You are on page 1of 14

Materials Research Express

PAPER

A bilayer GO/nanofibrous biocomposite coating to enhance 316L


stainless steel corrosion performance
To cite this article: Fatemeh Khosravi et al 2019 Mater. Res. Express 6 086470

View the article online for updates and enhancements.

This content was downloaded from IP address 154.59.124.171 on 19/07/2019 at 09:18


Mater. Res. Express 6 (2019) 086470 https://doi.org/10.1088/2053-1591/ab26d5

PAPER

A bilayer GO/nanofibrous biocomposite coating to enhance 316L


stainless steel corrosion performance
RECEIVED
27 April 2019
REVISED
31 May 2019
ACCEPTED FOR PUBLICATION
Fatemeh Khosravi1, Saied Nouri Khorasani1 , Erfan Rezvani Ghomi1, Mohsen Karimi Kichi2,
4 June 2019 Hamid Zilouei1, Mousa Farhadian2 and Rasoul Esmaeely Neisiany1,3
1
PUBLISHED Department of Chemical Engineering, Isfahan University of Technology, Isfahan, 8415683111, Iran
19 June 2019 2
Department of Material Engineering, Isfahan University of Technology, Isfahan, 8415683111, Iran
3
Division of Materials Science, Luleå University of Technology, SE-97187 Luleå, Sweden
E-mail: saied@cc.iut.ac.ir

Keywords: 316L stainless steel, graphene oxide, coating, nanofibers, corrosion resistance
Supplementary material for this article is available online

Abstract
A bilayer coating has been synthesized to be coated on the 316L stainless steel (SS) for bone implant
application. The first layer consisted of graphene oxide (GO) which was coated via the electrophoretic
deposition method. The second layer including Poly (ε-caprolactone) (PCL)/Gelatin-forsterite
nanofibers was electrospun on the first layer. The morphology of the bare 316L SS, GO-coated,
electrospun nanofibers, and nanofibers-coated samples were investigated using scanning electron
microscopy (SEM). The electrospun nanofibers were also characterized by Fourier transform infrared
spectroscopy (FTIR) and confirmed the presence of PCL, gelatin, and forsterite in the nanocomposite
coating. Furthermore, the morphological investigation of the nanofibers revealed that 80:20 weight of
PCL to gelatin did not show any beads, making them for coating on the GO coatings. In addition, the
corrosion behavior of the coated samples was assessed by potentiodynamic polarization and
electrochemical impedance spectroscopy (EIS). The samples coated with GO and GO/PCL-gelatin-
forsterite 1% showed the best corrosion resistance in comparison with other samples. Consequently,
the prepared bilayer biocomposite coating including 1 wt% forsterite nanoparticles can be a promising
candidate for orthopedic implants.

1. Introduction

Medical implants play an important role in bone fractures treatments. Nowadays, there is a considerable
growing demand for such implants due to the increasing number of orthopedic surgeries. Modern medical
implants are expected to have longer lifetimes and enhanced biological and mechanical performance [1, 2]. So
far, metal alloys have been highly noticed because of their mechanical and biocompatible properties [3, 4]. Bio-
inert metallic materials, as a substitute for high-bearing bones, are broadly employed due to their great
mechanical strength, acceptable corrosion resistance and biocompatibility [5, 6]. Among them, 316L SS is used
extensively in therapeutic applications, particularly as a temporary implant, due to its cost-efficient price and
preferable properties [7–9]. It should be noted that the release of ions such as molybdenum, nickel, chromium
after locating at the definite place, the relative poor corrosion resistance, and also far different strength between
metal and the body tissues, are important disadvantages of utilizing 316L SS in the body [10–13]. There are
several methods to overcome and diminish the drawbacks of 316L SS, such as ceramic and polymeric composite
coatings [14]. Composite coatings expected to improve corrosion performance owing to decreased crack
formation and defects which negates penetration of corrosive ions to the substrate [15]. Different bioactive
ceramics such as hydroxyapatite (HA), bioglass, akermanite, and forsterite [16–19] have been incorporated in
composite coatings to improve the coatings. Recently, researchers have focused on the medical application of
forsterite (chemical formula: Mg2SiO4) as an important bioactive ceramic. Forsterite is biocompatible in nature
with great mechanical properties which make it suitable for employing in metallic implants as well as scaffolds.

