Sunteți pe pagina 1din 11

Surface & Coatings Technology 206 (2012) 28592869

Contents lists available at SciVerse ScienceDirect

Surface & Coatings Technology


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

A comparative study on the direct and pulsed current electrodeposition of


hydroxyapatite coatings on surgical grade stainless steel
D. Gopi a, b,, J. Indira a, L. Kavitha b, c, d,
a

Department of Chemistry, Periyar University, Salem 636 011, Tamilnadu, India


Centre for Nanoscience and Nanotechnology, Salem 636 011, Periyar University, Tamilnadu, India
Department of Physics, Periyar University, Salem 636 011, Tamilnadu, India
d
The Abdus Salam International Centre for Theoretical Physics, Trieste, Italy
b
c

a r t i c l e

i n f o

Article history:
Received 10 May 2011
Accepted in revised form 5 December 2011
Available online 13 December 2011
Keywords:
Stainless steel
Pulsed electrodeposition
Hydrogen peroxide
Calcium phosphate
Crystallinity
Adhesion

a b s t r a c t
Hydroxyapatite [Ca10(PO4)6(OH)2, (HAP)] coatings were developed on 316L stainless steel substrate from the
electrolyte containing hydrogen peroxide (H2O2) with the concentration ranging from 600 to 3000 ppm by
both the direct and pulsed current electrodeposition methods. The effects of direct current density upon
the addition of H2O2 into the electrolyte on the phase purity and morphology of the as-deposited coatings
were reported. The inuence of pulsed parameters such as peak current density and pulse on and off time
on the deposit compositions was also examined and compared with direct continuous current deposition in relation to the crystallinity, microstructure and the corresponding phases. X-ray diffraction (XRD) and Fourier
transform infrared spectroscopic (FT-IR) techniques were performed in order to assure the purity, phase compositions of the coating and the morphology of the coating were characterized by scanning electron microscopic
(SEM) technique. The results showed that the coating consists of mixed phases of calcium phosphate (CaP)
in the absence of H2O2 in the electrolytic bath. Whereas the addition of H2O2 lowers the deposition current
with the formation of smooth and uniform layer comprised solely of HAP. It is highly benecial to increase the
peroxide concentration from 600 to 2000 ppm for the deposition of pure HAP. While increasing the peroxide
concentration to 3000 ppm, the coating morphology is not uniform as evidenced from the SEM result. Moreover,
the increased adhesion and crystallinity of the HAP coating were achieved by pulsed current electrodeposition
method at lower current density with longer pulse off time. The results of pulsed electrodeposition show that
the relaxation time of the pulse is benecial for the growth of HAP because it allows the diffusion of ions from
bulk solution to the surface of electrode and thus lowers the concentration polarization in the next pulse on
time. The combination of pulsed electrodeposition and addition of H2O2 into the electrolyte promisingly improve
the physico-chemical properties of HAP coating.
2011 Elsevier B.V. All rights reserved.

1. Introduction
Metallic implants such as stainless steel (SS), titanium and its alloys
and cobalt chromium alloys in the orthopedic prostheses have gained
signicant advantages in the recent years due to their immense mechanical features that satisfy the requirement of the human bone [1].
The need to reduce costs in public health services has compelled the
use of austenitic stainless steel of the type 316L SS as the most economical alternative for orthopedic implants respecting to the other usual Ti
or CoCr alloys [2]. It is traditionally used for implantation purposes

Correspondence to: D. Gopi, Department of Chemistry, Periyar University, Salem


636 011, Tamilnadu, India. Fax: + 91 427 2345124.
Correspondence to: L. Kavitha, Department of Physics, Periyar University, Salem
636 011, Tamilnadu, India. Fax: + 91 427 2345124.
E-mail addresses: dhanaraj_gopi@yahoo.com (D. Gopi), louiskavitha@yahoo.co.in
(L. Kavitha).
0257-8972/$ see front matter 2011 Elsevier B.V. All rights reserved.
doi:10.1016/j.surfcoat.2011.12.011

like bone screw/plate, intra-medullary rod, xation wire, HIP joint, and
knee joint in orthopedic surgery owing to their general corrosion resistance, good mechanical properties and acceptable biocompatibility
[35]. Moreover, in several countries it is customary to use 316L SS as
temporary implants. The stainless steels protect themselves by forming
passive oxide lms on their surfaces which provide the key to their high
resistance to corrosive attack. The low carbon content in the 316L SS
provides an improved corrosion resistance in simulated physiological
environments [6]. The presence of 2 to 4 wt.% Mo increases the resistance to pitting corrosion. Many surgeons, especially in developing
countries, prefer stainless steels as implant materials based on millions
of satisfactory clinical cases [7].
To further improve the biological performance and biocompatibility
of 316L SS alloy, ceramic CaP coating is often applied to the surface to
make the entire structure even more important. HAP has been recognized as being osteo-conductive, and is able to promote bone ingrowth and adhesion on the surface of the implant during the early
stages of the implantation [810].

