You are on page 1of 23

Microstructure design for biodegradable magnesium alloys basing on

ed
macroscopic and quasi-in-situ EBSD observations on biocorrosion behavior

Xiang Wanga, Chun Chena, Lingyu Lia, Jialin Niua, Shaokang Guanb , Hua Huanga,*,

iew
Hui Zengc,*, Guangyin Yuana,*

a National Engineering Research Center of Light Alloy Net Forming and Key State

Laboratory of Metal Matrix Composites, School of Materials Science and Engineering,

v
Shanghai Jiao Tong University, Shanghai, 200240, China;

re
b School of Materials Science and Engineering, Zhengzhou University, Zhengzhou,

450002, China; er
c National & Local Joint Engineering Research Center of Orthopaedic Biomaterials,
pe
Peking University Shenzhen Hospital, Shenzhen 518036, China;

* Corresponding authors: TEL: +86-21-34203051; FAX: +86-21-34202794;

Emails: huangh@sjtu.edu.cn (Hua Huang); zenghui_36@163.com (Hui Zeng);


ot

gyyuan@sjtu.edu.cn (Guangyin Yuan)

Abstract
tn

The influence of alloying elements, grain size and orientation on biocorrosion

resistance of Mg-2.1Nd-0.2Zn-0.5Zr (wt.%) were studied by immersion,


rin

electrochemical and quasi-in-situ electron back-scattering diffraction tests, to give a

better microstructure design. The corrosion rate of this alloy is lower than that of pure
ep

Mg due to the improved potential of the matrix, the presence of Nd2O3 in the corrosion

layer and smaller grains. Preparation of samples with an average grain size smaller than
Pr

10 μm and weakening basal texture was recommended. These results are useful for

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
uncovering the corrosion behavior of Mg alloys and designing microstructures for

ed
biodegradable implants.

Keywords: A. Magnesium alloy; C. Grain orientation; C. Grain size; C. Biocorrosion;

iew
B. Quasi-in-situ EBSD;

1. Introduction

Magnesium (Mg) alloys have attracted much attention as potential biodegradable

v
materials due to their good biocompatibility and appropriate mechanical properties [1-

re
4]. Unlike non-degradable materials, such as stainless steel and cobalt nickel alloy [5,

6], active Mg alloys can be gradually degraded after being implanted into the human
er
body. However, the degradation rates of Mg alloys are relatively fast for most clinical
pe
applications [7]. Sometimes the Mg alloy implants have been completely degraded but

the diseased region has not recovered [8, 9]. Therefore, the biocorrosion resistance of

Mg alloy is a subject of wide interest.


ot

For Mg alloys with certain composition, in terms of the microstructure, there are four

main factors affecting the biocorrosion rate including the solid solution of alloying
tn

elements in the Mg matrix [10-15], secondary phase precipitation [16-21], grain size

[22-30] and orientation [31-37]. The effect of the first factor on the corrosion properties
rin

of Mg alloys is ambiguous. Some solute atoms could increase the potential of the matrix,

while others decrease [10-15]. Hence, the biocorrosion resistance of the matrix may be
ep

improved or deteriorated after alloying. For the second factor, the influence of the

secondary phase on the corrosion properties of Mg alloys has been clarified. Generally
Pr

speaking, the potential difference between the secondary phase and the matrix is smaller,

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
the size of the secondary phase is smaller or the distribution is more even, and the

ed
corrosion resistance is better [16-21]. For the third factor, the results of grain size on

corrosion performance sometimes show an unclear tendency [22] or even the opposite

iew
law [23-25], which is related to the alloy systems. Further research is needed to clarify

the underlying mechanism [26-30]. Regarding the last factor, it is generally believed

that grain orientation has some influence on corrosion. Lower surface energy implies

v
that the diffusion rate of atoms is slow, which usually contributes to high corrosion

re
resistance [33].

In this study, the in vitro biocorrosion behavior of Mg-2.1Nd-0.2Zn-0.5Zr (wt.%)


er
(denoted as JDBM) alloy was studied in detail by immersion, electrochemical and
pe
quasi-in-situ electron back-scattering diffraction tests, to give a better microstructure

design strategy. This biodegradable alloy was chosen for this study because of its

combined good mechanical properties [38] and uniform corrosion performance [39].
ot

Pure Mg has also been studied as a reference metal. In order to study the effect of texture

at the macro level, two types of samples were prepared by cutting the pure Mg and
tn

JDBM alloy bars along the direction parallel or perpendicular to the extrusion direction.

In addition, unlike the previous papers, the corrosion resistance of grains with different
rin

orientations and sizes was directly compared at the micro level by quasi-in-situ EBSD

observations. Uncovering the corrosion behavior and underlying mechanism will be of


ep

significance for designing microstructures and fabrication processes of biodegradable

JDBM and other Mg alloy implants.


Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
2. Material and experimental details

ed
2.1 Material preparation

As-extruded Mg-2.1Nd-0.2Zn-0.5Zr (wt.%) (denoted as JDBM alloy) and pure Mg

iew
(99.9%) bars with the same dimension of 15 mm in diameter were used as raw materials.

These bars were obtained by hot extrusion with a ratio of 16:1. The extrusion

temperature for the former is 370 °C, while for the latter is 270 °C. Then to reduce the

v
influence of secondary phases, as-extruded JDBM alloy bars were solid solution treated

re
at 520 °C for 8 hours (hrs) [38, 40-42]. Samples for testing and characterization were

obtained by wire cutting, grinding and polishing. The four studied samples were taken
er
perpendicular or parallel to the extrusion direction of JDBM alloy and pure Mg bars,
pe
denoted as JDBM-A, JDBM-B, Mg-A and Mg-B, respectively.

