You are on page 1of 34

Journal Pre-proof

Effect of D-fructose on the in-vitro corrosion behavior of AZ31


Magnesium alloy in simulated body fluid

Durga Bhakta Pokharel, Liping Wu, Junhua Dong, Xin Wei, Ini-Ibehe
Nabuk Etim, Dhruba Babu Subedi, Aniefiok Joseph Umoh, Wei Ke

PII: S1005-0302(20)30610-1
DOI: https://doi.org/10.1016/j.jmst.2020.03.080
Reference: JMST 2371

To appear in: Journal of Materials Science & Technology

Received Date: 11 November 2019


Revised Date: 17 March 2020
Accepted Date: 30 March 2020

Please cite this article as: Pokharel DB, Liping W, Dong J, Wei X, Etim I-IbeheN, Subedi DB,
Umoh AJ, Ke W, Effect of D-fructose on the in-vitro corrosion behavior of AZ31 Magnesium
alloy in simulated body fluid, Journal of Materials Science and amp; Technology (2020),
doi: https://doi.org/10.1016/j.jmst.2020.03.080

This is a PDF file of an article that has undergone enhancements after acceptance, such as
the addition of a cover page and metadata, and formatting for readability, but it is not yet the
definitive version of record. This version will undergo additional copyediting, typesetting and
review before it is published in its final form, but we are providing this version to give early
visibility of the article. Please note that, during the production process, errors may be
discovered which could affect the content, and all legal disclaimers that apply to the journal
pertain.

© 2020 Published by Elsevier.


Research Article
Effect of D-fructose on the in-vitro corrosion behavior of AZ31 Magnesium alloy in
simulated body fluid
Durga Bhakta Pokharel1,2, Liping Wu1, Junhua Dong1,2,*, Xin Wei, Ini-Ibehe Nabuk Etim1,3,
Dhruba Babu Subedi 1,2, Aniefiok Joseph Umoh1,2,Wei Ke1
1
Shenyang National Laboratory for Materials Science (SYNL), Institute of Metal Research,
Chinese Academy of Sciences, Shenyang 110016, China
2
School of Material Science and Engineering, University of Science and Technology of China,
Hefei 230026, China
3
Department of Marine Biology, Akwa Ibom State University, Uyo, Nigeria

of
* Corresponding author.
E-mail address: jhdong@imr.ac.cn (J.H. Dong).

ro
[Received 11 November 2019; Received in revised form 17 March 2020; Accepted 30
March 2020]
-p
re
The effect of adding D-fructose to simulated body fluid (SBF) on the corrosion behavior of
lP

AZ31 magnesium (Mg) alloy at 37 oC and at a pH of 7.4 was studied by potentiodynamic

polarization (PDP), electrochemical impedance spectroscopy (EIS), potentiostatic polarization


na

and hydrogen (H2) collecting techniques, Raman spectroscopy technique, scanning electron

microscopy (SEM), energy dispersive spectroscopy (EDS), X-ray diffraction (XRD), X-ray
ur

photoelectron spectroscopy analysis (XPS) and Fourier transformed infrared (FTIR). The

results demonstrated that the addition of fructose enhanced the deposition of phosphates
Jo

forming thick and compact corrosion products, which inhibited the transmission of aggressive

ions into the Mg substrate. As a result, both the anodic dissolution of Mg and negative

difference effect (NDE) were suppressed. Thus, the corrosion resistance of AZ31 Mg alloy in

SBF was significantly improved.

1
Keywords: AZ31 Mg alloy, D-Fructose, SBF, EIS, Hydroxyapatite

1. Introduction

Over the years, several metals have been arbitrarily chosen for surgical fixation of broken

bones and suturing device for long-term implantation. Metallic bone plates such as stainless

steel and titanium-based alloys are commonly used as bone implants for a long period because

of their high degradation resistance characteristics [1]. Finding suggests that after the healing of

bone fracture, these implanted materials need to be removed to avoid osteopenia caused by

permanent stress shielding, resulting in the formation of wear debris and inflammation as well.

of
Hence, corrective surgery is required to address these problems, especially in young patients,

where removal of the metallic implant is mandatory to facilitate the natural bone growth [2].

ro
To resolve this issue, ceramics materials, their derivatives, polymeric materials and
-p
copolymers have been used as degradable biomaterials till dates [3,4]. However, these materials

are not well suited as implant materials due to their low mechanical characteristics [5]. Therefore,
re
it is essential to adopt the decisive biomaterials that would be completely biocompatible in
lP

every aspect of the implantation.

Interestingly, many researchers have reported that magnesium and its alloys are

promising biomaterials for the cardiovascular stent and bone implantation because of their low
na

density, inherent biocompatibility, adequate mechanical properties and higher fracture

toughness which are better than those of ceramics and polymers [6-9]. Furthermore, unlike other
ur

metallic implant materials (stainless steel and titanium), which cause stress-shielding
Jo

phenomenon, the elastic modulus of Mg alloys (40–45 GPa) is closer to that of the human
[10]
bones (10–40 GPa) . Meanwhile, mineral apposition rate of biodegradable magnesium

implants is higher as compared to commonly used polymeric implants, which is beneficial as

it leads to the formation of new bone at the periosteal and endosteal site without affecting the
[11]
surrounding tissues . In comparison with other commonly used metallic implants, the

2
degradation behavior of Mg alloys predicts the possibility of better physiological adjustment

and bone repair with minimal inflammatory response [12]. In addition, the gradual dissolution,

metabolism and absorption behavior of the Mg alloys in the physiological environments makes

it suitable to be used as the long-term implant materials, curtailing the need for a second

surgical operation after tissue healed completely [13-15]. However, it is known that insufficient

degradation resistance of Mg and its alloys in the physiological environment hinders their

clinical applications due to the fact that the released excess amount of hydrogen gas can damage

the surrounding tissues [13]. Meanwhile, it has been reported that the excess degradation rate of

of
magnesium alloy may induce the decrease of mechanical strength before the fractured bone is
[9,16]

ro
completely healed . Hence, keeping the degradation rate of magnesium alloys in an

appropriate level is the most essential precondition for their clinical applications. Previous
-p
literatures have verified that excess hydrogen evolution can be mitigated by modifying the

surface of Mg and its alloy through the application of coatings such as plasma spray/sol–gel
re
[17, 18] [19] [20-23]
composite coating , anodizing and polymeric coatings . These coatings can
lP

suppress the evolution of hydrogen by creating a durable barrier on the surface of the Mg alloy

and improve the active corrosion protection

Fundamentally, biodegradation of Mg and its alloys are directly related to the corrosion
na

behavior, which depends on the interaction of the surrounding physiological environments to

the interface of the implanted materials [24]. Blood plasma and intercellular fluid contain both
ur

inorganic ions and organic compounds such as Mg2+, Ca2+, Cl−, HCO3−, HPO4−2 and amino
Jo

acids, proteins, carbohydrates, respectively, which acts as neutral solution in the human body
[25]
. Previous studies have indicated that chloride ion can promote pitting corrosion, whereas

phosphates can play an effective role in reducing the corrosion rate [26-28]. Although carbonate
[22]
bolsters the dissolution of magnesium, it rapidly forms magnesium carbonate which acts as
[29,30]
a passive film on the surface of the metal . Other studies also demonstrated that organic