2860

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

Over the years, HAP coatings have been developed on metal surfaces
by variety of methods including plasma spraying [11,12], solgel
[1315], electrophoretic deposition [16,17], biomimetic method [18]
and ion beam deposition [19]. Though plasma spraying technique is
the only clinically accepted method to deposit HAP, it has several limitations. For example, it is very difcult to control the composition and
structure of the HAP coatings deposited, due to extremely high processing temperatures [20,21]. Secondly, this technique produces a nonhomogenous coating which has poor adhesion due to the delamination
of HAP from the substrate. Thirdly, this technique has a line-of-sight
requirement, which greatly limits its usage in the coating being developed for implants [22,23].
In view of the various limitations of the plasma spray technique,
research has been carried out on new methods of deposition. In recent
years, electrochemical cathodic deposition (ECD) has some evident
advantages such as the coating process at low temperature, controlled
chemical composition, simple set-up and relatively low expense
[2427]. However, pure HAP coating developed by traditional electrochemical deposition has several major drawbacks which could limit
the potential applications of this coating technique [2831].
During ECD, an electric current is passed through the electrodes submerged in an electrolyte, initially the OH ions are produced at the substrate (cathode) due to the electrochemical reaction which is usually
the reduction reaction of water, resulting in the formation of large
amount of H2 gas. H2 gas adheres itself to the surface of the metal
substrate, preventing further deposition of CaP, leading to a poor
adherence of the coating to the substrate [32]. This electrochemical
deposition often results in the deposition of monetite (CaHPO4) or
brushite (CaHPO4 2H2O) rather than HAP [28,29], and also pH generated at the substrate cathode, less favors the HAP formation. This problem
can neither be solved by the application of higher current density,
although this will lead to the generation of more OH ions at the surface
resulting in the formation of more hydrogen bubbles. This phenomenon
destroys the simple addition of base directly into the electrolyte as this
will cause course precipitation in the bulk solution rather than deposition on the substrate. Moreover, in the electrodeposition process at
cathode, upon the application of direct continuous current the loose,
porous and low adhesive coatings can easily be developed. The prime
reason may be due to the polarization of concentration differences
which is produced because of the lower mobility of ions diffused from
the main body of the solution to the surface of substrate.
To overcome these obstacles, several modications are proposed to
make on the direct current electrochemical deposition method more
suitable in the following way. At the outset, H2O2 was added into the
electrolyte during the deposition process and H2O2 can substitute for
H2O which in turn lowers the deposition current and change the mechanism of the whole lot of electrochemical reaction [33,34]. Due to this
peroxide addition, the effects of H2 evolution may be erased, and consequently, dense and uniform coatings may be formed [35]. Secondly, the
pulsed current is applied to electrodeposit the HAP on 316L SS substrate
which holds a relaxation time compared with direct continuous current.
In the pulsed electrodeposition process, as current is applied, the ions
closer to the cathode are adequately deposited. During the relaxation
time, the diffusion of ions from bulk solution to the surface of cathode
takes place, in such a way to improve the physico-chemical properties
of coatings [34]. D.J. Blackwood et al. [33] reported galvanostatic pulse
electrodeposition of HAP on titanium and demonstrated that periodic
pulsed current densities gave a better adhesion of HAP coatings deposited on the substrate surface when compared to the continuous current
density. However, the morphology of the coating deposited by pulsed
current has not been investigated. Moreover, the pulsed parameters
such as pulse off and on time and their effects on coating composition
and morphology have not been studied so far. Yu et al. [34] used pulsed
electrodeposition associated with H2O2 into the electrolyte to obtain
HAP coatings with improved adhesion strength between the coating
and substrate. However, they neither varied nor optimized H2O2

concentration and its effect on coating composition, morphology has


not even been demonstrated. Very recently, another method has been
proposed by Drevet et al. [36] who studied the effects of pulsed current
and thereby H2O2 concentration on the phase composition, morphology
of CaP coating deposited on Ti6Al4V. Even though H2O2 concentration was optimized, the morphology of the coating obtained from optimized conditions seems to be porous which is not much satisfactory.
Moreover the current density and H2O2 concentration used in the deposition were high such as 15 mA/cm 2 and 9% respectively.
The present paper deals with HAP coatings on the surface of 316L SS
substrate by the direct as well as pulsed current electrodeposition
method with hydrogen peroxide as an additive into the electrolyte.
The inuence of hydrogen peroxide concentration on the purity, phase
composition and morphology of as-deposited coatings was also investigated and optimized appropriately. The impact of pulsed parameters
such as peak current density, pulse on and off time on the crystallinity,
morphology and adhesiveness of HAP coatings deposited on the 316L
SS substrate was veried and compared to the continuous current density. In the present study, the current densities and the concentrations of
H2O2 used for the deposition of HAP were much lower than the previously reported one [36]. To the best of our knowledge, so far there is
no report on pulsed electrodeposition of HAP on 316L SS with H2O2
into the electrolyte.
2. Experimental procedure
2.1. Preparation of 316L SS specimen
Stainless steel substrates (purchased from Steel Authority of India,
Salem, Tamil Nadu, India) whose elemental composition (wt.%) is
C0.0222, Si0.551, Mn1.67, P0.023, S0.0045, Cr17.05, Ni11.65,
Mo2.53, Co0.136, Cu0.231, Ti0.0052, V0.0783, N0.0659 and
the rest Fe were used as cathode for electrodeposition. The substrates
are cut into dimensions of 1 cm 1 cm 0.3 cm and embedded in
epoxy resin with a working area of 1 cm2. Before deposition, their surfaces were abraded with different grades of SiC papers from 400 to
1200 grit to ensure the same surface roughness. The nal polishing was
done with coarse (6 m) and ne (1 m) diamond pastes in order to produce scratch-free mirror nish surface, and then degreased with acetone.
This was followed by ultrasonic cleaning in acetone for 10 min and then
the specimens were rinsed with deionized water, dried and used for
further studies.
2.2. Electrodeposition of HAP on 316L SS
In this typical experimental procedure, the electrolyte was prepared by mixing a solution of 0.042 mol l 1 of Ca(NO3)2 4H2O and
0.025 mol l 1 of NH4H2PO4 with varied concentrations of H2O2 such
as 600 ppm, 1000 ppm, 2000 ppm and 3000 ppm. The electrolyte was
de-aerated with N2 for 30 min prior to and during the experiments.
This is to reduce the amount of dissolved carbon dioxide and thus prevents the formation of CaCO3 deposits. The pH value of the electrolyte
was adjusted to 4.5 using dilute HNO3 and NH4OH. The temperature
of the electrolyte was maintained at 65 C by using an electric heater.
Magnetic stirring was controlled at a speed of 180 rpm to keep the
concentration of the electrolyte uniform. The electrodeposition was
performed in an individual cell using a regular three electrode conguration in which 316L SS alloy served as cathode and a platinum electrode acts as an anode. Along with this, saturated calomel electrode
(SCE) was used as the reference electrode. The deposition was carried
out in galvanostatic mode using CHI 760C (CH Instruments, USA) by
the direct and pulsed current electrodeposition methods. The current
density was varied from 0.5 to 3 mA/cm2 for the duration of 1 h. For
pulsed electrodeposition as indicated in Fig. 1, a constant pulsed on
time of ton = 1 s with a current density of j = 0.5 mA/cm 2, 3 mA/cm2
followed by the pulsed off time of toff = 1 s, 2 s with joff = 0 mA/cm2