2.2. Microstructure analysis


ot

For the observation of scanning electron microscopy (SEM, TESCAN-MIRA3), JDBM

and pure Mg samples were ground with #320, #1200, #3000, and #7000 grit SiC papers
tn

in turn, polished by MgO suspension and then chemically etched by solution (4.2 g

picric acid, 10 ml acetic acid, 70 ml ethanol and 10 ml distilled water) for 15 seconds
rin

(s). For the characterization of crystallographic orientation by SEM (TESCAN-GAIA3)

with an electron back-scattering diffraction (EBSD) detector, samples were also ground
ep

with those SiC papers in turn, mechanically ground by an automatic preparation system

(TegraPol-21, Struers) and then chemically etched by another solution (5 ml nitric acid,
Pr

15 ml acetic acid, 20 ml water and 60 ml absolute ethanol). The etch time only took 0.5

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
s, and the purpose is to eliminate the surface stress to obtain a higher EBSD resolution.

ed
2.3. Immersion tests

iew
Samples for immersion tests were discs with a dimension of 12 mm in diameter and 2

mm in thickness. These discs were ground with #320, #1200, and #3000 grit SiC papers

to remove the oxide layer and then cleaned up. Hank’s solution [43-45] was used and

v
these immersion tests were carried out for 14 days in a water bath pit with a constant

re
temperature of 37 ± 0.5 °C. The time interval for changing Hank’s solution was 24 hrs.

The weight of the sample before and after the immersion test was measured. The
er
corrosion rate of the sample was calculated by the following equation [18, 46]:

V = 87600∆𝑚/𝜌𝐴𝑡 (1)
pe
Where, ∆m is the weight loss of the samples (g), ρ is the density of Mg alloy (g/cm3),

A is the contact area between the sample and the solution (cm2), and t is the immersion
ot

time (h).

In order to monitor the dynamic changes in the corrosion process of samples, the
tn

amount of released hydrogen was recorded hourly in the first 24 hrs of the immersion

test, and once every 12 hrs in the following 13 days. After the immersion test, the
rin

corrosion morphology of samples was characterized by the stereoscopic microscope

(Leica-M125) and SEM (TESCAN-MIRA3) with energy-dispersive X-ray


ep

spectroscopy (EDS).

To observe the thickness and the element distribution of the corrosion layer, cross-
Pr

section samples studied by SEM (TESCAN-MIRA3) were cold-mounted into the epoxy

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
resin and also ground with SiC papers in turn. In addition, some immersed samples

ed
were further immersed in chromic acid solution (200g/L Cr2O3) for 5 minutes,

ultrasonically cleaned with anhydrous alcohol and dried in air to remove the corrosion

iew
products. These samples without corrosion layer were also characterized by SEM

(TESCAN-MIRA3). Moreover, Hank’s solution after soaking samples for 3, 6, 12, 18,

24, 72, 168, and 336 hrs were respectively taken for inductively coupled plasma (ICP,

v
Thermo-iCAP7600) tests to study the change of Mg2+ concentration. According to the

re
ASTM standard (ASTM G31-72), the ratio of Hank’s solution volume to the sample

surface was 40 ml/cm2. For the reliability of the corrosion data, three groups of parallel
er
tests were carried out for the same type of samples.
pe
2.4. Electrochemical tests

The four studied samples were ground by #320, #1200, and #3000 grit SiC papers and
ot

ultrasonic washed in ethanol and then immersed in Hank’s solution at 37 ± 0.5 °C for

electrochemical tests. There were three electrodes in this test, the platinum counter
tn

electrode, the reference electrode (a saturated calomel electrode), and the working

electrode (the samples installed in the special mold) [18]. The measurements of open
rin

circuit potential (OCP) were conducted for 3600 s to obtain a stable potential. After

that, the starting and ending frequencies of electrochemical impedance spectroscopy


ep

(EIS) measurements were 100 kHz and 10 mHz, respectively, and the data were

analyzed by the software ZSimpWin. The parameter of the dynamic polarization


Pr

measurement was set in the range of -0.5V to 0.5V, and corresponding results were

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
analyzed by the software DHElecChem.

ed
2.5. In-situ EBSD observation

iew
Typical JDBM and pure Mg samples were chosen for in-situ EBSD observation. EBSD

characterization was performed on the same position of the samples before and after

immersion in Hank’s solution for 1 and 3 days at 37 ± 0.5 °C. The software (channel 5)

v
was used to analyze the information on each grain, including the grain size, orientation,

re
and corrosion status. It is common to characterize the corrosion behavior of Mg alloys

by macroscopic characterization methods, such as weight loss, corrosion morphology,


er
etc., but it is rare and extremely necessary to characterize the corrosion behavior of Mg

alloys by microscopic characterization methods, such as quasi-in-situ EBSD


pe
observation. This microscopic characterization method can intuitively provide grain

information and corresponding corrosion conditions, and establish a direct relationship


ot

between grain information and corrosion.


tn

2.6 Statistical analysis

Each data in this paper was shown in the form of mean ± standard deviation of three
rin

experimental data of parallel tests. The standard with statistical significance was set as

p-value < 0.05.


ep

3. Results

3.1. Microstructure
Pr

Fig. 1 shows the SEM images of the four studied samples. The morphologies of the

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
samples JDBM-A and JDBM-B are similar and the average grain size is about 46 ± 1

ed
μm. The morphologies of the samples Mg-A and Mg-B are also similar but the average

grain size is about 65 ± 2 μm. Therefore, the two studied bars were recrystallization

iew
fully. In addition, as shown in Fig. 1 (a) and (b), a small amount of fine and dispersive

secondary phase particles could be found in JDBM samples, although solid solution

treatment was carried out. According to previous studies on JDBM, these secondary

v
phase particles should be Mg12Nd phase [11, 46, 47]. Furthermore, as shown in Fig. 1

re
(c) and (d), different from the JDBM samples, a small number of twins could be

observed in pure Mg samples. It may be introduced by the grinding and polishing


er
process during sample preparation since the pure Mg sample is relatively soft when the
pe
average grain size is relatively large.