3
compounds such as proteins, amino acids and carbohydrates play an important role in the
[31-34]
corrosion behavior of metals . To the best of our knowledge, significant effects of the

carbohydrates (glucose and fructose) on the corrosion behavior of magnesium and its alloys in

simulated body solutions have not been effectively elucidated. Based on artificial neural
[35]
networks, Willumeit reported that glucose does not significantly influence the corrosion

rate of magnesium with the presence of O2 and CO2 under atmospheric and cell culture
[36]
conditions. Whereas Zeng et.al. suggested that glucose retards the corrosion rate of pure

magnesium in Hank’s solution by forming the apatite on its surface. Also, Wang et al. [37]

of
indicated that glucose could restrain the HER process by its own cathodic reduction on the

ro
magnesium surface

Despite the usefulness of fructose in blood plasma and intercellular fluid of human body
-p
solution and increased concentration in some patients with hyperglycemia

investigation of its effect on the degradation behavior of magnesium Mg and its alloys in
[38-40]
, the
re
physiological solution are almost rare. Therefore, the main objective of this work is to assess
lP

the implications of simulated body fluid (SBF) with the presence of D- fructose on corrosion

behavior of AZ31 Mg alloy. It is anticipated that the mimic test results thus obtained would

contribute to understanding the biodegradability and to gain more insights into the feasibility
na

of using Mg and its alloy for medical applications.

2. Experimental methods
ur

2.1. Sample preparation


Jo

AZ31 magnesium alloy with chemical compositions of Al 2.98, Zn 0.88, Mn 0.38, Si

0.0135, Cu 0.001 and Mg balance was used as substrate. The samples were cut by electrical

discharge machining (EDM) into 11.3 mm in diameter and 5 mm in thickness, welded to copper

wire and encapsulated into epoxy resin exposing the area 0.785 cm2. The prepared samples

4
were ground with silicon carbide paper from 600 to 2000 grits then polished on an abrasive

finishing machine using 2.5 μm diamond polishing paste to the mirror face and cleaned with

absolute ethanol, and finally dried with cold air stream.

2.2. Immersion test

The polished samples were respectively immersed in simulated body fluid (SBF) with and

without 1 g/L D-fructose at 37 ± 0.5 °C and a pH of 7.40 ± 0.05 in a water bath, and removed

after 1, 3, 7, 14 and 28 days. The ratio of sample area to the solution volume was 1 cm 2: 250

of
[41]
mL . Afterwards, the corrosion products were scraped off from the samples for

characterization. The constituents of SBF were composed of 8.00 g/L NaCl, 0.35 g/L NaHCO3,

ro
0.20 g/L KCl, 0.12 g/L Na2HPO4·12H2O, 0.06 g/L KH2PO4, 0.14 g/L CaCl2 and 0.20 g/L

-p
MgSO4. During the immersion test, volume of hydrogen evolved was collected. Herein, the

SBF with and without D-fructose is referred to the term solution A and solution B, respectively.
re
The samples were removed from the solution, rinsed with distilled water and finally dried
lP

in air after the prescribed immersion periods. Surface and cross-section morphologies of each

sample were examined by using a Phillips FEI InspectTMF scanning electron microscopy
na

(SEM). The elemental compositions of the corrosion products were characterized using

matched energy-dispersive X-ray spectrometry (EDS). Phase compositions of the surface

corrosion products were obtained using X-ray diffractometry (XRD) (a Philips PW1700 X'Pert
ur

X-ray generator operated with Cu Kα radiation). Fourier transform infrared spectroscopy (FT-
Jo

IR, THENSOR-27) was used to identify the functional groups of the corrosion products in a

wave number ranging from 500 to 4000 cm-1. Before XRD and FTIR measurements, the

coating was scraped off from the specimen surface. The chemical compositions of the corrosion

products on the immersed samples were examined by using X-ray photoelectron spectroscopy

(XPS, ESCALAB 250). Corresponding binding energies were corrected according to the

5
adventitious C1s signal (284.6 eV) and probed after fitting the obtained data with XPS PEAK

4.1 software.

2.3.Electrochemical analysis

The potentiodynamic polarization, potentiostatic polarization and EIS were measured using

a Gamry work station (Reference 1000) in a three-electrode cell consisting of SBF solution. A

capsulated AZ31 magnesium alloy was used as the working electrode, a piece of platinum plate

as the counter electrode and a saturated calomel electrode (SCE) as the reference electrode,

of
respectively. The potentiodynamic polarization was performed at a scan rate of 0.5 mV s−1

with respect to the open circuit potential (OCP). The potentiostatic polarization was carried out

ro
applying -1.35 V with respect to SCE for 22 h, and simultaneously collected the unpolarized
-p
Raman spectra using LabRAM HR800 (Horiba Jobin-Yvon, France) equipped with an air-

cooled CCD array detector in the backscattering configuration as shown in Fig. 1. A He−Ne
re
laser (632.82 nm) with an incident power of around 8 mW was used as the excitation source.

The spot size of the laser was focused to 2 μm, where the spectral resolution was 0.65 cm-2.
lP

The distance between the sample and solution surface was kept at 1 mm, so that interference
[42]
of the solution with the Raman laser would be neglected during the test period . The EIS
na

measurements were conducted in a frequency range from 105 to 10-2 Hz with an applied

sinusoidal perturbation of ±10 mV at OCP and obtained data were fitted using the Zsimpwin
ur

software version 3.2.


Jo

3. Results and discussion

3.1. Surface morphology analysis

Fig. 2(a)-(e) and Fig. 2(f)-(j) show the surface morphologies of AZ31 magnesium alloy

immersed in solution A and B, respectively. Fig. 2(a')-(e'), (a") and (f')-(j'), (f") demonstrate

the corresponding EDS plots, where Fig. 2(a") and (f") represent the corresponding EDS

6
spectra of selected area B and D in Fig. 2(a) and (f), respectively. On the first day of immersion

as shown in Fig. 2(a), many scratches are observed on the surface (see the area A), while some

of the aggregated particles with irregular shape precipitated on the area B. At the same time,

Fig. 2(f) shows a cracked surface film structure on the area C and a less precipitation of

particles on the area D as compared with the surface of Fig. 2(a). The EDS plots of the areas A

and C show the composition of C, O, Mg, Al and Zn, as presented in Fig. 2(a') and (f'),

respectively. In Fig. 2(a") and (f"), areas B and D indicated the existence of C, O, Mg, Al, Zn,

