Microstructure and Corrosion Behavior of Advanced Alloys
Microstructure and Corrosion Behavior of Advanced Alloys
Microstructure and Corrosion Behavior of Advanced Alloys
and Corrosion
Behavior of
Advanced Alloys
Edited by
Marián Palcut
Printed Edition of the Special Issue Published in Materials
www.mdpi.com/journal/materials
Microstructure and Corrosion
Behavior of Advanced Alloys
Microstructure and Corrosion
Behavior of Advanced Alloys
Editor
Marián Palcut
MDPI • Basel • Beijing • Wuhan • Barcelona • Belgrade • Manchester • Tokyo • Cluj • Tianjin
Editor
Marián Palcut
Slovak University of Technology
in Bratislava
Slovakia
Editorial Office
MDPI
St. Alban-Anlage 66
4052 Basel, Switzerland
This is a reprint of articles from the Special Issue published online in the open access journal
Materials (ISSN 1996-1944) (available at: https://www.mdpi.com/journal/materials/special issues/
microstructure corrosion behavior advanced alloys).
For citation purposes, cite each article independently as indicated on the article page online and as
indicated below:
LastName, A.A.; LastName, B.B.; LastName, C.C. Article Title. Journal Name Year, Volume Number,
Page Range.
© 2022 by the authors. Articles in this book are Open Access and distributed under the Creative
Commons Attribution (CC BY) license, which allows users to download, copy and build upon
published articles, as long as the author and publisher are properly credited, which ensures maximum
dissemination and a wider impact of our publications.
The book as a whole is distributed by MDPI under the terms and conditions of the Creative Commons
license CC BY-NC-ND.
Contents
Nyambura Samuel Mbugua, Min Kang, Yin Zhang, Ndumia Joseph Ndiithi,
Gbenontin V. Bertrand and Liang Yao
Electrochemical Deposition of Ni, NiCo Alloy and NiCo–Ceramic Composite Coatings—A
Critical Review
Reprinted from: Materials 2020, 13, 3475, doi:10.3390/ma13163475 . . . . . . . . . . . . . . . . . . 27
Kweon-Hoon Choi, Bong-Hwan Kim, Da-Bin Lee, Seung-Yoon Yang, Nam-Seok Kim,
Seong-Ho Ha, Young-Ok Yoon, Hyun-Kyu Lim and Shae-Kwang Kim
Intergranular Corrosion and Microstructural Evolution in a Newly Designed Al-6Mg Alloy
Reprinted from: Materials 2021, 14, 3314, doi:10.3390/ma14123314 . . . . . . . . . . . . . . . . . . 87
Mien-Chung Chen, Ming-Che Wen, Yang-Chun Chiu, Tse-An Pan, Yu-Chih Tzeng
and Sheng-Long Lee
Effect of Natural Aging on the Stress Corrosion Cracking Behavior of A201-T7 Aluminum Alloy
Reprinted from: Materials 2020, 13, 5631, doi:10.3390/ma13245631 . . . . . . . . . . . . . . . . . . 101
Anna Dobkowska, Agata Sotniczuk, Piotr Bazarnik, Jarosław Mizera and Halina Garbacz
Corrosion Behavior of Cold-Formed AA5754 Alloy Sheets
Reprinted from: Materials 2021, 14, 394, doi:10.3390/ma14020394 . . . . . . . . . . . . . . . . . . . 111
Zhen Li, Zeyin Peng, Kai Qi, Hui Li, Yubing Qiu and Xingpeng Guo
Microstructure and Corrosion of Cast Magnesium Alloy ZK60 in NaCl Solution
Reprinted from: Materials 2020, 13, 3833, doi:10.3390/ma13173833 . . . . . . . . . . . . . . . . . . 125
v
Jannik Bühring, Maximilian Voshage, Johannes Henrich Schleifenbaum,
Holger Jahr and Kai-Uwe Schröder
Influence of Degradation Product Thickness on the Elastic Stiffness of Porous Absorbable
Scaffolds Made from an Bioabsorbable Zn–Mg Alloy
Reprinted from: Materials 2021, 14, 6027, doi:10.3390/ma14206027 . . . . . . . . . . . . . . . . . . 195
Sang-won Cho, Sang-Jin Ko, Jin-Seok Yoo, Yun-Ha Yoo, Yon-Kyun Song and Jung-Gu Kim
Effect of Cr on Aqueous and Atmospheric Corrosion of Automotive Carbon Steel
Reprinted from: Materials 2021, 14, 2444, doi:10.3390/ma14092444 . . . . . . . . . . . . . . . . . . 233
Peter Jurči, Aneta Bartkowska, Mária Hudáková, Mária Dománková, Mária Čaplovičová and
Dariusz Bartkowski
Effect of Sub-Zero Treatments and Tempering on Corrosion Behaviour of Vanadis 6 Tool Steel
Reprinted from: Materials 2021, 14, 3759, doi:10.3390/ma14133759 . . . . . . . . . . . . . . . . . . 269
Gaetano Palumbo, Kamila Kollbek, Roma Wirecka, Andrzej Bernasik and Marcin Górny
Effect of CO2 Partial Pressure on the Corrosion Inhibition of N80 Carbon Steel by Gum Arabic
in a CO2 -Water Saline Environment for Shale Oil and Gas Industry
Reprinted from: Materials 2020, 13, 4245, doi:10.3390/ma13194245 . . . . . . . . . . . . . . . . . . 293
Tomer Ron, Ohad Dolev, Avi Leon, Amnon Shirizly and Eli Aghion
Effect of Phase Transformation on Stress Corrosion Behavior of Additively Manufactured
Austenitic Stainless Steel Produced by Directed Energy Deposition
Reprinted from: Materials 2021, 14, 55, doi:10.3390/ma14010055 . . . . . . . . . . . . . . . . . . . 317
Patrik Šulhánek, Marián Drienovský, Ivona Černičková, LiborĎuriška, Ram ūnas Skaudžius,
Žaneta Gerhátová and Marián Palcut
Oxidation of Al-Co Alloys at High Temperatures
Reprinted from: Materials 2020, 13, 3152, doi:10.3390/ma13143152 . . . . . . . . . . . . . . . . . . 357
vi
About the Editor
Marián Palcut (b. 1981) graduated with honors in physical chemistry at the Comenius University
in Bratislava, Slovakia. He obtained his PhD in Materials Science and Technology at the Norwegian
University of Science and Technology (NTNU) in Trondheim in 2007. He completed his postdoctorate
at the Technical University of Denmark and the University of Oslo. At present, he holds a faculty
position at the Slovak University of Technology. He defended his habilitation thesis in May 2019.
Assoc. Prof. Marián Palcut has published 39 articles indexed in Web of Science and Scopus. His
papers have been cited more than 500 times (excluding self-citations). His current h index is 14.
Dr. Palcut is a regular editor of Advances in Materials Science and Engineering and a frequent
reviewer of international peer-reviewed journals in Materials Science and Technology. He has
participated in several national and international scientific projects focused on investigating the
corrosion resistance of advanced metallic materials and alloys. His current research interests include
corrosion studies of complex alloys, high entropy alloys, and lead-free solders.
vii
Preface to ”Microstructure and Corrosion Behavior of
Advanced Alloys”
In many industrial applications, metallic materials are exposed to harsh operating conditions.
Due to a combination of chemical and thermal stresses, the constructional and functional materials
are degraded and their utility properties are lost. These undesirable events are of a physicochemical
nature and are commonly known as ‘corrosion’. Several internal and external factors influence the
corrosion behavior of materials. Internal factors include the chemical composition, microstructure,
and phase constitution of the materials. External factors that can significantly contribute to an increase
in corrosion resistance include surface treatments and the use of various protective coatings. In this
Special Issue Book, 3 reviews and 18 original research papers focused on the complex relationships
between the microstructure, phase constitution, and corrosion behavior of metallic materials are
collected. Both high temperature and low temperature corrosion studies are included as they
investigate the physicochemical processes at the material interfaces. Furthermore, possibilities for
increasing the corrosion resistance of metallic materials are studied by means of surface modification
and application of protective layers. This Special Issue Book, Microstructure and Corrosion Behavior
of Advanced Alloys, displays the diversity and complexity of modern corrosion research. It is hoped
that it will become a valuable source of reference for corrosion scientists.
I am thankful to Ms. Hellen Hu from MDPI for providing valuable support during Special
Issue editing.
This work was supported by the Slovak Research and Development Agency under the Contract
no. APVV-20-0124.
Marián Palcut
Editor
ix
materials
Review
Duplex Steels Used in Building Structures and Their Resistance
to Chloride Corrosion
Mariusz Maslak 1, *, Marek Stankiewicz 1 and Benedykt Slazak 2
1 Faculty of Civil Engineering, Cracow University of Technology, Warszawska 24, 31-155 Cracow, Poland;
[email protected]
2 DAIKO SRL Welding Materials, Viale Felissent 84/D, 31100 Treviso, Italy; [email protected]
* Correspondence: [email protected]; Tel.: +48-501546577
Abstract: Welded structures made of duplex steels are used in building applications due to their
resistance to local corrosion attack initiated by chlorides. In this paper, the material and technological
factors determining the corrosion resistance are discussed in detail. Furthermore, recommendations
are formulated that allow, in the opinion of the authors, to obtain a maximum corrosion resistance
for welded joints. The practical aspects of corrosion resistance testing are also discussed, based on
the results of qualification tests. This work is of a review character. The conclusions and practical
recommendations are intended for contractors and investors of various types of structures made of
the duplex steel. The recommendations concern the selection and use of duplex steels, including the
issues of metallurgy, welding techniques, and corrosion protection.
Keywords: duplex steels; welded joints; corrosion resistance; corrosion tests; chloride environment
Figure 1. Chemical compositions and Pitting Resistance Equivalent Numbers (PREN) for various
categories of duplex steels—according to [21].
Figure 2. A properly balanced duplex steel microstructure containing approximately 50% of each
δ-ferrite (dark areas) and γ-austenite (light areas)—according to [21].
2
Materials 2021, 14, 5666
Balanced duplex stainless steels are theoretically located at a 50% ferrite content line
in Schaeffler–DeLong diagrams (Figure 4). However, in practice in properly balanced steel
−10%
ferrite
microstructure, the ferrite—austenite proportions range are wider, i.e., austenite ≈ 50%
50%+10%
.
The appropriate use of duplex steel demands the welding engineers perform thought-
ful actions based on well-established engineering knowledge and experience. A commonly
applied routine-based approach to welding admittedly leads to obtaining the expected
mechanical properties of welding joints. Nevertheless, it does not guarantee the concurrent
obtaining of required corrosion resistance for the joints. It is known that the corrosion
resistance of joints in the chloride environment impact zone reaches only 50–80% of the
parent material corrosion resistance [22].
3
Materials 2021, 14, 5666
4
Materials 2021, 14, 5666
Figure 5. Metallurgical transformations in duplex steels: (a) phase equilibrium diagram and austenitic transformation δ
→ δ + γ (according to [27]), (b) distribution coefficient of alloy additions Kδ/γ as a function of temperature (according
to [28]), (c) typical values of the partition coefficient of alloying elements Kδ/γ for supersaturated and water-cooled steels
(according to [28]), (d) influence of alloying elements on the size of electrochemical austenite potential in stainless steels
18/8 (according to [29]).
5
Materials 2021, 14, 5666
Table 1. Maximum solubility of alloy additives in ferrite and austenite (according to [26]).
Solubility (%wt.)
Alloy Additive Crystal Lattice
Ferrite Austenite
W 35 4.7 A2
Mo 31 1.7 A2
Mn 3.5 100 A1
Cr 100 12.5 A2
Cu 2.1 12 A1
Ni 6 100 A1
Si 11 1.7 A4
C 0.03 2.1 -
N 0.1 2.8 -
Figure 6. Influence of cooling rate on the content of δ-ferrite in duplex steels (according to [30]).
Left-hand frame: fast cooling, high δ-ferrite content; right-hand frame: slow cooling, high content of
γ-austenite precipitates.
6
Materials 2021, 14, 5666
tendency to secondary phase precipitations so that the cooling line does not cross the upper
Time-Temperature-Transformation (TTT) curve (Figure 7).
Table 2. Influence of various alloy additions and steel microstructure on duplex steel pitting and slotted corrosion resistance
(according to [32,33]).
7
Materials 2021, 14, 5666
5. The TTT Diagram, CPT Temperature, and the Ferrite Number Determined for
Duplex Steels
In high-alloy duplex steels of older generations (SDSS and HDSS), the precipitation
start time is short [34]. Exceeding the activation energy of the secondary phase precipita-
tions, expressed by the intersection of the cooling line with the TTT curve increases the
sensitivity to accumulation of heat exposure effects resulting from welding the steel. It is
because the energy of the precipitation reaction is always lower than the activation energy
of this process [35].
The formation of secondary phases in duplex steels, which are hard and brittle and
thus harmful, takes place in two temperature ranges (Figure 7). The upper curve, corre-
sponding to the range of 600–1050 ◦ C, shows the precipitation of nitrides, carbides, and
intermetallic phases as a result of prolonged thermal exposure of steel due to insufficiently
rapid cooling. Thus, the welding with high linear energies facilitates the transformation
of δ to δ + γ, which is favorable but also increases the probability of the secondary phase
formation within the ferrite at insufficiently fast cooling. The consequence of this is the
need to strictly adhere to the recommended linear energies of welding and to control
the cooling process of the joint during welding. The lower TTT curve, corresponding to
the temperature of 300–550 ◦ C, shows the remaining secondary changes in the steel mi-
crostructure. The most important is the change of δ-ferrite to acicular secondary α -ferrite,
significantly reducing the ductility and toughness of steel. The lower limit of the occurrence
of the unfavorable α -ferrite determines the highest temperature of the long-term thermal
exposure, which is about 300 ◦ C.
The advancement of harmful changes in the microstructure, lower ductility, and
corrosion resistance depends on the total heat exposure time to both critical ranges on
the TTT curves (Figure 8a,b). The exposure of the lowest alloyed LDSS to temperatures
between 800–1000 ◦ C may exceed 10 h without the release of harmful phases. However,
this time for the standard DSS 22% Cr steel is reduced to about 30–60 min, and for the
HDSS steel is even 5–10 min [36]. The presence of 0.5% secondary phases at the boundaries
of δ-ferrite causes a dramatic decrease in the breaking work (Figure 8a). The same volume
of the harmful and easily released Cr2 N phase lowers the critical pitting temperature (CPT)
by up to 20 ◦ C (Figure 8b).
Figure 8. Changes in the microstructure lowering ductility and corrosion resistance of duplex
steels, including (a) breaking energy KV as a function of intermetallic phase content in duplex steels
(according to [37]), (b) CPT (Critical Pitting Temperature) index as a function of the content of
intercrystalline Cr2 N precipitates (according to [38]).
8
Materials 2021, 14, 5666
Nitrogen in duplex steel increases the kinetics of austenite formation (Figure 9). With
increasing nitrogen content, the high-temperature equilibrium δ + γ area expands towards
lower Ni concentrations, and the temperature of austenite formation increases. It may
increase even to the temperature of liquids, which means crystallization of a small part
of austenite directly from the liquid metal. A significant increase in the transformation
rate δ → δ + γ occurs, which, with sufficiently rapid cooling, makes it possible to avoid
the effects of shifting the TTT curves on the time axis towards the origin of the coordinate
system. For this reason, nitrogen-rich steels can be cooled in the temperature range
of 800–1200 ◦ C faster, without the fear of exceeding the final ferrite content above the
permissible level (70%). The addition of N2 to the forming and shielding gases in Tungsten
Inert Gas Arc Welding (TIG, 141) is of fundamental importance for obtaining a properly
balanced weld microstructure.
Figure 9. TTT diagram for DSS 25% Cr duplex steels with different alloyed nitrogen content (accord-
ing to [28]).
