route, which involves hold- 2 h. Finally, all the HA–CT sintered samples were mirror
ing the powder compact at one/two intermediate temperatures polished for microstructural, dielectric and electrical character-
prior to the final sintering temperature, has been demonstrated ization.
to produce dense oxide and non-oxide ceramics.16–18 It can be The bulk density of the HA–CT composites was measured
reiterated here that in the context of densification of HA-based using Archimedes’s principle. The theoretical density of the
composites, SPS process inhibits grain growth, decomposi- composites has been calculated assuming the absence of any
tion of HA phase as well as restricts the sintering reactions reaction between the constituent phases. In order to deter-
between HA and reinforcement phase.19 Further, the dielec- mine the presence of different phases due to the dissociation
tric and electrical properties of spark plasma sintered HA–CT of HA or the reaction among constituent phases, the sintered
composites have been compared with those of the pressure- HA–CT composites were analyzed using XRD (Brucker Xpert
less sintered HA–CT composites to demonstrate the efficacy diffractometer) with Cu-K␣ radiation (λ = 1.54 Å). ´ The data
of SPS route in achieving a better combination of functional ◦
were collected over the 2θ range of 20–80 with a scan speed
properties. of 0.5◦ per min and a step size of 0.02◦ . The microstructure
of the etched samples was observed using scanning electron
2. Experimental microscope (SEM, FEI Quanta 200). The SPSed samples were
chemically etched using anhydrous citric acid and distilled water
2.1. Materials processing and microstructural in the ratio of 8:10 (by mass) with 2–3 drops of hydrofluoric
characterization acid (HF) for about 5 min. The conventionally sintered sam-
ples were thermally etched at a temperature of 100 ◦ C lower
The starting powders of HA [Ca10 (PO4 )6 (OH)2 ] and crys- than the respective sintering temperature. For transmission elec-
talline calcium titanate (CaTiO3 ) were synthesized using wet tron microscopy (TEM) analysis, electron transparent ultrathin
precipitation route 20 and mechanical activation of a mixture samples of 0.1 ␮m or lower thickness were machined using ultra-
of CaO and TiO2 (anatase), respectively.21 A range of HA–CT sonic disc cutter, which was then followed by dimpling and
composites were prepared by adding the varying amounts of precision ion polishing (PIPS, Gatan 691). The TEM (Tecnai
CaTiO3 (0, 20, 40, 60, 80,100 wt%) to HA. The powder mix- G2 ) was used at 200 kV.
ture with desired HA: CaTiO3 ratio was ball milled (Fritsch, In order to analyze the surface chemistry, selected SPSed
Pulverisette 1583, Germany) for 16 h using agate balls and ceramics were examined using X-ray photoelectron spec-
jars as grinding media, which allows homogeneous mixing as troscopy (XPS, Thermo Scientific, model – Multilab 2000).
well as reduces the initial particle size. The innovative multi- The measurement was performed using Al-K␣ radiation
stage spark plasma sintering technique (MSSPS) (Dr. Sinter, (1486.6 eV). The binding energies were calibrated with respect
Model 515S, SPS syntax Inc., Japan) was adopted to con- to C (1s) at 285 eV with a precision of (±0.1 eV). XPS results
solidate the composite powders. The powder compact was were analyzed on the basis of binding energies of Ti2p, O1s and
held for 5 min at each of the temperatures of 950 ◦ C, 1100 ◦ C Ca2s peaks.
and 1200 ◦ C, respectively, which was then followed by fur-
nace cooling (Fig. 1). However, the holding temperatures for 2.2. Dielectric and electrical characterization
pure HA were 850 ◦ C, 950 ◦ C and 1100 ◦ C, respectively with
similar holding time. The heating rate was 100 ◦ C/min. The For dielectric and electrical measurements, the polished
pressureless sintering of HA–CT composites was performed samples of thickness 1–2 mm were coated with silver paste
A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453 3447