P and Ca. With extending the immersion time to 3 days, the surface film morphology appears

of
as a cracked structure enwrapping bar-like and spherical particles, where some irregular

ro
particles start to embed, as presented in Fig. 2(b), while Fig. 2(g) shows almost the same

morphology besides a few more cracks as compared with Fig. 2(f). The EDS plots shown in
-p
Figs. 2(b') and (g') demonstrated the composition of C, O, Mg, Al, Zn, Ca and P. After 7 days

of immersion, the surface shown in Fig. 2(c) displays a similar morphology to Fig. 2(b) with
re
more deposition of bar-like and spherical particles. Fig. 2(h) shows similar morphology to Fig.
lP

2(g) with some cotton-like products covering the surface. The EDS plots presented in Fig. 2(c')

and (h') show the similar compositions with those obtained after 3 days of immersion. However,

after 14 days of immersion, the surface is covered with a uniform film composed of compact
na

spherical particles as shown in Fig. 2(d), meanwhile, Fig. 2(i) shows a similar morphology to

Fig. 2(b). The respective EDS plot displayed in Fig. 2(d') and (i') show the composition of C,
ur

O, Mg, P and Ca. After 28 days of immersion as shown in Fig. 2(e), the surface of the sample
Jo

is covered with porous coral-like particles, while Fig. 2(j) shows that the surface is covered by

some of the compact spherical particles. In addition, the EDS analysis results presented in Fig.

2(e') indicate the compositions of C, O, Mg, P and Ca, while that presented in Fig. 2(j') showed

the occurrence of Al and Cl aside the compositions which occurred in Fig. 2(e'). A comparison

of Fig. 2(a)-(e) with Fig. 2(f)-(j) and their corresponding EDS results indicate that content of

7
O, P and Ca is increased with extending the immersion period for both solutions. However, the

content of O, P and Ca on the surface immersed in the solution A is slightly higher than that of

solution B. The results obtained predicts that the addition of fructose enhances the formation

of apatite that contain O, Ca and P. Previous studies reported that the increase content of Ca

and P on magnesium surface when immersed in SBF indicates a decrease in the dissolution of
[25]
magnesium . The occurrence of Al and Zn might be from the substrate, which could be

possible due to the transmission of electron beam through the thin film layer into the substrate

during the EDS measurement.

of
Fig. 3(a)-(e) and (f)-(j) exhibit the cross-section images of AZ31 Mg alloy after being

ro
soaked in solutions A and B for 1, 3, 7, 14 and 28 days, respectively. Fig. 3(a) indicates that a

thin layer with a thickness of 0.8 μm is deposited on the surface, which corresponds to the area
-p
A as shown in Fig. 2(a), while, Fig. 3(f) shows a thin layer with 0.5 μm thickness, which

corresponds to the general surface as shown in Fig. 2(f). The cracked regions marked with an
re
arrow in Fig. 3(a) and (f) corresponds to the surface areas of B and D in Fig. 2(a) and (f),
lP

respectively. After 3 days of immersion, a two-layer film was identified on the cross-sectional

surface as presented in Fig. 3(b). The inner layer with a thickness of 1.33 μm is cracked, while

the outer layer is non-uniform relating to the aggregated particles on the surface as shown in
na

Fig. 2(b). However, a uniform single layer film with a thickness of 1.2 μm was observed for

the same day of immersion in solution B, as shown in Fig. 3(g). Fig. 3(c) shows that a compact
ur

inner layer with 1.56 μm in thickness and a cracked outer layer with 1.25 μm in thickness was
Jo

formed with the extension of the immersion time to 7 days. In contrast, a single zigzag layer

with the thickness of 1.8 μm, having less cracked morphology was seen for the same day of

immersion in solution B as shown in Fig. 3(h). When the immersion time increases to 14 days,

the formed film was comprised of a cracked single layer with 2.5 μm in thickness as shown in

Fig. 3(d), while Fig. 3(i) shows the formation of monolayer corrosion product with the

8
thickness of 2.2 μm in a uniform manner without forming any crack. After 28 days of

immersion, the surface film remains cracked but thickened with a thickness of 3.2 μm as shown

in Fig. 3(e). Unlike Fig. 3(e), non-uniform layer with a thickness of 2.8 μm, with severe surface

degradation was observed in Fig. 3(j), which predicts that surface film might be broken down

due to the presence of aggressive ions in solution B. The appearance of cracks might be caused

by scanning microscope vacuum process. It can be noticed that the thickness of the corrosion

products film increased with increasing the immersion period for both solutions A and B.

However, more compacted and thickened corrosion product is observed in solution A than

of
those formed in solution B over the immersion periods.

ro
Fig. 4 shows the FTIR spectra of the AZ31 Mg alloy immersed in the solutions for 28

days. The absorption bands at 875 cm-1 and at 1650 cm-1 can be assigned to the existence of
-p
HPO4-and H2O bending, respectively [43]. The strong broad absorption band at 3420 cm-1 can
re
be attributed to the vibration modes of hydroxyl (OH−) group [44] while the absorption bands at

1040 cm-1 and 562 cm-1 represented the bending mode of PO43- which normally associates with
lP

hydroxyapatite (HA). The absorption bands around 1420–1480 cm-1 can be attributed to CO32-
[45]
.
na

The fitting results of XPS spectra of the AZ31 magnesium alloy immersed in solution

A and B are presented in Fig. 5 (a)-(e) and (f)-(j), respectively. The peaks approximately at
ur

1302.3 eV and 1303 eV obtained at Mg 1s region can be attributed to the magnesium hydroxide
Jo

[41] [46]
and magnesium hydrogen phosphate (amorphous MgHPO4 or MgHPO4.xH2O) ,

respectively. The deconvoluted O 1s XPS spectra showed three clear components for both

solutions, the peak raised at 531.2 eV can be assigned to MgO or P=O or CO 32- bond, while

that raised at 532.5 eV indicates the existence of hydroxyl (OH-) group [47] and the peak raised

at 533.3 eV can be attributed to the P-OH [46]. A well resolved duplet was located at 347.3 eV

9
(Ca2p3/2) and 350.7 eV (Ca2p½) for the Ca 2p specific region, which are assigned as

hydroxyapatite [48]. Additionally, the deconvoluted spectra of P 2p clearly showed two peaks

at 132.4 and 133.7 eV, which represent the P–O bonding energies of PO43- and HPO42-,

respectively [49]. Subsequently, the main C 1s component appeared at 284.6 eV which can be

attributed to C-H/C-C bonds present as adventitious surface hydrocarbons, whereas the peak

raised at 286.6 eV indicates the existence of CO32- bonding [50], which might be bonded with

Mg2+ or Ca2+. Therefore, XPS analysis demonstrated the existence of Mg(OH)2, MgHPO4 and

hydroxyapatite (HA) as corrosion products.