9
Materials 2021, 14, 5666
Figure 10. Heat treatment of duplex steel, including (a) solubility of Nitrogen in ferrite and austenite
(according to [38]), (b) heat treatment of the thin-walled and thick-walled duplex steel elements
(according to [23]).
sorption of strongly austenitic-forming N from the shielding gases intensify the kinetics
of the δ → δ + γ transformation, thereby preventing the reduction in austenite content in
the weld microstructure. Therefore, the selection of the parent material with the highest
possible austenite content, within the limits of the correct balance of the initial duplex steel
microstructure, is important for the subsequent resistance of the HAZ to pitting corrosion
and its KV impact strength.
Figure 11. Pitting corrosion mechanism of duplex steel when exposed to chloride attack (according
to [21]).
The pitting corrosion is autocatalytic, self-accelerating the growth of pits due to low-
ering the pH of the corrosive solution inside the growing pit [40]. In the initial stage, the
pitting corrosion is of inter-crystalline character [42]. A balanced duplex steel microstruc-
−10%
ferrite
ture, within the limits of austenite ≈ 50%
50%+10%
reduces the tendency to harmful precipitations
of carbides and nitrides at the intergranular boundaries. This simultaneously reduces the
risk of developing inter-crystalline corrosion.
11
Materials 2021, 14, 5666
12
Materials 2021, 14, 5666
Figure 12. Chloride pits identified in the SDSS weld metal (1.4501, F55). Joint filling obtained
after automatic TIG welding in Ar shielding gas without N2 . In particular: (a) magnified optical
photography, (b) a photo taken with a scanning microscope, (c) Cl and Cr contents measured along
the white arrow marked in Figure 12b. The following locations are marked by vertical lines in
Figure 12c: solid lines—local maxima of the Cr content in the surface layer and the corresponding
local minima of the absorbed Cl content, dashed lines—local minima of the Cr content in the surface
layer, and the corresponding local maxima of the absorbed Cl content.
13
Materials 2021, 14, 5666
Figure 13. Influence of nitrogen content in a weld made with the TIG method on the volume fraction
of δ-ferrite in the steel microstructure (according to [52]).
The high vapor pressure of nitrogen at the welding temperature causes its migration
from the weld pool to the surrounding environment and, consequently, reduces the share
14
Materials 2021, 14, 5666
of austenite in the weld structure, thereby increasing the risk of losing corrosion resistance
and lowering the impact strength.
The nitrogen-poor weld contains up to 80% δ-ferrite. Since the electric arc does not
transfer the electrically neutral nitrogen atoms, N is not present in arc welding consumables.
Therefore, the only option to increase the N content in the weld metal during the welding is
the addition of N2 to the shielding and forming gases. The necessary amount of N2 depends
on its solubility limit in duplex steel and increases with increasing Ni concentration in the
steel. Nitrogen content in shielding gases for GTAW welding (TIG, 141) should be in the
range of 1–1.2% for DSS 22% steel and 2–2.5% for DSS 25%, both SDSS and HDSS steels.
The weld root is usually the area with the greatest risk of corrosion. Therefore, it is
important to use N2 -rich mixture as forming gas. However, the use of shielding gases
with an excessively high N2 concentration may result in exceeding the N solubility limit
in solid solution and the appearance of weld porosity, especially in thick-walled joints.
Since the corrosion resistance of stainless steel is primarily determined by the properties
of the surface layer, the beads of the multi-run welds can be welded in pure argon to
avoid porosity. In such cases, the pitting corrosion resistance tests should not include the
weld filling.
The addition of 20–40% of helium to shielding gas increases the thermal energy
supplied to the weld, and this allows increasing the welding efficiency with the GTAW
method (TIG, 141). Furthermore, the full control of the O2 content in welding gases prevents
its absorption in the weld pool, as well as a harmful O increase in the solid solution. In
addition, it allows a reduction in the thickness of the oxide layer above the welded joint
and thus the depth of the depletion of the steel surface layer in Cr and Mo. Therefore, it
is recommended to use welding gases with O2 content below 200 ppm for duplex steels
and to flush the pipes from the weld root side with forming gas in order to reduce the O2
content as much as possible. It is suggested to limit the O2 concentration to 25 ppm.
15
Materials 2021, 14, 5666
the weld metal is achieved by using the GTAW method (TIG, 141) and PAW (Plasma Arc
Welding) method (151), which is a GTAW method extension.
Figure 14. Breaking energy of the weld metal: (a) breaking energy KV determined at 20 ◦ C as
a function of oxygen content in the duplex weld metal, (b) breaking energy KV determined at
0–(−60) ◦ C, specified for various welding methods (according to [28]). The meaning of individual
abbreviations is explained in the Abbreviations of this paper.
The ease of GTAW (TIG, 141) made this method the primary choice. For the pipe
connections, a modified GMAW-STT method (Surface Tension Transfer, MIG-STT, 131-STT)
may be used. It provides a three to four times higher welding efficiency in rela-tion to the
conventional GTAW and gives a comparable resistance to pitting corrosion. Additionally,
it guarantees a satisfactory ductility of the material down to −40 ◦ C [53]. It should be
noted that the use of high-oxygen, slag arc welding methods for this type of welds, such as
SMAW (111), SAW (121), or FCAW (114), reduces the corrosion resistance of the joints and
also reduces the breaking energy of the weld metal. At the same time, it increases the lower
operating temperature threshold for welded joints use (Figure 14b). This is a consequence
16
Materials 2021, 14, 5666
of the high content of atomic oxygen dissolved in the solid solution and the presence of
oxide inclusions at the grain boundaries.
As most standards do not require pitting tests of the whole weld cross-section, any
flux-type welding (high-oxygen) methods can be applied for inner layers of thicker multi-
run joints. However, particular caution should be taken while welding SDSS and HDSS
steels due to their high yield strength and a stronger tendency to brittle fracture. In such a
situation, an application of the low-oxygen welding methods (both GTAW and PAW) at the
whole cross-section of the weld and precise balancing both of the weld microstructure and
the whole heat-affected zone (HAZ) is required for obtaining reasonable impact resistance.
The regulation of γ-austenite formation kinetics in duplex steel welds is dependent
on the proper shaping of the weld groove. The shapes of the grooves should be in general
analogous to those formed in acid-resistant austenitic steels. Nevertheless, minor discrep-
ancies are possible in the current situation. Examples of typical grooves recommended for
welding duplex steels are presented in [38]. For single-sided welding, the grooves should
be shaped to obtain a wider root gap, lower root face, and wider groove angle (bevel) [46].
The wider root gap and lower root face limit the weld metal and the parent material mixing
rate, which reduces the Ni content in the weld metal of the weld root. The root bead
should be massive enough to counteract the nitrogen deficiency at this welding step by
extending the cooling time in the austenite formation temperature range. The weld root
should be welded using high linear energy, within limits recommended by the weld metal
manufacturer. Under-heating of the weld root bead accelerates the cooling in the austenite
formation temperature range. Nevertheless, excessive overheating lengthens the cooling
time and stimulates the precipitation of harmful secondary phases. The consequence is a
reduction in pitting corrosion resistance and impact strength. The following filling beads
are often called “cold” runs. They should be welded with the linear energy reduced by up
to 75% and should not be massive so as not to cause changes in the microstructure of the
root run and in the HAZ, reaching directly under the passive layer. The following filling
runs are to be welded with recommended increased heat input energy, which in the face
layer reaches up to 150% of the heat input used for the weld root [54]. The thermal effects
of the successive layers of the weld, lying above the “cold” run, must in no way affect the
microstructure and properties of the weld root run.
The most problematic for maintaining the proper microstructure and sufficient pitting
resistance seems to be the single- and two-runs of the thin-walled welds. Such problems
appear in the seal welds of shell-and-tube heat exchangers [55]. Delicate girth welds
in-between massive perforated bottom and thin-walled pipe are often welded with an
intense mixing rate of weld and parent material of the pipe, and additionally, the cooling
rate is higher due to massive perforated bottom. The content of δ-ferrite usually exceeds
the limiting value of 70% even when using recommended welding material. In such cases,
the use of austenitic weld metal with a high Mo content allows obtaining ferrite amount
slightly smaller than 70%, which means a limited resistance to pitting corrosion. Further-
more, due to the cumulative effects of heat exposure of duplex steel and precipitation
of secondary phases, it is not advisable to cut the materials thermally. In order to avoid
a heat accumulation during welding, it is not recommended to preheat the steel, apart
from drying the surface at a temperature not exceeding 100 ◦ C. For the same reason, the
inter-pass weld temperature should be strongly limited.
The recommendations given above allow obtaining welded joints resistant to pitting
corrosion for standard duplex DSS 22% Cr steels. However, with increasing Cr content,
the necessary cooling rate must be controlled. An insufficient austenite content in the
weld can occur if the cooling rate is too high. A release of harmful secondary phases can
be observed if the cooling rate is too low. The welding of high Cr duplex steels can be
facilitated by using a combined welding method, for example, GTAW with assisted cooling
of the weld by a micro-jet injector and argon as a refrigerant [56,57]. The rapidly expanding
gas quickly removes heat from the weld, allowing for 2–3 times higher intensification of the
cooling process. The introduction of micro-jet cooling allows increasing the welding heat.
17
Materials 2021, 14, 5666
It ensures an increase in the share of austenite in the microstructure and thus increases the
pitting corrosion resistance of the joint. Moreover, the controlled and appropriately rapid
cooling below 1000 ◦ C avoids the formation of harmful secondary phases.
The implementation of the high-temperature heat treatment after the welding, carried
out to properly balance the microstructure of duplex steel joints, is possible only for small
objects that can be fully homogenized annealed at 1050–1150 ◦ C and then supersaturated in
water. Local heat treatment with the use of heating mats cannot be used due to degradation
of the steel microstructure at the edges and due to the impossibility of rapid cooling. The
large objects can be stress-relieved by tempering at temperatures below 300 ◦ C for about
10 h so as not to initiate the microstructural changes within the lower TTT curve, as shown
previously in Figure 7.
The corrosion resistance of welded joints made of duplex steel can be improved by
chemical etching. The etching removes the oxide layer formed above the weld and heat-
affected zone as a result of welding and re-establishes a more compact passive layer. The
original oxide layer on the joint and in HAZ can be thick up to 100 nm [54]. The layer is
enriched with Fe2 O3 and thus has a low resistance to pitting. The large depletion layer of
Cr and Mo further facilitates the development of pitting corrosion. The etched surfaces
are usually treated with highly oxidizing nitric acid, hydrofluoric acid, or peroxide [58].
The chemically formed passive layer is tighter, and the concentration of Cr2 O3 , MoO2 , and
MoO(OH) in the layer is higher.
11. Corrosion Resistance Tests of Welded Joints Made of Duplex Steel, Carried out for
the Chloride Environment
The American standard ASTM G48 [18] is the leading standard for corrosion resistance
testing of duplex alloys and their welded joints. It contains several fundamental test
procedures for assessing the resistance of stainless steels and related alloys in a ferric
chloride solution. However, in the ASTM G48 standard, the criteria for evaluating the
test results obtained after the experiments are not explicitly defined. Therefore, the results
should be interpreted in conjunction with other guidelines taken, for example, from the
Norwegian standard NORSOK M601 [19] or American standard ASTM A923 [20]. The
corrosive environment in these tests is an oxygenated aqueous 6% FeCl3 solution. This
salt partially hydrolyzes in water. The temperature of the solution increases the degree
of hydrolysis, which results in a more acidic solution. The FeCl3 salt does not introduce
18
Materials 2021, 14, 5666
foreign metal cations into the corrosive environment. The solution is not oxidizing. As
such, it does not passivate the metal surfaces and has a high penetration capacity for the
surface micro-damages [61].
The results of FeCl3 tests are affected by temperature and autocatalytic course of
pitting corrosion. According to the ASTM G48-method A standard procedures, the recom-
mended test temperature for duplex DSS 22% Cr steels is 22 ± 2 ◦ C and for SDSS steels is
35 ± 2 ◦ C. The use of thermostatic water baths with temperature stabilization at ±0.2 ◦ C is
recommended. If there is no consensus on the temperature conditions of tests, a deviation
from recommendations of the ASTM G48-method A standard is permissible, provided that
all the other elements of the standardized test procedure are followed. This deviation is
allowed since in less corrosive environments, such as, for example, NaCl solution, the test
temperature may be correspondingly higher. Due to the autocatalytic nature of the pitting
corrosion, the extension of the test duration is accompanied by an increase in the average
daily mass loss. Then, there is a gradual blurring of differences in corrosion losses between
materials of different resistance, and the probability of obtaining an unreliable test result in-
creases. For these reasons, the test time originally proposed in ASTM G48-method A (72 h)
has been reduced to 24 h in the NORSOK M601 and ASTM A923 standards. The standards
recommend using flat samples with dimensions of 25 mm × 50 mm or sections of tubular
surfaces which are equivalent to these sizes. Any unevenness caused by machining should
be smoothed and sharp edges rounded. Moreover, efforts should be made to minimize
the side surfaces of the samples. In the case of thick samples, taken, for example, from
multi-pass joints, cutting a thin sample from the weld root or weld face layer can be a good
option as it is responsible for the corrosion resistance of the entire joint. Ideally, the exposed
surface should be representative of the corrosion risks within the joint. It must therefore
encompass the joint itself, HAZ, and base material. It is also advisable to mirror the surface
roughness of the welded joint. The root and face weld surfaces should not be mechanically
polished [62]. In the comparative tests of pitting corrosion resistance of the basic materials,
a maximum standardization of the shape, dimensions, and surface conditions of samples
must be guaranteed. It is also necessary to round the sample edges.
The NORSOK M601 standard supplements the requirements with preliminary etching
in HNO3 and HF solutions. By etching, a thick and leaky passive layer above the weld and
heat-affected zone is removed. A re-passivation under free oxidation in the air requires
at least 24 h to obtain a sufficiently thick and tight passive layer. The NORSOK M601
standard suggests maintaining the time interval between the sample preparation and
the test itself. Direct corrosion testing immediately after etching may result in uniform
corrosion without visible pitting, exceeding the limit of the allowable weight loss. In such a
case, it is necessary to repeat the corrosion test by doing the preparations again and keeping
a 24 h interval between the HNO3 + HF pre-etch operation and the main FeCl3 test.
It is permissible and beneficial to replace the manual sample washing with an ultra-
sonic bath. This ensures more effective removal of corrosion products from the sample
surface and positively affects the reliability of mass measurements. The test should be
performed in a stationary medium with free air access to the FeCl3 solution. Cutting off
the access of oxygen and solution stirring causes the polarization of corrosion cells and the
electrochemical processes between the sample and the solution may cease.
The results of the pitting resistance test according to ASTM G48-method A are assessed
based on the weight loss measurements and visual inspection. Due to the small weight
loss of the samples during the test, analytical balances with a measurement accuracy of
1 mg should be used. The permissible maximum weight loss, according to NORSOK
M601, is 4 g/(m2 per day). It corresponds to a corrosion rate of ~0.2 mm/year. In the tests
associated with ASTM A901; however, a weight loss of 1 g/(m2 per day) is allowed, which
corresponds to a uniform corrosion rate of ~0.05 mm/year. In practice, these thresholds
are achieved at the time of few tiny pit appearances on the sample surface, noticeable at
20× magnification. The occurrence of a single pitting on the surface of a sample, noticeable
20× magnification, qualifies the test result as negative.