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

2861

Fig. 1. A schematic representation of the pulse current.

was applied. After the specimens were coated with calcium phosphate,
they were removed from the electrolytic solution and rinsed in distilled
water before dried off at 40 C for 4 h.
2.3. Characterization of the coatings
The phase and composition of the coatings were analyzed by X-ray
diffraction (XRD) using Bruker D-8 Advance-Germany Spectrometer,
with CuK radiation = 1.5406 generated at 35 kV and 25 mA.
Data were collected over the 2 range 2060 and also at low angles
(b20) with a step size of 0.010 and a count time of 0.2 s.
The fraction of crystallinity Xc, of the HAP coatings on metal
specimen was evaluated according to the following formula [37];
Xc 0:24=

where, is the corrected full width of the peak (plane) at half of the
maximum intensity (FWHM).

To identify the chemical bonding in the as-deposited coatings,


Fourier transform infrared spectra were recorded using Nicolet 380
FT-IR spectrophotometer over the range from 4000 to 400 cm 1
with a number of scans 32 and resolution of 4 cm 1. For this, small
amount of as-coated powder scrapped from the coating was blended
with KBr and then pressed into disks for the measurement.
The surface morphology of the coating deposited on 316L SS alloy
was examined by scanning electron microscopic (SEM) (JSM 840A
Scanning Microscope, JEOL-Japan) technique.
The thickness of coating was measured using a microprocessor
coating thickness gauge (MiniTest 4100, Elektrophysik, Germany).
The average thickness of each sample was obtained from 10 measurements at different positions.
The adhesion strength of the HAP coated 316L SS samples was
assessed by pull-out test according to American Society for Testing
Materials (ASTM) international standard F1044. The coated 316L SS
cylindrical specimens (25 mm in diameter by 25 mm in height)
were bonded to uncoated cylinder's surface with epoxy adhesive.
The test specimens were cured at 100 C for 60 min, and the pull-

Fig. 2. The XRD patterns of CaP coating deposited on 316L SS at various current densities without the addition of H2O2 into the electrolyte (a) 0.5 mA/cm2, (b) 1 mA/cm2, (c) 2 mA/cm2
and (d) 3 mA/cm2.

2862

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

Table 1
The crystallinity of the CaP coating obtained by electrodeposition method at various
current densities without H2O2 into the electrolyte.
Sample
conditions

OCP
Plane

FWHM

Crystallinity
(Xc)

0.5 mA/cm2

(0 1 0)

0.115

09.1

1 mA/cm2

(0 1 0)

0.105

11.9

(0 1 0)

0.102

13.0

3 mA/cm2

(0 1 0)

0.096

15.6

2 mA/cm


121


121


121


121

2H2 O 2e H2 2OH

DCPD
Plane

electrochemical and chemical reactions that might occur during the


electrolytic deposition are:

FWHM

Crystallinity
(Xc)

0.092

17.7

0.088

20.3

0.086

21.7

0.083

24.2

out test performed at a crosshead speed of 1 mm/min, using a universal testing machine (Model 5569, Instron). Five tests (n = 5) of adhesion strength were conducted for the same type of coating in order to
obtain the average value of bonding strength. The adhesive strength
of the coating was expressed by the unit Mega Pascals (MPa).

3. Results and discussion


3.1. X-ray diffraction
Fig. 2 displays the X-ray diffraction patterns of the as-deposited
coatings on 316L SS without H2O2 into the electrolyte. The phase
compositions of the coated sample at different current densities ranging from 0.5 to 3 mA/cm 2 are presented. The XRD patterns revealed
that at all current densities (Fig. 2(ad)) the main composition of
the coating layer was dicalcium phosphate dihydrate (DCPD or brushite, CaHPO4 2H2O) (International Centre for Diffraction Data (ICDD):
72-1240) with octacalcium phosphate (OCP) (ICDD): 26-1056) as
minor phases and few for the 316L SS substrates. The possible

2H2 PO4 2e 2HPO4 H2

2HPO4 2e 2PO4 H2 :
2

1
2
3

It has been known that the electrolytic deposition results in an


increase of pH at the interface due to hydroxyl ions and H2 gas
through the reduction reaction of water (Eq. (1)) [26,38]. Since, numerous bubbles of H2 gas and OH ions are produced around the
working electrode, OH ions themselves diffuse rapidly away from
cathode substrate to the bulk electrolyte and causing the pH change
around the metal/solution interface. Subsequently, the electron transfer at the substrateelectrolyte interface promotes the dissociation of
HPO42 (Eqs. (2) and (3)) [3941]. Therefore, at the cathodic current
density, HPO42 and H2PO4 also undergo reduction to yield PO43
anion. Even though the PO43 anions are present at the vicinity of
the cathode, there is no enough concentration of OH ion to facilitate
the formation of HAP on the metal specimen.
Other possible reactions for the formation of PO43 ion can also
occur chemically (Eqs. (4) and (5)) as the concentration of the OH
ions increases in the electrolyte [39,41].

H2 PO4 OH HPO4 H2 O

HPO4 OH PO4 H2 O
2

4
5

Since, DCPD is stable in the range of pH lower than 5 [42] and the
OH ions are not much enough to increase the pH up to 7 or more,

Fig. 3. The XRD patterns of CaP coating deposited on 316L SS at various current densities with the addition of 1000 ppm H2O2 into the electrolyte (a) 0.5 mA/cm2, (b) 1 mA/cm2,
(c) 2 mA/cm2 and (d) 3 mA/cm2.