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 086470 F Khosravi et al

This nanoparticle significantly enhanced the mechanical properties and corrosion resistance and of the implants
and scaffolds [20, 21]. Furthermore, various natural-based and synthetic polymers have been used as the
matrixes of coatings. Among them, gelatin has been applied widely in medical application and tissue engineering
due to its good adhesiveness, low cost, availability and hemostatic properties. In addition, gelatin originates from
collagen which is an effective material in bone regeneration [22–25]. It is proved that utilizing gelatin in coatings
leads to homogenous coatings with good adhesion [26], however, gelatin coatings show weak mechanical
properties [27, 28]. On the other hand, PCL is used as a great candidate in the field of tissue engineering due to its
biocompatibility, bioresorbable, and good mechanical properties [29]. However, PCL interaction with
biological environments is not strong enough, which leads to lower cell adhesion as well as less proliferation.
Consequently, PCL/gelatin blend is a promising candidate which provides the advantages of both synthetic and
natural polymers properties. It was reported that the 1:1 PCL/gelatin electrospun presented higher
compatibility and proliferation [30]. In addition, 50:50 wt/wt and 70:30 wt/wt PCL/gelatin were electrospun
and their degradation behaviors were evaluated [31]. The results confirmed that 50:50 PCL/gelatin showed
faster degradation rate and lower mechanical properties comparing to 70:30 PCL/gelatin.
Furthermore, in recent studies, bilayer coatings have been fabricated which are seeking superior corrosion
resistance, biological and mechanical properties [32, 33]. GO is a promising material which has been widely used
to enhance the corrosion resistance of different metallic structures [34]. The most common technique to deposit
GO on the surface of substrates is electrophoretic deposition (EPD) due to its environmental friendly feature. In
addition, EPD does not affect the microscopic properties of substrates because it is utilized at room temperature.
Furthermore, due to the existence of various functional groups in GO chemical structure, EPD is a feasible
technique to deposit GO from aqueous solutions [19, 35].
This study aims to prepare bilayer nanofibrous biocomposite coatings consisting of GO/PCL-gelatin-
forsterite with various weight ratios of gelatin and PCL as well as a different forsterite weight percent (0, 1 and 3).
The first layer (GO sheet) is coated on the surface of 316L SS using the EPD method. The corrosion behavior of
316L SS was significantly boosted by GO layer, while the functional groups, i.e., hydroxyl, and carboxylic,
provides a higher adhesion with the polymeric layer. The second layer (PCL-gelatin-Forsterite) was prepared
using the electrospinning method to achieve a better distribution of the forsterite nanoparticles on the first layer.
The effect of employing GO layer and the effects of various forsterite concentrations on corrosion resistance
performance of the bilayer nanofibrous biocomposite coating were investigated.

2. Materials and methods

2.1. Forsterite nanoparticles preparation


The solutions were firstly prepared by using Magnesium nitrate hexahydrate (Mg(NO3)2.6H2O, Merck Co.,
99.99% purity) and colloidal silica (SiO2, 34 wt% solid fraction, Sigma-Aldrich Co.) in a molar ratio of Mg/
Si=2/1 mol. The aqueous solution of magnesium nitrate prepared with a concentration of 0.284 mol l−1.
Subsequently, colloidal silica was poured into the prepared solution. Sucrose (Merck, 99.9% purity) solution was
prepared with sucrose to metal molar ratio of 4:1 and then poured to the precursor solution and homogenized
by magnetically stirring for 2 h. Afterward, poly (vinyl alcohol) (PVA) (Merck, molecular weight=72 000)
solution prepared and then added to the previous solution. The pH value of the final solution was adjusted to 1
by adding nitric acid. The final solution stirred for 2 more hours to achieve a homogenous solution. The
prepared solution was then heated to 80 °C for 2 h and preserved for 24 h and dried at 200 °C. Finally; the dried
gel was calcinated at 800 °C for 2 h. This procedure was also reported in previous research [36].

2.2. Surface preparation of the substrates and GO deposition


316L SS samples were cut into rectangular pieces with dimensions of 20×10×4 mm and used as substrates
and were mechanically polished from 80 to 320 grit SiC papers. Afterward, the samples were washed with
deionized (DI) water. Subsequently, the samples were sonicated in acetone in order to clean the surfaces
impurities and greasy materials. Finally, the 316L SS plates were dried at room temperature.
GO (Nanosany Corporation, Iran) nanopowder with a thickness of 3.4–7 nm and an average number of
layers of 6–10, was added to DI water to make various concentrations. This step followed by ultra-sonication to
prepare homogenous solutions. Several concentrations were examined to choose the best concentration for the
EPD. GO nanoparticles were deposited onto 316L SS electrophoretically while; it was employed as both the
counter and working electrode of the EPD. The anode and cathode were chosen to be the working and counter
electrodes, respectively. The electrodes were linked to a DC power supply and submerged in the EPD solution.
EPD was carried out at different voltages and times of deposition process. To adjust the applied voltage, the
deposition time was set at 5, 10 or 15 min. In the same way, to adjust the deposition time, the voltage was set at 3

2
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Table 1. The composition of the solutions.

Sample codes So.1 So.2 So.3 So.4 So.5 So.6 So.7 So.8 So.9

PCL to gelatin ratio (wt%) 100:0 100:0 100:0 80:20 80:20 80:20 60:40 60:40 60:40
Forsterite content (wt%) — 1 3 — 1 3 — 1 3

or 4 V. It should be noted the anode and cathode were located at 1 cm from each other. The coated samples were
dried at room temperature. This procedure was also reported in previous work [37].

