You are on page 1of 13

Applied Surface Science 410 (2017) 432–444

Contents lists available at ScienceDirect

Applied Surface Science


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

The effects of parametric changes in electropolishing process on


surface properties of 316L stainless steel
Zia ur Rahman a,b , K.M. Deen c,d , Lawrence Cano e , Waseem Haider a,b,∗
a
School of Engineering and Technology, Central Michigan University, Mt. Pleasant, MI 48859, USA
b
Science of Advanced Materials, Central Michigan University, Mt. Pleasant, MI 48859, USA
c
Department of Materials Engineering, University of British Columbia, Vancouver, BC V6T 1Z4, Canada
d
Department of Metallurgy and Materials Engineering, CEET, University of the Punjab, Lahore, 54590, Pakistan
e
Department of Mechanical Engineering, The University of Texas Rio Grande Valley, Edinburg, TX 78539 USA

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

Article history: Corrosion resistance and biocompatibility of 316L stainless steel implants depend on the surface features
Received 22 October 2016 and the nature of the passive film. The influence of electropolishing on the surface topography, surface free
Received in revised form 7 March 2017 energy and surface chemistry was determined by atomic force microscopy, contact angle meter and X-
Accepted 7 March 2017
ray photoelectron spectroscopy, respectively. The electropolishing of 316L stainless steel was conducted
Available online 12 March 2017
at the oxygen evolution potential (EPO) and below the oxygen evolution potential (EPBO). Compared to
mechanically polished (MP) and EPO, the EPBO sample depicted lower surface roughness (Ra = 6.07 nm)
Keywords:
and smaller surface free energy (44.21 mJ/m2 ). The relatively lower corrosion rate (0.484 mpy) and smaller
Electropolishing
Orthopedic
passive current density (0.619 ␮A/cm2 ) as determined from cyclic polarization scans was found to be
Impedance spectroscopy related with the presence of OH, Cr(III), Fe(0), Fe(II) and Fe(III) species at the surface. These species
Potentiodynamic assured the existence of relatively uniform passive oxide film over EPBO surface. Moreover, the relatively
Corrosion biomaterials large charge transfer (Rct ) and passive film resistance (Rf ) registered by EPBO sample from impedance
spectroscopy analysis confirmed its better electrochemical performance. The in vitro response of these
polished samples toward MC3T3 pre-osteoblast cell proliferation was determined to be directly related
with their surface and electrochemical properties.
© 2017 Elsevier B.V. All rights reserved.

1. Introduction rounding tissues is considered extremely important for enhanced


osseointegration at the early stage of implantation [6]. Dissolu-
Stainless steel is extensively used in various industrial appli- tion of ions from the implant surface and its accumulation at the
cations such as automotive parts, in semiconductor industry and interface could be one of the main reasons causing inflammatory
biomedical devices [1–3]. The low-carbon stainless steel is consid- reactions in the body [7].
ered biocompatible material and is mainly used in many orthopedic The presence of body fluids, minerals, chlorides, amino acids
devices and cardiovascular implants [4]. However, the mechani- and proteins in the human physiological environment could pro-
cal stability and corrosion resistance of 316L are the paramount mote corrosion of stainless steel implants. The ionic species such as
concerns for long-term applications. The in vivo utilization of this iron (Fe), chromium (Cr), nickel (Ni) and molybdenum (Mo) could
material therefore demands high quality surface, structural and release and accumulate within the surrounding tissues or may be
biological properties [5]. transported through the blood stream to the various parts of body.
In orthopedics, the bone/implant integration strongly depends These ions may initiate inflammatory reactions within surrounding
on the surface properties of the prosthetic device. Also the integrity tissues [7–10]. The build-up of metal debris in the soft tissues of the
of implant/tissue interface and the biological response of sur- body may lead to necrosis [11] and aseptic loosening of knee joint
(common example of implant failure), which can be life threatening
in older age [12]. Therefore, the formation of stable passive oxide
film over the implant surface is necessary which could limit the
∗ Corresponding author at: School of Engineering and Technology, Central Michi-
release of metallic ions.
gan University, Mt. Pleasant, MI 48859, USA.
E-mail address: haide1w@cmich.edu (W. Haider).

http://dx.doi.org/10.1016/j.apsusc.2017.03.081
0169-4332/© 2017 Elsevier B.V. All rights reserved.
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 433

For cardiovascular devices, the chemical composition, rough- Table 1


Chemical composition of 316L stainless steel (wt.%).
ness and surface energy are the most important properties of
stainless steel for its better thrombogenicity and electrochemical C Mn Cr Ni Mo N P S Fe
stability [13]. The non-homogenous surfaces could initiate adhe- 0.03 2.00 17.0 11.5 2.00 0.02 0.045 0.030 Balance
sion and aggregation of platelets which may activate the plasmatic
coagulation and immunological responses [14]. The intrinsic coag-
investigated to evaluate the proliferation behavior of MC3T3 pre-
ulation of proteins and platelets at the implant surface could induce
osteoblast cells.
clotting and thrombosis [15]. Furthermore, the surface characteris-
tics could influence neointimal hyperplasia, which may result in the
disruption of endothelial layer and triggers the cellular response 2. Experimental
at the wall of vascular muscle. Thus, rapid proliferation of the
smooth muscle cells could promote restenosis [16]. Therefore, in 2.1. Sample preparation
order to reduce inflammation and immunological response at the
implant/tissue interface, surface modification of implant materials The circular disks, 1.6 cm in diameter and 0.5 cm thickness were
(such as stainless steel) is considered extremely important [17]. cut from a commercial grade 316L stainless steel rod. The stainless
The electropolishing is an electrochemical surface treatment steel rod was purchased from onlinemetals.com and its chemical
process involving anodic dissolution of metal or alloy in an composition is given in Table 1 as provided by the supplier. The
®
appropriate electrolyte to restore defect free and smooth surface disks were ground sequentially on a Buehler abrasive belt grinder
[18]. During electropolishing, the simultaneous surface dissolu- by using silicon carbide papers of 240, 320, 400, 600, 800 and 1200
tion and brightening effect could generate smooth mirror-like grit size. These disks were washed in deionized water followed by
surface. In this way the deleterious influence of surface defects, ultrasonic cleaning in ethanol for 15 min. The ground disks were
microstructural variations and preferred crystallographic orienta- electropolished both below and at oxygen evolution potential at
®
tions on the electrochemical properties and biological response Electrobright (Macungie, PA, USA) facility. Briefly, 85% phosphoric
could also be minimized. It is therefore considered that the elec- acid (H3 PO4 ) mixed with 93% sulfuric acid (H2 SO4 ) in a volume ratio
tropolishing of 316L stainless steel implants could be an effective of 7:3 at 60 ◦ C was used for electropolishing at the oxygen evolution
method to improve corrosion resistance, biocompatibility and ser- potential (EPO). On the other hand, electropolishing below the oxy-
vice longevity [19]. gen evolution plateau (EPBO) was carried out in 100 ml methanol
In this research work, the electropolishing of 316L stainless steel (CH3 OH) containing 300 ml of 93% sulfuric acid solution at 25 ◦ C.
was carried out under optimized conditions. The current–voltage In both cases of electropolishing, 10 V DC potential was applied
(V–I) curves for electropolishing could be divided into two regions for 5 min and copper (Cu) was used as cathode. In the following
as ‘below oxygen evolution’ and ‘at oxygen evolution’ potential as discussion, mechanically polished and electropolished disks at and
shown in Fig. 1. Mechanistically, below the oxygen evolution below the oxygen evolution potentials are designated as MP, EPO
potential, the electrochemical reactions during electropolishing and EPBO samples, respectively.
are under pure kinetic controlled regime resulting in the syner-
gistic dissolution and passivation of the surface whereas at the 2.2. Surface characterization and X-ray photoelectron
oxygen evolution potential the electropolishing is carried out dur- spectroscopy
ing the dissociation of aqueous electrolyte [20]. The produced
surfaces were characterized to examine their topographical, chem- The surfaces of MP, EPO and EPBO samples were examined
ical, structural and electrochemical properties in the simulated by a scanning electron microscopy (SEM) (Sigma VP Carl Zeiss,
body environment. Furthermore, their biological response was also Germany). The surface roughness was measured by atomic force
microscopy (AFM) (Nanoscope IV MultiMode, Digital Instruments,
USA) operated under tapping mode and scanning 10 ␮m × 10 ␮m