To obtain more information on the grains, the texture of the four studied samples was

characterized by EBSD. As shown in Fig. 2 (a, b, c) and (d, e, f), the exposed plane of
ot

most surface grains for sample JDBM-A is close to (0001) plane, while for sample

JDBM-B is far from. In addition, the intensity of texture for samples JDBM-A and
tn

JDBM-B is relatively weak, namely 3.07 and 2.24 multiples of uniform density (mud),

respectively. As shown in Fig. 2 (g, h, i) and (j, k, l), the exposed plane of most surface
rin

grains for sample Mg-A is also close to the basal plane while for sample Mg-B is also

far from. The intensity of texture for samples Mg-A and Mg-B is also relatively weak,
ep

namely 2.21 and 3.33 mud, respectively. Therefore, the two studied bars both exhibits

weak basal texture. Therefore, in terms of texture, JDBM-A is similar to Mg-A, while
Pr

JDBM-B is similar to Mg-B.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
3.2. In vitro corrosion tests

ed
The calculated degradation rates of the four studied samples according to the mass loss

are presented in Fig.3. The corrosion rate of JDBM-A (0.0895 ± 0.0030 mm/year) is

iew
relatively smaller compared with JDBM-B (0.1129 ± 0.0105 mm/year). For pure Mg,

the corrosion rate of Mg-A and Mg-B changes more obviously. Mg-B present a larger

corrosion rate of 0.1605 ± 0.0196 mm/year compared with Mg-A of 0.1107 ±

v
0.0039mm/year. Therefore, no matter for JDBM or pure Mg, the samples containing

re
more grains with exposed plane close to (0001) show better corrosion resistance. In

addition, the JDBM shows the better corrosion resistance than pure Mg when they
er
exhibit the similar texture. The macro and micro corrosion morphology of the four
pe
studied samples with the corrosion layer are shown in Fig. S1 and S2 (Supporting

information), which correspond well to the corrosion rate results shown in Fig. 3.

According to the above results, JDBM-B and Mg-B samples show worse corrosion
ot

resistance than JDBM-A and Mg-A, respectively. Hence, the corrosion layer of JDBM-

B and Mg-B should be thicker and looser. To study the corrosion layer expediently,
tn

JDBM-B and Mg-B samples were further selected to obtain information on the cross

sections. As shown in Fig. 4, the thickness of the corrosion layer for samples JDBM-B
rin

and Mg-B is about 4.73 μm and 10.07 μm, respectively. Therefore, the corrosion rate

of Mg-B is faster than that of the JDBM-B, which is corresponding to the corrosion rate
ep

results shown in Figure 3. In addition, compared to the JDBM-B, the corrosion layer of

Mg-B has more cracks and the size of the crack is larger. Hence, the corrosion layer of
Pr

JDBM-B is more compact than that of the sample Mg-B. Two samples both mainly

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
contain carbon (C), calcium (Ca), phosphate (P), magnesium (Mg), and oxide (O), but

ed
sample JDBM-B contains a small amount of Nd element. At the same time, Ca and P

are mainly distributed on the surface of the corrosion layer, while O has a high content

iew
throughout the corrosion layer. Therefore, the outer layer maybe is the calcium-

phosphorus salt, while the inner layer maybe is MgO or Mg(OH)2.

Fig. 5 shows the surface SEM images of the studied samples after removing corrosion

v
products. The corrosion pits on the surface of samples JDBM-A and Mg-A both are

re
uniform and shallow.

To obtain more information about the dynamical characteristics of the corrosion, the
er
volume of hydrogen evolution and Mg2+ concentration in Hank’s solution during tests
pe
were collected, as shown in Fig. 6. Within the first few dozen hours, the corrosion rate

seems very slow. Then, the corrosion rate increased obviously and gradually reaches a

steady state. Hence, the corrosion layer on the surface of samples has protective effects.
ot

In addition, the volume of hydrogen evolution for JDBM-B is always greater than that

of JDBM-A and the concentration of Mg2+ is also evidently greater, which further
tn

confirms that the corrosion rate of JDBM-B is higher than that of the JDBM-A. These

results for pure Mg samples are just the same which means the corrosion rate of Mg-B
rin

is higher than that of the Mg-A.

3.3. Electrochemical tests


ep

Fig. 7 (a) and (d) show the open circuit potential (OCP) curves of the JDBM and pure

Mg samples. After immersion in Hank’s solution for 3600 s, a steady state was reached,
Pr

and the value for the JDBM-A, JDBM-B, Mg-A, and Mg-B samples is around -1.36, -

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
1.53, -1.51 and -1.65 VSCE, respectively. Therefore, the samples containing more grains

ed
with the exposed plane close to (0001) show higher open circuit potential. In addition,

the JDBM sample shows higher open circuit potential than that of the pure Mg when

iew
they exhibit a similar texture.

The potentiodynamic polarization curves of the four samples were also measured and

the results were shown in Fig. 7 (b) and (e). The corrosion parameters were determined

v
using the software DHElecChem and were listed in Table 1. The polarization resistance

re
Rp is inversely proportional to the corrosion rate, which can be obtained by Stern-Gray

formula [10]: er
𝑅p = 𝛽a × 𝛽c/[2.303·𝑖corr(𝛽a + 𝛽c)] (2)
pe
Where, βa, 𝛽c are the slopes of the anode and cathode of Tafel curves and 𝑖corr is the

corrosion current density. The Rp value for JDBM-A is 8.48×106 Ω·cm2, which is larger

than that value (1.89×106 Ω·cm2) for JDBM-B. The Rp value for Mg-B is 1.15×106
ot

μA·cm-2, which is smaller than that value (3.39×106 Ω·cm2) for Mg-A. Hence, the

sequence of testing samples for corrosion resistance is JDBM-A>Mg-A>JDBM-B>


tn

Mg-B, which is corresponding well to the corrosion rate results shown in Fig.3. In

addition, the 𝑖corr shown in Table 1 represents the same corrosion rule as the result of
rin

the polarization resistance.

Fig. 7 (c) and (f) show the EIS curves of the JDBM and pure Mg samples. The diameter
ep

of the loop for the JDBM-A is much larger than that of JDBM-B, which indicates that

JDBM-A presents a better corrosion resistance [18]. Similarly, the diameter of the loop
Pr

for the Mg-A is much larger than that of Mg-B, which means that the Mg-A is more

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
resistant to be corroded. The sequence of the loop diameter for these four studied

ed
samples is JDBM-A>Mg-A>JDBM-B>Mg-B, which also corresponds well to the

corrosion rate results shown in Fig. 3. Hence, as for the JDBM-A and Mg-A or the

iew
JDBM-B and Mg-B with similar texture, the corrosion resistance for JDBM alloy is

better than pure Mg.