samples were heated at a heating rate of 2 ◦ C/min. The dielec-


tric constant (εr ) was calculated using the following formula,

C×d
εr =
ε0 A
where C, d and A are the capacitance, thickness and area of
the sample, respectively and ε0 is the dielectric constant of the
free space (=8.854 × 10−12 F m−1 ). The dissipation factor (D)
has been recorded directly from the instrument. AC conductivity
(σ ac ) was calculated using the following relationship,
G×d
σac = ( cm)−1
A
where G (=ωCD), d and A are the conductance, thickness and
area of the sample, respectively.
Fig. 2. XRD analysis of the spark plasma sintered HA–CT composites as well
as pure HA and CT. 3. Results

followed by curing, which acts as electrodes. The capacitance 3.1. Densification and microstructure
and dielectric loss were measured as a function of frequency
(100 Hz–1 MHz) from room temperature (RT) to 200 ◦ C by The sinter density data and the respective sintering condi-
precision impedance analyzer (Wayne Kerr 6500B, UK). The tions are provided in Table 1. It is clear that multi-stage SPS

Fig. 3. SEM images of polished and chemically etched microstructure of spark plasma sintered ceramics (a) pure CT, (b) HA80CT, and (c) pressureless sintered
HA40CT composite (thermally etched).
3448 A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453

with total sintering duration of less than 30 min can result in


about 99% of theoretical density (ρth ) in the HA–CT composites
as well as in HA monolith. However, the pressureless sintering
of HA–CT composites revealed the maximum density of about
94% ρth at the optimal sintering temperature of 1400 ◦ C for
4 h.13 From the above, the efficacy of multi-stage SPS in den-
sifying HA–CT composites can be realized. Fig. 2 shows the
results of XRD analysis for all the SPSed samples. The traces
of TiO2 were found in the sintered samples, which might be due
to excess TiO2 , left during the mechanochemical synthesis of
CaTiO3 . Importantly, the XRD pattern of the SPSed samples
did not show any detectable presence of ␤-tricalcium phosphate
(␤-TCP). In all the composites, HA and CT are retained after
SPS processing. In contrast, the partial dissociation of HA to
␤-TCP as well as formation of other secondary phases, such as
CaO and TiO2 were observed in pressureless sintered HA–CT
composites.13 The pure HA generally decomposes into TCP
at temperature above 1300 ◦ C in air.22 However, the addition
of secondary phases as well as sintering in vacuum (such as
SPS) reduces the decomposition temperature of HA. The SPS
route inhibited the decomposition of HA into ␤-TCP due to
faster heating rate and lower sintering temperature. From SEM
images of polished and chemically etched spark plasma sin-
tered composites, as shown in Fig. 3(a) and (b), the presence of
grain sizes of 1–2 ␮m are observed in CT and HA80CT ceram-
ics. However, the SEM image [Fig. 3(c)] of thermally etched
microstructure of conventionally sintered HA40CT composite
revealed the biomodal distribution with grain sizes, character-
ized by coarser grains of 2–4 ␮m and finer grains of 1–2 ␮m.
The finer scale microstructural characteristic of HA–CT com-
posite was investigated using bright field (BF) TEM images.
Fig. 4 represents BF images of HA40CT and HA80CT compos-
ites. The predominant presence of HA and CaTiO3 is confirmed
using SAED pattern analysis of the microstructural phases (not
shown). The BF image acquired from CT grain in HA40CT com-
Fig. 4. Representative bright field TEM image of spark plasma sintered (a)
posite shows the presence of characteristics twins [see inset of
HA40CT and (b) HA80CT composites. The inset in (a) shows twinning mor-
Fig. 4(a)]. Irrespective of CT content, such twinning morphol- phology.
ogy is observed in all the HA–CT composites. It is also reported
that the twin width scales with grain size in HA-CaTiO3 sys-
tem. Considering SPS conditions, these twins are believed to of HA80CT, the dielectric relaxation seems to appear beyond
be deformation twins. It has been reported that the domain/twin our measuring frequency range. The other compositions show a
structure is a consequence of transformation from the parae- small variation in dielectric constant in the measured range of
lastic phase to the ferroelastic phase.23 The above observation temperature and frequency. In contrast, HA20CT, HA40CT and
therefore suggests that the CaTiO3 grains might be ferroelastic HA60CT composites follow a similar trend of decrease in the
in nature in HA–CT composites. dielectric loss at frequency below 1 kHz. At higher frequencies
(>10 kHz), almost frequency independent dielectric loss values
3.2. Dielectric properties are obtained.
Fig. 6 shows the variation of dielectric constant, dielectric
The variation of dielectric constant and loss with tempera- loss with temperature (RT-150 ◦ C) at a frequency of 1 MHz
ture and frequency for SPSed HA–CT composite samples are for pressureless sintered HA and HA–CaTiO3 composites.
shown in Figs. 5(a)–(d). The dielectric constant and loss val- Because of the differences in phase assemblage between spark
ues for pure CT and HA80CT are higher compared to pure HA plasma sintered and pressureless sintered samples, the different
and other grades of HACT composites. However, the dielectric polarization phenomena such as space charge and orientational
dispersion with frequency for CT and HA80CT suggested the polarizations also contribute at lower frequencies. Therefore, the
non-intrinsic origin of dielectric constants in these samples. The dielectric behaviour of the developed composites is compared
dielectric dispersion in CT is followed by the corresponding peak at 1 MHz frequency. Table 2 shows the comparison of dielectric
in loss curve, which represents the dielectric relaxation. In case behaviour for multistage spark plasma and pressureless sintered
A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453 3449