Fig. 1. Anodic polarization scan for 316L stainless steel showing two regions of electropolishing.
434 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

surface area. The interfacial free energy, contact angle, surface free
energy (SFE) and work of adhesion were calculated from sessile
drop method using Kyowa contact angle meter (DM-CE1, Japan).
The surface chemistry of each sample was analyzed by X-
ray photoelectron spectroscopy (XPS) (Thermo Scientific K-Alpha).
Themonochromatic Al K␣ radiation source (1486.67 eV) was used
and the photoelectron were detected at 90◦ take-off angle. The
base pressure was adjusted to 2 × 10−9 mbar. Survey spectra were
recorded in the binding energy scale from 0 to 1350.0 eV with a pass
energy of 200 eV, step size of 1.0 eV and dwell time of 10 ms. High
resolution spectra for O 1s, Cr 2p and Fe 2p were obtained at a fixed
analyzer transmission mode with a pass energy of 10.0 eV, step size
of 0.1 eV and dwell time of 50 ms. Binding energies in the survey
and high-resolution spectra were calibrated with respect to C 1s
(285.0 eV) adventitious peak. The acquired data was analyzed by
using the ThermoScientific Avantage XPS software package. Peak
fitting was performed using mixed 80% Gaussian–20% Lorentzian
deconvolution function after subtracting the background by Smart
method.

2.3. Electrochemical analysis

The electrochemical behavior of MP, EPO and EPBO was eval-


uated by using potentiodynamic cyclic polarization (PCP) and
electrochemical impedance spectroscopy (EIS) in phosphate buffer
saline (PBS) solution. The phosphate buffer saline solution was
prepared by dissolving one tablet (Part # P4417-50TAB, Sigma
®
Aldrich ) in 200 ml deionized water. The PBS solution was com-
posed of NaCl (0.8 g/L), Na2 HPO4 (0.06 g/L), NaHCO3 (0.35 g/L),
KH2 PO4 (0.06 g/L), MgSiO4 (0.2 g/L) and CaCl2 (0.14 g/L).
The three-electrode cell setup was used in this study for elec-
trochemical analysis. The saturated calomel electrode (SCE) and
graphite rod were used as reference and counter electrodes, respec-
tively. 316L stainless steel samples were used as working electrodes
in this setup. Both PCP and EIS experiments on this cell were
conducted under 5% CO2 environment in the incubator and tem-
perature was maintained constant at 37 ◦ C. The cell was connected
with Gamry potentiostate reference 3000 for electrochemical test-
ing.
Before each experiment, the open circuit potential (OCP) was
measured for 24 h in order to achieve 0.01 mV/s potential stability.
For PCP tests, the scan rate was adjusted to 1 mV/s. The surface
area of working electrode samples (MP, EPO, EPBO) exposed to
the electrolyte was 2.02 cm2 . To obtain EIS spectra, the poten-
tial perturbation of 10 mVrms (0 Vdc vs. OCP) was applied within
0.01 Hz–100 kHz frequency range.

2.4. Cell adhesion

The MC3T3-E1 (Subclone-4) cells were cultured in the ␣-


modified minimum essential medium (␣-MEM) (HyCloneTM ) Fig. 2. SEM micrographs of 316L stainless steel, (a) MP, (b) EPO and (c) EPBO.

containing 10% fetal bovine serum (FBS) (ThermoScientificTM ) and


®
1% Penicillin–Streptomycin (Sigma–Aldrich ). The cells were cul- 3. Result and discussion
tured in controlled humidified atmosphere having 5% CO2 at 37 ◦ C.
200 ␮l of growth media containing 15,000 cells was spilled over The surface morphology of mechanically polished and elec-
MP, EPO and EPBO samples to fully cover the surfaces. After 24 and tropolished 316L is shown in Fig. 2. The grinding marks over the MP
®
48 h of incubation periods, cell staining was done using NucBlue sample (Fig. 2a) were clearly visible as unidirectional parallel lines
Live Cell Stain Ready Probes (Invitrogen Inc.) for the nuclei, and across the entire surface. The mechanical grinding of stainless steel
ActinRedTM 555 Ready Probes Reagent (Thermofisher) for F-actin samples could increase the roughness, influences surface energy
filaments of cytoskeleton. Overlay fluorescent images of nuclei by distorting the subsurface microstructure [21]. These effects may
®
and cytoskeleton were captured by EVOS FL Cell Imaging System enhance the susceptibility of pitting corrosion at the metal surface
(AMF4300, Invitrogen Inc.). even in the mild aggressive environments. On the other hand, by
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 435

Fig. 3. XPS Survey spectra of 316L stainless steel samples, (a) MP, (b) EPO and (c) EPBO.