of
Fig. 6 shows the XRD patterns of the film scraped from AZ31 Mg alloy after immersion

ro
in solution A and B for 28 days. It indicates the presence of Mg(OH)2 and hydroxyapatite in

the films formed in solution A and B. However, the Mg signal was also detected, which may
-p
be caused by scraping the coating film from the substrate. The XRD pattern could not detect

the peak corresponding to MgHPO4 as described previously in Fig. 5(e) and (j), which might
re
be an indication of the formation of the amorphous MgHPO4 [46].
lP

3.2. Hydrogen evolution

The evolution of released hydrogen volume and corresponding hydrogen evolution rate
na

(HER) as a function of immersion time for the AZ31 Mg alloy in solutions A and B over a

period of 28 days are displayed in Fig. 7(a) and (b). At the initial 5 days of immersion for both
ur

solutions, it can be seen that the volume of H2 increased almost at the same rate. Hereafter, the
Jo

volume of hydrogen evolved in solution B increases quicker than that in solution A. The

corresponding HER decreases with increasing the immersion time for both solutions. However,

the HER for solution A is relatively lower compared to solution B, indicating that the protective

layer formed on the surface of the sample in solution A is gradually enhanced. It is suggested

that the HER at the corrosion potential of the magnesium alloy directly proportionate to the

10
[51,52]
amount of Mg dissolved, which reflect the instantaneous corrosion rate . Hence, more

reliable corrosion rate and high experimental efficiency was obtained from the HER curve over
[53]
the immersion periods . It shows that the addition of fructose into SBF solution restrained

the dissolution of Mg substrate.

3.3. Electrochemical analysis

3.3.1. Potentiodynamic polarization

Fig. 8 shows potentiodynamic polarization curves of AZ31 Mg alloy after immersion

of
in solution A and B for 28 days. The corrosion potential (Ecorr) in both solutions increases

ro
gradually with increasing the immersion time from 1 to 14 days, and then slightly drops with

a further increase in the immersion time to 28 days. The anodic dissolution of Mg dominates
-p
the current density of the anodic branch while the hydrogen evolution contributes to the current

density of the cathodic branch. In solution A, the cathodic current density decreases with
re
extending the immersion time from 1 to 3 days, and then it slightly increases for the remaining
lP

immersion duration. Meanwhile, the anodic current density obtained after 1, 3 and 7 days

continuously increases with increasing the immersion time, indicating the corrosion

acceleration of AZ31 Mg alloy. However, the anodic current density increases slowly from
na

corrosion potential to pitting potential (shown by arrow) and then rises fast with further positive

scanning after 14 and 28 days of immersion. The occurrence of pitting potential reveals the
ur

passive state of AZ31 Mg alloy. By comparison, the pitting potential increases with extending
Jo

the immersion time. In effect, the anodic branch presents a behavior transferring from pitting

corrosion to passive state. Initially, the pitting potential may be equal to the corrosion potential,

and it gradually increases with extending the immersion time. In solution B, the cathodic

current decreases with an increase in the immersion time from 1 to 7 days, and then it increases

with a rise of the immersion time to 28 days. In addition, the anodic current density also shows

11
active dissolution of AZ31 Mg alloy from 1 to 14 days and then passivation at 28 days. The

decrease in the anodic current is caused by the thickening of the corrosion film which can resist

the dissolution of Mg substrate. For the first day of immersion, the cathodic current for both

solutions is almost identical while the anodic current for solution A is lower compared to

solution B. As the immersion time increases to 3 and 7 days, the cathodic currents for solution

A is smaller than that for the solution B, while the anodic currents for both solutions are almost

equal. After 14 days of immersion, the cathodic currents for both solutions are the same, but

the anodic current for solution A is smaller compared to solution B. After 28 days of immersion,

of
the cathodic currents for both solutions are identical, nevertheless, the anodic current for

ro
solution A is smaller compared to solution B. By comparing corrosion potential, the anodic

current and cathodic current for both solutions at the same immersion time, it can be concluded
-p
that solution A mainly inhibits the anodic dissolution of the Mg substrate.
re
3.3.2. Potentiostatic polarization
lP

Fig. 9(a, b) shows the potentiostatic anodic current decay curves and corresponding

negative difference effect (NDE) of AZ31 Mg alloy in the solution A and B respectively. It is

considered that the total current density equals to the sum of the current densities induced by
na

[42, 54]
the dissolution of the Mg and the evolution of hydrogen (NDE) . The current density

corresponding to the hydrogen evolution as shown in Fig. 9(c) is calculated by Faraday’s law
ur

using the volume of hydrogen as shown in Fig. 9(b) [55], while Fig. 9(d) shows the current decay
Jo

corresponding to the dissolution of Mg obtained by subtracting the current density as shown in

Fig. 9(c) from the total current density shown in Fig. 9(a) [55]. Fig. 9(a) demonstrates that the

current densities of the samples for both solutions show an increase at the beginning, followed

by a decrease, and finally attained a steady-state value, and corresponding NDE and

magnesium dissolution for the solution A is suppressed as compared to the solution B. Hence,

12
under the same applied anodic potential, the total current density, volume of hydrogen and the

anodic dissolution of Mg in solution A are relatively lower than those in solution B, which

predicts the inhibition of the anodic dissolution of Mg with the addition of fructose. The

corresponding in-situ Raman spectra obtained after 22 h of potentiostatic polarization in both

the solutions A and B as presented in Fig. 9(e) demonstrates the existence of symmetric

stretching mode of the carbonate, which is characterized by peaks at 1086 and 731 cm-1 [56].

This might be attributed to the existence of calcite or magnesite as the corrosion product on the

AZ31 Mg alloy. Subsequently, the presence of v2 bonding of P-O-P, v4 bending of P-O-P and

of
v1 stretching of P-O, respectively, at 451 cm-1, 556 cm-1 and 983 cm-1, is assigned the
[57, 58]
. The band located at 916 cm-1 is

ro
characteristic vibration band of PO4 in HA crystal

assigned to the P-OH stretching of HPO42- [59]. It is noteworthy that the peak intensity of the
-p
phosphates in solution A is relatively higher compared to solution B indicating the higher

deposition of the phosphate-containing compounds on the surface of the AZ31 Mg alloy. This
re
shows the improvement of the corrosion resistance of the alloy in solution A.
lP

3.3.3. Electrochemical impedance spectroscopy (EIS)

Fig. 10 shows the electrochemical impedance spectra of AZ31 Mg alloy in solutions A


na

and B as a function of immersion time. Fig. 10(a) and (c) represent the |Z| vs. frequency plots

of AZ31 Mg alloy immersed in solutions A and B at different time, respectively. Fig. 10(b)
ur

and (d) show the phase angle vs frequency plots corresponding to Fig. 10(a) and (c),

respectively. All high frequency data of the bode plots indicate a high frequency phase shift
Jo

occurred during the EIS measurements. Fig. 10(a) shows the |Z| value at high frequency (|Z|HF)

after day 1 in solution A was about 30 Ω cm2, and at low frequency (|Z|LF) corresponding to

the corrosion resistance (Rct) was about 9.4 × 103 Ω cm2, while the corresponding phase angle-

frequency plot shows the symmetric pattern having a peak of 70° and centered at 49.8 Hz.