2.3. Preparation of the PCL -gelatin-forsterite coating


Prior to coating process, the solutions including three weight ratios of PCL (average Mw=80 000, Sigma) and
gelatin (type B bovine skin, Mw=50 000–100 000) with varied weight percent of forsterite nanoparticles were
provided employing formic acid and acetic acid (1:3 v/v) as a solvent. Full details of solution compositions are
summarized in table 1. Firstly, forsterite nanoparticles were dissolved in solvents by 2 h sonication. Then,
various amounts of polymer contents, table 1, were added to the solutions and magnetically stirred for 6 h. A
syringe was filled with the solutions and then the solutions were pumped by a syringe pump to control the feed
rate of the solutions during the electrospinning in a 15 cm distance of needle tip to the collector. The solutions
were moved into the needle at different flow rates and positive high voltages using a high voltage power supply to
achieve the optimum electrospinning parameters of the solutions.

2.4. Characterization and evaluation


2.4.1. Morphological and chemical characterizations
The morphology of the surface of coated specimens and prepared nanofibers was evaluated by scanning electron
microscopy (SEM, XL30, Philips Corporation). Furthermore, the chemical composition of the bare and GO-
coated 316L SS plates was examined by an energy dispersive x-ray spectrum (EDAX Silicon Drift). Prior to SEM
investigations, the specimens were coated by a thin layer of gold utilizing sputter coater (SCD 005). To recognize
the characteristic peaks in the gelatin, PCL, forsterite structures, as well as nanocomposite characteristic bonds,
Fourier transform infrared spectroscopy (Ray-Leigh WQF-510A FTIR) was employed in the range
4000–400 cm−1 using the KBr pellets method.

2.4.2. Electrochemical corrosion assessment


The area of the working electrode exposed to the PBS solution was 1 cm2. The 316L SS coated samples were
immersed into the PBS solution and allowed to be stabilized over 1 h. EIS measurements were performed due to
respective OCP of the related working electrode over the frequency range of 10–2 Hz to 10+5 Hz while the
sinusoidal amplitude was set at 10 mV. The impedance and potentiodynamic polarization were collected using a
PARSTAT2273 apparatus.
Electrochemical behavior of the samples was analyzed in standard Phosphate-buffered saline (PBS) at room
temperature and physiologic pH 7.4. A standard three-electrode setup arrangement was utilized. The working
electrode was 316L SS coated samples, while; a platinum sheet employed as the counter electrode and an Ag/
AgCl in saturated KCL used as the reference electrode. Firstly, the electrodes were polarized over the potential
range of±250 mV with respect to the OCP values of the coatings at a scan rate of 1 mV.s−1.

3. Results and discussion

3.1. Characterizations of bilayer coating


Figure 1 indicates the SEM images of bare 316L SS (A) and GO-coated (B) on 316L SS plates, fabricated under 4 V
in 10 min SEM images showed the surface of 316L SS substrate was coated by GO layer relatively homogenous
and thoroughly. The EDAX results confirmed that the elements appeared in figure 1(A) such as C, Fe, and Ni
were disappeared after GO coating (figure 2(B)). Furthermore figure 1(c) shows the SEM micrograph of GO
layer cross-section.
The PCL/gelatin solutions were electrospun at various concentrations as shown in table 1. In order to
choose the optimum voltage, each particular solution (So.1 to So.9) was spun at the range of 12–26 kV. By
making an analogy among all samples, it can be concluded that the concentration of 80 wt% of PCL (So.4 to
So.6) showed a thinner average of diameter nanofiber. It can be possibly attributed to the addition of gelatin to
the solutions. However, by increasing the percent of gelatin to 40 wt% (So.7 to So.9) the diameter of fibers
increased due to increases in the solution viscosity. figure S1 and S2 is available online at stacks.iop.org/MRX/
6/086470/mmedia show the SEM images and diameter distribution of So.1 to So.3 and So.7 to So.9,

3
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 1. The SEM images, EDAX micrographs and elemental composition of (A) 316L SS substrate (B) GO layer coated on 316L SS.
(C) SEM image of the cross-section of the GO layer.

4
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 2. The SEM image and nanofiber diameter distribution prepared from solutions of (A) So.4, (B) So.5, and (C) So.6.

respectively. This is quite surprising that the increase in the gelatin content of PCL/gelatin solutions caused an
increase in the diameter of the prepared nanofibers. This phenomenon can be justified by the presence of
emulsion which is more stable at higher gelatin contents [28]. It seems that gelatin to PCL ratio has an optimal
proportion and the content of PCL is bigger than gelatin. Moreover, all nanofibrous biocomposites containing 1
wt% of forsterite nanoparticles had a smaller average of diameter can be attributed to an increase in solution
conductivity after incorporation of the nanoparticles [38]. Based on the fibers’ morphology, the polymer
solutions with 80 wt% of PCL concentration with 15 cm needle tip-to-collector distance were selected to provide
bead-free nanofibers. Furthermore, the selected concentration revealed the minimum fiber diameter, more
uniformity, and no sign of beads, confirming the selected PCL/gelatin ratio as the best concentration for using as
the second layer and subsequently was evaluated amoung the furthur tests [39]. Figure 2 displays the SEM images
and the nanofibers’ diameter distribution of So.4 to So.6 electrospun nanofiber. For all samples, 50
measurements were performed to obtain the average, standard deviation, and diameter distribution of the
nanofiber’s diameter and the distance between the needle tip and collector was 15 cm. The mean value of the
So.4, So.5 and So.6 nanofiber’s diameters, at the feed rate of 0.1, 0.1 and 0.5 ml.h−1 and an applied voltage of 22,
13, 13 kV, were determined to be 167, 148 and 171 nm, respectively. The standard deviations of the
aforementioned nanofibers are determined to be 29, 36 and 43 nm, respectively. The rest SEM images of the
samples, as well as their diameter distribution, are presented in the supplementary file.