improving surface finish without disturbing surface or subsurface film. The surface species i.e. C, Na and Si could also be originated
microstructures such as electropolishing could be an effective way from the grinding and electropolishing process or during handling
to improve corrosion resistance. The dissolution characteristics of and/or storage as minor contaminants. The relatively higher Si con-
these surfaces are also strongly dependent on the stability of pas- tents on MP sample could be expected after grinding over silicon
sive film which in the adverse case could increase the chances of carbide papers.
toxic elements i.e. Fe, Cr and Ni release into the human body beyond Fig. 4 shows the high resolution XPS spectra of O1s, Cr2p and
the tolerance level. However, the significant improvement in the Fe2p species present at the surfaces of MP, EPO and EBPO samples.
surface smoothness could be observed after electropolishing (EPO The dashed lines represent the sum of the deconvoluted XPS spec-
and EPBO) as shown in Fig. 2b and c respectively. tra. In case of MP sample, three major peaks at 530.4 eV, 532.4 eV
and 533.5 eV were observed and associated with O1s in the form
of OH or some organic contaminants respectively. Peak “a” and
3.1. XPS analysis peak “b” could be assigned to the metal oxides and hydroxides,
whereas the peak “c” corresponded to organic carbonate impuri-
X-ray photoelectron spectroscopy analyses of the MP, EPO and ties phase at the surface [22]. On the other hand, in the case of EPO
EPBO surfaces were carried out to investigate and compare the and EBPO samples the satellite peak at 533.5 eV did not appear. The
compositional variations at the surfaces. XPS survey spectra of OH peak was intensified in EPO whereas in EPBO, the peak related
mechanically polished and electropolished 316L are shown in Fig. 3. with metal oxide was dominant. The formation of chromium oxide
The chemical composition of elements (wt.%) within the surface and/or iron oxides on the surface may be attributed to the for-
film is provided in Table 2. It was observed that the oxygen O1s mation stable oxide film after electropolishing. Furthermore, the
concentration was dominant within the surface film of all MP, EPO chromium (Cr2p) spectra of the MP sample indicated that the exis-
and EPBO samples. The slightly lower concentration of oxygen in tence of chromium Cr(0) but the dominancy of Cr(III) in the form
the EPO and EPBO when compared to MP could be related with the of Cr2 O3 and Cr(OH)3 was evident in EPO and EPBO samples. From
polishing procedure. The relatively stronger S2p and P2 peaks on the peaks analysis, it was also observed that the surface of EPO and
EPO sample when compared to EPBO might be attributed to the EPBO was enriched with hydrous chromium (III) oxide film and its
adsorption of phosphates and sulfates anions within the passive

Table 2
Elemental composition of 316L obtained from XPS survey spectra.

MP EP EPBO

Elem. Binding energy (eV) Con. (wt.%) Elem. Binding energy (eV) Con. (wt.%) Elem. Binding energy (eV) Con. (wt.%)

O1s 531.37 43.92 O1s 532.00 36.51 O1s 531.19 38.75


C1s 285.55 33.15 C1s 285.45 30.23 C1s 285.45 32.09
Si2p 102.61 3.16 Fe2p 723.99 10.51 N1s 398.48 3.68
Fe2p 710.67 13.89 Cr2p 586.97 7.55 Fe2p 721.16 11.5
N1s 400.06 1.90 Cr 44.32 5.65 Cr2p 586.56 8.55
P2p 133.18 0.54 N1s 399.81 3.33 Cr2p3 577.04 1.23
Mn2p3 640.90 0.76 P2s 191.4 0.44 P2p 191.4 0.47
Ni2p3 870.27 0.61 S2p 165.57 0.64 Ni2p3 853.36 1.28
Cr2p 586.38 1.86 S2p3 166.08 0.32 S2p 165.57 0.69
Mo3d3 231.80 0.15 Ni2p3 854.20 0.67 Mn2p3 640.17 1.25
S2p 165.57 0 Mn2p3 638.94 3.43 Si2p 101.08 0.00
436 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

Fig. 4. High resolution spectra of 316L samples before and after electropolishing.

intensity was found to be higher on EPBO sample. The formation of suggested the limited activity of PO4 2− and SO4 2− ionic species in
the Cr(OH)3 indicated the presence of thick passivated oxide film. the electrolyte under applied conditions [25,26].
Similarly, iron peaks at 706.88 eV and 710.58 eV were attributed to The phosphorus and sulfur in the metallurgical composition of
metallic state Fe(0) and Fe(III) (Fe2 O3 ) respectively [23]. The EPBO 316L stainless steel could possibly be oxidized at the surface dur-
has two satellite peaks at 714.78 eV and 716.38 eV indicating the ing electropolishing and the formed passive film may contain these
presence of both Fe(II) and Fe(III) species which corresponds to species. The deconvoluted P2p and S2p spectra were identified as
FeO and Fe2 O3 respectively. The iron oxide peaks dominancy in shown in Fig. 5. The splitting of P2p high resolution spectra into
both EPO (69.15 at.%) and EPBO (65.26 at.%) were attributed to the doublet peaks at 133.40 eV and 135.50 eV binding energies could
stable surface which could improve the corrosion resistance. Fur- be associated with the Cr(III)–PO4 and P2 O5 species respectively.
thermore, the EPBO sample showed relatively stronger Fe(0) peak Both EPO and EPBO showed similar behavior and indicated the
than other samples. This peak could influence the surface wettabil- formation of phosphate species. The prominent signatures of phos-
ity properties and protein adhesion [24]. The formation of PO4 2− phate species within the passive film could also be affiliated with
and SO4 2− complexes with the Cr(III), Fe(II) and Fe(III) species at the adsorption of H3 PO4 and or H2 PO4 − species present in the elec-
the electropolished samples could be possible. However, the elec- trolyte during electropolishing process. The sulfate species within
tropolishing at high applied DC potential (10 V) and at low pH also the surface passive film were also identified as presented in Fig. 5b.
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 437

Fig. 5. P2p and S2p XPS high resolution deconvoluted spectra of (a) EPO and (b) EPBO.