To further clarify the corrosion characteristics of the studied samples, an

v
electrochemical equivalent circuit model was constructed to fit the EIS spectra, as

re
shown in Fig. 8. The corresponding simulation parameters were listed in Table. 2. In

the fitting circuit diagram, Rs represents the solution resistance, R1 and CPE1 represent
er
the resistance and capacitance of the corrosion layer outside the substrate, and Rct and
pe
CPE2 represent the charge transfer resistance and capacitance. In general, the larger Rct

value indicates the lower dissolution rate of the Mg matrix and the lower CPE2 value

indicates a more compact surface [48, 49]. The sequence of the Rct is JDBM-A>Mg-
ot

A>JDBM-B>Mg-B and the opposite rule could be found for the CPE2, which are also

corresponding well with the corrosion rate results shown in Fig.3. Therefore, the
tn

exposed plane close to (0001) and alloying with Nd, Zn and Zr are both beneficial to

enhance the protective effect of the corrosion layer.


rin

3.4. In-situ observation


ep

Fig. 9 shows the EBSD results of JDBM and pure Mg samples before and after 1 or 3

days of immersion. Statistical analysis of the size, orientation, and degree of corrosion
Pr

for each grain in each sample was carried out. The size and orientation of each grain

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
could be extracted using the software channel 5, while the degree of corrosion needs to

ed
differentiate. Zero analytic point will appear in the EBSD result due to the presence of

corrosion products covering the Mg matrix. 5 grades were defined in this study

iew
according to the variation value of zero analytic point in the same grain before and after

corrosion, as shown in Fig. 10. Grade 0: the value does not change; grade 1: the value

is greater than or equal to 1 but less than 5; grade 2: the value is greater than or equal

v
to 5 but less than 10; grade 3: the value is greater than or equal to 10 but less than 15;

re
grade 4: the value is greater than or equal to 15, or the white spots are connected into

blocky areas. Hence, with the increase of the grade, the corrosion of the grain is more
er
severe.
pe
The data containing grain size, orientation angle and corresponding corrosion grade of

each grain in the studied samples were extracted from EBSD data. The corrosion grade

vs grain size or included angle of orientation is shown in Fig. 11. The included angle
ot

of orientation represents the included angle between the [0001] of grain and the normal

direction of the observation plane, which is the surface of the sample. As shown in Fig.
tn

11 (a) and (c), for both JDBM and pure Mg samples, the corrosion grade for those

grains with small grain sizes is usually low. Hence, the smaller the grain size, the more
rin

corrosion resistance. As shown in Fig. 11 (b) and (d), for the JDBM sample, there is no

obvious rule between the included angle of orientation and corrosion grade; for the pure
ep

Mg sample, the angle of the grains with relatively low corrosion grade is concentrated

between 45° and 75°, and even a few grains with an orientation angle close to 90° have
Pr

only corrosion grade 1. This phenomenon deviates from many reports [31-35], that is,

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
the basal plane of grains is more corrosion resistant. Besides, for the pure Mg sample

ed
corroded for 3 days, the corrosion grade of the grains in this sample is almost the grade

4 no matter how large the grain size and the orientation angle are. Maybe this result is

iew
closely related to the twinning since many twins can be found for pure Mg samples

shown in Fig. 9.

Combining all the data, Fig. 12 could be obtained. When the grain size is less than 10

v
μm, the influence of orientation is not obvious, and these grains show good corrosion

re
resistance. When the grain size is in the range of 10-60 μm, the influence of orientation

on corrosion is very obvious. When the angle is less than 45°, the surface is relatively
er
corrosion resistant, on the contrary, the surface is easy to be corroded. When the grain
pe
size is greater than 60 μm, the influence of orientation on corrosion is also not obvious

and the surface is not resistant to corrosion. In summary, grains with a smaller size and

a surface of which closer to the basal plane are both beneficial to improve corrosion
ot

performance.

4. Discussion
tn

In terms of the microstructure, the solid solution of alloying elements in the matrix,

secondary phase, grain size and orientation are the four main factors affecting the
rin

corrosion of Mg alloy. The influence of secondary phases on the corrosion property of

Mg alloy is mainly due to the galvanic corrosion between the secondary phases and
ep

matrix [50, 51]. In this study, to decrease the effects of the secondary phase, JDBM

samples were solid solution treated at 520 °C for 8 hrs. As shown in Fig. 1 (a) and (b),
Pr

fine and uniformly dispersed Mg12Nd phases still could be found in the JDBM samples

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
after solution treatment. In the previous research, it is found that fine and dispersed

ed
secondary phase has no obvious adverse effects on the corrosion property of JDBM

alloy [39, 42, 52]. Since the effect of the secondary phase on corrosion is very clear, in

iew
this study, a solid solution of alloying elements, grain size and orientation are the main

studied factors.

v
4.1. Solid solution of alloying elements

re
To illustrate the effects of alloying on corrosion clearly, thermodynamics and kinetics

of corrosion should be considered simultaneously. Corrosion potential representing the


er
activity of the metal belongs to the thermodynamics characteristics, while the protective

effects of the corrosion layer are the kinetics characteristics. Hence, the corrosion of
pe
Mg after alloying is closely related to the effects of solute atoms on the matrix and

corrosion layer, when the effects of secondary phase precipitation on corrosion were
ot

ignored. In this study, the potential of the Mg matrix increased slightly after alloying

with Nd, Zn and Zr simultaneously (as shown in Fig. 7), which implies that the alloying
tn

is beneficial. On the other hand, the corrosion layer of the JDBM sample is

more compact than that of the pure Mg sample (as shown in Fig. 4), and the protective
rin

effects are better (as shown in Fig. 6). Therefore, according to the thermodynamics and

kinetics, the corrosion rate of JDBM should be lower than that of the pure Mg, which
ep

corresponds well to the corrosion rate results shown in Fig. 3.