When the immersion period is extended to 3 days, the |Z|HF was almost equal to the first day of

13
immersion, while the |Z|LF increases to about 1.3 × 104 Ω cm2, and the respective phase angle

peak of 30° and 65° centered at 1.3 × 104 Hz and 40 Hz respectively. After 7 days of the

immersion, the |Z|HF is similar to that measured after 3 days, while the |Z|LF again increases to

about 2.7 × 104 Ω cm2, and the corresponding phase angle peaks of 41° and 65° can be seen at

1.2 × 104 Hz and 20.6 Hz, respectively. With extending the immersion period to 14 days, the

|Z|HF increases to about 100 Ω cm2 and the |Z|LF also increases to about 3.9 × 104 Ω cm2. The

corresponding phase angle reveals a shift to the low frequency region at first and exhibits a

valley of 32° and centered at 5 × 103 Hz, followed by an increase with a peak of 43° and

of
centered at 400 Hz, then slightly decreases with decreasing the frequency to form a second

ro
valley of 40° at 62 Hz. Subsequently, appears a highest peak of 48° and centered at 5 Hz.

Furthermore, a decrease of the |Z|HF can be seen after 28 days of immersion similar to day 1,
-p
while the |Z|LF increases to about 5.2 × 104 Ω cm2. The corresponding phase angle showed two

overlapped picks of 44° and 66° and centered at 104 Hz and 10 Hz respectively.
re
In Fig.10(c), it can be seen that the |Z|HF was about 30 Ω cm2 after day 1 of immersion
lP

in solution B, while the |Z|LF appeared at about 5.3 × 103 Ω cm2. Fig. 10(d) shows the

corresponding symmetrical phase angle peak of 65° centered at 70 Hz, which finally decreases
na

to -3.4° at 0.01 Hz. After 3 days of immersion, the |Z|HF value was almost similar to 1 day,

while the |Z|LF reaches to 1.07 × 104 Ω cm2 and the corresponding phase angle peak of 31° and
ur

60°, respectively, raised at 1.86 ×104 Hz and 40 Hz, and finally decreases to -6.1° at 0.01 Hz.

Moreover, the |Z|HF value after 7 days of immersion was similar to the 3rd day, while the |Z|LF
Jo

increases to 1.3 × 104 Ω cm2 and the respective phase angle shows the highest peak of 65°

centered at 15 Hz, while the second peak was raised at 1.2×104 Hz with a peak angle of 40°.

However, the |Z|HF increases to about 100 Ω cm2 after 14 days of immersion, while the |Z|LF

continuously increases to about 2.9 × 104 Ω cm2. The respective phase angle plot firstly shifts

to the low frequency region and creates a valley of 34° at about 6.3 × 103 Hz, and then it

14
increases to 40° and centered at 400 Hz. It gradually decreases and creates a second valley of

39° and centered at 79 Hz, which again increases for a second peak of 48° at 7.6 Hz. As the

immersion period is prolonged to 28 days, the |Z|HF decreases to about 60 Ω cm2 as compared

to 14 days of immersion, while the |Z|LF appears similar 14th day of immersion, and

corresponding phase angle becomes asymmetric with overlapped double peaks of 46° and 57°

and centered at about 4.8 × 104 Hz and 9 Hz respectively.

It is worth noting that the cathodic and anodic processes simultaneously occur during

of
the degradation of the Mg alloy. The anodic process is composed of two reactions: one is the

dissolution of AZ31 Mg alloy and the other is the formation of corrosion film, while the

ro
cathodic process is mainly dominated by the evolution of hydrogen. Aiming to obtain further

information from the EIS measurements, the impedance spectra are fitted according to the
-p
equivalent circuits shown in Fig. 11(a) and (b). The plots obtained after 1st and 3rd day of
re
immersion in solutions A and B are fitted by the equivalent circuit in Fig. 11(a), while the plots

obtained after 7, 14 and 28 days in solutions A and B are fitted with the equivalent circuit
lP

shown in Fig. 11(b). Rs represents the solution resistance, which depends on the conductivity

of the test medium [60]. QHFs represents the constant phase element (CPE) corresponding to the
na

high frequency shift. RH and QH represent the hydrogen evolution resistance and corresponding

CPE respectively. While Qf and Rf, represent the film formation capacitance and resistance on
ur

the surface of AZ31 Mg alloy. Rct represents the anodic charge transfer resistance, while Qdl

represents the electrical double layer capacity at the interface of Mg substrate and electrolyte
Jo

solution. RL and L represent the inductance resistance and inductor, respectively, and are used

to describe the low frequency inductance caused by the localized corrosion of the AZ31

magnesium alloy [61]. The RL and L circuit parameters are added in the circuits representing 1st

and 3rd day of immersion. This could be possible due to the inhomogeneous distribution of

electrochemical current density on the AZ31 Mg alloy caused by the non-uniform film

15
formation. The disappearance of inductance after 3 days of immersion indicates the formation

of thick and compacted corrosion product which restrains the transfer of aggressive ions into

the surface of the Mg alloy.

In the equivalent circuits as shown in Fig. 11a and b, the impedance of CPE (ZCPE) can

be described by Eq. (1) as follows:

𝑍CPE = 𝑌0−1 (j𝑤)−𝑛 (1)

of
where ZCPE represents the impedance of CPE, Y0 a proportional factor (CPE), j the imaginary

unit, w the angular frequency (in rad/s) and n is a CPE exponent whose value ranges from 0 to

ro
1 [62].

-p
Table 1 shows that the charge transfer resistance (Rct) in electric double layer of AZ31

Mg alloy in solution A increases from 801.2 Ω cm2 to 4895.6 Ω cm2, indicating the retardation
re
of the dissolution of Mg as the immersion time prolongs from 1 to 28 days. The Y0-dl
lP

corresponding to Qdl decreases with extending the immersion time. This suggests a reduced

electrolytic activity at the metal/solution interface due to the formation of compact layer
na

corrosion product. As the immersion time prolongs, the values of Y0-F corresponding to QF

decreases from 5.95×10-6 to 1.56×10-6 Ssn cm-2. The RF increases from 482.1 to 2201.5 Ω cm2
ur

indicating the formation of a compact layer of corrosion product which restricted the transfer

of ions to the surface of Mg alloy. Subsequently, an increase of nF with extending the


Jo

immersion time exhibits the increment of homogeneous distribution of the corrosion product

on the surface of AZ31 Mg alloy. The simulated value of Y0-H corresponding to the hydrogen

evolution region also decreases from 6.57 × 10-6 to 3.33 × 10-6 Ssn cm-2 and corresponding RH

value increases from 265 to 1781.4 Ω cm2 with prolonged immersion, signifying the retardation

of the evolution of hydrogen with time. Moreover, the increase of RL after 1 and 3 days of

16
immersion is related to the improvement of the film stability on the AZ31 Mg alloy with

extending the immersion period.