Fig. 5. The frequency and temperature dependent dielectric constant (a) and (b) and dielectric loss (c) and (d) for spark plasma sintered HA–CT composites as well
as pure CT and HA. A comparison of dielectric constant of HA–CT composites with Wiener bounds and logarithmic mixture rule at 1 MHz of frequency and room
temperature is also shown (e).

Table 2
Comparison of dielectric constant, loss and AC conductivity (measured at 27 ◦ C and 1 MHz) of pressureless (PS) and multistage spark plasma (SPS) sintered HA–CT
composites.
Composition Dielectric constant Dielectric loss AC conductivity ( cm)−1

PS SPS PS SPS PS SPS

HA 11.1 8.9 6 × 10−5 0.008 – –


HA-40 wt% CaTiO3 22.4 24.9 10−4 0.004 3.5 × 10−10 1.2 × 10−7
HA-60 wt% CaTiO3 34.2 22.4 0.004 0.007 8.0 × 10−8 3.2 × 10−7
HA-80 wt% CaTiO3 95.5 155.0 0.02 0.780 8.0 × 10−8 4.0 × 10−5
CaTiO3 ∼170(33) 490 ∼10−4(34) 1.006 ∼10−9(34) 10−3

HA–CT composites at room temperature (27 ◦ C) and at 1 MHz dielectric constant value in the pressureless sintered composites
frequency. Similar to the SPSed samples, the pressureless does not vary significantly in the measured temperature range.
sintered HA80CT composite shows the significantly higher In contrast, the dielectric loss for these samples increases with
dielectric constant (95.5) and loss (0.02) values as compared an increase in temperature. However, for HA80CT, it appears
to other pressureless sintered composites. Except HA80CT, the to be constant in the temperature range of 50–150 ◦ C.
3450 A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453

Fig. 6. Variation of dielectric constant (a) and dielectric loss (b) with temperature for pressureless sintered HA–CT composites. A comparison of dielectric constant
of HA–CT composites with Wiener bounds and logarithmic mixture rule at 1 MHz of frequency and room temperature is also shown (c).