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869


Table 2
The crystallinity of the CaP coating obtained by electrodeposition method at various
current densities with 1000 ppm of H2O2 into the electrolyte.
Sample
conditions

0.5 mA/cm2
1 mA/cm2
2 mA/cm

3 mA/cm

DCPD

HAP

Plane





1 21
1 21
1 21
1 21






FWHM

Crystallinity
(Xc)

Plane

FWHM

Crystallinity
(Xc)

0.101

13.4

(0 0 2)

0.098

14.7

0.103

12.6

(0 0 2)

0.095

16.1

0.112

09.8

(0 0 2)

0.093

17.2

0.118

08.4

(0 0 2)

0.090

18.9

the DCPD easily starts to deposit more than other phases of calcium
phosphate. So, the deposition of DCPD is formed on the cathode surface as expressed by the following reaction
2

Ca

HPO4 2H2 OCaHPO4 2H2 O:


2

2863

the OH ion was produced more in the electrodeposition bath it is unfortunately not sufcient enough to derive the HAP formation in the
as-deposited coating. The intensity of these OCP and DCPD peaks increases remarkably when the current density increases from 0.5 to
3 mA/cm2 suggesting an increase in the crystallinity of the
 coatings.
The diffraction peaks at (1 0 0) plane for OCP and 1 2 1 plane for
DCPD were chosen for the calculation of crystallinity since they are
sharper and isolated from others and the results are presented in Table 1.
The XRD patterns of as-deposited coatings on 316L SS specimen in
the presence of H2O2 (1000 ppm) into the electrolyte at various current densities are shown in Fig. 3. When the current density was
0.5 mA/cm 2, the peaks are attributed to both HAP (ICDD: 09-0432)
and DCPD (Fig. 3(a)). On increasing the current density to 1 mA/cm 2
(Fig. 3(b)) the same trend was retained as before except intensity.
On further increasing the current density towards 2 and 3 mA/cm 2
formation of HAP along with minor phases of DCPD is evident from
Fig. 3(c and d). From the gure it is revealed that the addition of
H2O2 into the electrolyte provides an alternative electrochemical
source of hydroxyl ion at the cathode surface by the following reduction reaction (7):

OCP could be co-precipitated as the empirical formula of


Ca8(HPO4)(PO4)4 5H2O [43]. Generally, at acidic condition, DCPD and
OCP were often precipitated as a precursor for HAP precipitation
[38,43]. In our present study, it is worth noting that at all current densities DCPD as well as OCP is deposited on metal surfaces and this phenomenon can be explained as follows: When the current density was
low such as 0.5 and 1 mA/cm2, the concentration of hydroxyl ions is insufcient to convert the HPO42 into PO43 by reaction (4). In the light of
the above process, reactions (2), (4) and (6) will be the dominating ones
and therefore DCPD starts to deposit faster than OCP. When the pH value
is increased due to OH ion generation, reactions (3) and (5) seem to
occur as the minor ones. On the contrary, OCP is deposited easier than
DCPD when the current density is increased to 2 and 3 mA/cm2. Though,

H2 O2 2e 2OH :

When excess OH ion is produced the formation of HAP is favored


on the cathode surface by the following reaction (8):
2

10Ca

6PO4 2OH Ca10 PO4 6 OH2 :


3

The intensity of DCPD peaks decreases on increasing the current


density to 2 and 3 mA/cm 2 as shown in Fig. 3(c and d) reveals the
decreased crystallinity of the OCP coatings and their values are
presented in Table 2. Therefore, in these conditions reactions
(2)(5), (7) and (8) will be the major reactions whereas (2), (4)
and (6) will be the minor reactions. Thus, the addition of H2O2 into

Fig. 4. The XRD patterns of CaP coating deposited on 316L SS at 0.5 mA/cm2 with various concentration of H2O2 into the electrolyte (a) 600 ppm (b) 1000 ppm (c) 2000 ppm and
(d) 3000 ppm.

2864

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

Table 3
The crystallinity of the CaP coating obtained by electrodeposition method at 0.5 mA/cm2
with various concentrations of H2O2 into the electrolyte.
Sample
conditions

600 ppm
1000 ppm
2000 ppm
3000 ppm

DCPD

HAP

Plane



1 21
1 21




FWHM

Crystallinity
(Xc)

Plane

FWHM

Crystallinity
(Xc)

0.099

14.2

(0 0 2)

0.097

15.1

0.101

13.4

(0 0 2)

0.095

16.1

(0 0 2)
(0 0 2)

0.092
0.089

17.7
19.6

the electrolyte favors the formation of HAP without generating hydrogen gas bubbles which can destroy the as-deposited coatings.
Moreover, formation of DCPD with HAP is due to lack of hydroxyl
ion produced by 1000 ppm H2O2 to react with all the calcium and
phosphate ions in the electrolytic bath. From the results it is portrayed that the HAP formation is favored at low current density of
0.5 mA/cm 2 which is considered as an optimum and taken for further
investigations in order to obtain the phase pure HAP solely. Therefore,
there is no need to attempt at higher current densities for HAP
deposition.