Table 2 displays that the charge transfer resistance (Rct) in electric double layer of AZ31

Mg alloy in solution B increases from 523 to 3115.7 Ω cm2, showing the hindrance of the

dissolution of Mg during the whole immersion period. Y0-dl corresponding to Qdl decrease with

the prolongation of the immersion time, advocating the low ionic absorption at the interface of

the AZ31 Mg alloy. Furthermore, as the immersion time increases, the values of Y0-F

corresponding to QF continuously diminishes from 6.99 × 10-6 to 2.69 × 10-6 Ssn cm-2, while the

of
value of RF raises from 381 to 1869 Ω cm2 and value of nF increases with extending the

ro
immersion time attributing the progress of homogeneous distribution of the corrosion product

on the surface of AZ31 Mg alloy. Moreover, the simulated value of Y0-H analogous to the
-p
hydrogen evolution region also decreases from 8.77 × 10-6 to 4.17 × 10-6 Ssn cm-2 and respective

RH value enhances from 259.4 to 1615.3 Ω cm2 with increasing the immersion time, indicating
re
the retardation of cathodic evolution of hydrogen with time. Furthermore, the increased value
lP

of RL for the immersion of 1 and 3 days is related to the declining of the localized corrosion.

The electrochemical reactivity of the AZ31 Mg alloy in two solutions is fundamentally

analyzed based on charge transfer resistance (Rct) values. Despite the increasing trend of Rct
na

with increasing the immersion time in solution A and B, the simulated Rct value for solution A

is always higher compared to solution B at all the immersion durations. The higher values of
ur

Rct indicates weak chemical activity and confesses the low dissolution rate of Mg alloy [63-65],
Jo

which will be less prone to be activated and dissolve in the electrolyte at the corresponding test

potential. In addition, the hydrogen evolution resistance and protective film resistance for

solution A has higher magnitude compared to solution B for the same day of immersion. This

could be due to the formation of Mg(OH)2 layer which is deposited by a precipitation reaction

when the concentration of the Mg2+ ions around the surface of the sample gets saturated [66]
.

17
The deposited MgHPO4 and Mg(OH)2 film on the surface of magnesium alloy is considered

to give a favorable environment for hydroxyapatite nucleation [67]. Consequently, tremendous

hydroxyapatite nuclei form and grow spontaneously on the surface film by accumulating the

calcium and phosphate ions from the surrounding solution [67]. The resulting layer of compact

and thick corrosion product hinders ion transfer and inhibits corrosion of the substrate

underneath, thereby restricting the continuous activation and dissolution of magnesium alloy

and simultaneously restrains the evolution of hydrogen. The above results ascertained that

solution A can induce the more rapid formation of corrosion products and more effective

of
protective layer tends to form in SBF. This result indicates that the corrosion resistance of the

ro
AZ31 Mg alloy immersed in solution A is superior to that of the sample in solution B.

4. Conclusion -p
In synopsis, the initial stage of the immersion period characterized with weaker film on the
re
surface shows that AZ31 Mg alloy promoted the localized corrosion for both solutions with
lP

and without D-Fructose possibly due to the penetration of aggressive ions into the corrosion

film, and subsequently enhanced the volume of hydrogen evolution. As for the prolonged
na

immersion time, content of O, P and Ca for both solutions were increased. However, deposition

of phosphates (PO43-) in solution with D-Fructose is comparatively higher. As a result, thicker

and compacted corrosion products layer was formed on the surface of AZ31 Mg alloy in
ur

solution with D-Fructose than in solution without D-Fructose, which in turn suppressed the
Jo

anodic dissolution of Mg substrate as well as NDE. Furthermore, thick and compact film

formed on the Mg substrate in solution with D-Fructose restricted the transmission of

aggressive ions into the surface, which caused the lowering of the evolution of hydrogen gas

as compared to solution without D-Fructose with increasing the immersion period. The overall

study showed that corrosion resistance of the AZ31 Mg alloy in solution with D-Fructose was

18
enhanced with increasing the immersion time compared to solution without D-Fructose.

Consequently, the results obtained from the present study shows better understanding of the

degradation behavior of AZ31 magnesium alloy in a physiological environment and could

provide a constructive reference to a better extent for clinical applications.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of

China (No. 51701221) and the National Key Research and Development program of China

of
(No. 2017YFB0702302).

ro
References:

[1] W.F. Ng, K.Y. Chiu, F.T. Cheng, Mater. Sci. Eng. C 30 (2010) 898–903.

[2]
-p
K.Y. Chiu, M.H. Wong, F.T. Cheng, H.C. Man, Surf. Coatings Technol. 202 (2007)
re
590–598.

[3] M. Navarro, A. Michiardi, O. Castan, J.A. Planell, J. R. Soc. Interface. 5 (2008) 1137–
lP

1158.

[4] A.R. Amini, J.S. Wallace, S.P. Nukavarapu, J. Long. Term. Eff. Med. Implants 21 (2011)
na

93–122.

[5] Z. Dewei, W. Frank, F. Lu, W. Jiali, L. Junlei, Q. Ling, Biomaterials 112 (2016) 287–
ur

302.

[6] B. Heublein, Heart 89 (2003) 651–656.


Jo

[7] G. Mani, M.D. Feldman, D. Patel, C.M. Agrawal, Biomaterials 28 (2007) 1689–1710.

[8] P. Zartner, R. Cesnjevar, H. Singer, M. Weyand, Catheter. Cardiovasc. Interv. 66 (2005)

590–594.

[9] M.P. Staiger, A.M. Pietak, J. Huadmai, G. Dias, Biomaterials 27 (2006) 1728–1734.

[10] L. Xu, G. Yu, E. Zhang, F. Pan, K. Yang, J. Biomed. Mater. Res. Part A 83 (2007) 703–

19
711.

[11] F. Witte, V. Kaese, H. Haferkamp, E. Switzer, A. Meyer-Lindenberg, C.J. Wirth, H.

Windhagen, Biomaterials 26 (2005) 3557–3563.

[12] A. Loos, R. Rohde, A. Haverich, S. Barlach, Macromol. Symp. 253 (2007) 103–108.

[13] A. Zomorodian, F. Brusciotti, A. Fernandes, M.J. Carmezim, T. Moura, J.C.S.

Fernandes, M.F. Montemor, Surf. Coat. Technol. 206 (2012) 4368–4375.

[14] J.W. Erdman, I.A. Macdonald, S.H. Zeisel (Eds.), Present Knowledge in Nutrition, 10th

ed., Wiley-Blackwell, 2012, pp. 459-474.

of
[15] J.McCarthy, R. Kumar, in: R. Schrier (Ed.), Atlas of Diseases of the Kidney, Wiley-

ro
Blackwell, 1999, 4.1-4.12,

-p
[16] M. Zhang, S. Cai, F. Zhang, G. Xu, F. Wang, N. Yu, X. Wu, J. Mater. Sci. Mater. Med.