The high activity of cationic species (H+ ) and aggressively oxidizing Table 3
Average values of contact angle in degrees (n = 3).
conditions during electropolishing could also promote the dissolu-
tion and oxidation of Fe into Fe(II) and Fe(III). These species could Contact angles
physically adsorb at the surface and could possibly form sulfates as
MP EPO EPBO
confirmed from the S2p peak splitting analyses. The following elec-
DI water 67.16 ± 5.99 66.35 ± 0.39 53.51 ± 2.44
trochemical reactions (1) and (2) could occur at the surface during
Ethylene glycol 26.30 ± 4.74 36.07 ± 0.40 36.63 ± 4.48
electropolishing [25]. Diiodomethane 25.40 ± 11.95 27.43 ± 0.11 30.04 ± 7.33

4Fe + [4H2 SO4 ]ads → [3FeSO4 ]ads + FeS + 4H2 O (1)


− + −
2Fe + [3HSO4 ]ads → [Fe2 (SO4 )3 ]ads + 3H + 6e (2)
influence the cell proliferation due to protein adhesion which could
The XPS high resolution spectra of S2p shows strong peaks of influence the implant/tissue integration. Therefore, the surface
Fe2 SO4 , Fe2 (SO4 )3 and other metal sulfides in both cases i.e. EPO energetics and wettability are considered as the key factors influ-
and EPBO samples. Furthermore, SO2 peaks might be attributed encing the integrity of implant materials in the aggressive biological
to the partly immerged anions in the passive oxide layer during environment. The quantitative detail about contact angle (CA) of
electropolishing. MP, EPO and EPBO sample is displayed in Table 3. The CA measured
from water droplet at the surface of MP (67.16◦ ) and EPO (66.35◦ )
3.2. Wettability and surface roughness samples was almost similar, whereas EPBO sample represented rel-
atively small contact angle (53.51◦ ). The statistical analysis showed
The solid surface and liquid molecular interaction was studied that there was no significant difference in wettability of MP and
in order to investigate the wettability of surfaces after mechani- EPO whereas; EPBO was significantly different than MP and EPO
cal and electro-polishing. Wettability of the implant surface could (P < 0.05). Table 3 also displays the contact angles with non-polar
438 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

Table 4
Surface free energy (SFE) components measured by using acid–base theory.

LW AB Acid (+) Base (−) Total SFE (mJ/m2 ) Fractional polarity, AB/(LW + AB)

MP 48.50 ± 0.50 3.31 ± 0.07 0.3 9.12 ± 0.20 51.81 ± 0.70 0.06
EPO 45.25 ± 0.16 2.29 ± 0.05 0.1 12.70 ± 0.37 47.52 ± 0.18 0.05
EPBO 44.21 ± 0.09 0 0 28.58 ± 0.16 44.21 ± 0.09 0

(diiodomethane) and dipolar (ethylene glycol) which are the essen- 3.3. Electrochemical analysis
tial reagents for calculating the total surface free energy (SFE). The
total surface free energy was calculated from the polar components The open circuit potential trends of MP, EPO and EBPO samples
(AB) and dispersive component (LW). are shown in Fig. 6. It was determined that upon immersion of MP
The detailed experiments to separate individual contribution of in the PBS solution presented more negative value (−0.3 V vs. SCE)
surface polarity and dispersive component (LW) were conducted compared to EPO and EBPO. The rapid shift in OCP toward more
based on the acid/base theory and quantitative results are given in positive (noble) direction in initial 4 h could be associated with the
Table 4. It was evident that the major contribution to the total SFE relatively higher surface roughness of MP sample. On the other
was arising because of LW component as their values were almost hand, EPO sample represented similar trend but attained a con-
close to the total SFE. High energy surfaces could supports better stant potential value (−0.145 V vs. SCE) in the initial 2 h. The OCP of
cell adhesion and mineralization compared to low energy surfaces MP and EPO approached to the constant values after 16 h whereas
[27]. The results indicated that MP has comparatively higher SFE EBPO remained constant during immersion. The negligible change
(51.81 mJ/m2 ) when comparison with EPO (47.52 mJ/m2 ) and EPBO in potential of MP and EPO samples was observed in the following
(44.21 mJ/m2 ). However, the surface polarity was very low for both 8 h. However, it could be seen that initial OCP of EBPO observed
MP and EPO and became zero for EPBO. The low value of surface in the PBS solution was more noble (112.8 mV) than MP and EPO
polarity is considered beneficial for optimal cellular adhesion and samples. Finally, after 24 h all the polished samples achieved stable
growth [28]. It has been reported previously that fibroblast attach- OCP values and followed the EBPO > EPO > MP sequence represent-
ment and their spreading was more pronounced over hydrophilic ing higher to the least noble corrosion tendency in the PBS. The
surfaces compared to hydrophobic ones [29]. Hence the surface stabilization of OCP and slight difference in the final values of
energetics could significantly influence the performance of perma- mechanically and electrochemically polished samples suggested
nent prosthetic devices. On the other hand, the hydrophobic (large that electropolishing above the oxygen evolution regime could oxi-
contact angle) surfaces could limit electrochemical charge transfer dize the elemental phosphorus and sulfur present at the surface.
reactions and the challenge of achieving better cell adhesion and Also the presence of phosphate ions (H2 PO4 − , HPO4 2− and PO4 3− )
corrosion resistance properties at the same time is still an active in the electropolishing electrolyte could modify the properties of
area of research. Here it could be suggested that the modification passive film as discussed above. These ions may greatly influence
of implant surface should be based on the targeted function and the electrochemical properties of stainless steel sample whereas in
location of implant in the body. In bone prosthetics, hydrophilic case of EBPO sample it was expected that electropolishing condi-
(small contact angle) surfaces are favorable as these could support tions remained within water stability region without dissociation
cell adhesion and osseointegration. In case of coronary stent and into O2 which could promote the formation of homogeneous and
blood contacting devices, hydrophobic surfaces may promote pro- uniform surface passive film at the sample surface. To confirm this
tein adsorption and blood coagulation which may ultimately lead behavior, the PCP scans were initiated from very negative cathodic
to thrombosis [30]. potential (−1.0 V vs. SCE) (reducing conditions) to eliminate the
On the other side, wettability is sensitive to surface rough- possible effect of inherent passive film and to determine purely,
ness therefore, the long term integrity of implant materials depend the electrochemical response related with surface topography. The
on the compromise between osseointegration and electrochemi- relatively higher current density below −0.8 V vs. SCE observed in
cal dissolution of metal ions at the bone/implant interface [31–33]. case of MP compared to EPO and EBPO was in support with higher
AFM micrographs and the roughness parameters measured for MP, surface roughness. The lower current density (about 3 order of
EPO and EBPO are provided in Table 5. The MP sample presented decades) observed in case of EBPO sample at relatively negative
relatively large average roughness (Ra = 33.51 nm) compared to potential could be associated with the limited dissolution of pas-
EPO (Ra = 11.50 nm) and EBPO (Ra = 6.07 nm) samples which can sive film compared to higher current density observed in case of
be expected depending upon the surface preparation method. The MP and EPO samples. This could be due to the presence of rela-
topographical features (Ra and Rq) were found to be directly related tively thick passive oxide film over EBPO sample which may have
with the contact angle as discussed above. This behavior could be not been dissolved during cathodic polarization. Also the compro-
associated with the competition between the surface tension of mised or corrosion potential (Ecorr ) was more negative (active) in
the stainless steel samples and cohesive forces of the interacting case of MP than EPO and EBPO as shown in Fig. 7. The intersection
electrolyte. The higher micro scale surface roughness is consid- of cathodic polarization curves of EPO and EBPO with the anodic
ered beneficial for cell adhesion, osseointegration and for improved trend of MP in forward scan at the same point could be associated
biomechanical integration. Although, increased surface roughness with the preferential dissolution of active sites within the passive
could enhance the localized corrosion behavior of the surface. In film. The estimation of kinetic parameter in the Tafel region of each
other words, the relatively smooth surfaces (EBPO) would promote curve also validated the higher corrosion rate (1.866 mpy) of MP
the spreading of electrolyte (lower contact angle). However, the than EPO (0.717 mpy) and EBPO (0.484 mpy). The corrosion rate of
active passive nature of stainless steel and complex electrochem- MP, EPO and EBPO samples under similar conditions were relatively
ical reactions would also depend on the surface free energy and lower than the values reported elsewhere [34,35].
composition of the interacting species in the electrolyte. Therefore, In the real biological environment number of other factors such
PCP and EIS tests were conducted to evaluated the effect of surface as effect of oxidant (i.e. dissolved oxygen, low pH etc.) and presence
properties on their electrochemical performance in the simulated of chloride ions could also affect the stability of passive film. To
biological environment. investigate the stability of passive film in the PBS the anodic polar-
ization scans up to 2.5 V vs. SCE or current density limit 10 mA/cm2
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 439