The evaluation of phase composition and crystallinity of the


as-deposited coatings on 316L SS at a constant current density of
0.5 mA/cm2 with various concentrations of H2O2 into the electrolyte is
demonstrated by the XRD patterns shown in Fig. 4. As can be seen in
Fig. 4(a) the pattern consists of mixed phases of CaP when the concentration of H2O2 was 600 ppm (below 1000 ppm). In a similar fashion,
(in Fig. 3) the formation of DCPD with HAP is due to the lack of OH
ions produced by 600 ppm H2O2 to react with all the calcium and phosphate ions into the electrolyte. Hence, the intensity of the DCPD peaks
seems to be higher than those peaks obtained with 1000 ppm of H2O2
(Fig. 4(a and b)) which was conrmed by the calculated fraction of crystallinity values shown in Table 3. On increasing the H2O2 concentration
to 2000 ppm, all the prominent peaks that are indexed to HAP are found
to be well consistent with the ICDD card No. 09-0432 (Fig. 4(c)). On further increasing the concentration of H2O2 to 3000 ppm, there is no
abrupt change in the composition as previous one except for intensity
as shown in Fig. 4(d). The intensity of HAP peaks increases as the concentration of H2O2 is increased thus suggesting the increased crystallinity of the coatings as shown in Table 3. The diffraction peak at plane
(0 0 2) was chosen for the calculation of crystallinity of the HAP
coatings.
Fig. 5 shows the XRD patterns of HAP coatings deposited at two different current densities with two different pulses off time. In all the

Fig. 5. The XRD patterns of HAP coating deposited on 316L SS by pulsed electrodeposition method (a) low current density with small off time, (b) low current density with longer
off time, (c) high current density with small off time and (d) high current density with longer off time.

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869


Table 4
The crystallinity of the HAP coating obtained by pulsed current electrodeposition
method at two current densities with two pulses off times.
Sample conditions

HAP

0.5 mA/cm with 1 s


0.5 mA/cm2 with 2 s
3 mA/cm2 with 1 s
3 mA/cm2 with 2 s

Plane

FWHM

Crystallinity (Xc)

002
002
002
002

0.093
0.091
0.088
0.085

17.1
18.3
20.3
22.5

cases, the obtained peaks can be attributed to HAP (ICDD: 09-0432).


However, when the pulsed off time was as small as 1 s, the intensity
of the peaks as shown in Fig. 5(a), is not prominent at the lower current
density as compared to the higher one (Fig. 5(c)) suggesting less crystalline in nature and the calculated fraction of crystallinity values is
shown in Table 4. If the pulsed off time is increased to 2 s for both current densities, the intensity of the XRD peaks slightly increased for low
current density (Fig. 5(b)) and drastically showed high prole for high
current density (Fig. 5(d)) thus revealing the increased crystallinity of
the as-deposited coatings (Table 4). A comparison of the XRD proles
shown in Fig. 5 demonstrates that pulsed electrodeposition at high current density of 3 mA/cm 2 with longer off period of 2 s gives rise to
higher degree of crystallinity than those obtained from smaller current
density. Generally, it has been known that high degree of crystallinity

2865

results in the nondegradability of pure HAP coating when implanted


into the organ which leads to poor effect for implant usage [4446].
Apart from that according to the very well known Scherrer's equation,
the particle size increases if the crystallinity of the coating is increased.
Therefore, from the XRD results one can conclude that the best coating
is the one deposited by a lower current density with longer off time.

3.2. Fourier-transform infrared spectroscopy (FT-IR)


Representative FT-IR spectra of as-deposited CaP coatings on
316L SS in the absence and presence of an optimized concentration
of H2O2 into the electrolyte are shown in Fig. 6. The FT-IR spectra
for the coatings deposited in the absence of hydrogen peroxide
(Fig. 6(a)) show a vibration at 1224 cm 1 which is due to a signicant peak for brushite. Moreover, a broad stretching peak at
3455 cm 1 as well as a bending peak at 1630 cm 1 was observed
for adsorbed water which is an indicative of brushite. Apart from
these, two asymmetric P\O stretching (3) peaks were present at
1088 and 1018 cm 1. One asymmetric P\O stretching (1) peak is
present at 964 cm 1 and three asymmetric O\P\O bending (4)
peaks were detected at 633, 600 and 568 cm 1 respectively. Lastly
a symmetrical O\P\O bending (2) peak is present at 471 cm 1.
Whereas, on concerning the FT-IR spectra shown in Fig. 6(b) for the
coatings in the presence of H2O2, a signicant concentration of hydroxyl groups remains in the structure as evident from the intensity

Fig. 6. The FT-IR spectra of as-deposited CaP coating on 316L SS (a) electrolyte without H2O2 (b) electrolyte with 2000 ppm of H2O2.

2866

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

Fig. 7. The SEM micrographs of CaP coatings deposited on 316L SS at 0.5 mA/cm2 with various concentration of H2O2 to the electrolyte (a) 600 ppm, (b) 1000 ppm, (c) 2000 ppm
and (d) 3000 ppm.

of stretching and bending vibration peaks at 3571 and 632 cm 1 respectively. The vibration peaks showed typical apatite characteristics,
with PO43 peaks at around 1090, 1020, 602 and 568 cm 1. The split
bands mainly at 1020 and 1090 cm 1 seem to agree with the formation

of a well crystallized apatite. A more detailed analysis permits one to


deduce the presence of carbonate as a peak at 1460 cm 1 was observed
for both FT-IR spectra, because CO2 in air dissolved in the electrodeposition process. So the comparison of the FT-IR results demonstrates

Fig. 8. The SEM micrographs of CaP coating deposited on 316L SS by pulsed electrodeposition method (a) low current density with small off time, (b) low current density with
longer off time, (c) high current density with small off time and (d) high current density with longer off time.

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

that the presence of an optimized concentration of H2O2 favors the HAP


formation in the electrodeposition process.