19
Materials 2021, 14, 5666
In Figures 15 and 16, typical visual assessment results are shown [21]. The tests were
carried out according to procedures recommended for use in the ASTM G48-method A
standard. The photos presented in Figure 15 show the classic DSS 22% Cr steel, grade
2205 (1.4462, F51), hand-welded with the GTAW method (TIG, 141), shielded with Ar + N2
mixture, with 2209 wire. A well-balanced parent material with δ-ferrite content of ~51%
was used for welding. As there is a relatively long initiation time of secondary phase
precipitation, it is possible to extend the thermal exposure time in the range of austenite
forming temperature by introducing more heat to the weld. As a result, ~42% δ-ferrite
content has been obtained in the root of the weld, and ~55% in its face. The thermal
welding cycle increased the ferrite content in the HAZ to ~62%. Due to the relatively
low temperature, the amount was still within safe limits. The cooling rate below 1050 ◦ C
was sufficiently high, and no harmful secondary phases were released at the ferrite grain
boundaries. The corrosion resistance test performed according to the ASTM G48-method
A procedure, in conditions typical for the DSS duplex steel, showed a weight loss of less
than 1 g/(m2 per day).
Figure 15. Authors’ analysis of welded connection made of duplex DSS steel with high resistance to
pitting corrosion identified for the environment of chlorides.
20
Materials 2021, 14, 5666
Figure 16. Authors’ analysis of welded connection made of super duplex SDSS steel with a low
resistance to pitting corrosion identified for the environment of chlorides.
21
Materials 2021, 14, 5666
showed a weight loss of over 10 g/(m2 per day). This example shows the sensitivity of
SDSS steels to welding technology errors, resulting in a lack of resistance to pitting.
Author Contributions: Conceptualization, M.M., M.S., and B.S.; methodology, M.S. and B.S.; vali-
dation, M.S.; formal analysis, M.S. and B.S.; resources, M.S. and B.S.; data curation, M.S. and B.S.;
writing—original draft preparation, M.M. and M.S.; writing—review and editing, M.M. and M.S.;
visualization, M.M., M.S., and B.S.; supervision, M.M. All authors have read and agreed to the
published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
22
Materials 2021, 14, 5666
Abbreviations
References
1. Kahar, S.D. Duplex stainless steels—An overview. Int. J. Eng. Res. Appl. 2017, 7, 27–36. [CrossRef]
2. Marrow, T.J.; Sunoog, K.; Oh, J.T. The effect of microstructure on brittle fracture and stress corrosion cracking in duplex stainless
steel. In Proceedings of the Stainless Steel World 99 Conference on Corrosion-Resistant Alloys, The Hague, The Netherlands,
16–18 November 1999.
3. Levkov, L.; Shurygin, D.; Dub, V.; Kosyrev, K.; Balikoev, A. New generation of super duplex steels for equipment gas and oil
production. In Proceedings of the International Conference on Corrosion in the Oil and Gas Industry, St. Petersburg, Russia,
22–24 May 2019.
4. Perren, R.A.; Sutter, T.A.; Uggowitzer, P.J.; Weber, L.; Magdowski, R.; Böhni, H.; Speidel, M.O. Corrosion resistance of super
duplex stainless steel in chloride ion containing environments: Investigations by means of a new microeletrochemical method. I.
Precipitation-free states. Corros. Sci. 2001, 43, 707–726. [CrossRef]
5. Perren, R.A.; Sutter, T.A.; Solenthaler, C.; Gullo, G.; Uggowitzer, P.J.; Böhni, H.; Speidel, M.O. Corrosion resistance of super
duplex stainless steel in chloride ion containing environments: Investigations by means of a new microeletrochemical method. II.
Influence of precipitates. Corros. Sci. 2001, 43, 727–745. [CrossRef]
6. Singh Raman, R.K.; Siew, W.H. Role of nitrite addition in chloride stress corrosion cracking of a super duplex stainless steel.
Corros. Sci. 2010, 52, 113–117. [CrossRef]
7. Vignal, V.; Zhang, H.; Delrue, O.; Heintz, O.; Popa, I.; Peultier, J. Influence of long-term ageing in solution containing chloride
ions on the passivity and the corrosion resistance of duplex stainless steels. Corros. Sci. 2011, 53, 894–903. [CrossRef]
8. Kim, S.-T.; Lee, I.-S.; Kim, J.-S.; Jang, S.-H.; Park, Y.-S.; Kim, K.-T.; Kim, Y.-S. Investigation of the localized corrosion associated
with phase transformation of tube-to-tube sheet welds of hyper duplex stainless steel in acidified chloride environments. Corros.
Sci. 2012, 64, 164–173. [CrossRef]
9. Yang, Y.Z.; Jiang, Y.M.; Li, J. In situ investigation of crevice corrosion on UNS S32101 duplex stainless steel in sodium chloride
solution. Corros. Sci. 2013, 76, 163–169. [CrossRef]
10. Zanotto, F.; Grassi, V.; Balbo, A.; Monticelli, C.; Zucchi, F. Stress corrosion cracking of LDX 2101 duplex stainless steel in chloride
solutions in the presence of thiosulphate. Corros. Sci. 2014, 80, 205–212. [CrossRef]
11. Schmid, A.; Moria, G.; Hönig, S.; Weil, M.; Strobl, S.; Haubner, R. Comparison of the high-temperature chloride-induced corrosion
between duplex steel and Ni based alloy in presence of H2S. Corros. Sci. 2018, 139, 76–82. [CrossRef]
12. Silva, R.; Kugelmeier, C.L.; Vacchi, G.S.; Martins Junior, C.B.; Dainezi, I.; Afonso, C.R.M.; Mendes Filho, A.A.; Rovere, C.A.D. A
comprehensive study of the pitting corrosion mechanism of lean duplex stainless steel grade 2404 aged at 475 ◦ C. Corros. Sci.
2021, 191, 109738. [CrossRef]
13. Sasikumar, C.; Sundaresan, R.; Merlin Medona, C.; Ramakrishnan, A. Corrosion study on AA-TIG welding of duplex stainless
steel. Mater. Today: Proc. 2021, 45, 3383–3385. [CrossRef]
14. Shi, J.; Ming, J.; Sun, W. Passivation and chloride-induced corrosion of a duplex alloy steel in alkali-activated slag extract solutions.
Constr. Build. Mater. 2017, 155, 992–1002. [CrossRef]
23
Materials 2021, 14, 5666
15. Briz, E.; Biezma, M.V.; Bastidas, D.M. Stress corrosion cracking of new 2001 lean–duplex stainless steel reinforcements in chloride
contained concrete pore solution: An electrochemical study. Constr. Build. Mater. 2018, 192, 1–8. [CrossRef]
16. Alonso, M.C.; Luna, F.J.; Criado, M. Corrosion behavior of duplex stainless steel reinforcement in ternary binder concrete exposed
to natural chloride penetration. Constr. Build. Mater. 2019, 199, 385–395. [CrossRef]
17. Welding procedures and welders. In Guidelines of Approval Testing, WG 01, 3rd ed.; Safety Assessment Federation (SAFed):
London, UK, 2020.
18. ASTM G48-11; Standard Test Method for Pitting and Crevice Corrosion Resistance of Stainless Steels and Related Alloys by Use of Ferric
Chloride Solutions; ASTM: West Conshohocken, PA, USA, 2011.
19. NORSOK M601; Welding and Inspection of Piping; NORSOK: Lysaker, Norway, 2008.
20. ASTM A923-03; Standard Test Methods for Detecting Detrimental Intermetallic Phase in Duplex Austenitic/Ferritic Stainless Steels;
ASTM: West Conshohocken, PA, USA, 2003. [CrossRef]
21. Stankiewicz, M.; Slazak, B. Resistance to pitting corrosion of duplex steel and its welded joints. In The Importance of Steel Selection,
Modeling Methods and Experimental Tests in the Design of Building Structures; Maslak, M., Ed.; Selected Issues; Publishing House of
the Cracow University of Technology: Cracow, Poland, 2020; pp. 47–83. (In Polish)
22. Calderon-Uriszar-Aldaca, I.; Briz, E.; Garcia, H.; Matanza, A. The weldability of duplex stainless steel in structural components to
withstand corrosive marine environments. Metals 2020, 10, 1475. [CrossRef]
23. Knyazeva, M.; Pohl, M. Duplex Steels. Part I: Genesis, Formation, Structure and Part II: Carbides and Nitrides. Metallogr.
Microstruct. Anal. 2013, 2, 113–121, 343–351. [CrossRef]
24. Craig, B. Clarifying the applicability of PREN equations: A short focused review. Corrosion 2021, 77, 382–385. [CrossRef]
25. Peguet, L.; Gaugain, A. Localized corrosion resistance of duplex stainless steels: Methodology and properties: A review paper.
Revue de Métallurgie 2011, 108, 231–243. [CrossRef]
26. Berns, H.; Theisen, W. Ferrous Materials–Steel and Cast Iron; Springer: Berlin, Germany, 2008; pp. 47, 322. [CrossRef]
27. Nilsson, J. Overview–super duplex stainless steels. J. Mater. Sci. Technol. 1992, 8, 685–700. [CrossRef]
28. Gunn, R.N. (Ed.) Duplex Stainless Steels. Microstructure, Properties and Applications; Abington Publishing: Nashville, TN, USA,
1997. [CrossRef]
29. Speidel, M.O. Corrosion science of stainless steels. In Proceedings of the International Conference on Stainless Steels, Chiba
Japan, 10–13 June 1991; Volume 1, pp. 25–34.
30. Holloway, G. Effective welding of duplex & super duplex stainless steels. In Proceedings of the Welding Society Meeting,
Singapore, 29 October 2003.
31. Charles, J. Super duplex stainless steels: Structure and properties. In Proceedings of the Conference on Duplex Stainless Steels,
Beaune, Bourgogne, France, 28–30 October 1991; Volume 1, pp. 3–48.
32. Bernhardsson, S. The corrosion resistance of duplex stainless steels. In Proceedings of the Conference Duplex Stainless Steels,
Beaune, Bourgogne, France, 28–30 October 1991; Volume 1, p. 185.
33. Elhoud, A.M.; Renton, N.C.; Deans, W.F. The effect of manufacturing variables on the corrosion resistance of a super duplex
stainless steel. Int. J. Adv. Manuf. Technol. 2011, 52, 451–461. [CrossRef]
34. Chail, G.; Kangas, P. Super and hyper duplex stainless steels: Structures, properties and applications. In Proceedings of the 21st
European Conference on Fracture (ECF21), Catania, Italy, 20–24 June 2016.
35. Hosseini, V.; Karlsson, L.; Engelberg, D.; Wessman, S. Time-temperature-precipitation and property diagrams for super duplex
stainless weld metals. Weld. World 2018, 62, 517–533. [CrossRef]
36. Pohl, M.; Storz, O.; Glogowski, T. Effect of intermetallic precipitations on the properties of duplex stainless steel. Mater. Charact.
2007, 58, 67–71. [CrossRef]
37. Barbosa, C.A.; Sokolowski, A. Development of UNS S 32760 super-duplex stainless steel produced in large diameter rolled bars.
Revista Escola de Minas 2013, 66, 201–208. [CrossRef]
38. Practical Guidelines for the Fabrication of Duplex Stainless Steel. Available online: http://www.edelstahl.de/fileadmin/user_
upload/ess_duplex_steel.pdf (accessed on 2 July 2021).
39. Lippold, J.C.; Kotecki, D.J. Welding Metallurgy and Weldability of Stainless Steels; John Willey & Sons Inc.: New York, NY, USA, 2015.
40. Akpanyung, K.V.; Loto, R.T. Pitting corrosion evaluation: A review, International Conference on Engineering for Sustainable
World. J. Phys. Conf. Ser. 2019, 1378, 022088. [CrossRef]
41. Örnek, C.; Davut, K.; Kocabaş, M.; Bayatli, A.; Ürgen, M. Understanding corrosion morphology of duplex stainless steel wire in
chloride electrolyte. Corros. Mater. Degrad. 2021, 2, 397–411. [CrossRef]
42. Kotecki, D.J. Ferrite determination in stainless steel welds: Advances since 1974. Weld. J. 1997, 76, 24–37.
43. Olsson, C.-O.A.; Landolt, D. Passive films on stainless steels–chemistry, structure and growth. Electrochim. Acta 2003, 48,
1093–1104. [CrossRef]
44. Zhang, B.; Wei, X.X.; Wu, B.; Wang, J.; Shao, X.H.; Yang, L.X.; Zheng, S.J.; Zhou, Y.T.; Jin, Q.Q.; Oguzie, E.E.; et al. Chloride attack
on the passive film of duplex alloy. Corros. Sci. 2019, 154, 123–128. [CrossRef]
45. Souier, T.; Martin, F.; Bataillon, C.; Cousty, J. Study of the passive film on duplex stainless steels and its breakdown by AFM
and CS-AFM. In Proceedings of the International Conference EUROCORR, Nice, France, 6–10 September 2009; Volume 5,
pp. 2869–2880.
46. Alvarez-Armas, I.; Degallaix-Moreuil, S. (Eds.) Duplex Stainless Steels; ISTE Ltd.: London, UK, 2009. [CrossRef]
24
Materials 2021, 14, 5666
47. Wiktorowicz, R.; Crouch, J. Shielding gas developments for TIG welding of duplex and super duplex stainless steels. Weld. Metal.
Fabr. 1994, 62, 379–383.
48. Messer, B.; Oprea, V.; Wright, A. Duplex stainless steel welding: Best practices. Stainless Steel World 2007, 53, 53–63.
49. Babish, F. Repair Welding of Stainless Steels; Sandvik Materials Technology: Clarks Summit, PA, USA, 2014.
50. Stankiewicz, M.; Slazak, B. Arc welding of duplex steels in terms of maximizing the resistance of welded joints to pitting corrosion
in the chloride environment. Przeglad ˛ Spawalnictwa 2018, 90, 142–150. (In Polish) [CrossRef]
51. Patrick, C.W.; Cox, M.A. Challenges welding duplex and super duplex stainless steel, American Fuels & Petrochemical Manufac-
turers. In Proceedings of the Reliability & Maintenance Conference, San Antonio, TX, USA, 20–23 May 2014. Available online:
http://www.gowelding.com/met/duplex.html (accessed on 10 July 2021).
52. Sato, Y.S.; Kokawa, H.; Kuwana, T. Effect of nitrogen on sigma transformation in duplex stainless steel weld metal. Sci. Technol.
Weld. Join. 1999, 41, 41–49. [CrossRef]
53. Van Nassau, L.; Meelker, H.; Neessen, F.; Hilkes, J. Welding Duplex and Superduplex Stainless Steel. An Update of the Guide for
Industry. Available online: http://www.lasgroepzuid.com/documenten/Update_guide_Welding_duplex_and_superduplex_
stainless_steel.pdf (accessed on 12 July 2021).
54. Baxter, C.; Young, M. Practical aspects for production welding and control of duplex stainless steel pressure and process
plants. In Proceedings of the Conference Duplex America 2000; DA2 032; Stainless Steel World, KCI Publishing BV: Zuthphen,
The Netherlands, 2000; pp. 129–139.
55. Faes, W.; Lecompte, S.; Zaaquib, Y.A.; Van Bael, J.; Salenbien, R. Corrosion and corrosion prevention in heat exchangers. Corros.
Rev. 2019, 2, 131–155. [CrossRef]
56. Wegrzyn, T.; Piwnik, J.; Lazarz, B.; Wieszala, R.; Hadrys, D. Parameters of welding with micro-jet cooling. Arch. Mater. Sci. Eng.
2012, 54, 86–92.
57. Wegrzyn, T.; Piwnik, J.; Hadrys, D.; Wszolek, L. Low alloy steel structures after welding with micro-jet cooling. Arch. Metall.
Mater. 2017, 62, 115–118. [CrossRef]
58. Ward, I. Report on Weld Cleaning Methods, Document IW 150807; Sandvik Australia Pty Ltd.: Mulgrave, VIC, Australia, 2007.