5
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 3. FTIR spectra of neat PCL, neat gelatin, neat forsterite nanoparticles and prepared nanofibers from So.4, So.5, and So.6.

Figure 3 represents the FTIR spectra of neat PCL nanofiber, neat gelatin powder, neat forsterite
nanoparticles and prepared nanofibers from So.4, So.5, and So.6. For neat PCL FTIR spectrum, two basic bands
at 1180 and 1730 cm−1 are attributed to the ether groups (C–O–C) stretching vibrations and the carboxyl (C=O)
stretching vibrations, respectively. Furthermore, the peaks at 2860 cm−1 and 2950 cm−1 are attributed to
symmetric and symmetric C–H stretching groups in PCL chemical structure which are also observed in FTIR
spectra of the prepared nanofibers from So.4, So.5, and So.6 solutions. The FTIR spectrum of the neat gelatin
shows the characteristic peaks at 1650 cm−1 (C=O stretching), 1555 cm−1 (N–H bend and C–H stretching), and
3430 cm−1 (N–H stretching vibration) corresponded to amide I, amide II and amide A, respectively [40]. These
peaks can be also detected in the FTIR spectra of the prepared nanofibers from So.4, So.5, and So.6. Similar to
PCL, the characteristic peaks at 2855 cm−1 and 2925 cm−1 attributed to symmetric and symmetric C–H
stretching groups in the neat gelatin and the nanofibers containing gelatin. The incorporation of forsterite
nanoparticles can be investigated from FTIR spectra. The absorption broad peaks of forsterite can be discerned
at 1000–800 cm−1 related to Si–O–Si bond [41] in the neat forsterite FTIR spectrum, and proving the forsterite
presentation in the nanocomposite coating. Furthermore, there is a sharp peak at 3700 cm−1 attributed to OH
band in forsterite nanoparticles [15].

3.2. Electrochemical test results


The electrochemical analyses were performed in the PBS solution at room temperature for one-layer coating of
GO-coated substrate and nanofibrous biocomposites and for the bilayer coating of GO/ nanofibrous
biocomposite coatings. Similar to the aforementioned characterizations, the nanofibers from solutions of So.4,
So.5, and So.6 were coated on the surfaces of bare 316L SS and Go-coated substrates. Potentiodynamic
polarization curves are presented in figure 4 and the potentiodynamic polarization parameters are summarized
in table 2. In order to better comparison among the results, the obtained results were shown in three figures but
the same scales.
From figure 4, it can be observed that the corrosion potential (Ecorr) of the coated samples shifted toward
more positive potential values comparing to neat 316L SS substrate with Ecorr of −342.13 mV which confirms
the electrochemical reactions would occur harder [42]. In other words, the electrochemical reaction of the
bilayer GO/nanofibrous coatings was the least feasible reaction among the other coated samples. In addition,
increasing the amount of forsterite nanoparticles caused an increase in Ecorr , confirming a higher corrosion
resistance of them. Meanwhile, the determined corrosion current density, Icorr, which is summarized in table 2
confirms a decrease when the substrate was protected by the coatings in comparison to the results for bare 316L
SS substrate with Icorr of 0.59 μA cm−2. In other words, the charge transfer (Icorr) showed the lower amounts
when the coating acts as a barrier versus corrosive media which is attacking the 316L SS substrate [43]. In fact, the
addition of forsterite nanoparticles until 1 wt% caused a decrease in the Icorr due to filling the existing micro-
cracks of the samples. However, the Icorr enhanced again by 3 wt% incorporation of forsterite nanoparticles
which increases the roughness and hydrophilicity of the coatings which makes them susceptible to pitting

6
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 4. Potentiodynamic polarization curves of 316L SS substrate, GO-coated, electrospun coated and GO/nanofibrous
biocomposite coating containing (A) 0%, (B) 1 wt%, and (C) 3 wt% of forsterite nanoparticles.

corrosion [42]. Therefore, the most corrosion resistant coatings among 7 coatings are GO and GO/PCL-
Gelatin-forsterite 1% by Icorr of 0.006 and 0.003 μA cm−2 which are presented in table 2.
The Nyquist diagrams for the coatings were plotted and well fitted with the three various equivalent electrical
circuits (EEC) models using the Zview version 3.10 software. Figure 5(A) is related to fit samples of PCL-Gelatin-
forsterite 0, 1, and 3%, GO/PCL-Gelatin-forsterite 0, 1% because these coatings were fitted truly in a one-loop

7
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Table 2. Determined parameters from potentiodynamic polarization


curves.