From Figs. 5 and 6, the dielectric behaviour of spark plasma as highest value of conductivity due to high dissipation factor,
well as pressureless sintered HA and HA40CT samples appears which increases further with increase in temperature.
to be comparable. However, pressureless sintered HA60CT
ceramic shows a higher dielectric constant value as compared to 4. Discussion
spark plasma sintered HA60CT composite. In contrast to spark
plasma sintered samples, the phase analysis of pressureless sin- The dielectric properties of the composites can be explained
tered HA60CT revealed the formation of secondary phases, such in the light of the existing theoretical models as well as in refer-
as CaO and ␤-TCP.13 The significantly higher values of dielec- ence to the structural characteristics of the constituent ceramic
tric constant and loss have been obtained for SPSed HA80CT phases. HA is an ionic crystal with hydroxyl ion (OH− ) sublat-
composite as compared to the pressureless sintered HA80CT tice as a series of one dimensional chain in the crystal lattice.
sample. There are two types of calcium sublattices in HA lattice with one
having three fold rotational axis and other having equilateral
triangles about the c-axis with OH− ions at the centre of the
3.3. AC conductivity property triangles. The ionic transport within the HA structure occurs by
OH− ion vacanc ies and diffusion of Ca2+ and PO4 3− ions under
Fig. 7 shows the variation of AC conductivity with temper- the influence of applied electric field and temperature. This
ature at frequency of 1 MHz for SPSed as well as pressureless causes a relative displacement of OH− ions with centre of posi-
sintered HA–CT composites. While the conductivity of SPSed tive (Ca2+ ) and negative (PO4 3− ) ions forming a chain segment.
HA linearly increases in the measured temperature range, Consequently, the displacement of OH− ions within HA lattice
the conductivity of HA20CT and HA40CT samples appear creates a dipole parallel to the c-axis, which is greater than the
to be independent of temperature. For HA60CT, the change displacement of either Ca2+ or PO4 3− ions. However, the parallel
in conductivity with temperature is more significant for pres- and perpendicular displacement of PO4 3− ions along the c-axis
sureless sintered sample. The comparison of AC conductivity is of the order of magnitude smaller than that of Ca2+ ions. All
for pressureless and multistage spark plasma sintered HA–CT these contribute towards the randomness of displacement in the
composites are shown in Table 2. For a given composite calcium sublattice as well as in structure of HA lattice.24 From
composition, the conductivity of the SPSed samples is found the above, it can be realized that the polarization of HA depends
to higher than the respective pressureless sintered samples. on the number of lattice sites occupied by the OH− vacancies,
The spark plasma sintered HA80CT ceramic demonstrated the applied electric field, temperature, charge on the vacancy and the
A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453 3451

plasma sintering in vacuum. It is expected that [V]/[O] of SPSed


HA would be more than that in a conventional sintered HA in
ambient atmosphere. As a result, dielectric loss of SPSed HA is
more than that sintered in air (see Table 2).
Further, the experimental values of dielectric constant were
compared with the theoretical estimations using parallel and
series Wiener bounds 27 as well as the logarithmic mixture
rule,28

εc = υHA εHA + υCT εCT (3)

1 υ υ
= HA + CT (4)
εc εHA εCT

log εc = υHA log εHA + υCT log εCT (5)