In addition, compared to pure Mg, the protective effects for the corrosion layer of
Pr

JDBM alloy is better, which is probably due to the presence of Nd2O3 in the corrosion

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
inner layer (as shown in Fig. S2) [53-55]. Protective effects of the corrosion layer are

ed
closely related the Pilling-Bedworth (P-B) ratio. The MgO layer suffers tension stress

due to the low P-B ratio of MgO (~0.81) [56], which would further lead to cracks and

iew
then decreased its protective effects. However, the P-B ratio of Nd2O3 is about 1.16 [57,

58], which is larger than that of MgO. Hence, introducing Nd2O3 into the corrosion

layer of Mg alloys (Fig. S2) will increase the P-B ratio, and then decrease the tension

v
stress and cracks (Fig. 4). Hence, as the corrosion layer is doped with Nd2O3, the

re
protective effects of the corrosion layer could be improved.

Comprehensively, the corrosion rate of JDBM is lower than that of pure Mg partly due
er
to the improved potential of the matrix and the presence of Nd2O3 in the corrosion layer.
pe
4.2. Grain size

There are many reported studies about the influence of grain size on the corrosion
ot

property of Mg alloy by adjusting the average grain size, but these results are extremely

contradictory. Some scholars have indicated that grain boundaries act as


tn

crystallographic defects [27, 28] possessing a higher element diffusion rate [59] and

thus facilitating corrosion. Hence, they believe that samples with small grains are prone
rin

to be corroded due to containing more grain boundaries. Others hold the opposite

opinion because they believe that samples with small grain sizes represent better
ep

corrosion resistance. The dense corrosion network is easier to form at grain boundaries,

which could act as a corrosion barrier and retard corrosion kinetics [29, 30]. Actually,
Pr

corrosion is complicated progress and the above two opinions were both reasonable. In

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
different alloy systems, the dominant factor might be different. In this study, as

ed
mentioned above, the protective effects for the corrosion layer of JDBM alloy are better

than that of pure Mg due to the presence of Nd in the protective corrosion layer. Hence,

iew
retard corrosion kinetics is probably the dominant factor. In addition, grain refinement

is beneficial for improving the corrosion resistance of JDBM alloys, which is also

confirmed in this study. As shown in Fig. 10, 11, and 12, grains with smaller sizes (less

v
than 10 μm) exhibit a lower corrosion rate while grains with a size larger than 60 μm

re
are more easily corroded. Hence, the JDBM alloy samples show better corrosion

resistance than pure Mg, which may also be related to the smaller grain size. On the
er
other hand, as we know, grain refinement is also beneficial to the improvement of
pe
mechanical properties based on the typical Hall-Petch theory. For biodegradable Mg

implants, superior mechanical properties and corrosion resistance are expected.

Therefore, grain refinement below 10 μm is recommended for biomedical Mg alloys


ot

and their implants.


tn

4.3. Grain orientation

Usually, higher atomic coordination is associated with a more-closely packed plane, a


rin

lower surface energy and a higher activation energy for dissolution [31-33]. Hence, a

crystal plane with higher atomic coordination shows better corrosion resistance. The
ep

atomic density of the (0001) basal plane, (1120) and (1010) non-basal planes, is

1.13×1019, 6.94×1018, and 5.99×1018 atoms·m−2, respectively [33]. Correspondingly,


Pr

the surface energy is 1.54×104, 3.04×104, and 2.99×104 J/mol [31, 32], which implies

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
the (0001) basal plane is more difficult to dissolve and corrode. The theoretical

ed
dissolution rates of crystallographic planes (1120) and (1010) are about 18-20 times

higher than that of the basal plane (0001) [31]. In this study, as shown in Fig. 2, the

iew
JDBM-A and Mg-A samples possess more surface grains with the exposed plane

closing to the basal plane and are more resistant to be corroded. In addition, as shown

in Fig. 10 and Fig. 12, when the grain size is in the range of 10-60 μm, the influence of

v
grain orientation on corrosion is very obvious. When the orientation angle is less than

re
45°, the grain shows relatively better corrosion resistance, otherwise, the grain is easier

to corrode. When the grain size is smaller than 10 μm or larger than 60 μm, the influence
er
of grain orientation on corrosion becomes less obvious because grain size maybe is the
pe
main influence factor. Although grain orientation has an obvious influence on the

corrosion of Mg alloy, usually, it is not preferred to process samples with strong

textures. The texture will cause anisotropy of samples, such as tension-compression


ot

yield strength asymmetry [60, 61], which is harmful for later plastic deformation. Hence,

many studies have been carried out on the texture randomization of Mg alloys [62, 63].
tn

Comprehensively, grain orientation has an obvious influence on the corrosion of Mg

alloys, and the grain with the exposed plane closing to the basal plane shows better
rin

corrosion resistance.

5. Conclusions
ep

In this study, as-extruded pure Mg and JDBM alloy bars were used to study the in vitro

biocorrosion behavior by various corrosion tests and characterization. The following


Pr

research results can be obtained:

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
1) The corrosion rate of JDBM is lower than that of pure Mg slightly due to the

ed
improved potential of the matrix, the presence of Nd2O3 in the corrosion layer and the

smaller grains.

iew
2) Grain size has an obvious influence on the corrosion resistance of JDBM alloy and

pure Mg. The grain size is smaller, and the corrosion resistance of the sample is better.

Grain refinement below 10 μm is recommended for JDBM alloys and their implants.

v
3) Grain orientation has a certain effect on the corrosion resistance of JDBM alloy and

re
pure Mg. JDBM-A and Mg-A samples containing more grains with exposed plane close

to (0001) show better corrosion resistance.


er
CRediT authorship contribution statement
pe
Xiang Wang: Conceptualization, Investigation, Methodology, Writing-original draft,

Writing-review & editing. Chun Chen: Conceptualization, Investigation. Lingyu Li:

Methodology, Investigation. Jialin Niu: Resources, Investigation. Shaokang Guan:


ot

Formal analysis. Hua Huang: Methodology, Supervision, Funding acquisition,

Writing-review & editing. Hui Zeng: Resources, Funding acquisition. Guangyin


tn

Yuan: Methodology, Supervision, Funding acquisition, Writing-review & editing.