Sample Icorr (μA cm−2) Ecorr (mV)

316L SS substrate 0.59 −342.13


GO 0.003 101.26
PCL-gelatin 0.046 −103.74
PCL-gelatin-forsterite 1% 0.037 −107.20
PCL-gelatin-forsterite 3% 0.094 −203.64
GO/PCL-gelatin 0.012 97.53
GO/PCL-gelatin-forsterite 1% 0.006 7.37
GO/PCL-gelatin-forsterite 3% 0.017 135.79

Figure 5. EECs used to fit Nyquist and bode plots. (A) 316L SS substrate, PCL-Gelatin-forsterite 0, 1, and 3%, GO/PCL-Gelatin-
forsterite 0, 1%, (B) GO/PCL-Gelatin-forsterite 3%, (C) GO.

EES model. Figure 5(B) is attributed to GO/PCL-Gelatin-forsterite 3% coating which has one completed loop
and a half loop with 45-degree slope with the horizontal axis which is considered as Warburg element. The GO
coating showed a two-loop EES after fitting the experimental plot which is presented in figure 5(C). The Nyquist
plots and Bode modulus are shown in figures 6 and 7, respectively. The determined EIS parameters are
summarized in table 3 and defined as follows:
Rs: the resistance of the electrolyte,
Rct: the electrode-electrolyte interface charge transfer resistance,
CPEdl : the double-charge layer constant phase element
Rcoat: the resistance of the outer layer of the prepared coating,
CPEcoat : the capacity of the outer layer of the prepared coating,
RP : given from Rct+Rcoat as polarization resistance.
Simulating of EEC by lowest errors needs a replacement of the constant phase element by double-layer
capacitance element (CPEdl) which is attributed to the coating inhomogeneity. The maximum amount of Rp
indicates the most resistant of the coating to corrosion. In this research, the GO layer showed the highest
resistant coating due to its maximum amount of RP, as shown in table 3. The bilayer capacitance shows the
accumulated charge capacity of the coatings. The increase in the CPEdl leads to a surface to be vulnerable to the
corrosion reaction that occurred on the surface. Hence, a lower amount of CPEdl in EIS parameters could be a
sign of better corrosion resistance. Therefore, a bilayer of GO accompanying with nanocomposite significantly
increased the corrosion resistance of the coating. According to table 3, the 3 bilayer coating samples showed a
higher total of RP than the 3 other samples without GO coating, which is also observed in figure 6. In this figure,
the loop length of the three bilayer coating samples are larger in general and the largest loop belongs to the
sample of GO/PCL-Gelatin-forsterite 1% due to the performance of GO coating and the sufficient amount of
forsterite nanoparticles incorporation [44, 45]. Furthermore, the increase in the amount of forsterite to 3 wt%
practically led to the appearance of an increment of susceptive pitting corrosion zone. Thus, the GO/PCL-
Gelatin-forsterite 1% showed the most corrosion resistant performance.
In addition, according to figure 7, the amount of |Z| for GO/PCL-Gelatin-forsterite 1%sample is higher
than the others which proves that 1 wt% incorporation in the amount of forsterite causes an increase in the
corrosion performance [46]. Furthermore, it can be observed that by increasing forsterite nanoparticles in the
GO/PCL-gelatin-forsterite 3 wt% sample, the Warburg Impedance occurred which can be seen in figure 6(c). It
should be mentioned that the low-frequency area belongs to the double layer or interface of the nanofibrous
biocomposite layer with GO coating. It is inferred from EIS plots that corrosive media is infiltrated to first in
contact coating (nanofibrous biocomposite) which possibly attributed to non-uniformity distribution of
forsterite nanoparticles and the high content of incorporation (3 wt%) which was also reported by other
researches [47–49]. It made the coating susceptible to corrosion [46] and subsequently involved by the second
coating (GO coating). This phenomenon generally claims the diffusion of the electrolyte solution creates

8
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 6. Nyquist curves of 316L SS substrate, GO-coated, electrospun coated, and GO/nanofibrous biocomposite coating containing
(A) 0 wt%, (B) 1 wt%, and (C) 3 wt% of forsterite nanoparticles.

impedance which is mentioned as Warburg Impedance. In addition, the Nyquist plot in figure 6(c) reveals that
the GO/nanofibrous biocomposite containing 3 wt% of forsterite nanoparticles in low frequency has a linear
shape with an angle of approximately 45° by the horizontal line which confirms the great corrosion resistance of
the second layer (GO layer). It should be noted that equation (1) can be employed to infinite the Warburg
Impedance at the lower frequencies.