3.3. Scanning electron microscopy (SEM)


Fig. 7 shows the evaluation of the surface morphologies of asdeposited coatings at 0.5 mA/cm 2 with various concentrations of
H2O2 into the electrolyte. When the concentration of H2O2 was
600 ppm, as shown in Fig. 7(a) the crystallized coating consists of
ake like morphology which grows outward and disorderly. Upon increasing the H2O2 concentration to 1000 ppm, the coating morphology seems to be broken like glass piece (Fig. 7(b)) as the sample still
contains brushite. However, the formation of ake like crystals loosely
stacked with each other bearing pores in between them was observed
at 2000 ppm of H2O2 as presented in Fig. 7(c). At a level of 3000 ppm
concentration of H2O2, no improvement in the coating to the substrate
was observed (Fig. 7(d)). A gel like coating was formed that slipped
off from the 316L SS substrate when rinsed with distilled water. At
this concentration, the rate of hydroxyl ion generation is faster than
the rate of OH ion consumption by the hydrolysis reaction at the
electrode so that, the OH ions move away from the electrode surface
towards the electrolyte, this is the reason why the poor adhesion of
the deposits resulted [47]. Although, the coating morphology was better
below 2000 ppm of H2O2 concentration, mixed phases of CaP were
obtained as also detected by XRD results previously discussed (see
Fig. 4). Therefore based on the XRD and SEM results the optimum concentration of H2O2 for obtaining the phase pure and uniform morphology of HAP was found to be 2000 ppm.
Fig. 8 illustrates the surface morphology of as-deposited HAP coatings obtained at a pulsed off-time either 1 s or 2 s at two different
peak current densities of 0.5 mA/cm 2 and 3 mA/cm 2. It can be seen
in Fig. 8(a) that at low current density and low pulsed off-time
(0.5 mA/cm 2, 1 s), the coating exhibited a plate-like morphology
interweave with each other and the visual porosity level seems to
be dramatically lower than those obtained from direct continuous
current electrodeposition method (See Fig. 7(c)). While, the pulsed
off time was incremented to 2 s, the surface of the coating became
uniform, smooth and coherent and the morphology of the coating
consists of plate-like crystals stacking rigidly with other to the alloy
substrate as shown in Fig. 8(b). As can be seen in Fig. 8(c) for a higher
peak current density and lower off-time (3 mA/cm 2, 1 s), the coating
that consists of needle shaped crystals was formed with pores in
between. The micrograph in Fig. 8(d) showed that for a longer offtime an increase in the peak current density resulted in a less compact
surface as the structure showed irregular akes. Additionally, small
grains that lay on the top of the akes revealed a re-crystallization process. If the micrographs are compared in Fig. 8, then the inuence of the
peak current density on the surface morphology of the HAP deposits can
be seen. The micrograph obtained by lower current density 0.5 mA/cm 2
shows a more compact uniform plate-like morphology than those
observed at a higher current density 3 mA/cm2.
If the pulsed off time is considered as the period the adsorption
and desorption of several species, originating from the electrolyte,
determine the surface diffusion of the new ions, which affects
their incorporation of new crystallites. This is actually supported
by the fact that an increase of the pulsed off time to 2 s leads to
the formation of a more compact uniform layer of HAP that has
resulted in crack free coatings as seen in Fig. 8 (micrograph (b)
and (d)). In particular, it has been observed that the pulsed deposition at low current density with longer pulse off time gives deposits
of smaller crystallite size and of more uniform morphology, as compared to the direct continuous current cases. Hence, it is the best
coating one can obtain from pulsed electrodeposition at a lower current density with the longer pulse off time such as 0.5 mA/cm 2, 2 s
respectively.

2867

The thickness measurement performed by the microprocessor


coating thickness gauge equipment for the HAP coated sample at a
lower pulse off time of 1 s resulted in 22.36 m. While the pulse off
time was increased to 2 s for the coating of HAP, the thickness of
the coating was found to be 24.23. Hence the coating obtained by
pulsed electrodeposition method with longer pulse off time of 2 s
gives a little thicker coating than the one that can be obtained from
the lower pulse off time 1 s.

Fig. 9. Chronopotentiometric curves for (a) galvanostatically deposited CaP coating on


316L SS in the absence and presence of 2000 ppm of H2O2 into the electrolyte, (b) pulsed
electrodeposited HAP on 316L SS for 1 h with 1 s on time at j = 0.5 mA/cm2 and 1 s off
time at j = 0 mA/cm2, (b1) the enlarged view of curve b up to 30 s only, (c) pulsed electrodeposited HAP on 316L SS for 1 h with 1 s on time at j = 0.5 mA/cm2 and 2 s off time at
j = 0 mA/cm2 and (c1) the enlarged view of curve c up to 30 s only.

2868

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

3.4. Adhesion results


Adhesion of the HAP coatings to the substrate is of the utmost importance for the implant to function properly in physiological conditions. Failure of the implant takes place when decohesion of the
coating occurs. The measured adhesive strength for the HAP coating
by direct continuous current electrodeposition method with optimized H2O2 into the electrolyte is 12.6 0.7 MPa, whereas for the
coating obtained in the pulsed electrodeposition method with longer
pulse off time for both lower and higher current densities (2 s,
0.5 mA/cm 2 and 3 mA/cm 2), the measured adhesive strength were
found to be 17.5 0.7 MPa and 16.8 0.3 MPa respectively. This
shows that pulsed power not only benets the formation as well as
growth of HAP, but also improves the adhesion strength between
coatings and alloy substrate without any pre treatment. This nding
is complementary with SEM results as discussed earlier.

3.5. Chronopotentiometric studies


The potential evolution versus time during the electrodeposition of
CaP coatings on 316L SS at a constant current density of 0.5 mA/cm 2
with and without the addition of H2O2 into the electrolyte is shown in
Fig. 9(a). In the absence of H2O2 into the electrolyte, the potential was
found to be 1.4 V. So, at this potential there is a possibility to form
H2 gas bubbles by the reduction reaction of water (Eq. (1)) as discussed
earlier. When H2O2 was added into the electrolyte the polarization potential makes a shift towards noble direction as detected. This phenomenon proved the fact that the addition of H2O2 into the electrolyte
increases the polarization potential by the reduction reaction of H2O2
(Eq. (7)) and thereby favoring the OH ion generation without the formation of H2 gas bubbles.
Fig. 9(b) shows the chronopotentiometric curve for the pulsed
electrodeposition of HAP for 1 h with 1 s pulse on time followed by
1 s pulse off time. The curve consists of 1800 cycles in which each

cycle is represented by 1 s on time (j = 0.5 mA/cm 2) and 1 s pulse


off time (j = 0 mA/cm 2) and magnication of this curve only up to
30 s is also presented in Fig. 9(b1) for a clear view. On increasing
the pulse off time to 2 s for the pulsed electrodeposition of HAP, the
chronopotentiometric curve and its corresponding magnied curve
only up to 30 s are depicted in Fig. 9(c and c1). In this curve each
cycle is represented by 1 s pulse on time followed by 2 s pulse off
time and totally comprised 1200 cycles (Fig. 9(c)) and the potential
variation as a function of time is similar to that of Fig. 9(b).