28 (2017) 1–9.
re
[17] Y. Chen, X. Lu, S.V. Lamaka, P. Ju, C. Blawert, T. Zhang, F. Wang, M.L. Zheludkevich,

Appl. Surf. Sci. 504 (2020) 144462.


lP

[18] Q. Miao, C.E. Cui, J.D. Pan, Surf. Coat. Technol. 201 (2007) 5077-5080

[19] H.M. Mousa, K.H. Hussein, H. Raj, H.M. Woo, C. Hee, C. Sang, Colloid Surf. A-
na

Physicochem. Eng. Asp. 488 (2016) 82–92.


ur

[20] P. Tian, X. Liu, Regen. Biomater. 2 (2015) 135–151.

[21] Z. Zhang, R. Zeng, C. Lin, L. Wang, X. Chen, D. Chen, J. Mater. Sci. Technol. 41 (2020)
Jo

43–55.

[22] L. Guo, C. Gu, J. Feng, Y. Guo, Y. Jin, J. Tu, J. Mater. Sci. Technol. 37 (2020) 9–18.

[23] L. Wu, C. Wang, D.B. Pokharel, I.N. Etim, L. Zhao, J. Dong, W. Ke, N. Chen, , J. Mater.

Sci. Technol. 34 (2018) 2084–2090.

20
[24] Y. Jang, B. Collins, J. Sankar, Y. Yun, Acta Biomater. 9 (2013) 8761–8770.

[25] A. Yamamoto, S. Hiromoto, Mater. Sci. Eng. C 29 (2009) 1559–1568.

[26] R.C. Zeng, Y. Hu, S.K. Guan, H.Z. Cui, E.H. Han, Corros. Sci. 86 (2014) 171–182.

[27] L. Wang, T. Shinohara, B.P. Zhang, J. Alloy. Compd. 496 (2010) 500–507.

[28] R. Rettig, S. Virtanen, J. Biomed. Mater. Res. A 88 (2009) 359–369.

[29] Y. Xin, K. Huo, H. Tao, G. Tang, P.K. Chu, Acta Biomater. 4 (2008) 2008–2015.

[30] F. El-Taib Heakal, A. Mohammed Fekry, M. Ziad Fatayerji, Electrochem. 39 (2009)

583–591.

of
[31] L. Yang, N. Hort, R. Willumeit, F. Feyerabend, Corros. Eng. Sci. Technol. 47 (2012)

ro
335–339.

[32] H. Hornberger, F. Witte, N. Hort, W.D. Mueller, Mater. Sci. Eng. B 176 (2011) 1746–

1755.
-p
[33] B. El Ibrahimi, A. Jmiai, L. Bazzi, S. El Issami, Arab. J. Chem. 13 (2020) 740-771
re
[34] Y. Wang, C.S. Lim, C.V. Lim, M.S. Yong, E.K. Teo, L.N. Moh, Mater. Sci. Eng. C 31
lP

(2011) 579–587.

[35] R. Willumeit, F. Feyerabend, N. Huber, Acta Biomater. 9 (2013) 8722–8729.

[36] R.C. Zeng, X.T. Li, S.Q. Li, F. Zhang, E.H. Han, Sci. Rep. 5 (2015) 1–14.
na

[37] B. Wang, Biodegradability Behavior of Magnesium Alloy AZ31 in Simulated

Physiological Environment Fluids and the Associated Micro Corrosion Mechanism,


ur

Ph.D. Thesis, Institute of Metal Research, University of Chinese Academy of Sciences,


Jo

2016. (in Chinese)

[38] T. Kawasaki, N. Ogata, H. Akanuma, T. Sakai, H. Watanabe, K. Ichiyanagi, T.

Yamanouchi, Metabolism 53 (2004) 583–588.

[39] T. Kawasaki, H. Akanuma, T. Yamanouchi, Diabetes Care 25 (2002) 353–357.

[40] J.J. Hwang, L. Jiang, M. Hamza, F. Dai, R. Belfort-DeAguiar, G. Cline, D.L. Rothman,

21
G. Mason, R.S. Sherwin, JCI Insight 2 (2017) 1–6.

[41] B. Wang, D. Xu, J. Dong, W. Ke, J. Mater. Sci. Technol. 32 (2016) 646–652.

[42] S. Xu, J. Dong, W. Ke, Int. J. Corros.2010 (2010) 1–6.

[43] S. Raynaud, E. Champion, Biomaterials 23 (2002) 1073–1080.

[44] J.K. Han, H.Y. Song, F. Saito, B.T. Lee, Mater. Chem. Phys. 99 (2006) 235–239.

[45] J. Liu, X. Ye, H. Wang, M. Zhu, B. Wang, H. Yan, Ceram. Int. 29 (2003) 629–633.

[46] L. Wu, L. Zhao, J. Dong, W. Ke, N. Chen, Electrochim. Acta 145 (2014) 71–80.

[47] A. Boyd, M. Akay, B.J. Meenan, Surf. Interface Anal. 35 (2003) 188–198.

of
[48] S. Kačiulis, G. Mattogno, L. Pandolfi, M. Cavalli, G. Gnappi, A. Montenero, Appl. Surf.

ro
Sci. 151 (1999) 1–5.

[49] W. Zhang, B. Tian, K.Q. Du, H.X. Zhang, F.H. Wang, Int. J. Electrochem. Sci. 6 (2011)

5228–5248.
-p
[50] S.W. Ha, M. Kirch, F. Birchler, K.L. Eckert, J. Mayer, E. Wintermantel, C. Sittig, I.
re
Pfund-Klingenfuss, M. Textor, N.D. Spencer, M. Guecheva, H. Vonmont, J. Mater. Sci.:
lP

Mater. Med. 8 (1997) 683–690.

[51] G. Song, Adv. Eng. Mater. 7 (2005) 563–586.


na

[52] G.L. Song, Corros. Sci. 51 (2009) 2063–2070.

[53] G. Song, A. Atrens, D. StJohn, Magnes. Technol. 2001 (2013) 254–262.


ur

[54] G.L. Song, A. Atrens, Adv. Eng. Mater. 1 (1999) 11–33

[55] J. Chen, J. Dong, J. Wang, E. Han, W. Ke, Corros. Sci. 50 (2008) 3610–3614.
Jo

[56] W.J.B. Dufresne, C.J. Rufledt, C.P. Marshall, J. Raman Spectrosc. 49 (2018) 1999–2007.

[57] U. Posset, E. Lo, R. Thull, W. Kiefer, J. Biomed. Mater. Res. 40 (1997) 640–645.

[58] K.L. Rogers, J. Cosmidis, K. Benzerara, F. Guyot, F. Skouri-panet, Front. Earth Sci. 3

(2015) 1–20.

[59] B.O. Fowler, M. Markovi, W.E. Brownlb, Chem. Mater. 5 (1993) 1417–1423.