59. Alvarez, T.; Guttemberg, C.S.; Pardal, J.M.; Tavares, S. Influence of interpass temperature on the properties of duplex stainless
steel during welding by submerged arc welding process. Weld. Int. 2016, 30, 348–358. [CrossRef]
60. Pan, J. Studying the passivity and breakdown of duplex stainless steel at micrometer and nanometer scales–the influence of
microstructure. Front. Mater. 2020, 7, 133. [CrossRef]
61. Woollin, P. Ferric chloride testing for weld procedure qualification of duplex stainless steel weldments. In Proceedings of the
International Conference UK Corrosion and EUROCORR, Bournemouth, UK, 31 October–3 November 1994.
62. Rajaguru, J.; Arunachalam, N. Effect of machined surface integrity on the stress corrosion cracking behaviour of super duplex
stainless steel. Eng. Fail. Anal. 2021, 125, 105411. [CrossRef]
25
materials
Review
Electrochemical Deposition of Ni, NiCo Alloy and
NiCo–Ceramic Composite Coatings—A
Critical Review
Nyambura Samuel Mbugua, Min Kang *, Yin Zhang, Ndumia Joseph Ndiithi,
Gbenontin V. Bertrand and Liang Yao
College of Engineering, Nanjing Agricultural University, Nanjing 210031, China;
[email protected] (N.S.M.); [email protected] (Y.Z.); [email protected] (N.J.N.);
[email protected] (G.V.B.); [email protected] (L.Y.)
* Correspondence: [email protected]; Tel.: +86-25-5860-6578
Abstract: In recent years, alloy and alloy-ceramic coatings have gained a considerable attention owing
to their favorable physicochemical and technological properties. In this review, we investigate Ni,
NiCo alloy and NiCo–ceramic composite coatings prepared by electrodeposition. Electrodeposition is
a versatile tool and cost-effective electrochemical method used to produce high quality metal coatings.
Surface finish and tribological properties of the coatings can be further improved by the addition of
suitable agents and control of deposition operating conditions. In this review, Ni, NiCo alloy and
NiCo–ceramic composite coatings prepared by electrodeposition are reviewed by critically evaluating
previous researches. The use of the coatings and their potential for future research and development
are discussed.
1. Introduction
Materials are a fundamental pillar in engineering technology. The materials’ electrochemical,
thermal and mechanical interaction begins on the surface. However, material surfaces face the constant
threat of wear and corrosion resulting in massive losses in industry. The use of surface enhancement
technology to prevent or mitigate the loss has therefore become inevitable. Over the last few decades,
major scientific development in the fields of metallurgy and materials has occurred, giving rise to new
engineering materials with superior properties. Developing suitable processing to produce desired
materials is complex and requires altering the inherent properties of the materials. Consideration is
made of the overall economic perspective and its environmental impact.
Electrodeposition is an electrochemical process that is used to modify the surface structure. Use
of electrodeposition in surface engineering can be traced to nearly 200 years ago based on some
hypotheses [1]. The galvanic cell, invented in early 1800 s, paved way for use of electric current to
produce coatings as a more cost-effective technique. The electrodeposition technique possesses an
edge over other coating techniques due to several reasons [2]:
(v) It is an easy concept that can be replicated in industry and laboratories with minimal
technological barriers.
(vi) Electrodeposited Ni coatings have exhibited superior density and lower porosity.
Electrodeposition is based on the principle that a layer of coating, either single or multilayer, forms
as a result of the electrode–electrolyte electrochemical reactions occurring leading to electrodeposition
of ions contained in electrolyte. It utilizes the properties of materials in their metastable state as a
result of reduction of the grain size on the nano scale. In such a state, the grain boundaries contain
a proportion of atoms that is higher or equivalent to those inside the grains [3]. These new types
of materials consisting of nanoparticles are referred to as nanocomposite materials and they exhibit
superior properties compared to traditional grain sized materials and sometimes offer completely
new properties altogether. As a result, nanocomposite materials with grain sizes below 100 nm
have received considerable attention from scientists and researchers all over the world. Several new
electrodeposition synthesis methods have been devised over the years to increase production efficiency
of new materials and minimize cost. This has ranged from using the basic direct current set ups to
more ambitious procedures such as: pulse plating, jet electrodeposition, and the pulse reversal current
technique [4–6].
Pure Ni is one of the most widely used alloying metals in the world owing to its superior corrosion
and wear resistance [7]. Electrodeposited Ni coatings have uses in decoration, functional uses, as well
as in engineering for surface protection [8]. Superplasticity in metals (materials exhibiting increased
elongations to failure of >500%) is dependent on elevated testing temperatures and fine grain sizes of
<10 μm [9]. Prasad and Chokshi [9] reported that electrodeposited nanocrystalline Ni is characterized
by good superplasticity properties and is used in studying the phenomenon in electrodeposited metals
owing to the ease of synthesizing coatings with small grain sizes. Ni coatings exhibit improved
resistance to localized corrosion and this makes them perfect for use as anti-corrosive coatings [2].
Furthermore, research shows that thin films of different materials coated with epitaxial thin Ni coatings
of a few nanometers have similar hardness to bulk nickel, such that the wear resistance of micro/nano
electro mechanical system devices can be greatly improved if coated with a thin Ni coating [10]. Studies
have shown that nano-sized nickel can aid the diamond yielding process. Ni atoms have a three
dimension absent state which serves to attract electrons in the carbon fullerenes. This in turn causes
the sp2 fullerene to be transformed into a diamondlike sp3 structure. This transformation can also
be attributed to high reactivity which effectively aids the change at an impulse during shock wave
loading [2].
Ni based alloy coatings can also be produced using electrodeposition. Choice of the alloying
metal depends on the properties desired, which can range from good electrical conductivity, good
wear and corrosion resistance, soft magnetic properties to special optical properties. Over the years,
many different types of metals have been electrodeposited with Ni to form alloys: Co, Fe, Cu and W.
Ni–Co is one such alloy. There exist several different synthesis techniques for Ni–Co alloy coatings
and they include: radio frequency magnetron sputtering [11], electrodeposition [12] and vacuum
evaporation [13,14]. The electrodeposition technique has several advantages over the other two
methods and they include: low cost, simplicity, scalability and manufacturability [15]. Furthermore,
electrodeposition can be used to grow a wider range of materials.
Research shows that electrodeposited Ni–Co alloy coatings exhibit better properties compared
to pure Ni coatings [12]. Ni–Co alloy exhibits higher hardness, better adhesion, excellent magnetic
properties, high wear and corrosion resistance as well as good stability at high temperature [16–19].
Wang et al. [20] researched the effect of cobalt content on mechanical and microstructural properties on
Ni–Co alloy coatings. It was reported that Ni–Co alloy coatings exhibited approximately double the
microhardness when compared to pure Ni coatings. It was also reported that the Ni–49Co coatings
exhibited a decreased rate of wear in comparison to pure Ni coatings. Hassani et al. [21] researched
on low temperature superplasticity of nanocrystalline electrodeposited Ni–Co alloy with an average
grain size of 20 nm. It was reported that a maximum elongation of 279% at a temperature of 773 K was
28
Materials 2020, 13, 3475
obtained. Ni–Co alloys are considered to be the best suited materials for replacing hard chromium [22].
Research shows that microhardness in Ni–Co alloys increases gradually with increase in Co content up
to an optimum level, after which the microhardness decreases with further increase in Co content [16].
To further improve the properties of Ni–Co alloy coatings, nanoparticles have been suspended
in electrolyte and they become embedded into the electro-formed solid phase layer during
electrodeposition [23]. In such materials, the inherent properties of the nanoparticles have been
found to significantly influence the overall properties of the nanocomposite coatings. Many different
types of nanoparticles have been electrodeposited with Ni–Co alloy including SiC, Al2 O3 , SiO4 , ZrO2 ,
Cr2 O3 , Si3 N4 and TiO2 [24]. These nano particles used in electrodeposition can be classified as either
hard materials or soft materials depending on the desired properties. Soft materials such as graphite
offer properties which include lower the friction coefficient to reduce wear and the coefficient of
friction between shearing surfaces [25]. Hard materials such as Al2 O3 improve the microhardness and
wear resistance of surfaces [26]. The matrix phase microstructure coupled with nanoparticle content
and distribution in the metallic matrix phase significantly influences the properties of nanoparticle
reinforced Ni–Co matrix nanocomposite coatings. From past research, it is clear that addition of
nanoparticles has served to improve the properties of Ni–Co coatings. In some instances, however,
addition of Co has been observed to improve the overall properties of Ni-nanocomposite coatings such
as hardness and residual stress [27]. Addition of Co2+ in Ni/diamond plating baths has been seen
to greatly improve the deposition of diamond in the coatings resulting in higher diamond content,
enhance bonding between matrix and particles, as well as more uniformly dispersed particles in the
metal matrix, resulting in improved wear resistance and hardness properties [28]. The properties of
Ni–Co alloys and their composite coatings have also been previously reported on by Karimzadeh
A et al. [29]. This review aims takes a different approach to the wide research field of Ni–Co and
Ni–Co-based composite coatings. Factors such as the effect of electrode orientation and forces existing
in the electrolyte bath on the deposition process have been considered. Moreover, additives such as
boric acid have been extensively explored, and the intricate working of the Watts solution has been more
deeply discussed for easier understanding. Properties such as adhesion between substrate and coating
have been extensively discussed, comparisons between techniques drawn, and recommendations for
further adhesion improvement have been presented.
2. Electrodeposition Methods
29
Materials 2020, 13, 3475
1
f = (1)
TOFF + TON
TON
γ= ∗ 100 (2)
TOFF + TON
TON
Iavg ∗I = Ipeak (3)
TOFF + TON peak
where γ is the duty cycle, f the frequency, Ipeak the peak current density, and Iavg the average
current density.
In pulse electrodeposition, peak current density has a significant effect on microhardness, crystallite
size, surface morphology, microstructure, composition, and tensile strength of PC deposited Ni–Co
alloys and their nanocomposites [34]. Crystallization and growth in turn determine the microstructure
of the nanocomposite. The texture of the coatings is determined by both peak current density and
organic surfactants [35]. The quantity of adatoms located at the surface is higher due to the applied
higher current density as compared to DC electrodeposition. This results in smaller grain sizes [5].
Padmanabhan [36] reported that pulse electrodeposition is used as an effective method for the reduction
of grain sizes to nanoscale [5].
Although it is a relatively low-cost synthesis technique with simple implementation, pulse
electrodeposition is ideal for production of full density nanocrystalline materials [33]. Yang and
Cheng [32] reported that the morphology of the deposited coatings changed from nodular to acicular,
and a finer grain size was observed with increasing pulse frequency and decreasing duty cycle. It has
also been reported that at low current densities, smoother surface morphologies are observed [37],
but an increase in peak current density produces distinct colony-like morphologies characterized by
clearer colony boundaries [34].
Pulse electrodeposition exhibits several advantages over DC electrodeposition, such as improved
wear resistance and hardness, particle distribution, structure, morphological structure and the ability
to control the grain sizes of the deposits [38]. PC electrodeposition has higher instantaneous current
density compared to DC electrodeposition, thereby increasing its effectiveness in agitating the
adsorption–desorption processes that occur at the Ni electrolyte interface. This makes it possible
to control the electrodeposited Ni coating microstructure [39,40]. It has also been reported that PC
deposits contain a higher nanoparticle content compared to coatings deposited by the DC deposition
technique [31]. Yang and Cheng [32] concluded that the high microhardness of deposited Ni–Co–SiC
coatings was primarily associated with increasing SiC content in the coatings with increasing pulse
frequency and decreasing duty cycle. This significantly increased the microhardness as a function of
the dispersion strengthening mechanism. Watts type baths used for pulse plating of Ni based coatings
have been reported to produce the best coating results [2].
The PC deposition technique suffers from a drawback called the double layer capacitance effect.
The double charging layer of the electrodes is subjected to charging and discharging occurrences. In a
case where the ON- and OFF-times are much shorter than the charging and discharging times, the PC
would revert to DC current. As such, care is taken to select a high frequency thus ensuring the effect of
the double layer capacitance effect is negligible.
30
Materials 2020, 13, 3475
As the plating solution flows, electric current is transferred along the stream of fluid to the substrate
surface, thereby enabling deposition to occur on the cathode surface where the jet flows over [41]. Jet
electrodeposition is a high-speed electroplating technique that offers a wide range of advantages [4].
These include: (i) a higher deposition rate compared to other conventional electrodeposition techniques,
and (ii) a more efficient grain size refining effect. This is attributed to the cathode in the jet technique
having a larger overpotential that can be used simultaneously with higher current densities. In
this technique, Co content in the coatings increases with increases in electrolyte jet speed, Co2+ ion
concentration in the electrolyte, and cathodic current density.
31
Materials 2020, 13, 3475
Ip−
x= (11)
Ip+
1
f = (12)
T
Figure 2 shows the schematic diagram of a typical pulse-current wave form when Iav = 15 A·dm2 ,
λ = 0.5, x = 0.5, and f = 10 Hz.
Figure 2. Schematic diagram of pulse waveform when Iav = 15 A·dm2 , λ = 0.5, x = 0.5, and f = 10 Hz.
Pulse reversal current (PRC) electrodeposition has been used over the years owing to its unique
ability to enhance nanoparticle incorporation in electrodeposited composite coatings, with the highest
reported content being 23 wt% [44–46]. Ni–Co coatings produced using the PRC synthesis technique
have been reported to exhibit higher hardness, better anti-wear properties, more compact surfaces,
lower residual macro stress and better compact surfaces than those produced using DC and PC
electrodeposition techniques [6,40]. In particular, the hardness improvement has been attributed to
increased distribution of nanoparticles in the deposited Ni–Co matrix. As for lower residual macro
stress, PRC deposits are characterized by uniform coatings since the technique hinders thickening
of the corners and edges, as is common with DC electrodeposition. It can also be postulated that
adsorption of H atoms by the deposited coatings is suppressed by pulse intervals and reversal current
in PRC technique [40]. PRC deposited Ni–Co nanocomposite coatings possess smoother surfaces, finer
Ni matrix crystals, and smaller grain sizes compared to DC technique deposited coatings [40].
The smooth surfaces coupled with high hardness increase the load-carrying capacity of
Ni–Co nanocomposite coatings and thereby improves their wear resistance properties. The lower
macro-residual stress also plays a great part in wear resistance, where it causes lower brittleness
and higher toughness in the coatings, and this decreases the rate of nanoparticle flaking and metal
matrix peeling during wear testing [40]. PRC technique can be used in ultrasonic power conditions
and this presents desirable effects on deposited Ni–Co coatings [47,48]. Overall, it presents several
advantageous effects on the electrochemical process:
(i) Hindering sundries adsorption thereby altering the reaction mechanism,
(ii) An increase in exchange current density,
(iii) Lowering cathodic polarization,
(iv) Current efficiency and yield improvements, and
(v) (Strengthening conversion and diffusion.
It has been reported that Ni–Co nanocomposite coatings electrodeposited using the PRC technique
subject to ultrasonic conditions present finer grained, compact, and uniform coatings [6].
32
Materials 2020, 13, 3475
33
Materials 2020, 13, 3475
density range of 100–200 A·dm−2 was suggested for grain sizes ranging from 15–20 nm, a 7%–8%
cobalt content, 590–600 kg mm−2 microhardness and 118–1200 MPa tensile strength. When using the
PC technique to electrodeposit Ni–Co, it was reported that the lowest current densities achieved the
most compact layer of alloys [5].
Extremely high current density Jp however has been reported to have a drastic effect on
microhardness and tensile strength of electrodeposited Ni–Co coatings [34]. This has also held
true for Ni–Co alloys electrodeposited using jet electrodeposition technique [4]. It could be suggested
that this is because of the decrease in cobalt content with increase in the current density. Li et al. [34]
researched the effects of peak current density on mechanical properties of Ni–Co alloys and reported
that Co content decreased with peak current density increase. One of the key aspects in electrodeposited
Ni–Co alloys and nanocomposite coatings is solid solution strengthening where the Ni and Co in
electrolyte combine to form a solid solution. A decrease in Co content will therefore yield less solid
solution strengthening for coatings electrodeposited at higher current densities. Figure 3 shows the
relationship between peak current density and cobalt content [34].