Table 5
AFM 3D surface topography of 316L Stainless steel.

Surface AFM analysis

Topography Surface roughness, Ra (nm) RMS of roughness, Rq (nm)

MP 33.51 ± 5.54 48.28 ± 8.65

EPO 11.50 ± 1.61 14.79 ± 1.90

EPBO 6.07 ± 0.73 7.91 ± 0.73

Fig. 6. Open circuit potential of 316L before and after electropolishing.


440 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

Fig. 7. Cyclic potentiodynamic polarization scans of 316L before and after electropolishing.

Table 6
Quantitative analysis of potentiodynamic polarization scans and comparison with literature.

Icorr (A/cm2 ) Ecorr vs. ref (mV) Passive potential, Breakdown Protection Corrosion rate Corrosion rates
Ipass (mV) potential, Eb potential, Ep (mpy)a from other
literature (mpy)

MP 5.610e−6 −410.0 −297.4 239.9 −149.7 1.866 51.18 [34] b


EPO 2.370e−6 −353.0 −248.6 319.8 −73.9 0.717 4.069 [35] c
EPBO 1.290e−6 −288.0 −187.2 308.4 −60.31 0.484
a
This work.
b
Mechanically polished 316LSS (corrosion rate was measured by weight loss).
c
Electropolished 316LSVM SS.

were obtained. The potential was reversed to the negative (ini- mined. It was calculated that pitting resistance of EPO (672.8 mV)
tial) potential to evaluate the pitting tendency of samples. It was was slightly higher than the MP (649.9 mV) and EBPO (596.4 mV)
observed that passive current density (ipass ) of MP (4.436 ␮A/cm2 ) sample. The increase in potential or the presence of oxidants in the
was higher than EPO (3.24 ␮A/cm2 ) and EBPO (0.6194 ␮A/cm2 ) as electrolyte could also enhance the pitting susceptibility of stain-
given in Table 6. All the samples represented large positive hys- less steel by breakdown of passive film at ‘Eb ’ corresponding to the
teresis during reverse of anodic polarization scan suggesting strong following reaction (6) and the behavior is evident from the large
pitting susceptibility of samples. The XPS analyses also ensured the positive hysteresis loop (Fig. 7) [36].
presence of chromium (III) and oxygen species at the surface of
these samples which are the principle ingredient of passive film Cr2 O3 + 5H2 O → 2CrO4 2− + 10H+ + 6e− (6)
on stainless steel and may hydrolyzed in aqueous environment
Another useful parameter which determines the pitting sus-
(reactions (3) and (4)).
ceptibility was also calculated for these samples from PCP scans.
The possible reactions which may occur at the breakdown
The comparable value exhibited by EBPO (227.69 mV) than MP
potential (Eb ) of passive film and may initiate pitting in the presence
(260.3 mV) and EPO (279.1 mV) samples was attributed to relatively
of chloride ions could provide insight of the corrosion process.
to their higher pitting susceptibility in PBS solution. This situation
Cr2 O3 + 3H2 O → 2Cr(OH)3 (3) could be simulated with the shift of surface potential toward Epit
from Ecorr and may only arise in the presence of oxidants such as
Cr2 O3 + H2 O + 2H+ → 2Cr(OH)2 + (4) dissolved oxygen and Fe(III) ions in the electrolyte.
The electrochemical behavior of inherent passive film devel-
Cr(OH)2 + + Cl− → Cr(OH)2 Cl (5)
oped over MP, EPO and EBPO samples at OCP was also determined
Reaction (3) could also be possible reaction at the defect site by EIS. The small potential amplitude (10 mVrms ) was applied to
(high surface energy region) where Cr2 O3 could hydrolyzed to avoid the occurrence of any faradaic irreversible reactions at the
Cr(OH)2 + . This could further promote the ingress of chloride ions surface and to evaluate the intrinsic characteristics of inherent pas-
toward the defect site (reaction (5)) and my initiate the pit forma- sive film. The Nyquist and Bode plots of the samples are shown
tion. in Figs. 8 and 9 respectively. The impedance spectra were ana-
To compare the pitting resistance (Eb − Ecorr ) and pitting sus- lyzed and simulated with the equivalent electrical circuit models
ceptibility (Ep − Ecorr ) quantitatively, the passive film breakdown (EEC) as shown in Fig. 8(b). The impedance as a function of fre-
potential (Eb ) and protection potential (Ep ) values were also deter- quency was dependent on the current response associated with
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 441

Fig. 8. (a) EIS Nyquist plots of 316L before and after electropolishing and (b) electrical equivalent circuit (EEC) diagram.

Fig. 9. EIS bode plots of 316L before and after electropolishing.

the reversible charge transfer characteristics when small AC poten- response. The high frequency regime of impedance spectra could
tial perturbation was applied. The distribution of charge could be be related with the surfacial charge relaxation in the double layer
divided into the double layer and interfacial charge transfer pro- represented by Ydl and transport through Rct . The intermediate and
cesses associated with the adsorption/desorption and passive film low frequency behavior could be affiliated with the intrinsic char-
442 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

Table 7
Kinetic parameters obtained after simulating of experimental EIS spectra with EEC models.