3.6. Mechanism of pulsed electrodeposition


The mechanism of pulsed electrodeposition for HAP coatings on
316L SS is shown in Fig. 10. As shown in Fig. 10(a) when a pulsed current density was applied, H2O2 is rst reduced into OH ions according to Eq. (7) and as discussed earlier leads to an increase in pH at the
interface between the alloy and electrolyte. Subsequently, Ca 2+ and
PO43 ions in the vicinity of the alloy surface formed by reactions
(2)(5) react with OH ions to form HAP directly on the substrate
surface (Fig. 10(b)) according to the Eq. (8). With the HAP forming,
the concentration of PO43 and Ca 2+ ions near the electrode begins
to decrease, but there is no change in the concentration of OH
ions which causes the difference of concentration of PO43 and Ca 2+
ions between the bulk and vicinity of the electrode as shown in
Fig. 10(b). The difference can lead to the polarization of concentration
difference of alloy substrates. During the pulsed off time, i.e., in the
relaxation time Ca 2+ and PO43 ions diffused from bulk solution to
the vicinity of the alloy (Fig. 10(c)) resulting in the lowering of concentration polarization. Again pulsed power is applied, Ca 2+ and
PO43 reacts with OH ions produced on the surface of alloy to form
HAP as represented in Fig. 10(d). Suppose if continuous current is applied, the difference in concentration polarization makes the HAP
away from the surface of the alloy, thereby leaving the coating to become loose and porous with the poor adhesive strength between the

Fig. 10. The schematic illustration for the mechanism of pulsed electrodeposition of HAP. The ,
represent the Ca2+, PO43 ions present in the electrolyte and the bar denotes
316L SS substrate, during the application of pulsed on time (a) and (b), application of pulsed off time (c) and again the application of pulsed on time (d) respectively.

D. Gopi et al. / Surface & Coatings Technology 206 (2012) 28592869

coating and the substrate. Therefore, the pulsed electrodeposition


retains the polarization and improves the adhesion properties of
as-deposited HAP coatings on the alloy surface. Moreover, the pulsed
deposition gives deposits of smaller grain size and of more uniform
morphology, as compared to the direct continuous current cases.

References

4. Conclusions

[6]
[7]
[8]
[9]
[10]

Coating of HAP on 316L SS has successfully been achieved by


galvanostatic direct and pulsed current electrodeposition methods
without any post treatment. Hydrogen peroxide with various concentration ranges from 600 to 300 ppm was added into the electrolyte for
the requirement of higher alkaline conditions for the deposition, and
thereby opposing the need to evolve hydrogen gas bubbles which
destroys the coating. From the experimental results it was found
that the addition of an optimized concentration of 2000 ppm H2O2
into the electrolyte in the direct continuous current electrodeposition
imparts pure HAP at a lower current density of 0.5 mA/cm 2 but the
adhesion to the substrate surface is poor. Moreover, the XRD, SEM
results revealed that the pulsed current inuences the crystallinity
and adhesion of HAP lms deposited on 316L SS substrate than one
can obtain from direct continuous current method. Furthermore, it
is believed that the off part of the cycle in pulsed electrodeposition
method gives Ca 2+ and PO43 ions in the bulk solution sufcient
time to diffuse to the vicinity of the 316L SS substrate maintaining
more favorable conditions for HAP deposition. Microscopic studies
of the as-deposited HAP coating by pulsed method demonstrated
that the most compact uniform layer was obtained at low current
density of 0.5 mA/cm 2 with longer off time of 2 s. Increasing peak
current density to 3 mA/cm 2 or decreasing pulsed off time to 1 s led
to a nonuniform akes like rougher surface with pores in between.
Therefore the coatings obtained by pulsed method were better adherent and of more uniform compared to those made of direct continuous
current electrodeposition. Hence, the best conditions for the preparation of HAP coatings on 316L SS by pulsed electrodeposition method
are applied current density of jon = 0.5 mA/cm2, joff = 0 mA/cm 2
ton = 1 s, toff = 2 s and H2O2 content of 2000 ppm and the highest adhesive strength of HAP coating is 17.5 0.7 MPa.
Acknowledgments
One of the authors D. Gopi acknowledges the major nancial
support from the Indian Council of Medical Research (ICMR, IRIS ID
No. 2010-08660, Ref. No: 5/20/11(Bio)/10-NCD-I), Department of
Science and Technology, New Delhi, India (DST-SERC, Ref. No: SR/
FTP/ETA-04/2009 and DST-TSD, Ref. No.: DST/TSG/NTS/2011/73)
and Tamilnadu State Council for Science and Technology (TNSCST),
Tamilnadu in the form of major research projects. Another author
(J. Indira) wishes to thank the Council of Scientic and Industrial
Research (CSIR), New Delhi, India for the award of Senior Research
Fellowship (CSIR-SRF). Also L. Kavitha acknowledges the nancial
support from ICTP, Italy in the form of Junior Associateship.