Figure 3. Cobalt content in Ni–Co alloy coatings with varying peak current density [34]. Reprinted
from Applied Surface Science, 254, Yundong Li, Hui Jiang, Weihua Huang, Hui Tian, Effects of peak
current density on the mechanical properties of nanocrystalline Ni–Co alloys produced by pulse
electrodeposition/Pages No. 6865–6869, Copyright (2020), with permission from Elsevier.
In the case of Ni–Co, electrodeposition of the Co element is influenced and controlled by diffusion
as compared to that of Ni which is predominantly controlled by activation. In light of this, an increase
in cathodic current density results in an increase in cathodic overpotential. Concurrently, activation of
the electrode reaction increases and, as a result, the rate of deposition of Ni element into the coating
increases significantly [34].
In the case of Ni–Co nanocomposites, it has been reported that thinner coatings resulting from
smaller amounts of electricity tend to have a higher content of nanoparticles compared to thicker
coatings that are formed with larger amounts of electricity. This suggests that nanoparticles become
adsorbed during the early phase of deposition and become unevenly distributed with increases in
thickness of coatings [55].
34
Materials 2020, 13, 3475
are several interrelated factors that influence incorporation and distribution of nanoparticles into
the Ni–Co coatings. These factors can be broadly classified as (i) electrolyte composition (reagents,
pH, additives); (ii) nanoparticle characteristics (size, shape, type); (iii) deposition parameters (current
density, bath temperature, concentration of nanoparticles and rate of electrolyte agitation); (iv) cobalt
content in the coating where the Co2+ cations that are adsorbed on the particle surface increase
the incorporation of nano particles into the coating deposits, causing an increase in nanoparticle
deposition with increase in cobalt content [28,51,56]; and (v) electrode orientation, which determines
the incorporation efficiency of the nanoparticles. Increase in nano particle content on Ni–Co nano
composite coatings can be improved by using the sediment deposition technique (SCD) as opposed
to the conventional electrodeposition technique. In the SCD technique, the electrodes immersed in
the electrolyte are placed horizontally and parallel to each other. As such, electrodeposition in this
technique takes advantage of gravitational pull coupled with the electrophoresis force resulting in better
incorporation of nano particles [30,49]. In conventional electrodeposition, only the electrophoresis
force is utilized. Figure 4 shows the electrodeposition setups depending on electrode orientation in the
electrolyte. Research shows that nanoparticle content in electrodeposited Ni–Co coatings increases
steadily with increase in nanoparticle concentration to a given maximum value beyond which the
nanoparticle content in the deposit decreases. This increase in nanoparticle content with increase
in concentration can be attributed to increased transportation of the nanoparticles to the cathode
surface where more and more nanoparticles can be engulfed in the growing Ni–Co matrix. At high
concentrations of nanoparticles, the interaction equilibrium between the suspended nanoparticles and
the embedded ones is exceeded, beyond which the surface of the cathode becomes covered such that
more suspended nanoparticles cannot be embedded into the coatings. Moreover, there is increased
mechanical collisions between the nanoparticles and this reduces their transportation efficiency across
the electrolyte bulk. The content of nanoparticles in the coatings therefore decreases.
Figure 4. Schematic image of the deposition setups: (A) DC power supply, (B) Epoxy cover, (C) plating
solution, (D) anode, (E) cathode, (F) magnetic bar and (G) external pH–temperature probe [30].
35
Materials 2020, 13, 3475
the coating roughness and improves uniformity [53]. Furthermore, increasing the stirring speed of
the electrolyte during the deposition process increases the content of nanoparticles deposited in the
Ni–Co coatings up to a given maximum level beyond which the content reduces [26,57]. At low
agitation rates, the concentration of nanoparticles surrounding the cathode may reduce, resulting in the
feed rate of the nanoparticles being lower than their adsorption into the Ni–Co matrix. Additionally,
incomplete dispersion as a result of insufficient convection may cause agglomeration of nanoparticles
and gravity settling. The surface energy of nanoparticles is greatly reduced when nanoparticles
agglomerate, and this lowers the content in the deposited coatings. At higher agitation rates, the
volume of nanoparticles reaching the cathode surface (mass transfer) increases thereby increasing
the overall content of nanoparticles in the coatings. Goto et al. [55] reported that for the range of the
experiment conducted, the nano diamond (ND) content in the coatings increased with increase in
the stirring speed as shown in Figure 5. Where pulse electrodeposition is conducted using sediment
deposition technique, agitation is a factor of TON /TOFF ratio, where TON and TOFF represent the time
intervals when the agitation is on and off. In this technique, the nanoparticles settle on the horizontal
cathode with aid from gravitational force. As such, the lower the TON /TOFF ratio, the higher the rate
of nanoparticle incorporation into the growing Ni–Co matrix [57].
1'FRQWHQWRIFRDWLQJ PDVV
6WLUULQJVSHHG USP
Figure 5. Relation between stirring speed and nano diamond (ND) content of the coatings [55].
Excessive rotation is however detrimental to coating quality. Vigorous hydrodynamic forces are
generated at high agitation rates and these forces pluck out nanoparticles from the cathode surface
before they become successfully embedded in the growing matrix surface, leading to lower nanoparticle
content in the deposited coatings. This conclusion was reported by [26,58,59] who related the increase
in corrosion resistance up to a maximum value to an increase in electrolyte agitation rate, beyond
which it decreased with further increase in agitation rate.
3.5. Temperature
Shi et al. [60] postulated that there is a two-fold effect with respect to temperature. When
electrolyte temperatures are low, there is an increase in nanoparticle kinetic activity with a rise in the
temperature of the bath, and this boosts adsorption of nanoparticles into the metal matrix. Increasing
the temperature decreases the density and viscosity of the bath, thereby improving the mobility
of ions within the electrolyte. As such, coatings deposited at higher temperatures exhibit superior
properties owing to thicker coatings and higher nanoparticle contents than those deposited at lower
temperatures [53].
Temperature variation has also been reported to alter the size of crystals in the deposited coatings.
Prabu and Wang [53] reported that increasing temperature from 20 to 60 ◦ C resulted in larger and
36
Materials 2020, 13, 3475
sharper crystals. This was attributed to increase in rate of reducible ion diffusion to the cathode with
an increase in temperature which decreased the polarization resistance. Research on the effect of
temperature on Ni–Co coatings shows that the deposition rate increases with increases in temperature
of the electrolytic bath [4]. Idris et al. [54] reported that, for Ni–Co coatings electrodeposited using high
speed jet electrodeposition, an increase in temperature from 55 to 65 ◦ C (at 1 A/cm2 current density)
resulted in a subsequent increase in thickness of the electrodeposited coatings from 61.4 to 71.7 μm,
respectively. Similar increasing trends of the thickness with rise in temperature were observed at
current densities of 0.1, 0.3 and 0.5 A/cm2 . This can be attributed to grain growth as a result of a free
growth mode of Ni resulting from the temperature rise.
When the temperature exceeds certain limits, thermodynamic ion movement is enhanced greatly,
and the nanoparticle’s kinetic energy increases. As a result, less nanoparticles become adsorbed into
the metal matrix. This conforms to Langmuir’s adsorption theory, where temperature increase beyond
certain levels has a negative effect on nanoparticle absorbability. As a result, the electric field and the
overpotential of the cathode are decreased, making it harder for nanoparticles to be embedded into the
coating. As such, lower contents of nanoparticles are observed in the deposited coatings.
3.6. Electrolyte pH
According to past research, it can be seen that the composition and structures of Ni–Co alloys
and their nanocomposites can significantly affect their physiochemical properties. The effect of pH
on Ni–Co deposits is predominantly dominated by three factors [61]: (i) the acidic environment in
the electrolyte dissolving newly deposited metal atoms on the cathode surface, (ii) metal hydroxide
formation and adsorption on the surface of the electrode, and (iii) normal electrodeposition of metals.
When the electrolyte pH is low, the newly deposited metal becomes dissolved at a faster rate and
formation and adsorption of metal hydroxides becomes depressed. In this case, the electrodeposition
process is mainly dominated by normal electrodeposition of metals and this results in lower Co2+
in the bath. At higher pH values however, the formation and adsorption of metallic hydroxides is
promoted, and the newly deposited metal dissolution becomes suppressed. For higher pH values, the
electrodeposition process is dominated by formation and adsorption of Co hydroxides on the cathode
surface and this produces higher Co contents in deposited coatings [62,63]. Tian et al. [61] reported a
gradual increase in Co content from 9.4% to 19.6% with an increase in the value of pH from 2.0 to 5.4.
pH value in the electrolyte has also been observed to affect current efficiency. An increase in current
efficiency from 52.1% to 81.2% was also reported with increase in pH from 2.0 to 5.4. Research linking
pH value to hydrogen evolution has also been reported. Increase in pH in Ni–Co-based deposits has
been associated with an increase the in hydrogen evolution rate, followed by creation of trace amounts
of Ni and Co hydroxides which hinder the growth of crystals [64].
37
Materials 2020, 13, 3475
Figure 6. Alloy composition as a function of Co2+ concentration in the electrolyte [16]. Reprinted from
Surface and Coatings Technology, 201, Meenu Srivastava, V. Ezhil Selvi, V.K. William Grips, K.S. Rajam,
Corrosion resistance and microstructure of electrodeposited nickel-cobalt alloy coatings/Pages No.
3051–3060, Copyright (2020), with permission from Elsevier.
Many models have been proposed concerning co-deposition of nanoparticles. Celis [69] suggested
that it was a five-step process which took into consideration the creation of an ionic cloud that engulfed
the nanoparticles, movement of the nanoparticles through the electrolyte and the diffusion layer. These
steps can be classified as: (i) formation of an ionic cloud on the surface of the nanoparticles from the
adsorbed ions; (ii) the charged nanoparticles are then transferred through the electrolyte bulk until
they reach the hydrodynamic boundary layer; (iii) through diffusion, the nanoparticles are transferred
en-masse to electrode surface; (iv) adsorption of electroactive ions and the free ions occurs on the
particles on the cathode; and (v) electroreduction of adsorbed ions occurs followed by incorporation
38
Materials 2020, 13, 3475
of particles into the growing metal matrix [70]. Guglielmi’s model proposed that the adsorption
mechanism onto the cathode followed a two-step process for the charged nanoparticles. Firstly, the
charged nanoparticles are loosely adsorbed while still being engulfed in a film of adsorbed cations [69].
6. Additives
Additives are added into the electrolyte bath during electrodeposition to increase the range of
current density, change physical and mechanical properties, reduce the size of crystallites by reducing
growth, increase the coating’s luster, reduce nanoparticle agglomeration, and reduce internal stresses
generated during the deposition process [35].
39
Materials 2020, 13, 3475
acid in Nickel plating baths during electrodeposition has been researched over the years and the results
can be summed up into six views:
(a) Boric acid suppresses oxygen evolution. Gadad and Harris [77] researched on oxygen
incorporation in electrodeposited Mi, Fe and Ni–Fe alloy coatings. It was reported that an
increase in applied current density resulted in an increase in the content of oxygen in Ni coatings
and this posed a detrimental effect to the magnetic and electrical properties of the coatings.
Addition of boric acid reduced the oxygen incorporation in all three electrodeposition systems,
with less than 2 wt% oxygen observed in all cases. The buffering effect is not attributed directly to
the boric acid but more to the complexing ability of boric acid with metal ions in the electrolyte.
(b) Boric acid promotes deposition of Nickel by acting as a catalyst. The adsorptive interaction of
boric acid has also been observed in Ni–Zn alloy coatings where boric acid increased the current
efficiency of the system at lower Zn (II) concentrations and increased the Ni content of the coatings
at higher Zn (II) concentrations [78]. Significant change in primary nucleation rate coupled with
suppressed secondary nucleation on coatings was also reported. Cyclic voltammetric deposition
results have shown that the hydrogen evolution rate (HER) increases relative to the increase in
boric acid concentration.
(c) Boric acid as a pH buffer. In electrodeposition, the practical buffer range is given at pKa ± 1,
but this value is much higher in the case of boric acid (9.23 ± 1 at 25 ◦ C). This is an anomaly
considering the pH of the Ni electrolyte is 4.0. The anomaly can be attributed to formation of
weak bound complexes between nickel ions and boric acid, such that the said complexes act as pH
buffers [79,80]. The presence of these complexes however, has yet to be confirmed experimentally.
This pH buffering phenomena has been found to be significantly influenced by the applied current
density. Tsuru et al. [76] reported that at lower current densities (below 1.0 A dm−2 ), the pH
buffering properties of boric acid were exhibited.
(d) Suppression of hydrogen evolution by boric acid. During electrodeposition, electric current
flowing through the system causes an increase in pH and as a result, hydrogen gas is produced at
the cathode. Hydrogen evolution at the cathode is detrimental to the reduction of metal ions, and
therefore boric acid is added into the plating bath solution to prevent electrode surface passivation
as well as act as a surface agent which acts as a selective membrane to block passage of the
reduction of Nickel, while permitting the reduction of iron in a retarded state. Improving the
electrodeposition current density range thereby minimizes the effect [81]. Yin et al. [82] suggested
that boric acid acted like a surfactant which was adsorbed onto the surface and hinders hydrogen
evolution. The adsorbed boric acid interferes with the alloy nucleation process thereby reducing
the hydrogen evolution rate in Ni-enriched phases [82]. It should be noted that the hydrogen
evolution suppressing properties of boric acid have only been observed in presence of nickel ions
and this suggests that there exists a mutual interaction between nickel and boric acid [76].
(e) Reduction of passive film formation by boric acid during Ni electrodeposition. Boric acid
was found to significantly deter surface passivation on Ni reduction in Fe–Ni setups [82].
Tsuru et al. [76] suggested that, by acting as a surface agent, boric acid hindered passivation of
the electrode surface during reduction of nickel.
(f) Accelerating growth rates of deposits. Boric acid improves the lateral as well as the outward
growth rate during deposition of nickel [83].
It has also been reported that coatings electrodeposited in electrolytes containing boric acid exhibit
better appearance coupled with reduced brittleness [81].
6.2. Surfactants
Nanoparticle stabilization in the electrolyte and non-agglomerated nanoparticle dispersion are key
elements of producing Ni–Co nanocomposite coatings with excellent properties. Surfactants such as
cetyltrimethylammonium bromide (CTAB) [84–86], sodium dodecyl sulfate (SDS) [87,88], and sodium
40
Materials 2020, 13, 3475
lauryl sulfate (SLS) [89] are added into the Ni–Co nanocomposite coating baths to prevent nanoparticle
agglomeration by reducing the electrolyte’s surface tension to create smaller hydrogen bubbles thereby
limiting pitting effect [35]. Surfactants of cationic and anionic nature are commonly used owing to their
significant influence on ceramic-based nanoparticle dispersibility. Surfactants added into the bath work
by changing the cathode polarization potentials, thereby changing the adhesion, smoothness, grain
growth rate, and grain size of the deposited coatings [90]. Several researchers have used optimum
quantities of different surfactants and reported improvement in corrosion resistance and mechanical
properties [91].
Different surfactants have different effects on different nanoparticles. It was reported that CTAB
surfactant offered a better alternative over SDS and triton X when they were compared in deposition of
SiC nanoparticles using the PRC technique [92] and the PC technique [35]. However, the advantage
offered by surfactant addition is not limitless. When the concentration of a surfactant exceeds a certain
limit, it becomes counter-productive. Ger M [93] observed that excessive CTAB surfactant increased
nanoparticle’s adhesive force which resulted in deposition of coarser SiC nanoparticles.