where εc is the dielectric constant of the HA–CT composite,


υHA , υCT , εHA and εCT are the volume fractions and dielectric
constants of HA and CT, respectively. From Figs. 5(e) and 6(c),
it can be concluded that independent of the processing route,
the dielectric constant of the composites follow closely with the
series Weiner bound for CT content of ≤60 wt% in HA. One of
the possible reasons for this resemblance can be the lack of con-
nectivity among the high dielectric constant CaTiO3 phases.29
Also, the secondary phases formed (e.g., CaO, TiO2 , ␤-TCP
etc.) during the conventional processing of these composites
contribute to the effective dielectric constant of the composite.
The difference in conductivity property will now be
explained. With an increase in temperature, the dehydration
of chemically bound water causes ionic conduction due to the
movement of OH− ions and H+ proton along the columnar chan-
Fig. 7. Variation of AC conductivity with temperature for spark plasma sin- nels in HA crystals.30 The conductivity increases due to the
tered (a) as well as pressureless sintered (b) HA–CT composites at 1 MHz of presence of surface bound water and hydroxyl ions in HA, while
frequency. conductivity of pure CT and HA80CT composite is dominated
by ionic polarization.
interstice separation of chain segment along c-axis. The dielec- In case of monolithic CT, the mobility of O2− ions, located
tric constant, as measured with HA–CT composites, can at the face centred position in the crystal lattices of CaTiO3
therefore be related to the polarization of HA and this can be along c-axis contributes as the dominant conduction mecha-
expressed as25 : nism. It is particularly due to the presence of Ti3+ ions in SPSed
4πP CaTiO3 , as confirmed by XPS results (Fig. 8). Since CaTiO3 is
ε + iε = 1 + (1) retained predominantly in all SPSed samples and also because
E
SPSed samples exhibit better conductivity and dielectric prop-
where E is the periodic applied field, P is total polarization of erties, XPS analysis is carried out only on monolithic CaTiO3 .
the crystals, ε and ε are the real and imaginary components of It can be observed that the 2p3/2 peak at a binding energy of
dielectric constant, respectively. about 459 eV and a small shoulder peak at about 465 eV corre-
Also, the relative dielectric loss, ε of HA can be written sponds to the Ti3+ 2p1/2 .31 Moreover, no peak of the 2p3/2 and
in terms of OH− ion vacancy and oxygen concentrations in 2p1/2 appears at binding energies of 458 eV and 463 eV, which
hydroxyl ion sublattice as: corresponds to Ti4+ 2p peaks. This can be correlated with the
  fact that the colour of the CT samples in as-SPSed conditions
 A [V]
τ/2 [O]
ε = (2) were black, indicating that Ti4+ has been reduced to Ti3+ dur-
T [O] 1 + (
τ/2 [O])2
ing spark plasma sintering.32 Such a reduction of Ti4+ state to
where A is a constant, [V] and [O] are the vacancy and oxygen Ti3+ state generates vacancy in O2− sublattice. In the CaTiO3
impurity concentrations in OH– ion sublattice, respectively, ω structure, O2− ions occupy fourfold and eight fold positions.
and τ are the angular frequency and relaxation time for vacancy As a result, Ti3+ ions produce oxygen vacancies on eight fold
displacement. It is reported that high [V]/[O] (∼103 ) is respon- position sites, which lie in the same lattice plane. However, the
sible for the large values of dielectric loss in HA.26 It is to be four fold position of oxygen ions is independent of Ti3+ ion. It
noted that HA was sintered at 1100 ◦ C with multi stage spark is expected that O2− ion vacancies in both HA and CT lattices
3452 A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453

Fig. 8. Plot of XPS results showing the binding energy of Ti2p, Ca2s, and O1s peaks, obtained from spark plasma sintered CT ceramic. Along Y-axis, XPS counts
(arbitrary units) are plotted.

are the dominant conduction mechanisms in the spark plasma addition to HA has significantly improved both, the dielec-
sintered composites. tric constant (155) and AC conductivity [∼10−5 ( cm)−1 at
As a concluding note, the present study unambiguously 1 MHz frequency] as compared to the lower CaTiO3 contain-
confirms that the electrical transport properties of HA–CT com- ing HA composites. A comparison between the spark plasma
posites are significantly influenced with the addition of more and pressureless sintered HA–CT composite revealed a better
than 60 wt% CaTiO3 or more. For lower CT containing com- efficacy of SPS process to develop HA–CaTiO3 composites
posites, the functional properties are similar or marginally better with better electrical conductivity properties.
than HA matrix. The implication of the present experimental b. A comparison with the existing theoretical models revealed
results can be realized in terms of the significant increase in the that irrespective of processing route, the dielectric constant of
dielectric constant or conductivity of the spark plasma sintered HA–CaTiO3 composites follow closely with Weiner (series)
HA80CT composites as compared to the respective convention- model. The apparently large dielectric constant as well as
ally sintered composites. It may be worthwhile to mention that conductivity has been explained by the presence of Ti3+ ion
the substrate conductivity is found to modulate the myoblast cell and oxygen vacancy in the spark plasma sintered materi-
functionality (both proliferation and differentiation) on the spark als. Overall, spark plasma sintering is found to be a superior
plasma sintered HA-CaTiO3 composites in vitro.14 Also, spark processing route to develop HA–CaTiO3 composites with a
plasma sintered HA80CT composite is reported to exhibit early better combination of functional properties.
stage osteogenesis than sintered HA in rabbit animal model.35
The results presented in this paper together with our recently
published papers therefore establish the motivation to develop Acknowledgements
electroconductive composites for biomedical applications.
The use of precision impedance analyzer in the laboratory of
Prof. K. B. R. Varma is gratefully acknowledged. The authors
5. Conclusions
also thank Department of Science and Technology as well as
Department of Biotechnology for financial support.
Based on the experimental work reported in the present study,
the following conclusions can be drawn,
References
a. From the impendence analysis, it is observed that the dielec-
tric constant and AC conductivity of HA–CaTiO3 composites 1. Basu B, Katti D, Kumar A. Advanced biomaterials: fundamentals,
significantly increase for the higher amount (≥60 wt%) processing and applications. USA: John Wiley & Sons, Inc.; 2009. ISBN:
of CaTiO3 addition to HA. Importantly, 80 wt% CaTiO3 978-0-470-19340-2.
A.K. Dubey et al. / Journal of the European Ceramic Society 33 (2013) 3445–3453 3453