Rs ( cm2 ) Ydl( ␮S sn /cm2 ) n1 Rct ( cm2 ) Yf (␮S sn /cm2 ) n2 Rf ( cm2 ) Goodness of fit

MP 93.40 53.51 0.889 1.57 × 10 6


0.00773 0.624 1.45 × 104 4.859 × 10−3
EPO 104.2 55.13 0.920 279.7 21.63 0.909 8.097 × 106 7.404 × 10−3
EPBO 100.8 27.13 0.923 1.316 × 106 2.396 0.4248 1.242 × 1010 7.204 × 10−3

acteristics of passive film. The EEC model is presented in Fig. 8(b) The relatively high Rct values of both MP and EBPO but higher
which describes the individual parameters and their quantitative charge transport through the double layer (Ydl ) in the former sam-
information is given in Table 7. ple further validated the adverse effect by surface roughness. It can
The impedance spectra of MP, EPO and EBPO represented be deduced from the impedance analyses that the nature of pas-
differentiable potential distribution across the interface and elec- sive film and surface roughness are the critical parameters to ensure
trochemical response was dependent on the charge transfer electrochemical integrity of the 316L stainless steel in the PBS solu-
process (faradaic reactions) associated with the possible adsorp- tion. The impedance analyses confirmed the resistive properties of
tion/desorption of electrolyte species i.e. PO4 3− , Cl− , H+ and OH− passive film developed over EBPO sample and the results were in
etc. which may either help in the healing of defects in the pas- agreement with its low corrosion rate and relatively smaller passive
sive film or may enhance localized dissolution. The very high film current density in PBS solution as determined from PCP curves.
resistance (Rf ) (12420 M cm2 ) was presented by EBPO sample
compared to MP and EPO. This could be related with the better
barrier properties of passive film developed after electropolish- 3.4. Cell proliferation study
ing below the oxygen evolution regime. Similarly, the relatively
larger Rf of (EPO (8.097 M cm2 ) sample than MP 14.5 k cm2 also The proliferation and morphology of MC3T3 pre osteoblast
suggested the beneficial effects of electropolishing compared to cells were investigated on 316L stainless surfaces. The study was
mechanical polishing. But very small charge transfers resistance intended to examine topographic and wettability features of the
(Rct ) provided by EPO sample could be attributed to the instability passive film on cell proliferation. Topographic alterations could
of passive film in the PBS solution. This behavior may be correlated affect the roughness, wettability, surface energetics and surface
with the defective structure of the passive film on EPO sample. The chemistry, which could further affect the protein adsorption behav-
existence of Fe(II) and Fe(III) species in the passive film of elec- ior over the surface [38]. These parameters could also affect the cell
tropolished samples was confirmed from the XPS analysis and these morphology, organization, cell adhesion and osseointegration [39].
could be beneficial for the development of non-crystalline passive Generally, cell attachment and its standard morphology is the ini-
film on 316L stainless steel. Also the surface smoothness could sig- tial step in a cascade of cell–implant interaction and determine the
nificantly influence the stability of passive film. The relatively lower biocompatibility of the implant.
surface roughness of EBPO compared to EPO and MP could be the In Fig. 10, the cell proliferation rate on MP, EPO and EPBO after
reason for the its higher Rct and Rf values. Also the deficiency of iron 24 h was almost equal, however it can be observed that on MP, the
species (as found in the XPS spectra) in the passive film developed cells morphology showed peculiar structure in comparison with
over MP sample could also be the possible cause for its lower Rf the cells on EPO and EPBO. This might be attributed to the rel-
in PBS solution. The relatively higher constant phase element (Yf ) ative higher surface roughness (33.51 ± 5.54) of the MP sample.
for the MP (7.73 × 10−3 S sn /cm2 ) and EPO (2.16 × 10−5 S sn /cm2 ) It has been previously studied that rough surfaces could release
than EBPO (2.396 × 10−6 S sn /cm2 ) samples also indicated the large higher amount of ions in comparison with smooth surface [40].
amount of charge relaxation in the least resistive oxide passive film The leached ions from biocompatible material at low concentra-
which could be due to their higher surface roughness and defec- tion are considered non-toxic but they could significantly influence
tive structure respectively [37]. The existence of higher oxidation the cell proliferation [41]. It is worthy to mention that the surface
state Fe(II) and Fe(III) species could also promote the adsorption of roughness could provide better focal contact [42] and may lead to
OH− ions at the surface by forming FeOOH and/or Fe(OH)3 . These better adhesion while surfaces with lower roughness below 10 ␮m
species could enhance the barrier properties of passive film. Dur- may influence the biological interactions and protein adsorption
ing polishing samples above the oxygen evolution plateau (as EPO [43]. Hence, the EPO and EPBO presented better cell proliferation
sample), the electropolishing solution could be depleted with OH− and morphology in initial 24 h incubation period compared to MP.
ionic species by reaction (7), and may produce defective oxide pas- The cells also indicated the healthy and stellar morphology with
sive film. This behavior was confirmed by the very low Rct value pronounced protrusion at the surface of EPO and EPBO samples.
(279.7  cm2 ) provided by the EPO and could be simulated with the This could be due to the limited release of metal ions which could
rapid ingress of anionic species (OH− , PO4 3− etc.) available in the adversely affect the cell morphology and proliferation as is the
PBS solution toward the film. The higher charge transfer through case with MP sample. The low corrosion rate registered by EBPO
double layer via adsorption processes is found to be dependent on than MP and EPO also validated this behavior and PCP scans were
the nature of passive film. The higher values of Ydl and Yf as calcu- in agreement with the cell proliferation character which followed
lated from the simulation of impedance spectra of MP and EPO were the EPBO > EPO > MP trend. The relatively higher corrosion rate of
directly related with the charge transport through the distributed MP sample adversely affected the cell morphology as shown in
double layer and defective structure of the passive film respectively Fig. 10. The surface wettability is also another important factor
as given in Table 7. The very low Rct and relatively higher Rf , by the which could also influence the cell response. It has been previously
EPO than MP sample could therefore be associated with the pref- investigated that SFE was an important surface characteristic for
erential adsorption of ionic species and the thickening of the oxide cellular adhesion strength and proliferation [44]. From our wetta-
passive film respectively. bility data, it is calculated that the acidic components are lower
than basic components. These polar components could influence
the cell adhesion and proliferation. The lower fractional polarity
4OH− → 2H2 O + O2 + 4e− (7) (AB/(LW + AB)) might be expected to provide better cell prolif-
Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444 443

Fig. 10. MC3T3 Cell proliferation on 316L SS for 24 h and 48 h.

eration and optimal cytocompatibility [28]. In comparison EPBO relatively higher corrosion rate of MP sample and release of metal
provided lowest fractional polarity which could be attributed to its ions could be the possible reasons of its limited cell proliferation
better response toward cell proliferation than MP and EPO sam- characteristics.
ples. This could also be the reason for enhanced bioactive behavior
of electropolished surfaces than the mechanically polished one.
After 48 h of incubation the relatively higher proliferation rate was Acknowledgement
observed on all samples with no obvious difference and the surfaces
are almost covered by the cell monolayers. Z. Rahman would like to acknowledge financial support from
Science of Advanced Materials program at Central Michigan Uni-
4. Conclusions versity.