6.3. Saccharin
Addition of saccharin results in an increase in alloy deposition overvoltage which promotes
deposition of Ni while hindering that of Co, and as such the resulting Ni–Co coatings exhibit reduced
Co content [4,94]. Research shows that surface morphologies of Ni–Co coatings exhibited colony-like
morphologies which consist of grain morphologies where several grain colonies converge to form one
larger colony. As such, grain size is greatly reduced owing to the grain refinement phenomena. This is
achieved by inhibition of pyramidal growth by saccharin thereby leading to the production of shiny,
smooth surfaces [94]. This concurs with Weil and Cook [95] who reported that addition of organic
additives such as coumarin and thiourea into the Ni–Co electrolytes hindered the growth of pyramids,
caused surface roughness reduction, grain size reduction and increased surface brightness.
Increase in saccharin content in the coatings up to a certain value also improves the microhardness
of Ni–Co coatings beyond which the microhardness reduces with increase in saccharin content.
Li et al. [94] reported that increase in saccharin content to 3 g/L, 4 g/L and 5 g/L resulted in increased
microhardness values of 456 kg/mm2 , 507 kg/mm2 and 554 kg/mm2 , respectively. This conclusion was
also reached by Wang et al. [96]. The increase in microhardness with increase in saccharin content to a
certain value can be attributed to grain refining effect, and the decrease in microhardness beyond that
level of saccharin can be attributed to the inverse Hall–Petch relationship when refining of grain size
reaches a certain level [97]. The drawback to adding saccharin to the electrolyte is the reduced ductility
of the resulting Ni–Co coatings owing to the sulphur and carbon impurities that are usually present in
saccharin laden nanocrystalline coatings. The said impurities separate into grain boundaries thereby
preventing the efficient sliding of grain boundaries and hence the low ductility [98].
Saccharin has been reported to act as an internal stress reliever in electrodeposited Ni–Co coatings
and this has been attributed to grain refinement. Internal stresses are developed within the coating layer
during the electrodeposition process and they cause oriented resultant strain in deposited coatings.
Hydrogen ion reduction occurs at the cathode, and the small sized H+ promote favorable conditions
for diffusion to the coating’s active centers. As the H+ become transformed into H molecules, internal
stresses are developed as the volume changes. They are classified into three main categories [99–102]:
(i) Macroscopic stresses. These are caused by inhomogeneity in the deposited coatings. These
comprise of either compressive or tensile stresses and they occur in galvanic cells.
(ii) Microscopic stresses. These originate at grain boundaries and at locations where dislocations
accumulate.
(iii) Sub-microscopic stresses.
While internal stresses have been known to improve hardness and abrasion resistance of deposited
coatings, at high levels, these stresses increase the coating’s brittleness. Brittle coatings develop
41
Materials 2020, 13, 3475
extensive microcracks which expose the substrate surface to corrosive attack and degradation when
exposed to a corrosive medium. The saccharin molecules are reversibly adsorbed on active sites
thereby hindering growth of crystals and impeding surface diffusion of adatoms. As such the volume
of grain boundaries increases and this dissipates the energy created by internal stresses [103].
Saccharin has been employed in tensile research of electrodeposited Ni–Co nanocomposite
coatings. Wang et al. [96] reported that PC deposited Ni–Co/Al2 O3 nanocomposites containing
saccharin exhibited low-temperature superplasticity, where a maximum elongation of 632% was
achieved at a strain of 1.67 × 10−3 s−1 and a temperature of 823 K. The dominant superplastic
accommodation process was taken to be dislocation glide.
7.1. Microstructure
As stated earlier, Ni–Co coatings deposition is anomalous. This can be explained as a function of
local pH increase resulting in creation and adsorption of metallic hydroxides, as well as a faster rate
of cobalt hydroxide adsorption [104], deposition of Co cations (first step) followed by Ni deposition
(second step) in a two-step process [105], and preferential Co element deposition which causes the
diffusion layer to become depleted [4]. Ni–Co alloys with up to 58 wt% cobalt content exhibit the
single phase of Ni matrix with FCC type of phase structure. When the cobalt content is in the range
between 64 wt% to 80 wt% the phase structure becomes a combination of FCC and HCP as shown in
Figure 7 [68].
Figure 7. X-ray diffraction (XRD) patterns of Ni–Co films deposited from the electrolytes containing
different Co concentrations [68]. Reprinted from Applied Surface Science, 258, Ali Karpuz, Hakan
Kockar, Mursel Alper, Oznur Karaagac, Murside Haciismailoglu, Electrodeposited Ni–Co films from
electrolytes with different Co contents/Pages No. 4005–4010, Copyright (2020), with permission
from Elsevier.
42
Materials 2020, 13, 3475
Ni–Co alloys with Co content above 80 wt% exhibit complete HCP phase structure [16,40,68].
Ni–Co alloy surface morphologies are significantly influenced by the coating’s chemical composition.
Rafailovic et al. [106] researched the mechanical properties of Ni–Co alloys deposited on Cu substrates.
It was reported that a platelet structured morphology was formed for coatings with a Ni2+ /Co2+ ratio
of 0.25 at 65 mA cm−2 . The surface morphology exhibited enhanced dendritic growth when the ratios
were 0.5 and 2. At the highest Ni2+ /Co2+ ratio of 4, the surface morphology exhibited cauliflower
structure as shown in Figure 8.
(a) (b)
(c) (d)
(e)
Figure 8. SEM micrographs of Ni–Co deposits obtained at a current density 65 mA cm−2 from an
electrolyte with different Ni2+ /Co2+ concentration ratio: (a) 0.25, (b) 0.5, (c) 1, (d) 2 and (e) 4 [106].
Reprinted from Materials Chemistry and Physics, 120, L.D. Rafailovic, H.P. Karnthaler, T. Trisovic, D.M.
Minic, Microstructure and mechanical properties of disperse Ni–Co alloys electrodeposited on Cu
substrates/Pages No. 409–416, Copyright (2020), with permission from Elsevier.
43
Materials 2020, 13, 3475
advantageous and detrimental effect depending on the desired output of the process. Hydrogen
offers a promising versatile, efficient, and clean candidate for use as an energy source to replace
commonly used fossil fuels which cause CO2 emissions that are harmful to the environment [107–109].
Hydrogen can be successfully generated using the less efficient (higher operating cost) alkaline water
electrolysis [110,111], or using low hydrogen evolution reaction (HER) overpotential electrodes [112].
Ni-based alloys and compounds form such electrode materials owing to their low cost coupled
with high catalytic activity and stability [113]. In the case of depositing quality coatings however,
the evolution of hydrogen gas is detrimental to the structure, hence the critical need to control its
production. The synthesized hydrogen gas attaches to the surface of the base metal creating a blanket
of air that inhibits nucleation and deposition of the coatings and this causes poor adherence leading
to non-uniform coatings [114]. The hydrogen evolution phenomenon has been reported to be more
significant at higher current densities, owing to lower hydrogen overpotential where numerous gas
pits are formed on the coating surface as a result of the hydrogen produced [112].
Figure 9. The effect of current density on microhardness of Ni–Co/SiC and Ni/SiC coatings [115].
Reprinted from Surface and Coatings Technology, 206, S.M. Lari Baghal, M. Heydarzadeh Sohi,
A. Amadeh, A functionally gradient nano-Ni–Co/SiC composite coating on aluminum and its tribological
properties/Pages No. 4032–4039, Copyright (2020), with permission from Elsevier.
44
Materials 2020, 13, 3475
law, which provides that the sliding wear volume loss is directly proportional to friction coefficient
and inversely proportional to the hardness of the material [117]. This relationship can be seen in
Figure 10 [60].
Figure 10. Microhardness and wear rate of the nanocomposite coating vs. weight percentage of
co-deposited SiC particulates in the nanocomposite coating [60]. Reprinted from Applied Surface
Science, 252, Lei Shi, Chufeng Sun, Ping Gao, Feng Zhou, Weimin Liu, Mechanical properties and wear
and corrosion resistance of electrodeposited Ni–Co/SiC nanocomposite coating/Pages No. 3591–3599,
Copyright (2020), with permission from Elsevier.
In some instances, however, the wear rate decreases with decrease in microhardness, a deviation
from Archard’s Law. This is concurrent with findings reported by Wang et al. [20] where the decrease
in wear rate beyond 49 wt% with a concurrent decrease in microhardness from 462 HV to 298 HV
was attributed to changes in the phase structure. As Co content increases, the phase structure of the
deposited Ni–Co coatings changes from solely FCC structure, to FCC coupled with HCP structure,
and when the Co content goes beyond 80 wt% (See Figure 7), the phase structure is transformed to a
predominantly HCP structure. This transformation in phase structure to a higher ratio of HCP causes
a decrease in the coefficient of friction (COF) of the deposited Co-rich coatings and therefore decreased
wear loss [20]. As such, the decreasing wear rate was associated with the decreasing coefficient of
friction (COF). This is shown in Figure 11.
Magnetic measurements done on electrodeposited Ni–Co coatings show that Co content has
significant influence on the magnetic and structural properties of the coatings. Increase in Co content in
the coatings results in a gradual increase in the saturation magnetization. This conclusion is concurrent
with Karpuz et al. [68] where the highest in-plane saturation magnetization of 1000 emu/cm3 was
achieved at the highest contents of Co (80%). The same trend was also reported by [118].
Coatings improve the performance of the component by isolating the material’s structure from
the environment. Different substrate materials have been used for deposition of Ni, Ni–Co, and
Ni–Co-nanocomposite based coatings ranging from steel, aluminum, to copper [106,115]. Failure
of systems is usually associated with substrate–coating interface failure owing to the differences in
mechanical and physical properties. Coating adhesion plays a key role in a surface’s wear resistance
and it is measured through friction testing where the friction abruptly changes when the coating
breaks, also known as the critical load point. As such, a larger critical load indicates stronger coating
adhesion [119]. Different mechanisms aimed at improving the critical load have been researched over
the years. Wei et al. [120] reported that use of the magnetic jet electrodeposition technique (MJE)
yielded a higher adhesion compared to traditional jet electrodeposition technique (TJE). At 4 g/L, TJE
45
Materials 2020, 13, 3475
technique had a maximum adhesion of 23.58 N compared to 33.20 N under MJE technique as shown in
Figure 12.
Figure 11. Friction coefficient as function of Co content in the Ni–Co alloy deposit [20]. Reprinted from
Applied Surface Science, 242, Liping Wang, Yan Gao, Qunji Xue, Huiwen Liu, Tao Xu, Microstructure
and tribological properties of electrodeposited Ni–Co alloy deposits/Pages No. 326–332, Copyright
(2020), with permission from Elsevier.
Figure 12. Adhesion of the composite coatings [120]. Reprinted from Journal of Alloys and
Compounds, 791, Wei Jiang, Lida Shen, Mingyang Xu, Zhanwen Wang, Zongjun Tian, Mechanical
properties and corrosion resistance of Ni–Co–SiC composite coatings by magnetic field-induced jet
electrodeposition/Pages No. 847–855, Copyright (2020), with permission from Elsevier.
46
Materials 2020, 13, 3475
The adhesion of Ni–Co binary alloys can be further improved by incorporating nanoparticles into
the matrix such that the effective area of contact between the substrate and coatings is increased, and the
dispersion strengthening mechanism of the nanoparticles improves the coating’s adhesion [121,122].
Another concept that has been considered for adhesion improvement is functionally graded materials
(FGMs), whereby interfacial problems are mitigated by controlling progressive changes in structure
and properties [123].
Ni–Co-nanocomposite coatings deposited at high current densities have exhibited higher surface
roughness. This can be traced to adsorption of nanoparticles into the coating surface coupled with
formation of pits and crevices as a result of an increase in hydrogen evolution rate at high current
densities. Similar observations were made by Dheeraj et al. [35] where a surface roughness of 2.31 ±
1.78 μm was reported for sample S3 which represented coatings deposited at higher current densities.
Figure 13. The potentiodynamic polarization curves of the Ni–Co alloy coating as a function of the
cobalt content [126]. Reprinted from Applied Surface Science, 307, Babak Bakhit, Alireza Akbari, Farzad
Nasirpouri, Mir Ghasem Hosseini, Corrosion resistance of Ni–Co alloy and Ni–Co/SiC nanocomposite
coatings electrodeposited by sediment codeposition technique/Pages No. 351–359, Copyright (2020),
with permission from Elsevier.
The phase structures of Ni–Co binary alloy coatings have been observed to consist of FCC
single-phase solid solutions [12,126]. In the case of the coating’s materials microstructure, single phase
structures have proven more corrosion resistant than two-phase structures. Corrosion attacks usually
47
Materials 2020, 13, 3475
occur along grain boundaries (between phases) owing to the galvanic cells that are created between the
phases and their higher levels of energy compared to parts located within the crystal itself. Moreover,
base centered cubic (bcc) phases have lower corrosion resistance compared to face centered cubic (FCC)
phases as a result of their lower packing factor.
As a factor of preferred orientation, it has been reported that Zn–Ni alloys exhibit high corrosion
resistance owing to crystallographic planes being predominantly present with higher packing
densities [127]. Babak et al. reported that Ni–17Co alloy coatings exhibited high corrosion resistance as
a function of predominantly (111) preferred orientation. Lupi et al. [66] found that electrodeposited
Ni–Co alloys containing 40%–50% Co content provide the best catalytic properties of the alloy for the
reaction leading to evolution of hydrogen in alkaline media.
8. NiCo–Ceramic Composites
Nanoparticles have been added to the Ni–Co matrix to further improve the properties. Ni–Co
nanocomposite coatings exhibit superior hardness and wear resistance to alloy coatings. This can
be attributed to inherent hardness of nanoparticles used, coupled with nanoparticle strengthening
through Orowan mechanism [30]. According to this mechanism, the adsorbed nanoparticles hinder
the formation and propagation of dislocations as the metallic matrix carries the load.
In some instances, the increase in Co concentration in the bath has been reported to improve
nanoparticle incorporation thereby improving coating properties such as corrosion resistance [30]. This
effect of nanoparticles can be linked to several factors:
(i) Reduction of exposed area open to corrosive media. Nanoparticles used in Ni–Co nanocomposite
electrodeposition are usually ceramics. When these nanoparticles are uniformly distributed in
the Ni–Co matrix, they minimize the metallic area that is exposed to corrosive attacks and, as a
result, the corrosion potential is shifted to nobler values [128].
(ii) SiC nanoparticles acting as physical barriers that hinder creation and propagation of corrosive pits.
(iii) Nanosized SiC particles which offer greater corrosion resistance than micro-sized particles when
used to deposit Ni–Co/SiC nanocomposites. Owing to their smaller sizes, such nanoparticles can
access structural defects such as porosities and cracks thereby mitigating the corrosive effect at
such locations.
(iv) Formation of micro-galvanic cells. The metallic matrix acts as an anode while the nanoparticles
act as cathodes when the Ni–Co nanocomposite coatings are exposed to corrosive media. Where
the metallic matrix’s electrochemical potential is less positive than that of the nanoparticles, the
corrosion mechanism of the micro–galvanic cells is transformed to uniform corrosion from pitting
and localized corrosion [126].
Other factors that may be associated with metallic alloy nanocomposite coatings include:
reinforcing phase induced hardening, texture evolution, grain refinement of the matrix, and solid
solution strengthening depending on selected matrix [129–131]. Wear resistance of Ni–Co alloy coating
is greatly improved by incorporation of nanoparticles in the matrix. This increase can be attributed
to the dispersion strengthening effect of adsorbed hard nanoparticles coupled with grain-refining
tendencies. The uniform dispersion in the matrix and to a small extent particle agglomeration may be
linked to improved wear resistance in Ni–Co nanocomposites. Similar conclusions were reached by
Shi et al. [60].