9
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

Figure 7. Bode modulus curves of 316L SS, GO-coated, electronspun coated, and GO/nanofibrous biocomposite coating containing
(A) 0 wt%, (B) 1 wt%, and (C) 3 wt% of forsterite nanoparticles.

zw = 2 s / w1 / 2 (1)

Where w is radial frequency and σ is the Warburg coefficient, which can be obtained from the Warburg curve
slope, or by fitting to an EEC model including Warburg Impedance (figure 5(b)). Other parameters calculated
from EIS, and summarized in table 3, are Tw and Pw. Tw=L2/D has a unit of second (s) by means of system
diffusion time, L is the effective diffusion thickness, and D is the diffusion coefficient by the unit of cm2 s−1. Pw is

10
Mater. Res. Express 6 (2019) 086470
Table 3. The impedance parameters obtained by fitting the experimental data using the above-mentioned EEC model.

Sample Rs (Ω.cm2) Rct (Ω.cm2) CPEdl (sn.Ω−1.cm−2) nct Rcoat (Ω.cm2) ZW (Ω.s−1/2 .cm2) CPEcoat (sn.Ω−1.cm−2) TW (second) ndl PW
11

316L SS 23 1.1×105 9.31×10–5 0.79 N/A N/A N/A N/A N/A N/A
GO-coated 137 7.20×108 1.22×10–5 0.93 9461 N/A 1.78×10–5 N/A 0.75 N/A
Elctrospun coated 135 3.97×105 4.57×10–5 0.82 N/A N/A N/A N/A N/A N/A
GO/Elctrospun coated 126 8.82×105 2.26×10–5 0.81 N/A N/A N/A N/A N/A N/A
Elctrospun (1%) coated 169 1.1×106 8.48×10–6 0.71 N/A N/A N/A N/A N/A N/A
GO/Elctrospun (1%) coated 146 1 ×108 2.16×10–6 0.60 N/A N/A N/A N/A N/A N/A
Elctrospun (3%) coated 132 8.83×105 2.12×10–5 0.82 N/A N/A N/A N/A N/A N/A
GO/Elctrospun (3%) coated 179 1.4×105 4.97×10–11 0.99 N/A 9.70×105 N/A 46.94 N/A 0.62

F Khosravi et al
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

the exponential factor of diffusion by the value of 0.5 for finite thickness systems such as coatings by the value of
0.6 [50].

4. Conclusions

In conclusion, the novel bilayer nanocomposite GO/biocomposite coatings including forsterite nanoparticles
were coated on 316L SS substrate employing EPD and electrospinning methods. FTIR and SEM were used to
investigate the chemical structure and morphology of nanofibers, respectively. The results confirmed the
existence of forsterite nanoparticles in the nanocomposite electrospun coating, and also the best weight ratio of
PCL to gelatin to be 80:20 to achieve the thinnest nanofibers without any bead. Moreover, the corrosion
evaluation tests verified excellent corrosion resistance of the bilayer GO/biocomposite coated samples and one-
layer GO coating sample compared to the bare 316L SS substrate. To summarize, the bilayer GO/biocomposite
with 1 wt% forsterite nanoparticle coating can be offered as a novel bilayer corrosion resistant coating to be more
investigated to make them a promising candidate as a bone implant application.