Electropolishing of 316L stainless steel at the oxygen evolu-


tion and below the oxygen evolution potentials was carried out References
to improve its surface roughness, topography and corrosion resis-
[1] S. Morais, J.P. Sousa, M.H. Fernandes, G.S. Carvalho, J.D. De Bruijn, C.A. Van
tance. The oxidation of 316L samples in the presence of H+ , H3 PO4 Blitterswijk, Effects of AISI 316L corrosion products in in vitro bone
and H2 SO4 during electropolishing could significantly modify the formation, Biomaterials 19 (1998) 999–1007.
surface passive film. The XPS analysis of the core peaks of O1s, Cr2p, [2] H. Hendra, R. Dadan, J.R.P. Djuansjah, Metals for biomedical applications, in:
R. Fazel (Ed.), Biomedical Engineering – From Theory to Applications, InTech
Fe2p, P2p and S2p elements confirmed the existence of Cr (III) Fe (II) Publisher, 2011, pp. 411–430.
and Fe(III) species at the surface of MP, EPO and EPBO samples. The [3] H. Hocheng, P.S. Kao, Y.F. Chen, Electropolishing of 316L stainless steel for
phosphate and sulfate species within the passive film were found anticorrosion passivation, J. Mater. Eng. Perform. 10 (2001) 414–418.
[4] C.-C. Shih, C.M. Shih, Y.-Y. Su, L.H.J. Su, M.-S. Chang, S.-J. Lin, Effect of surface
to be related with the possible oxidation of phosphorus or sulfur
oxide properties on corrosion resistance of 316L stainless steel for biomedical
content in the material and/or due to the adsorption of H3 PO4 , applications, Corros. Sci. 46 (2004) 427–441.
H2 PO4 − , H2 SO4 and HSO4 − species during the electropolishing pro- [5] Z. Rahman, L. Pompa, W. Haider, Influence of electropolishing and
magnetoelectropolishing on corrosion and biocompatibility of titanium
cess. The relatively higher surface free energy provided by MP
implants, Mater. Eng. Perform. 23 (2014) 3907–3915.
(51.81 mJ/m2 ) than EPO (47.52 mJ/m2 ) and EPBO (44.21 mJ/m2 ) was [6] Z.U. Rahman, W. Haider, L. Pompa, K.M. Deen, Electrochemical & osteoblast
in accordance with the decrease in surface roughness in the order adhesion study of engineered TiO2 nanotubular surfaces on titanium alloys,
of MP > EPO > EPBO as measured by AFM. The relatively positive Mater. Sci. Eng. C 58 (2016) 160–168.
[7] Z. ur Rahman, I. Shabib, W. Haider, Surface characterization and cytotoxicity
OCP, lower corrosion rate (0.484 mpy) and significantly large pas- analysis of plasma sprayed coatings on titanium alloys, Mater. Sci. Eng. C 67
sive film resistance (12430 M cm2 ) provided by EPBO sample than (2016) 675–683.
EPO and MP in the PBS solution confirmed the improvement in the [8] M.K. Lei, X.M. Zhu, In vitro corrosion resistance of plasma source ion nitrided
austenitic stainless steels, Biomaterials 22 (2001) 641–647.
electrochemical stability of passive film. [9] R. Rokicki, W. Haider, T. Hryniewicz, Influence of sodium hypochlorite
In comparison, EPBO also provided the lowest fractional polarity treatment of electropolished and magnetoelectropolished nitinol surfaces on
and surface roughness which could be beneficial for the cell prolif- adhesion and proliferation of MC3T3 pre-osteoblast cells, J. Mater. Sci. Mater.
Med. 23 (2012) 2127–2139.
eration and to increase the bioactivity of the surface. The enhanced [10] M. Rizwan, A. Ahmed, K.M. Deen, W. Haider, Electrochemcial behavior and
protrusion and stellar morphology of the cell was observed on the biological response of mesenchymal stem cells on cp-Ti after N-ions
EPO and EPBO samples during initial 24 h incubation period. The implantation, Appl. Surf. Sci. 320 (2014) 718–724.
444 Z. ur Rahman et al. / Applied Surface Science 410 (2017) 432–444