48
Materials 2020, 13, 3475
trend was reported by Borkar [31], with a maximum Al2 O3 content being achieved with 40 g L−1
nanoparticle concentration, beyond which the content decreased. Ni–Co alloy coatings exhibit a face
centered cubic (FCC) crystal structure and the same has been reported for Ni–Co/Al2 O3 . However, the
crystal orientation of the resulting Ni–Co/Al2 O3 nanocomposite coating undergoes transformation
from crystal face (200) lattice to (111) lattice [40].
Like with most Ni–Co nanocomposite coatings, adsorption of Al2 O3 nanoparticles into the Ni–Co
matrix results in a subsequent increase in the coating’s hardness and wear to a certain maximum,
beyond which the coating becomes brittle and spalls off. Tian [26] concluded that an increase in Al2 O3
nanoparticle concentration in the bath caused a subsequent increase in the corrosion resistance of the
Ni–Co/Al2 O3 nanocomposite coatings up to a certain limiting value, suggesting that optimal operating
conditions and parameters are key to corrosion resistance maximization. This may be attributed to
uniform dispersion of Al2 O3 nanoparticles in the Ni–Co matrix.
It has been suggested that Al2 O3 nanoparticles may also improve deposition of Ni and Co
elements in the coatings [26]. Wear resistance is also higher for Ni–Co/Al2 O3 compared to their alloy
counterparts and this too increases with increases in nanoparticle content.
9. Applications
Electrodeposited Ni–Co alloys and nanocomposites exhibit unique properties and, as such, they
are used for a wide variety of industrial applications. Their combined reduced localized corrosion and
49
Materials 2020, 13, 3475
microhardness greatly improves protective coating performance. These coatings can therefore be used
to protect less wear resistant and softer substrate surfaces for use in industry.
Karpuz et al. [68] reported that Ni–Co coatings electrodeposited from baths containing nickel
sulfamate, boric acid and cobalt sulfate have potential for application in magnetic sensors. These
magnetic properties of deposited Ni–Co coatings and their nanocomposites offers attractive potential
to serve as soft magnets for motors, power supplies and high-efficiency transformers [3].
Ni–Co hydroxide nanosheets have been identified as candidates for pseudocapacitor application
to meet the ever-growing demand for new energy storage devices. Pu J, et al. [139] researched on
Ni–Co layered double hydroxides (LDHs) nanosheets and reported that the nanosheets exhibited
excellent specific capacitance of 1734 F g−1 at 6 A g−1 . The nanosheets also exhibited better stability
with a capacitance retention of 86% in the galvanostatic charge–discharge test after 1000 cycles.
(i) Corrosion resistance behavior in varying environments, such as steady and dynamic conditions.
(ii) Tribological properties such as dry and wet abrasive behavior under controlled loads.
(iii) Electroless deposition of Ni–Co coatings, which offers a more competitive and specialized option.
(iv) Thermal oxidation resistance of Ni–Co alloy matrices.
(v) More efficient electrolyte agitation techniques such as submerged jet impingement and flow cells,
jet eductors, and ultrasound.
(vi) Extensive research on hydrogen evolution mitigation in electrodeposited Ni-based coatings by
using pulse electrodeposition and additives.
(vii) Use of response surface methodology to optimize the Ni–Co electrodeposition process and
increase accuracy of the desired properties and also predict tested properties.
Hydrogen evolution affects the coating structure of deposited coatings. Several approaches have
been used with different materials to great success. Kannan and Wallipa [114] coated a magnesium alloy
with calcium phosphate using constant-potential and pulse-potential methods and analyzed the in vitro
corrosion resistance properties. It was reported that the polarization resistance of pulse-potential
deposited coatings was three times higher than that exhibited by constant-potential deposited coatings,
and this was associated with the calcium phosphate particles being closely packed for pulse-potential
coatings. This provides an interesting approach that can be researched on using Ni and Ni–Co
based coatings.
Several other hypotheses have been postulated for the mitigation of hydrogen evolution in the
electrodeposition process in different coatings. In past research, an organic solvent (ethanol) was
added to the electrolyte bath to slow down hydrogen evolution on a magnesium alloy by decreasing
conductivity of the plating solution, resulting in decreased hydrogen bubble bursting rate and hence
a highly dense coating was deposited [142]. This approach of slowing down hydrogen evolution
by decreasing the conductivity, however, results in lower deposition rates. Metal deposition utilizes
the OH− ions generated during H2 O breakdown whereby the metallic ions are reduced to form the
metal. The mechanisms for hydrogen generation and metal-hydroxyl ion adsorption are shown in
Equations (14)–(18) [143,144].
2H2 O + 2e− = H2 + 2OH− (14)
2H + 2e− = H2 (15)
2+ − +
M + OH = M(OH) (16)
+ +
M(OH) → M(OH)ads (17)
+ − −
M(OH)ads + 2e = M + OH (18)
50
Materials 2020, 13, 3475
where M can be Ni or Co ions. As such, a balance must be struck between the rate of hydrogen evolution
and the deposition rate, and this offers an interesting area for research and application in Ni and Ni–Co
based electrodeposited coatings. Other additives used include polyethylene glycol and di-sodium
ethylenediamine tetraacetic acid (EDTA) in pulse copper deposition, and it has been reported that this
improves the throwing power, current efficiency, and thickness of deposited coatings [145]. These
additives can also be considered for electrodeposited Ni and Ni–Co based coatings.
It is advisable to use larger sample sizes because they hinder the manifestation of edge effects that
are common in smaller sample sizes. This is especially common where the current density used is high
with respect to the substrate’s surface area.
The adhesive force that exists between the substrate and the coatings plays a major role in the
wear resistance of the material. As such, ensuring a good bond exists between the deposit and
substrate surface is key. Copper electrodes tend to exhibit better adhesive properties compared to
their steel counterparts, but they are also more costly. Using a pre-treatment step offers the chance
to improve this bond, especially where the substrate is made of steel. From personal experience, a
triple immersion procedure of electronic cleaning can be used for better adhesion. This comprises
of degreasing using electro-hydrostatic fluid, then removal of oxide layer by passing the substrates
through a strong activating solution, and finally removal of carbon-black by passing the substrates
through a weak activating solution [49]. A similar pre-treatment process has been used in other
coatings like Ni–W nanocomposite coatings with good adhesion translating to superb wear resistance
achieved [146]. This pre-treatment process provides interesting possibilities for future use in Ni–Co
alloys and nanocomposites. De-ionized water should be used to clean the substrate surface after each
pre-treatment step.
Research shows that orientation of electrodes in the electrolyte plays a major role in the deposition
process. Results obtained from Ni–Co deposition suggest that the sediment deposition technique (SCD)
is more favorable compared to conventional deposition technique. Ni–Co alloy and nanocomposite
coatings deposited from the SCD technique have exhibited superior properties of higher Co content
and higher nanoparticle content which translate to better microhardness, and improved wear and
corrosion resistance of the deposited coatings. As such, selection of SCD in DC electrodeposition of
Ni–Co composites should be considered for further research. Based on current trends, it can be seen
that owing to their exceptional wear resistance and corrosion properties, deposited Ni–Co alloys and
their nanocomposites are strong contenders for further application in the aviation industry, for use in
jet engine fabrication, automotive engineering, textiles and general engineering. In recent years, a keen
interest has developed in specialized engineering where deposited coatings consist of mixed functional
properties, as well as deposition of superhydrophobic surface coatings which exhibit excellent wear
and corrosion resistance, better self-cleaning and good tribological properties.
In essence, nanoparticles can be selected to match the desired properties of any coating, and with
such capacity for discovery coupled with the ever-growing need of better material properties, the
possibilities for future applications are endless.
Author Contributions: Conceptualization, N.S.M. and M.K.; methodology, N.S.M., and Y.Z.; validation,
N.S.M. and M.K.; formal analysis, N.S.M. and L.Y.; investigation, N.S.M., Y.Z., and N.J.N.; resources, M.K.;
writing—original draft preparation, N.S.M.; G.V.B., and N.J.N.; writing—review and editing, N.S.M., Y.Z., and
M.K.; visualization, N.S.M.; G.V.B., and L.Y.; supervision, M.K.; project administration, M.K.; funding acquisition,
M.K. All authors have read and agreed to the published version of the manuscript.
Funding: This research was funded by the Technology development programmer for the Northern Jiangsu area,
grant number BN2014019.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design of the
study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to
publish the results.
51
Materials 2020, 13, 3475
References
1. Martin, P. Introduction to Surface Engineering and Functionally Engineered Materials; John Wiley & Sons:
New York, NY, USA, 2011.
2. Shriram, S.; Mohan, S.; Renganathan, N.G.; Venkatachalam, R. Electrodeposition of nanocrystalline nickel—A
brief review. Int. J. Surf. Eng. Coat. 2000, 78, 194–197. [CrossRef]
3. Gurrappa, I.; Binder, L. Electrodeposition of nanostructured coatings and their characterization—A review.
Sci. Technol. Adv. Mat. 2008, 9, 1–11. [CrossRef] [PubMed]
4. Qiao, G.; Jing, T.; Wang, N.; Gao, Y.; Zhao, X.; Zhou, J.; Wang, W. High-speed jet electrodeposition and
microstructure of nanocrystalline Ni–Co alloys. Electrochim. Acta 2005, 51, 85–92. [CrossRef]
5. Tury, B.; Lakatos-Varsányi, M.; Roy, S. Ni–Co alloys plated by pulse currents. Surf. Coat. Technol. 2006, 200,
6713–6717. [CrossRef]
6. Chang, L.; Guo, H.; An, M. Electrodeposition of Ni–Co/Al2 O3 composite coating by pulse reverse method
under ultrasonic condition. Mater. Lett. 2008, 62, 3313–3315. [CrossRef]
7. Wielage, B.; Lampke, T.; Zacher, M.; Dietrich, D. Electroplated nickel composites with micron-to nano-sized
particles. Key Eng. Mater. 2008, 384, 283–309. [CrossRef]
8. Schlesinger, M.; Paunovic, M. Modern Electroplating; John Wiley & Sons: New York, USA, 2011.
9. Prasad, M.; Chokshi, A. Superplasticity in electrodeposited nanocrystalline nickel. Acta Mater. 2010, 58,
5724–5736. [CrossRef]
10. Saraev, D.; Miller, R.E. Atomic-scale simulations of nanoindentation-induced plasticity in copper crystals
with nanometer-sized nickel coatings. Acta Mater. 2006, 54, 33–45. [CrossRef]
11. Kim, D.; Kim, M.K.; Son, J.T.; Kim, H.G. Effect of target properties on deposition of lithium nickel cobalt
oxide thin-films using RF magnetron sputtering. J. Power Sources. 2002, 108, 239–244. [CrossRef]
12. Bakhit, B.; Akbari, A. Nanocrystalline Ni–Co alloy coatings: Electrodeposition using horizontal electrodes
and corrosion resistance. J. Coat. Technol. Res. 2013, 10, 285–295. [CrossRef]
13. Rashkova, V.; Kitova, S.; Konstantinov, I.; Vitanov, T. Vacuum evaporated thin films of mixed cobalt and
nickel oxides as electrocatalyst for oxygen evolution and reduction. Electrochim. Acta 2002, 47, 1555–1560.
[CrossRef]
14. Koikeda, T.; Fujiwara, S.; Chikazumi, S. Perpendicular anisotropy of evaporated magnetic iron-nickel and
cobalt-nickel thin films. J. Phys. Soc. Jpn. 1966, 21, 1914–1921. [CrossRef]
15. Dharmadasa, I.; Haigh, J. Strengths and advantages of electrodeposition as a semiconductor growth technique
for applications in macroelectronic devices. J. Electrochem. Soc. 2006, 153, G47–G52. [CrossRef]
16. Srivastava, M.; Selvi, V.E.; Grips, V.K.W.; Rajam, K.S. Corrosion resistance and microstructure of
electrodeposited nickel–cobalt alloy coatings. Surf. Coat. Technol. 2006, 201, 3051–3060. [CrossRef]
17. Yang, X.; Li, Q.; Zhang, S.; Gao, H.; Luo, F.; Dai, Y. Electrochemical corrosion behaviors and corrosion
protection properties of Ni–Co alloy coating prepared on sintered NdFeB permanent magnet. J. Solid. State.
Electr. 2010, 14, 1601–1608. [CrossRef]
18. Ranjith, B.; Kalaignan, G.P. Ni–Co–TiO2 nanocomposite coating prepared by pulse and pulse reversal
methods using acetate bath. Appl. Surf. Sci. 2010, 257, 42–47. [CrossRef]
19. Wang, G.; Chan, K.; Zhang, K. Low temperature superplasticity of nanocrystalline electrodeposited Ni–Co
alloy. Scr. Mater. 2006, 54, 765–770. [CrossRef]
20. Wang, L.; Gao, Y.; Xue, Q.; Liu, H.; Xu, T. Microstructure and tribological properties of electrodeposited
Ni–Co alloy deposits. Appl. Surf. Sci. 2005, 242, 326–332. [CrossRef]
21. Hassani, S.; Raeissi, K.; Golozar, M. Effects of saccharin on the electrodeposition of Ni–Co nanocrystalline
coatings. J. Appl. Electrochem. 2008, 38, 689–694. [CrossRef]
22. Hansal, W.E.G.; Tury, B.; Halmdienst, M.; Varsányi, M.L.; Kautek, W. Pulse reverse plating of Ni–Co alloys:
Deposition kinetics of Watts, sulfamate and chloride electrolytes. Electrochim. Acta 2006, 52, 1145–1151.
[CrossRef]
23. Gogotsi, Y. Nanomaterials Handbook; CRC Press: Florida, FL, USA, 2006.
24. Wang, C.; Chan, K. Enhanced low-temperature superplasticity of Ni–Co alloy by addition of nano-Si3 Ni4
particles. Mat. Sci. Eng. A-Struct. 2008, 491, 266–269. [CrossRef]
25. Afshar, A.; Ghorbani, M.; Mazaheri, M. Electrodeposition of graphite-bronze composite coatings and study
of electroplating characteristics. Surf. Coat. Technol. 2004, 187, 293–299. [CrossRef]
52
Materials 2020, 13, 3475
26. Tian, B.; Cheng, Y. Electrolytic deposition of Ni–Co–Al2 O3 composite coating on pipe steel for corrosion/
erosion resistance in oil sand slurry. Electrochim. Acta 2007, 53, 511–517. [CrossRef]
27. Cai, F.; Jiang, C.; Fu, P.; Ji, V. Effects of Co contents on the microstructures and properties of electrodeposited
NiCo–Al composite coatings. Appl. Surf. Sci. 2015, 324, 482–489. [CrossRef]
28. Wang, L.; Gao, Y.; Liu, H.; Xue, Q.; Xu, T. Effects of bivalent Co ion on the co-deposition of nickel and
nano-diamond particles. Surf. Coat. Technol. 2005, 191, 1–6. [CrossRef]
29. Karimzadeh, A.; Aliofkhazraei, M.; Walsh, F.C. A review of electrodeposited Ni-Co alloy and composite
coatings: Microstructure, properties and applications. Surf. Coat. Technol. 2019, 372, 463–498. [CrossRef]
30. Bakhit, B.; Akbari, A. Effect of particle size and co-deposition technique on hardness and corrosion properties
of Ni–Co/SiC composite coatings. Surf. Coat. Technol. 2012, 206, 4964–4975. [CrossRef]
31. Borkar, T. Electrodeposition of Nickel Composite Coatings; Oklahoma State University: Oklahoma, OK, USA, 2010.
32. Yang, Y.; Cheng, Y.F. Fabrication of Ni–Co–SiC composite coatings by pulse electrodeposition—Effects of
duty cycle and pulse frequency. Surf. Coat. Technol. 2013, 216, 282–288. [CrossRef]
33. Qu, N.S.; Zhu, D.; Chan, K.C.; Lei, W.N. Pulse electrodeposition of nanocrystalline nickel using ultra narrow
pulse width and high peak current density. Surf. Coat. Technol. 2003, 168, 123–128. [CrossRef]
34. Li, Y.; Jiang, H.; Tian, H. Effects of peak current density on the mechanical properties of nanocrystalline
Ni–Co alloys produced by pulse electrodeposition. Appl. Surf. Sci. 2008, 254, 6865–6869. [CrossRef]
35. Dheeraj, P.R.; Patra, A.; Sengupta, S.; Das, S.; Das, K. Synergistic effect of peak current density and nature
of surfactant on microstructure, mechanical and electrochemical properties of pulsed electrodeposited
Ni-Co-SiC nanocomposites. J. Alloy. Compd. 2017, 729, 1093–1107. [CrossRef]
36. Padmanabhan, K. Grain boundary sliding controlled flow and its relevance to superplasticity in metals,
alloys, ceramics and intermetallics and strain-rate dependent flow in nanostructured materials. J. Mater. Sci.