[1]
[2]
[3]
[4]
[5]

[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]
[24]
[25]
[26]
[27]
[28]
[29]
[30]
[31]
[32]
[33]
[34]
[35]
[36]
[37]
[38]
[39]
[40]
[41]
[42]
[43]
[44]
[45]
[46]
[47]

2869

S. Kannan, A. Balamurugan, S. Rajeswari, Electrochim. Acta 50 (2005) 2065.


C. Garcia, S. Cere, A. Duran, J. Non-Cryst. Solids 348 (2004) 218.
M. Vallet-Regi, I. Izquierdo-Barba, F.J. Gil, J. Biomed. Mater. Res. 67A (2003) 674.
G. Rondelli, B. Vicentini, A. Cigada, Br. Corros. J. 32 (1997) 193.
Feng-Huei Lin, Yao-Shan Hsu, Shih-Hsun Lin, Jui-Sheng Sun, Biomaterials 23
(2002) 4029.
T.M. Sridhar, N. Eliaz, U. Kamachi Mudali, B. Raj, Corros. Rev. 20 (2002) 255.
A. Cigada, G.D. Santis, A.M. Gatti, A. Roos, D. Zaffe, J. Appl. Biomater. 4 (1993) 39.
F.H. Jones, Surf. Sci. Rep. 42 (2001) 75.
X. Liu, P.K. Chu, C. Ding, Mater. Sci. Eng. R 47 (2004) 49.
W. Weng, S. Zhang, K. Cheng, H. Qu, P. Du, G. Shen, J. Yuan, G. Han, Surf. Coat.
Technol. 167 (2003) 292.
K.A. Gross, C.C. Berndt, R. Dinnebier, P. Stephens, J. Mater. Sci. Mater. Med. 33
(1998) 3985.
F. Brossa, A. Cigada, R. Chiesa, L. Paracchini, C. Consonni, J. Mater. Sci. Mater. Med.
5 (1994) 855.
A. Balamurugan, G. Balossier, S. Kannan, S. Rajeswari, Mater. Lett. 60 (2006) 2288.
D.M. Liu, Q. Yang, T. Troczynski, Biomaterials 23 (2002) 691.
D. Gopi, V. Collins Arun Prakash, L. Kavitha, Mater. Sci. Eng. C 29 (2009) 955.
T.M. Sridhar, U. Kamachi Mudali, M. Subbaiyan, Corros. Sci. 45 (2003) 237.
I. Zhitomirsky, L. Gal-Or, J. Mater. Sci. Mater. Med. 8 (1997) 213.
X. Fan, J. Chen, J.M. Ruan, Z.C. Zhou, J.P. Zou, Trans. Nonferrous Metals Soc. China
19 (2009) 347.
F.Z. Cui, Z.S. Luo, Q.L. Feng, J. Mater. Sci. Mater. Med. 8 (1997) 403.
Shinn-jyh Ding, Tsui-hsien Huang, Chia-tze Kao, Surf. Coat. Technol. 165 (2003)
248.
D.J. Blackwood, A.W.C. Chua, K.H.W. Seah, R. Thampuran, S.H. Teoh, Corros. Sci. 42
(2000) 481.
R.B. Heimann, Surf. Coat. Technol. 201 (2006) 2012.
C.E. Wen, Y. Yamada, K. Shimojima, Y. Chino, T. Asahina, M. Mabuchi, J. Mater. Sci.
Mater. Med. 13 (2002) 397.
L.Y. Huang, K.W. Xu, J. Lu, J. Mater. Sci. Mater. Med. 11 (2000) 667.
M. Shirkhanzadeh, J. Mater. Sci. Lett. 10 (1991) 1415.
M. Manso, C. Jimenez, C. Morant, P. Herrero, J.M.M. Duart, Biomaterials 21 (2000)
1755.
D. Gopi, V. Collins Arun Prakash, L. Kavitha, S. Kannan, P.R. Bhalaji, E. Shinyjoy, J.M.F.
Ferreira, Corros. Sci. 53 (2011) 2328.
M.H.P. Silva, J.H.C. Lima, G.A. Soares, C.N. Elias, M.C.D. Andrade, S.M. Best, I.R. Gibson,
Surf. Coat. Technol. 137 (2001) 270.
M. Kumar, H. Dasarathy, C. Riley, J. Biomed. Mater. Res. 45 (1999) 302.
S. Ban, S. Maruno, J. Biomed. Mater. Res. 42 (1998) 387.
C.M. Ocampo, D. Villegas, L. Veleva, J. Electrochem. Soc. 152 (2005) C692.
J.S. Chen, H.Y. Juang, M.H. Hon, J. Mater. Sci. Mater. Med. 9 (1998) 297.
D.J. Blackwood, K.H.W. Seah, Mater. Sci. Eng. C 30 (2010) 561.
C.X. Yu, Z.Z. Wei, C.A. Liang, L.H. Gui, Trans. Nonferrous Metals Soc. China 17
(2007) 617.
Y.Q. Zhai, K.Z. Li, H.J. Li, C. Wang, H. Liu, Mater. Chem. Phys. 106 (2007) 22.
R. Drevet, H. Benhayoune, L. Wortham, S. Potiron, J. Douglade, D.L. Maquin, Mater.
Charact. 61 (2010) 786.
N. Degirmenbasi, D.M. Kalyon, E. Birinci, Colloids Surf., B: Biointerfaces 48 (2006)
42.
M.C. Kuo, S.K. Yen, Mater. Sci. Eng. C 20 (2002) 153.
E.A. Abdel-Aal, D. Dietrich, S. Steinhaeuser, B. Wielage, Surf. Coat. Technol. 202
(2008) 5895.
Y. Song, S. Zhang, J. Li, C. Zhao, X. Zhang, Acta Biomater. 6 (2010) 1736.
Q. Yuan, T.D. Golden, Thin Solid Films 518 (2009) 55.
J.C. Elliot, Structure and Chemistry of Apatite and Other Calcium Orthophosphates, Elsevier, Amstedam, 1994, p. 4.
M.S.A. Johnson, G.H. Nancollas, Crit. Rev. Oral Biol. Med. 3 (1992) 61.
Y.F. Fu, D.M. Chen, J. Oral Tissue Eng. 2 (2005) 76.
C.M. Mardziah, I. Sopyan, S. Ramesh, Trends Biomater. Artif. Organs 23 (2009)
105.
L. Yingguang, Y. Zhuoru, C. Jiang, J. Rare Earths 25 (2007) 452.
I. Zhitomirsky, Adv. Colloid Interface Sci. 97 (2002) 279.

S-ar putea să vă placă și