ORCID iDs

Saied Nouri Khorasani https://orcid.org/0000-0002-8093-6735

References
[1] Mishnaevsky L et al 2014 Nanostructured titanium-based materials for medical implants: modeling and development Materials Science
and Engineering: R: Reports 81 1–19
[2] Salek N, Hadizadeh M, Hosseini S A, Daneshkazemi A R and Kouhi M 2018 An investigation into the three-point bending properties
and the vickers microhardness of dental composites reinforced with nylon 66 nanofibers Mater. Res. Express 5 105401
[3] Devi K B, Singh K and Rajendran N 2011 Sol–gel synthesis and characterisation of nanoporous zirconium titanate coated on 316L SS
for biomedical applications Journal of Sol-Gel Science Technology 59 513
[4] Srinivasan A, Ranjani P and Rajendran N 2013 Electrochemical polymerization of pyrrole over AZ31 Mg alloy for biomedical
applications Electrochim. Acta 88 310–21
[5] Sutha S, Kavitha K, Karunakaran G and Rajendran V 2013 In-vitro bioactivity, biocorrosion and antibacterial activity of silicon
integrated hydroxyapatite/chitosan composite coating on 316L stainless steel implants Materials Science and Engineering: C 33 4046–54
[6] Tiyyagura H R, Fuchs-Godec R, Gorgieva S, Arthanari S, Mohan M K and Kokol V 2018 Biomimetic gelatine coating for less-corrosive
and surface bioactive Mg–9Al–1Zn alloys J. Mater. Res. 33 1449–62
[7] Srinivasan A and Rajendran N 2015 Surface characteristics, corrosion resistance and MG63 osteoblast-like cells attachment behaviour
of nano SiO2–ZrO2 coated 316L stainless steel RSC Adv. 5 26007–16
[8] Edathazhe A B and Shashikala H D 2018 Corrosion resistance and in-vitro bioactivity of BaO containing Na2O–CaO–P2O5 phosphate
glass-ceramic coating prepared on 316 L, duplex stainless steel 2205 and Ti6Al4V Mater. Res. Express 5 035404
[9] Munis A, Zheng M and Zhao T 2019 Effect of sulfate and meta-silicate ions on pitting corrosion of stainless steel-316 in chloride
containing simulated coal gasifier aqueous effluents Mater. Res. Express 6 076541
[10] Subramanian B, Ananthakumar R, Kobayashi A and Jayachandran M 2012 Surface modification of 316L stainless steel with magnetron
sputtered TiN/VN nanoscale multilayers for bio implant applications J. Mater. Sci., Mater. Med. 23 329–38
[11] Le V Q, Cochis A, Rimondini L, Pourroy G, Stanic V, Palkowski H and Carradò A 2013 Biomimetic calcium–phosphates produced by
an auto-catalytic route on stainless steel 316L and bio-inert polyolefin RSC Adv. 3 11255–62
[12] Suresh R, Shruthi P, Kumar R S, Siva J, Ananth M P and Ramesh R 2014 Experimental investigation of nano-composite coated stainless
steel (316L) surfaces under unidirectional sliding Applied Mechanics and Materials 440 37–41
[13] Madhan Kumar A, Nagarajan S, Ramakrishna S, Sudhagar P, Kang Y S, Kim H, Gasem Z M and Rajendran N 2014 Electrochemical and
in vitro bioactivity of polypyrrole/ceramic nanocomposite coatings on 316L SS bio-implants Materials Science and Engineering: C 43
76–85
[14] Madhan Kumar A and Rajendran N 2013 Electrochemical aspects and in vitro biocompatibility of polypyrrole/TiO2 ceramic
nanocomposite coatings on 316L SS for orthopedic implants Ceram. Int. 39 5639–50
[15] Frajkorová F, Molero E and Ferrari B 2015 Electrophoretic deposition of gelatin/hydroxyapatite composite coatings onto a stainless
steel substrate Key Eng. Mater. 654 195–9
[16] Kheirkhah M, Fathi M, Salimijazi H R and Razavi M 2015 Surface modification of stainless steel implants using nanostructured
forsterite (Mg2SiO4) coating for biomaterial applications Surf. Coat. Technol. 276 580–6
[17] Enayati M S, Neisiany R E, Sajkiewicz P, Behzad T, Denis P and Pierini F 2019 Effect of nanofiller incorporation on thermomechanical
and toughness of poly (vinyl alcohol)-based electrospun nanofibrous bionanocomposites Theor. Appl. Fract. Mech. 99 44–50
[18] Kouhi M, Jayarama Reddy V, Fathi M, Shamanian M, Valipouri A and Ramakrishna S 2019 Poly (3-hydroxybutyrate-co-3-
hydroxyvalerate)/fibrinogen/bredigite nanofibrous membranes and their integration with osteoblasts for guided bone regeneration J.
Biomed. Mater. Res. Part A 107 1154–65 107 1154-65
[19] Farrokhi-Rad M, Beygi Khosrowshahi Y, Hassannejad H, Nouri A and Hosseini M 2018 Preparation and characterization of
hydroxyapatite/titania nanocomposite coatings on titanium by electrophoretic deposition Mater. Res. Express 5 115004
[20] Tavangarian F and Emadi R 2011 Improving degradation rate and apatite formation ability of nanostructure forsterite Ceram. Int. 37
2275–80
[21] Naghiu M A, Gorea M, Mutch E, Kristaly F and Tomoaia-Cotisel M 2013 Forsterite nanopowder: structural characterization and
biocompatibility evaluation Journal of Materials Science & Technology 29 628–32
[22] Thomazine M, Carvalho R A and Sobral P J A 2005 Physical properties of gelatin films plasticized by blends of glycerol and sorbitol
J. Food Sci. 70 E172–6