22
[60] Y. Zhang, C. Yan, F. Wang, W. Li, Corros. Sci. 47 (2005) 2816–2831.

[61] J. Liang, P.B. Srinivasan, C. Blawert, W. Dietzel, Electrochim. Acta 55 (2010) 6802–

6811.

[62] X. Hao, J. Dong, I.I.N. Etim, J. Wei, W. Ke, Corros. Sci. 110 (2016) 296–304.

[63] Y. Song, E.H. Han, K. Dong, D. Shan, C.D. Yim, B.S. You, Corros. Sci. 88 (2014) 215–

225.

[64] I. N. Etim, J. Wei, J. Dong, D. Xu, N. Chen, X. Wei, M. Su, W. Ke, Biofouling 34

(2019) 1–17.

of
[65] I.N. Etim, J. Dong, J. Wei, C. Nan, D. B. Pokharel, A.J. Umoh, D. Xu, M. Su, W. Ke,

ro
Effect of organic silicon quaternary ammonium salts on mitigating corrosion of

reinforced steel induced by SRB in mild alkaline simulated concrete pore solution, J.
-p
Mater. Sci. Technol. (2019), https://doi.org/10.1016/j.jmst.2019.10.006
re
[66] M.C. Zhao, P. Schmutz, S. Brunner, M. Liu, G. ling Song, A. Atrens, Corros. Sci. 51
lP

(2009) 1277–1292.

[67] T. Kokubo, Thermochim. Acta 280/281 (1996) 479–490.


na
ur
Jo

23
of
ro
Table list

-p
Table 1: Fitting results of the EIS data for solution A

Time YHFs (S sn ns RS (Ω Y0-H (S sn nH RH (Ω Y0-F (S sn nF RF (Ω Y0-dl (S sn ndl Rct (Ω RL (Ω L (Ω cm-2)

re
(days) cm-2) cm2) cm-2) cm2) cm-2) cm2) cm-2) cm2) cm2)
1 7.94×10-6 0.668 28.1 6.57×10-6 0.807 265 5.95×10-6 0.755 482.1 7.05×10-6 0.801 801.2 553 3.2×104

P
3 6.14×10-6 0.726 27.7 6.31×10-6 0.871 558.9 5.38×10-6 0.897 501 6.31×10-6 0.887 1138 625 5.0×104

7 6.02×10-6 0.763 30 6.85×10-6 0.887 818.9 4.74×10-6 0.860 1116.6 6.17×10-6 0.852 1901.2 _ _
al
14 5.18×10-6 0.804 70.1 5.88×10-6 0.830 1230. 3.31×10-6 0.892 1773.2 5.35×10-6 0.911 3245.3 _ _
2
n
28 4.56×10-6 0.744 34.6 3.33×10-6 0.706 1781. 1.56×10-6 0.845 2201.5 4.57×10-6 0.899 4895.6 _ _
4
ur
Jo

24
Table 2 Fitting results of the EIS data for solution B

of
Time YHFs (S sn ns RS (Ω Y0-H (S sn nH RH (Ω Y0-F (S sn nF RF (Ω Y0-dl (S sn ndl Rct (Ω RL (Ω L (Ω cm-2)
(days) cm-2) cm2) cm-2) cm2) cm-2) cm2) cm-2) cm2) cm2)

ro
1 8.79×10-6 0.707 29.1 8.77×10-6 0.788 259.4 6.99×10-6 0.735 381 7.13×10-6 0.815 523 401 2.2×104

3 8.57×10-6 0.793 31.1 7.75×10-6 0.791 532.1 4.63×10-6 0.812 454 6.84×10-6 0.830 930.1 512.2 1.3×104

-p
7 7.60×10-6 0.843 30.5 6.34×10-6 0.866 735.4 4.92×10-6 0.825 998.5 5.31×10-6 0.851 1420.5 _ _

3.10×10-6 5.56×10-6 3.12×10-6 4.99×10-6

re
14 0.896 77.1 0.874 960.2 0.858 1470. 0.887 2990.3 _ _
5
28 2.76×10-6 0.819 40 .3 4.17×10-6 0.867 1615. 2.69×10-6 0.836 1869 4.15×10-6 0.828 3115.7 _ _
3

P
n al
ur
Jo

25
Figure list:

of
ro
Fig. 1. Schematic representation of In-situ Laser Raman Spectrum measurement during the

potentiostatic polarization.
-p
re
lP
na
ur
Jo

26
of
ro
-p
re
lP
na

Fig. 2. Surface morphology of the AZ31 Mg alloy immersed in solution A (a)-(e) and solution

B (f)-(j) and corresponding EDS for different time intervals: 1, 3, 7, 14 and 28 days. Images (a'),
ur

(a")-(e') and (f'), (f")-(j') are the EDS results of the corresponding SEM images, respectively.
Jo

27
Fig. 3. Cross- section morphology of the AZ31 Mg alloy immersed in the solution A (a)-(e)

of
and solution B (f)-(j) for different time intervals: 1, 3, 7, 14 and 28 days.

ro
-p
re
lP
na

Fig. 4. FTIR spectra of the AZ31 Mg alloy immersed in solutions A and B for 28 days.
ur
Jo

28
Fig. 5. Typical XPS deconvoluted spectra of surface film for the AZ31 Mg alloy immersed in

solution A (a)-(e) and solution B (f)-(j) for 28 days: (a) Mg-1s, (b) O-1p and (c) Ca-2p, (d) P-

of
2p and (e) C-1s and solution B (f) Mg-1s, (g) O-1p and (h) Ca-2p and (i) P-2p and (j) C-1s,

respectively.

ro
-p
re
lP
na

Fig. 6. XRD patterns of the AZ31 Mg alloy immersed in solutions A and B for 28 days.
ur
Jo

29
Fig. 7. Curves of volume of hydrogen released (a) and corresponding hydrogen evolution rate

of
(HER) (b) in solutions A and B as function of immersion time.

ro
-p
re
lP
na
ur
Jo

30
of
Fig. 8. Potentiodynamic polarization curves of the AZ31 Mg alloy after immersion in the

ro
solutions A and B for 28 days.

-p
re
lP
na
ur
Jo

31
of
ro
-p
re
lP
na
ur

Fig. 9. Comparison of total current density (a), volume of hydrogen (b), Hydrogen release

corresponds to current density-time curve (c), Mg dissolution during the anodic polarization
Jo

for 22 h in solutions A and B (d), and corresponding In-situ Raman spectra after 22 h of

potentiostatic polaraizaton with the anodic potential of -1.35 V (e).

32
of
ro
-p
re
Fig. 10. Electrochemical impedance spectra of the AZ31 Mg alloy in solutions A and B. (a)

and (c) Bode plots, (b) and (d) Phase angle for solution A and B, respectively.
lP
na
ur

Fig. 11. Equivalent circuits of the electrochemical impedance spectroscopy (EIS) spectra.
Jo

33

You might also like