Declaration of Competing Interest


rin

The authors declare that they have no known competing financial interests or personal

relationships that could have appeared to influence the work reported in this paper.
ep

Acknowledgments

This work was funded by grants from the National Natural Science Foundation of China
Pr

(No.52130104 and 52273318) , The National Key Research and Development Program

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
of China( No. 2021YFC2400701 and 2020YFE020210) , and the Science and

ed
Technology Innovation Commission of Shenzhen Municipality (No.

JCYJ20220818102815033)

iew
Data Availability

Data will be made available on request.

References

v
[1] N.M. Bexiga, M.M. Alves, M.G. Taryba, S.N. Pinto, M.F. Montemor, Early biomimetic degradation

re
of Mg-2Ca alloy reveals the impact of β-phases at the interface of this biomaterial on a micro-scale level,
Corros Sci 207 (2022) 110526.
[2] A.A. Oliver, M. Sikora-Jasinska, A.G. Demir, R.J. Guillory, 2nd, Recent advances and directions in
the development of bioresorbable metallic cardiovascular stents: Insights from recent human and in vivo
studies, Acta Biomater 127 (2021) 1-23.
er
[3] X. Wang, B. Zhou, H. Huang, J. Niu, S. Guan, G. Yuan, Extraordinary ductility enhancement of Mg-
Nd-Zn-Zr alloy achieved by electropulsing treatment, J Magnes Alloy https:// doi.org/ 10.1016/
j.jma.2022.07.007.
pe
[4] L. Lei, Z. Cui, H. Pan, K. Pang, X. Wang, H. Cui, Effect of extrusion on the microstructure and
corrosion behavior of Mg-Zn-Mn-(0, 1.5)Sr alloys in Hank’s solution, Corros Sci 195 (2022) 109975.
[5] S.H. Im, D.H. Im, S.J. Park, Y. Jung, D.-H. Kim, S.H. Kim, Current status and future direction of
metallic and polymeric materials for advanced vascular stents, Prog Mater Sci 126 (2022) 100922.
[6] E. Vahabli, J. Mann, B.S. Heidari, M. Lawrence-Brown, P. Norman, S. Jansen, E. De-Juan-Pardo, B.
ot

Doyle, The Technological Advancement to Engineer Next-Generation Stent-Grafts: Design, Material,


and Fabrication Techniques, Adv Healthc Mater 11(13) (2022) 2200271.
[7] D. Bian, X. Zhou, J. Liu, W. Li, D. Shen, Y. Zheng, W. Gu, J. Jiang, M. Li, X. Chu, L. Ma, X. Wang,
tn

Y. Zhang, S. Leeflang, J. Zhou, Degradation behaviors and in-vivo biocompatibility of a rare earth- and
aluminum-free magnesium-based stent, Acta Biomater 124 (2021) 382-397.
[8] L.S.M. Kerkmeijer, R.Y.G. Tijssen, S.H. Hofma, R.J. van der Schaaf, E.K. Arkenbout, A. Weevers,
R.P. Kraak, Y. Onuma, P.W. Serruys, J.J. Piek, J.G.P. Tijssen, J.P.S. Henriques, R.J. de Winter, J.J.
rin

Wykrzykowska, Three-year clinical outcomes of the absorb bioresorbable vascular scaffold compared to
Xience everolimus-eluting stent in routine PCI in patients with diabetes mellitus-AIDA sub-study,
Catheter Cardio Inte 98(4) (2021) 713-720.
[9] M.M. Schoos, P. Clemmensen, G.D. Dangas, Second-generation drug-eluting stents and
ep

bioresorbable vascular scaffolds in patients with diabetes, JACC Cardiovasc Interv 7(5) (2014) 494-6.
[10] L.G. Bland, A.D. King, N. Birbilis, J.R. Scully, Assessing the Corrosion of Commercially Pure
Magnesium and Commercial AZ31B by Electrochemical Impedance, Mass-Loss, Hydrogen Collection,
and Inductively Coupled Plasma Optical Emission Spectrometry Solution Analysis, Corrosion 71(2)
Pr

(2015) 128-145.
[11] J.-W. Chang, P.-H. Fu, X.-W. Guo, L.-M. Peng, W.-J. Ding, The effects of heat treatment and
zirconium on the corrosion behaviour of Mg–3Nd–0.2Zn–0.4Zr (wt.%) alloy, Corros Sci 49(6) (2007)

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
2612-2627.
[12] B. Zberg, P.J. Uggowitzer, J.F. Loffler, MgZnCa glasses without clinically observable hydrogen

ed
evolution for biodegradable implants, Nat Mater 8(11) (2009) 887-91.
[13] A. Atrens, G.-L. Song, M. Liu, Z. Shi, F. Cao, M.S. Dargusch, Review of Recent Developments in
the Field of Magnesium Corrosion, Adv Eng Mater 17(4) (2015) 400-453.
[14] A. Bahmani, S. Arthanari, K.S. Shin, Achieving a high corrosion resistant and high strength

iew
magnesium alloy using multi directional forging, J Alloys Compd 856 (2021) 158077.
[15] M. Esmaily, J.E. Svensson, S. Fajardo, N. Birbilis, G.S. Frankel, S. Virtanen, R. Arrabal, S. Thomas,
L.G. Johansson, Fundamentals and advances in magnesium alloy corrosion, Prog Mater Sci 89 (2017)
92-193.
[16] L. Sun, H. Ma, C. Guan, J. Wang, P. Zhang, P. Jin, F. Wei, Y. Peng, Roles of the heat treatment on

v
the Mg-Nd phases and corrosion mechanisms of Mg-4Nd alloy under sulfuric acid type acid rain, Corros
Sci 208 (2022) 110610.