12
Mater. Res. Express 6 (2019) 086470 F Khosravi et al

[23] Karim A A and Bhat R 2008 Gelatin alternatives for the food industry: recent developments, challenges and prospects Trends in Food
Science & Technology 19 644–56
[24] Mendieta-Taboada O, Sobral P J d A, Carvalho R A and Habitante A M B Q 2008 Thermomechanical properties of biodegradable films
based on blends of gelatin and poly(vinyl alcohol) Food Hydrocolloids 22 1485–92
[25] Rezvani Ghomi E, Khalili S, Nouri Khorasani S, Esmaeely Neisiany R and Ramakrishna S 2019 Wound dressings: current advances and
future directions J. Appl. Polym. Sci. 136 47738
[26] Xu X, Lu P, Guo M and Fang M 2010 Cross-linked gelatin/nanoparticles composite coating on micro-arc oxidation film for corrosion
and drug release Appl. Surf. Sci. 256 2367–71
[27] Torkaman R, Darvishi S, Jokar M, Kharaziha M and Karbasi M 2017 Electrochemical and in vitro bioactivity of nanocomposite gelatin-
forsterite coatings on AISI 316 L stainless steel Prog. Org. Coat. 103 40–7
[28] Saadatkish N, Nouri Khorasani S, Morshed M, Allafchian A-R, Beigi M-H, Masoudi Rad M, Esmaeely Neisiany R and
Nasr-Esfahani M-H 2018 A ternary nanofibrous scaffold potential for central nerve system tissue engineering J. Biomed. Mater. Res.
Part A 106 2394-401
[29] She H, Xiao X and Liu R 2007 Preparation and characterization of polycaprolactone-chitosan composites for tissue engineering
applications J. Mater. Sci. 42 8113–9
[30] Yao R, He J, Meng G, Jiang B and Wu F 2016 Electrospun PCL/Gelatin composite fibrous scaffolds: mechanical properties and cellular
responses J. Biomater. Sci. Polym. Ed. 27 824–38
[31] Ghasemi-Mobarakeh L, Prabhakaran M P, Morshed M, Nasr-Esfahani M-H and Ramakrishna S 2008 Electrospun poly(ε-
caprolactone)/gelatin nanofibrous scaffolds for nerve tissue engineering Biomaterials 29 4532–9
[32] Pourhashem S and Afshar A 2014 Double layer bioglass-silica coatings on 316L stainless steel by sol–gel method Ceram. Int. 40
993–1000
[33] Wang Y, Zhang S, Lu Z, Wang P, Ji X and Li W 2018 Preparation and performance of electrically conductive Nb-doped
TiO2/polyaniline bilayer coating for 316L stainless steel bipolar plates of proton-exchange membrane fuel cells RSC Adv. 8 19426–31
[34] Yu Z, Lv L, Ma Y, Di H and He Y 2016 Covalent modification of graphene oxide by metronidazole for reinforced anti-corrosion
properties of epoxy coatings RSC Adv. 6 18217–26
[35] Raza M A, Rehman Z U, Ghauri F A, Ahmad A, Ahmad R and Raffi M 2016 Corrosion study of electrophoretically deposited graphene
oxide coatings on copper metal Thin Solid Films 620 150–9
[36] Kharaziha M and Fathi M H 2009 Synthesis and characterization of bioactive forsterite nanopowder Ceram. Int. 35 2449–54
[37] Park J H and Park J M 2014 Electrophoretic deposition of graphene oxide on mild carbon steel for anti-corrosion application Surf. Coat.
Technol. 254 167–74
[38] Sharifi A, Khorasani S N, Borhani S and Neisiany R E 2018 Alumina reinforced nanofibers used for exceeding improvement in
mechanical properties of the laminated carbon/epoxy composite Theor. Appl. Fract. Mech. 96 193–201
[39] Denis P, Dulnik J and Sajkiewicz P 2015 Electrospinning and structure of bicomponent polycaprolactone/gelatin nanofibers obtained
using alternative solvent System International Journal of Polymeric Materials and Polymeric Biomaterials 64 354–64
[40] Khalili S, Khorasani S N, Razavi S M, Hashemibeni B and Tamayol A 2019 Nanofibrous scaffolds with biomimetic composition for skin
regeneration Appl. Biochem. Biotechnol. 187 1193–203
[41] Kim G, Son J, Park S and Kim W 2008 Hybrid process for fabricating 3D hierarchical scaffolds combining rapid prototyping and
electrospinning Macromol. Rapid Commun. 29 1577–81
[42] Jokar M, Darvishi S, Torkaman R, Kharaziha M and Karbasi M 2016 Corrosion and bioactivity evaluation of nanocomposite PCL-
forsterite coating applied on 316L stainless steel Surf. Coat. Technol. 307 324–31
[43] Duraipandy N, Syamala K M and Rajendran N 2018 Antibacterial effects, biocompatibility and electrochemical behavior of zinc
incorporated niobium oxide coating on 316L SS for biomedical applications Appl. Surf. Sci. 427 1166–81
[44] Rekha M Y, Kamboj A and Srivastava C 2017 Electrochemical behaviour of SnZn-graphene oxide composite coatings Thin Solid Films
636 593–601
[45] Wang M-H, Li Q, Li X, Liu Y and Fan L-Z 2018 Effect of oxygen-containing functional groups in epoxy/reduced graphene oxide
composite coatings on corrosion protection and antimicrobial properties Appl. Surf. Sci. 448 351–61
[46] Kharaziha M, Fathi M and Edris H 2013 Development of novel aligned nanofibrous composite membranes for guided bone
regeneration J. Mech. Behav. Biomed. Mater. 24 9–20
[47] Wang J, Li D, Liu Q, Yin X, Zhang Y, Jing X and Zhang M 2010 Fabrication of hydrophobic surface with hierarchical structure on Mg
alloy and its corrosion resistance Electrochim. Acta 55 6897–906
[48] Bommersbach P, Alemany-Dumont C, Millet J P and Normand B 2005 Formation and behaviour study of an environment-friendly
corrosion inhibitor by electrochemical methods Electrochim. Acta 51 1076–84
[49] Dong Y, Ma L and Zhou Q 2013 Effect of the incorporation of montmorillonite-layered double hydroxide nanoclays on the corrosion
protection of epoxy coatings J. Coat. Technol. Res. 10 909–21
[50] Taylor S R and Gileadi E 1995 Physical interpretation of the warburg impedance Corrosion 51 664–71

13

You might also like