2. Young-M K, Chang-J B, Su-Hee L, Hae-W K, Hyoun-Ee K. Improvement 18. Jain D, Reddy KM, Mukhopadhyay A, Basu B. Achieving uniform
in biocompatibility of ZrO2 –Al2 O3 nano-composite by addition of HA. microstructure and superior mechanical properties in ultrafine grained
Biomaterials 2005;26:509–17. TiB2 TiSi2 composites using innovative multi stage spark plasma sintering.
3. Volceanov A, Volceanov E, Stoleriu S. Hydroxiapatite–zirconia composites Mater Sci Eng: A 2010;528(1):200–7.
for biomedical applications. J Optoelctron Adv Mater 2006;8(2):585–8. 19. Gu Y, Loh N, Khor K, Tor S, Cheang P. Spark plasma sintering of hydroxy-
4. Rapacz-Kmita A, Ślósarczyk A, Paszkiewicz Z. Mechanical properties of apatite powders. Biomaterials 2002;23(1):37–43.
HAp–ZrO2 composites. J Eur Ceram Soc 2006;26(8):1481–8. 20. Santos MH, Oliveira M, Souza LPF, Mansur HS, Vasconcelos WL. Synthesis
5. Suchanek W, Yashima M, Kakihana M, Yoshimura M. control and characterization of hydroxyapatite prepared by wet precipitation
Hydroxyapatite/hydroxyapatite–whisker composites without sinter- process. Mater Res 2004;7(4):625–30.
ing additives: mechanical properties and microstructural evolution. J Am 21. Nikolić MV, Pavlović VP, Pavlović VB, Labus N, Stojanović B. Application
Ceram Soc 1997;80(11):2805–13. of the master sintering curve theory to non-isothermal sintering of BaTiO3
6. Nath S, Dubey AK, Basu B. Mechanical properties of novel cal- ceramics. Mater Sci Forum 2005;494:417–22.
cium phosphate-mullite biocomposites. J Biomater Appl 2011;27(1): 22. Evis Z, Doremus RH. Hot-pressed hydroxylapatite/monoclinic zirco-
67–78. nia composites with improved mechanical properties. J Mater Sci
7. Fukada E, Yasuda I. On the piezoelectric effect of bone. J Phys Soc Jpn 2007;42:2426–31.
1957;12(10):1158–62. 23. Orlovskaya N, Browning N, Nicholls A. Ferroelasticity in mixed conduct-
8. Teng N, Nakamura S, Takagi Y, Yamashita Y, Ohgaki M, Yamashita K. A ing LaCoO3 based perovskites: a ferroelastic phase transition. Acta Mater
new approach to enhancement of bone formation by electrically polarized 2003;51(17):5063–71.
hydroxyapatite. J Dent Res 2001;80(10):1925–9. 24. Den Hartog H, Welch D, Royce B. The Diffusion of calcium, phosphorous,
9. Ohgaki M, Kizuki T, Katsura M, Yamashita K. Manipulation of selec- and OD− ions in fluorapatite. Phys Status Solidi B 1972;53(1):201–12.
tive cell adhesion and growth by surface charges of electrically polarized 25. Crenshaw ME. Electromagnetic energy in dispersive magnetodielectric lin-
hydroxyapatite. J Biomed Mater Res 2001;57(3):366–73. ear media. J Phys B: At Mol Opt Phys 2006;39:17–25.