re
[17] A. Bahmani, S. Arthanari, K.S. Shin, Improvement of corrosion resistance and mechanical
properties of a magnesium alloy using screw rolling, J Alloys Compd 813 (2020) 152155.
[18] H. Miao, H. Huang, Y. Shi, H. Zhang, J. Pei, G. Yuan, Effects of solution treatment before extrusion
on the microstructure, mechanical properties and corrosion of Mg-Zn-Gd alloy in vitro, Corros Sci 122
(2017) 90-99.
er
[19] H. Miao, D. Zhang, C. Chen, L. Zhang, J. Pei, Y. Su, H. Huang, Z. Wang, B. Kang, W. Ding, H.
Zeng, G. Yuan, Research on Biodegradable Mg-Zn-Gd Alloys for Potential Orthopedic Implants: In
Vitro and in Vivo Evaluations, ACS Biomater Sci Eng 5(3) (2019) 1623-1634.
pe
[20] M. Taheri, J.R. Kish, N. Birbilis, M. Danaie, E.A. McNally, J.R. McDermid, Towards a Physical
Description for the Origin of Enhanced Catalytic Activity of Corroding Magnesium Surfaces,
Electrochimica Acta 116 (2014) 396-403.
[21] C. Zhang, C. Liu, X. Li, K. Liu, G. Tian, J. Wang, Quantifying the influence of secondary phases
on corrosion in multicomponent Mg alloys using X-ray computed microtomography, Corros Sci 195
ot

(2022) 110010.
[22] C. op't Hoog, N. Birbilis, Y. Estrin, Corrosion of PureMg as a Function of Grain Size and Processing
Route, Adv. Eng. Mater. 10 (2008) 579-582.
tn

[23] K.D. Ralston, N. Birbilis, Effect of Grain Size on Corrosion: A Review, Corrosion 66(7) (2010)
075005.
[24] N. Shrestha, V. Utgikar, K.S. Raja, The Effect of Grain Size on the Corrosion Behavior of Mg-RE
Alloy ZE10A, ECS Transactions 85(13) (2018) 671-682.
rin

[25] K.V. Kutniy, I.I. Papirov, M.A. Tikhonovsky, A.I. Pikalov, S.V. Sivtzov, L.A. Pirozhenko, V.S.
Shokurov, V.A. Shkuropatenko, Influence of grain size on mechanical and corrosion properties of
magnesium alloy for medical implants, Materialwiss Werkstofftech 40(4) (2009) 242-246.
[26] J.W. Seong, W.J. Kim, Development of biodegradable Mg-Ca alloy sheets with enhanced strength
ep

and corrosion properties through the refinement and uniform dispersion of the Mg(2)Ca phase by high-
ratio differential speed rolling, Acta Biomater 11 (2015) 531-42.
[27] G.-L. Song, Z. Xu, The surface, microstructure and corrosion of magnesium alloy AZ31 sheet,
Electrochimica Acta 55(13) (2010) 4148-4161.
Pr

[28] T. Zhang, Y. Shao, G. Meng, Z. Cui, F. Wang, Corrosion of hot extrusion AZ91 magnesium alloy:
I-relation between the microstructure and corrosion behavior, Corros Sci 53(5) (2011) 1960-1968.
[29] M. Alvarez-Lopez, M.D. Pereda, J.A. del Valle, M. Fernandez-Lorenzo, M.C. Garcia-Alonso, O.A.

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
Ruano, M.L. Escudero, Corrosion behaviour of AZ31 magnesium alloy with different grain sizes in
simulated biological fluids, Acta Biomater 6(5) (2010) 1763-71.

ed
[30] G.R. Argade, S.K. Panigrahi, R.S. Mishra, Effects of grain size on the corrosion resistance of
wrought magnesium alloys containing neodymium, Corros Sci 58 (2012) 145-151.
[31] G.-L. Song, R. Mishra, Z. Xu, Electrochem. Commun. 12 (2010) 1009-1012.
[32] M. Gao, K. Yang, L. Tan, Z. Ma, J Magnes Alloy 10(8) (2022) 2147-2157.

iew
[33] B. Jiang, Q. Xiang, A. Atrens, J. Song, F. Pan, Corros. Sci. 126 (2017) 374-380.
[34] Z. Pu, G.L. Song, S. Yang, J.C. Outeiro, O.W. Dillon, D.A. Puleo, I.S. Jawahir, Grain refined and
basal textured surface produced by burnishing for improved corrosion performance of AZ31B Mg alloy,
Corros Sci 57 (2012) 192-201.
[35] G.-L. Song, Z. Xu, Corros Sci 54 (2012) 97-105.

v
[36] P. Du, D. Mei, T. Furushima, S. Zhu, L. Wang, Y. Zhou, S. Guan, J Magnes Alloy 10(5) (2022)
1286-1295.

re
[37] M. Yamasaki, Z. Shi, A. Atrens, A. Furukawa, Y. Kawamura, Influence of crystallographic
orientation and Al alloying on the corrosion behaviour of extruded α-Mg/LPSO two-phase Mg-Zn-Y
alloys with multimodal microstructure, Corros Sci 200 (2022) 110237.
[38] W. Lu, R. Yue, H. Miao, J. Pei, H. Huang, G. Yuan, Mater Lett 245 (2019) 155-157.
[39] L. Mao, L. Shen, J. Niu, J. Zhang, W. Ding, Y. Wu, R. Fan, G. Yuan, Nanophasic biodegradation
er
enhances the durability and biocompatibility of magnesium alloys for the next-generation vascular stents,
Nanoscale 5(20) (2013) 9517-22.
[40] F. Penghuai, P. Liming, J. Haiyan, C. Jianwei, Z. Chunquan, Effects of heat treatments on the
pe
microstructures and mechanical properties of Mg–3Nd–0.2Zn–0.4Zr (wt.%) alloy, Mater Sci Eng A
486(1-2) (2008) 183-192.
[41] F. Penghuai, P. Liming, J. Haiyan, Z. Zhenyan, Z. Chunquan, Fracture behavior and mechanical
properties of Mg–4Y–2Nd–1Gd–0.4Zr (wt.%) alloy at room temperature, Mater Sci Eng A 486(1-2)
(2008) 572-579.
ot

[42] X. Zhang, G. Yuan, L. Mao, J. Niu, P. Fu, W. Ding, Effects of extrusion and heat treatment on the
mechanical properties and biocorrosion behaviors of a Mg-Nd-Zn-Zr alloy, J Mech Behav Biomed Mater
7 (2012) 77-86.
tn

[43] J.K.E. Tan, N. Birbilis, S. Choudhary, S. Thomas, P. Balan, Corrosion protection enhancement of
Mg alloy WE43 by in-situ synthesis of MgFe LDH/citric acid composite coating intercalated with 8HQ,
Corros Sci 205 (2022) 110444.
[44] C. Cai, M.M. Alves, R. Song, Y. Wang, J. Li, M.F. Montemor, Non-destructive corrosion study on
rin

a magnesium alloy with mechanical properties tailored for biodegradable cardiovascular stent
applications, J Mater Sci Technol 66 (2021) 128-138.
[45] D.B. Pokharel, L. Wu, J. Dong, A.P. Yadav, D.B. Subedi, M. Dhakal, L. Zha, X. Mu, A.J. Umoh,
W. Ke, Effect of glycine addition on the in-vitro corrosion behavior of AZ31 magnesium alloy in Hank’s
ep

solution, J Mater Sci Technol 81 (2021) 97-107.