[11] B. Ollivere, C. Darrah, T. Barker, J. Nolan, M.J. Porteous, Early clinical failure of [29] K. Webb, V. Hlady, P.A. Tresco, Relative importance of surface wettability and
the Birmingham metal-on-metal hip resurfacing is associated with metallosis charged functional groups on NIH 3T3 fibroblast attachment, spreading, and
and soft-tissue necrosis, J. Bone Joint Surg. Br. 91 (2009) 1025–1030. cytoskeletal organization, J. Biomed. Mater. Res. 41 (1998) 422–430.
[12] S. Landgraeber, M. Jäger, J.J. Jacobs, N.J. Hallab, The pathology of orthopedic [30] L.C. Xu, C.A. Siedlecki, Effects of surface wettability and contact time on
implant failure is mediated by innate immune system cytokines, Mediators protein adhesion to biomaterial surfaces, Biomaterials 28 (2007) 3273–3283.
Inflamm. 2014 (2014) 9. [31] Hideo Nakae, Ryuichi Inui, Yosuke Hirata, Hiroyuki Saito, Effects of surface
[13] H. Nygren, Initial reactions of whole blood with hydrophilic and hydrophobic roughness on wettability, Acta Mater. 46 (1998) 2313–2318.
titanium surfaces, Colloids Surf. B: Biointerfaces 6 (1996) 329–333. [32] R.C.C. Wang, M.C. Hsieh, T.-M. Lee, Effects of nanometric roughness on surface
[14] C. Pulletikurthi, N. Munroe, D. Stewart, W. Haider, S. Amruthaluri, R. Rokicki, properties and fibroblast’s initial cytocompatibilities of Ti6Al4V,
M. Dugrot, S. Ramaswamy, Utility of magneto-electropolished ternary nitinol Biointerphases 6 (2011) 87.
alloys for blood contacting applications, J. Biomed. Mater. Res. B: Appl. [33] Y.J. Lim, Y. Oshida, C.J. Andres, M.T. Barco, Surface characterizations of
Biomater. 103 (2015) 1366–1374. variously treated titanium materials, Int. J. Oral Maxillofac. Surg. 16 (2001)
[15] N. Mackman, Triggers, targets and treatments for thrombosis, Nature 451 333.
(2008) 914–918. [34] S. Karimi, T. Nickchi, A.M. Alfantazi, Long-term corrosion investigation of AISI
[16] S. Habibzadeh, L. Li, D. Shum-Tim, E.C. Davis, S. Omanovic, Electrochemical 316L, Co-28Cr-6Mo, and Ti-6Al-4V alloys in simulated body solutions, Appl.
polishing as a 316L stainless steel surface treatment method: towards the Surf. Sci. 258 (2012) 6087–6096.
improvement of biocompatibility, Corros. Sci. 87 (2014) 89–100. [35] P. Sojitra, C. Engineer, D. Kothwala, A. Raval, H. Kotadia, G. Mehta,
[17] P. Gill, V. Musaramthota, N. Munroe, A. Datye, R. Dua, W. Haider, A. McGoron, Electropolishing of 316LVM stainless steel cardiovascular stents: an
R. Rokicki, Surface modification of Ni–Ti alloys for stent application after investigation of material removal, surface roughness and corrosion
magnetoelectropolishing, Mater. Sci. Eng. C 50 (2015) 37–44. behaviour, Trends Biomater. Artif. Organs 23 (2010) 115–121.
[18] K.B. Hensel, Electropolishing, Met. Finish 100 (2002) 425–433. [36] K.M. Deen, M.A. Virk, C.I. Haque, R. Ahmad, I.H. khan, Failure investigation of
[19] K. Rokosz, J. Lahtinen, T. Hryniewicz, S. Rzadkiewicz, XPS depth profiling heat exchanger plates due to pitting corrosion, Eng. fail Anal. 17 (4) (2010)
analysis of passive surface layers formed on austenitic AISI 304L and AISI 316L 886–893.
SS after high-current–density electropolishing, Surf. Coat. Technol. 276 (2015) [37] X.Y. Wang, Y.S. Wu, L. Zhang, Z.Y. Yu, Atomic force microscopy and X-ray
516–520. photoelectron spectroscopy study on the passive film for type 316L stainless
[20] S.J. Lee, J.J. Lai, The effects of electropolishing (EP) process parameters on steel, Corrosion 57 (2001) 540–546.
corrosion resistance of 316L stainless steel, J. Mater. Process. Technol. (2003) [38] S. Chatterjee, S. Sarkar, S. Bhattacharya, Toxic metals and autophagy, Chem.
206–210. Res. Toxicol. 27 (2014) 1887–1890.
[21] R.N. Wenzel, Resistance of solid surfaces to wetting by water, J. Ind. Eng. [39] C.N. Elias, Y. Oshida, J.H.C. Lima, C.A. Muller, Relationship between surface
Chem. (Washington, D.C.) 28 (1936) 988–994. properties (roughness, wettability and morphology) of titanium and dental
[22] A. Latifi, M. Imani, M.T. Khorasani, M.D. Joupari, Electrochemical and chemical implant removal torque, J, Mech. Behav. Biomed. Mater. 1 (2008) 234–242.
methods for improving surface characteristics of 316L stainless steel for [40] I. Degasne, M.F. Baslé, V. Demais, G. Huré, M. Lesourd, B. Grolleau, L. Mercier,
biomedical applications, Surf. Coat. Technol. 221 (2013) 1–12. D. Chappard, Effects of roughness, fibronectin and vitronectin on attachment,
[23] M.C. Biesinger, B.P. Payne, A.P. Grosvenor, L.W.M. Lau, A.R. Gerson, R.S.C. spreading, and proliferation of human osteoblast-like cells (Saos-2) on
Smart, Resolving surface chemical states in XPS analysis of first row transition titanium surfaces, Calcif Tissue Int. 64 (1999) 499–507.
metals, oxides and hydroxides: Cr, Mn, Fe, Co and Ni, Appl. Surf. Sci. 257 [41] J.Y. Wang, D.T. Tsukayama, B.H. Wicklund, R.B. Gustilo, Inhibition of T and B
(2011) 2717–2730. cell proliferation by titanium, cobalt, and chromium: role of IL-2 and IL-6, J.
[24] J.S. Chen, S.P. Lau, Z. Sun, G.Y. Chen, Y.J. Li, B.K. Tay, J.W. Chai, Effect of surface Biomed. Mater. Res. 32 (1996) 655–661.
properties on the wettability of iron containing amorphous carbon films, Int. [42] K.T. Bowers, J.C. Keller, B.A. Randolph, D.G. Wick, C.M. Michaels, Optimization
J. Mod. Phys. B 16 (2002) 1031–1040. of surface micromorphology for enhanced osteoblast responses in vitro, Int. J.
[25] W.Y. Lai, W.Z. Zhao, Z.F. Yin, J. Zhang, EIS and XPS studies on passive film of Oral Maxillofac. Implants 7 (1992) 302–310.
AISI 304 stainless steel in dilute sulfuric acid solution, Surf. Interface Anal. 44 [43] J.I. Rosales-Leal, M.A. Rodríguez-Valverde, G. Mazzaglia, P.J.
(2012) 418–425. Ramón-Torregrosa, L. Díaz-Rodríguez, O. García-Martínez, M.
[26] J.N. Butler, Ionic Equilibrium: Solubility and pH Calculations, John Wiley & Vallecillo-Capilla, C. Ruiz, M.A. Cabrerizo-Vílchez, Effect of roughness,
Sons, Inc., New York, 1998, pp. 162–178. wettability and morphology of engineered titanium surfaces on
[27] J.Y. Lim, M.C. Shaughnessy, Z. Zhou, H. Noh, E.A. Vogler, H.J. Donahue, Surface osteoblast-like cell adhesion, Colloids Surf. A: Physicochem. Eng. Asp. 365
energy effects on osteoblast spatial growth and mineralization, Biomaterials (2010) 222–229.
29 (2008) 1776–1784. [44] N.J. Hallab, K.J. Bundy, K. O’Connor, R.L. Moses, J.J. Jacobs, Evaluation of
[28] L. Ponsonnet, K. Reybier, N. Jaffrezic, V. Comte, C. Lagneau, M. Lissac, C. metallic and polymeric biomaterial surface energy and surface roughness
Martelet, Relationship between surface properties (roughness, wettability) of characteristics for directed cell adhesion, Tissue Eng. 7 (2001) 55–71.
titanium and titanium alloys and cell behaviour, Mater. Sci. Eng. C 23 (2003)
551–560.

You might also like