10. Yamashita K, Oikawa N, Umegaki T. Acceleration and deceleration of bone- 26. Royce BSH. Field-induced transport mechanisms in hydroxyapatite. Ann N
like crystal growth on ceramic hydroxyapatite by electric poling. Chem Y Acad Sci 1974;238(1):131–8.
Mater 1996;8(12):2697–700. 27. Karkkainen K, Sihvola A, Nikoskinen K. Analysis of a three-dimensional
11. Kobayashi T, Nakamura S, Yamashita K. Enhanced osteobonding by nega- dielectric mixture with finite difference method. IEEE Trans Geo and Rem
tive surface charges of electrically polarized hydroxyapatite. J Biomed Mater Sens 2001;39(5):1013–8.
Res 2001;57(4):477–84. 28. Kingery WD, Bowen HK, Uhlmann DR. Introduction to ceramics. New
12. Webster TJ, Ergun C, Doremus RH, Lanford WA. Increased osteoblast adhe- York, NY: John Wiley and Sons; 1976.
sion on titanium-coated hydroxylapatite that forms CaTiO3 . J Biomed Mater 29. Newnham R, Skinner D, Cross L. Connectivity and
Res Part A 2003;67(3):975–80. piezoelectric–pyroelectric composites. Mater Sci Bull 1978;13(5):525–36.
13. Dubey AK, Tripathi G, Basu B. Characterization of 30. Nakamura S, Takeda H, Yamashita K. Proton transport polarization and
hydroxyapatite–perovskite (CaTiO3 ) composites: phase evaluation depolarization of hydroxyapatite ceramics. J Appl Phys 2001;89:5386–93.
and cellular response. J Biomed Mater Res Part B: Appl Biomater 31. Stefanov P, Shipochka M, Stefchev P, Raicheva Z, Lazarova V, Spassov L.
2010;95(2):320–9. XPS characterization of TiO2 layers deposited on quartz plates. J Phys Conf
14. Greeshma T, Mallik PK, Basu B. Substrate conductivity dependent mod- Series 2008;100(012039):1–4.
ulation of cell proliferation and differentiation in vitro. Biomaterials 32. Mei A, Jiang Q-H, Lin Y-H, Nan C-W. Lithium lanthanum titanium
2013;34(29):7073–85. oxide solid-state electrolyte by spark plasma sintering. J Alloys Compd
15. Basu B, Balani K. Advanced structural ceramics. USA: John Wiley & Sons, 2009;486(1):871–5.
Inc.; 2011. [33].Lemanov VV, Sotnikov AV, Smirnova EP, Weihnacht M, Kunze R.
16. Reddy KM, Kumar N, Basu B. Innovative multi-stage spark plasma sin- Perovskite CaTiO3 as an incipient ferroelectric. Solid State Commun
tering to obtain strong and tough ultrafine-grained ceramics. Scr Mater 1999;110:611–4.
2010;62(7):435–8. [34].Rao TS. Structure and dielectric properties of CaTiO3 ceramics. Balkan
17. Reddy KM, Mukhopadhyay A, Basu B. Phys Lett 2004;12(4):195–201.
Microstructure–mechanical–tribological property correlation of mul- [35].Mallik PK, Basu B. Better early stage osteogenesis of electroconductive
tistage spark plasma sintered tetragonal ZrO2 . J Eur Ceram Soc HA-CaTiO3 composites in a rabbit animal model. J Biomed Mat Res A
2010;30(16):3363–75. 2013 (in Press).

You might also like