[46] X. Zhang, G. Yuan, J. Niu, P. Fu, W. Ding, Microstructure, mechanical properties, biocorrosion
behavior, and cytotoxicity of as-extruded Mg–Nd–Zn–Zr alloy with different extrusion ratios, J. Mech.
Behav. Biomed. Mater. 9 (2012) 153-162.
Pr

[47] F. Liu, C. Chen, J. Niu, J. Pei, H. Zhang, H. Huang, G. Yuan, Mater. Sci. Eng., C 48 (2015) 400-
407.
[48] Y.S. Huang, X.T. Zeng, X.F. Hu, F.M. Liu, Corrosion resistance properties of electroless nickel

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689
composite coatings, Electrochimica Acta 49(25) (2004) 4313-4319.
[49] S.D. Wang, D.K. Xu, X.B. Chen, E.H. Han, C. Dong, Effect of heat treatment on the corrosion

ed
resistance and mechanical properties of an as-forged Mg–Zn–Y–Zr alloy, Corros Sci 92 (2015) 228-236.
[50] P.-P. Wu, G.-L. Song, Y.-X. Zhu, Z.-L. Feng, D.-J. Zheng, The corrosion of Al-supersaturated Mg
matrix and the galvanic effect of secondary phase nanoparticles, Corros Sci 184 (2021) 109410.
[51] A. Bahmani, S. Arthanari, K.S. Shin, Formulation of corrosion rate of magnesium alloys using

iew
microstructural parameters, J Magnes Alloy 8(1) (2020) 134-149.
[52] X. Zhang, G. Yuan, L. Mao, J. Niu, W. Ding, Biocorrosion properties of as-extruded Mg–Nd–Zn–
Zr alloy compared with commercial AZ31 and WE43 alloys, Mater Lett 66(1) (2012) 209-211.
[53] H. Ardelean, A. Seyeux, S. Zanna, F. Prima, I. Frateur, P. Marcus, Corrosion processes of Mg–Y–
Nd–Zr alloys in Na2SO4 electrolyte, Corros Sci 73 (2013) 196-207.

v
[54] Y.L. Song, Y.H. Liu, S.R. Yu, X.Y. Zhu, S.H. Wang, Effect of neodymium on microstructure and
corrosion resistance of AZ91 magnesium alloy, J Mater Sci 42(12) (2007) 4435-4440.

re
[55] Y. Zhang, Y. Huang, F. Feyerabend, C. Blawert, W. Gan, E. Maawad, S. You, S. Gavras, N.
Scharnagl, J. Bode, C. Vogt, D. Zander, R. Willumeit-Romer, K.U. Kainer, N. Hort, Influence of the
amount of intermetallics on the degradation of Mg-Nd alloys under physiological conditions, Acta
Biomater 121 (2021) 695-712.
[56] K.A. Unocic, H.H. Elsentriecy, M.P. Brady, H.M. Meyer, G.L. Song, M. Fayek, R.A. Meisner, B.
er
Davis, Transmission Electron Microscopy Study of Aqueous Film Formation and Evolution on
Magnesium Alloys, J Electrochem Soc 161(6) (2014) C302-C311.
[57] A. Karagoz, V. Craciun, G.B. Basim, Characterization of Nano-Scale Protective Oxide Films:
pe
Application on Metal Chemical Mechanical Planarization, ECS J Solid State Sci Technol 4(2) (2014)
P1-P8.
[58] C. Xu, W. Gao, Pilling-Bedworth ratio for oxidation of alloys, Mater Res Innovations 3(4) (2016)
231-235.
[59] S. Zhang, Z. Jiang, H. Li, B. Zhang, S. Fan, Z. Li, H. Feng, H. Zhu, Precipitation behavior and phase
ot

transformation mechanism of super austenitic stainless steel S32654 during isothermal aging, Mater
Charact 137 (2018) 244-255.
[60] Z. Yang, A. Ma, B. Xu, J. Jiang, J. Sun, Revealing the tensile anisotropy, tension–compression
tn

asymmetry, and strain-hardening behavior of a high-performance Mg–Gd–Ag alloy, J Alloys Compd


868 (2021) 159238.
[61] D.D. Yin, C.J. Boehlert, L.J. Long, G.H. Huang, H. Zhou, J. Zheng, Q.D. Wang, Tension-
compression asymmetry and the underlying slip/twinning activity in extruded Mg–Y sheets, Int J Plast
rin

136 (2021) 102878.


[62] J. Li, A. Zhang, H. Pan, Y. Ren, Z. Zeng, Q. Huang, C. Yang, L. Ma, G. Qin, Effect of extrusion
speed on microstructure and mechanical properties of the Mg-Ca binary alloy, J Magnes Alloy 9(4) (2021)
1297-1303.
ep

[63] S. Lyu, G. Li, R. Zheng, W. Xiao, Y. Huang, N. Hort, M. Chen, C. Ma, Reduced yield asymmetry
and excellent strength-ductility synergy in Mg-Y-Sm-Zn-Zr alloy via ultra-grain refinement using simple
hot extrusion, Mater Sci Eng A 856 (2022) 143783.
Pr

This preprint research paper has not been peer reviewed. Electronic copy available at: https://ssrn.com/abstract=4327689

You might also like