2009, 44, 2226–2238. [CrossRef]
37. Chung, C.K.; Chang, W. Effect of pulse frequency and current density on anomalous composition and
nanomechanical property of electrodeposited Ni–Co films. Thin Solid Film. 2009, 517, 4800–4804. [CrossRef]
38. Gyftou, P.; Pavlatou, E.; Spyrellis, N. Effect of pulse electrodeposition parameters on the properties of
Ni/nano-SiC composites. Appl. Surf. Sci. 2008, 254, 5910–5916. [CrossRef]
39. Tury, B.; Lakatos-Varsányi, M.; Roy, S. Effect of pulse parameters on the passive layer formation on pulse
plated Ni–Co alloys. Appl. Surf. Sci. 2007, 253, 3103–3108. [CrossRef]
40. Chang, L.M.; An, M.Z.; Guo, H.F.; Shi, S.Y. Microstructure and properties of Ni–Co/nano-Al2 O3 composite
coatings by pulse reversal current electrodeposition. Appl. Surf. Sci. 2006, 253, 2132–2137. [CrossRef]
41. Karakus, C.; Chin, D.T. Metal distribution in jet plating. J. Electrochem. Soc. 1994, 141, 691. [CrossRef]
42. Wang, W.; Hou, F.Y.; Wang, H.; Guo, H.T. Fabrication and characterization of Ni–ZrO2 composite
nano-coatings by pulse electrodeposition. Scr. Mater. 2005, 53, 613–618. [CrossRef]
43. Chandrasekar, M.S.; Pushpavanam, M. Pulse and pulse reverse plating−Conceptual, advantages and
applications. Electrochim. Acta 2008, 53, 3313–3322. [CrossRef]
44. Podlaha, E.; Landolt, D. Pulse-Reverse Plating of Nanocomposite Thin Films. J. Electrochem. Soc. 1997, 144,
L200. [CrossRef]
45. Vidrine, A.; Podlaha, E. Composite Electrodeposition of Ultrafine γ-Alumina Particles in Nickel Matrices;
Part I: Citrate and chloride electrolytes. J. Appl. Electrochem. 2001, 31, 461–468. [CrossRef]
46. Xiong-Skiba, P.; Engelhaupt, D.; Hulguin, R.; Ramsey, B. Effect of pulse plating parameters on the composition
of alumina/nickel composite. J. Electrochem. Soc. 2005, 152, C571–C576. [CrossRef]
47. Sáez, V.; Gonzalez, V.; Iniesta, J.; Frías, A.; Aldaz, A. Electrodeposition of PbO2 on glassy carbon electrodes:
Influence of ultrasound frequency. Electrochem. Commun. 2004, 6, 757–761. [CrossRef]
48. Touyeras, F.; Hihn, J.Y.; Bourgoin, X.; Jacques, B.; Hallez, L.; Branger, V. Effects of ultrasonic irradiation on
the properties of coatings obtained by electroless plating and electro plating. Ultrason. Sonochem. 2005, 12,
13–19. [CrossRef] [PubMed]
49. Mbugua, N.S.; Kang, M.; Li, H.; Liu, Y.; Joseph, N.; Zhang, Y. The Influence of Co Concentration on the
Properties of Conventionally Electrodeposited Ni–Co–Al2 O3 –SiC Nanocomposite Coatings. Prot. Met. Phys.
Chem. 2020, 56, 94–102. [CrossRef]
50. Popov, K.; Grgur, B.; Djokić, S.S. Fundamental Aspects of Electrometallurgy; Kluwer Academic Publishers:
Moscow, Russia, 2007.
53
Materials 2020, 13, 3475
51. Wu, G.; Li, N.; Wang, D.L.; Zhou, D.R.; Xu, B.Q.; Mitsuo, K. Effect of α-Al2 O3 particles on the electrochemical
codeposition of Co–Ni alloys from sulfamate electrolytes. Mater. Chem. Phys. 2004, 87, 411–419. [CrossRef]
52. Yari, S.; Dehghanian, C. Deposition and characterization of nanocrystalline and amorphous Ni–W coatings
with embedded alumina nanoparticles. Ceram. Int. 2013, 39, 7759–7766. [CrossRef]
53. Prabu, S.; Wang, H.W. Factors Affecting the Electrodeposition of Aluminum Metal in an Aluminum
Chloride–Urea Electrolyte Solution. J. Chin. Chem. Soc-Taip. 2017, 64, 1467–1477. [CrossRef]
54. Idris, J.; Christian, C.; Gaius, E. Nanocrystalline Ni-Co alloy synthesis by high speed electrodeposition. J.
Nanomater. 2013, 2013, 1–8. [CrossRef]
55. Goto, Y.; Kamebuchi, Y.; Hagio, T.; Kamimoto, Y.; Ichino, R.; Bessho, T. Electrodeposition of copper/carbonous
nanomaterial composite coatings for heat-dissipation materials. Coatings 2018, 8, 5. [CrossRef]
56. Srivastava, M.; Grips, V.W.; Rajam, K. Influence of Co on Si3 N4 incorporation in electrodeposited Ni. J. Alloy.
Compd. 2009, 469, 362–365. [CrossRef]
57. Bakhit, B.; Akbari, A. Synthesis and characterization of Ni–Co/SiC nanocomposite coatings using sediment
co-deposition technique. J. Alloy. Compd. 2013, 560, 92–104. [CrossRef]
58. Bercot, P.; Pena-Munoz, E.; Pagetti, J. Electrolytic composite Ni–PTFE coatings: An adaptation of Guglielmi’s
model for the phenomena of incorporation. Surf. Coat. Technol. 2002, 157, 282–289. [CrossRef]
59. Lozano-Morales, A.; Podlaha, E. The Effect of Al2 O3 Nanopowder on Cu Electrodeposition. J. Electrochem.
Soc. 2004, 151, C478–C483. [CrossRef]
60. Shi, L.; Sun, C.; Gao, P.; Zhou, F.; Liu, W. Mechanical properties and wear and corrosion resistance of
electrodeposited Ni–Co/SiC nanocomposite coating. Appl. Surf. Sci. 2006, 252, 3591–3599. [CrossRef]
61. Tian, L.; Xu, J.; Xiao, S. The influence of pH and bath composition on the properties of Ni–Co coatings
synthesized by electrodeposition. Vacuum. 2011, 86, 27–33. [CrossRef]
62. Gomez, E.; Pane, S.; Valles, E. Electrodeposition of Co–Ni and Co–Ni–Cu systems in sulphate–citrate medium.
Electrochim. Acta 2005, 51, 146–153. [CrossRef]
63. Oriňáková, R.; Orinak, A.; Vering, G.; Talian, I.; Smith, M.R.; Arlinghaus, H. Influence of pH on the electrolytic
depositon of Ni–Co Film. Thin Solid Film. 2008, 516, 3045–3050. [CrossRef]
64. Ma, C.; Wang, S.C.; Low, C.T.J.; Wang, L.P.; Walsh, F.C. Effects of additives on microstructure and properties
of electrodeposited nanocrystalline Ni–Co alloy coatings of high cobalt content. Int. J. Surf. Eng. Coat. 2014,
92, 189–195. [CrossRef]
65. Puippe, J.C.; Leaman, F. Theory and Practice of Pulse Plating; Amer Electroplaters Soc.: Orlando, FL, USA, 1986.
66. Lupi, C.; Dell’Era, A.; Pasquali, M.; Imperatori, P. Composition, morphology, structural aspects and
electrochemical properties of Ni–Co alloy coatings. Surf. Coat. Technol. 2011, 205, 5394–5399. [CrossRef]
67. Fan, C.; Piron, D. Study of anomalous nickel-cobalt electrodeposition with different electrolytes and current
densities. Electrochim. Acta 1996, 41, 1713–1719. [CrossRef]
68. Karpuz, A.; Kockar, H.; Alper, M.; Karaagac, O.; Haciismailoglu, M. Electrodeposited Ni–Co films from
electrolytes with different Co contents. Appl. Surf. Sci. 2012, 258, 4005–4010. [CrossRef]
69. Celis, J.P.; Roos, J. Kinetics of the deposition of alumina particles from copper sulfate plating baths. J.
Electrochem. Soc. 1977, 124, 1508–1511. [CrossRef]
70. Ahmad, Y.H.; Mohamed, A. Electrodeposition of nanostructured nickel-ceramic composite coatings: A
review. Int. J. Electrochem. Sci. 2014, 9, 1942–1963.
71. Gomez, E.; Pane, S.; Alcobe, X.; Vallés, E. Influence of a cationic surfactant in the properties of cobalt–nickel
electrodeposits. Electrochim. Acta 2006, 51, 5703–5709. [CrossRef]
72. Di Bari, G.A. Electrodeposition of nickel. In Modern Electroplating, 5th ed.; Schlesinger, M., Paunovic, M.,
Eds.; John Wiley & Sons, Inc.: Hoboken, NJ, USA, 2000; Volume 3, pp. 79–114.
73. Gezerman, A.O.; Corbacioglu, B.D. Analysis of the characteristics of nickel-plating baths. Int. J. Chem. 2010,
2, 124–137. [CrossRef]
74. Ispas, A.; Matsushima, H.; Bund, A.; Bozzini, B. A study of external magnetic-field effects on nickel–iron alloy
electrodeposition, based on linear and non-linear differential AC electrochemical response measurements.
Electroanal. Chem. 2011, 651, 197–203. [CrossRef]
75. Ramazani, A.; Asgari, V.; Montazer, A.H.; Kashi, M.A. Tuning magnetic fingerprints of FeNi nanowire arrays
by varying length and diameter. Curr. Appl. Phys. 2015, 15, 819–828. [CrossRef]
76. Tsuru, Y.; Nomura, M.; Foulkes, F. Effects of boric acid on hydrogen evolution and internal stress in films
deposited from a nickel sulfamate bath. J. Appl. Electrochem. 2002, 32, 629–634. [CrossRef]
54
Materials 2020, 13, 3475
77. Gadad, S.; Harris, T.M. Oxygen Incorporation during the Electrodeposition of Ni, Fe, and Ni-Fe Alloys. J.
Electrochem. Soc. 1998, 145, 3699. [CrossRef]
78. Karwas, C.; Hepel, T. Influence of Boric Acid on Electrodeposition and Stripping of Ni-Zn Alloys. J.
Electrochem. Soc. 1988, 135, 839. [CrossRef]
79. Saubestre, E.B. The Chemistry of Watts Nickel Plating Solutions. Plating 1958, 45, 927–936.
80. Tilak, B.; Gendron, A.; Mosoiu, M. Borate buffer equilibria in nickel refining electrolytes. J. Appl. Electrochem.
1977, 7, 495–500. [CrossRef]
81. Šupicová, M.; Rozik, R.; Tmkova, L.; Oriňáková, R.; Gálová, M. Influence of boric acid on the electrochemical
deposition of Ni. J. Solid State Electr. 2006, 10, 61–68. [CrossRef]
82. Yin, K.M.; Lin, B.T. Effects of boric acid on the electrodeposition of iron, nickel and iron-nickel. Surf. Coat.
Technol. 1996, 78, 205–210. [CrossRef]
83. Abyaneh, M.; Hashemi-Pour, M. The effect of the concentration of boric acid on the kinetics of
electrocrystallization of nickel. Int. J. Surf. Eng. Coat. 1994, 72, 23–26. [CrossRef]
84. Narasimman, P.; Pushpavanama, M.; Periasamyb, V. Effect of surfactants on the electrodeposition of Ni-SiC
composites. Port. Electrochim. Acta 2012, 30, 1–14. [CrossRef]
85. Li, Q.; Fu, W.; Mu, Y.; Zhang, W.; Lv, P.; Zhou, L.; Yang, H.; Chi, K.; Yang, L. The effects of CTAB concentration
on the properties of electrodeposited cadmium telluride films. CrystEngComm 2014, 16, 5227–5233. [CrossRef]
86. Kılıc, F.; Gul, H.; Aslan, S.; Alp, A.; Akbulut, H. Effect of CTAB concentration in the electrolyte on the
tribological properties of nanoparticle SiC reinforced Ni metal matrix composite (MMC) coatings produced
by electrodeposition. Colloid Surf. A. 2013, 419, 53–60. [CrossRef]
87. Sabri, M.; Sarabi, A.A.; Kondelo, S.M.N. The effect of sodium dodecyl sulfate surfactant on the
electrodeposition of Ni-alumina composite coatings. Mater. Chem. Phys. 2012, 136, 566–569. [CrossRef]
88. Baghal, S.M.L.; Amadeh, A.; Sohi, M.H.; Hadavi, S.M.M. The effect of SDS surfactant on tensile properties of
electrodeposited Ni–Co/SiC nanocomposites. Mat. Sci. Eng. A-Struct. 2013, 559, 583–590. [CrossRef]
89. Sen, R.; Bhattacharya, S.; Das, S.; Das, K. Effect of surfactant on the co-electrodeposition of the nano-sized
ceria particle in the nickel matrix. J. Alloy. Compd. 2010, 489, 650–658. [CrossRef]
90. Gamburg, Y.D.; Zangari, G. Theory and Practice of Metal Electrodeposition; Springer Science & Business Media:
Berlin, Germany, 2011.
91. Guo, C.; Zuo, Y.; Zhao, X.; Zhao, J.; Xiong, J. Effects of surfactants on electrodeposition of nickel-carbon
nanotubes composite coatings. Surf. Coat. Technol. 2008, 202, 3385–3390. [CrossRef]
92. Pradhan, A.K.; Das, S. Pulse reverse electrodeposition of Cu-SiC nanocomposite coating: Effects of surfactants
and deposition parameters. Met. Mater. Trans. A 2014, 45, 5708–5720. [CrossRef]
93. Ger, M.D. Electrochemical deposition of nickel/SiC composites in the presence of surfactants. Mater. Chem.
Phys. 2004, 87, 67–74. [CrossRef]
94. Li, Y.; Jiang, H.; Wang, D.; Ge, H. Effects of saccharin and cobalt concentration in electrolytic solution on
microhardness of nanocrystalline Ni–Co alloys. Surf. Coat. Technol. 2008, 202, 4952–4956. [CrossRef]
95. Weil, R.; Cook, H. Electron-Microscopic Observations of the Structure of Electroplated Nickel. J. Electrochem.
Soc. 1962, 109, 295. [CrossRef]
96. Wang, G.; Jiang, S.; Zhen, L.U.; Zhang, K. Preparation and tensile properties of Al2 O3 /Ni-Co nanocomposites.
T Nonferr. Met. Soc. 2011, 21, s374–s379. [CrossRef]
97. Koch, C. Optimization of strength and ductility in nanocrystalline and ultrafine grained metals. Scr. Mater.
2003, 49, 657–662. [CrossRef]
98. Yin, W.M.; Whang, S.H.; Mirshams, R. Effect of interstitials on tensile strength and creep in nanostructured
Ni. Acta Mater. 2005, 53, 383–392. [CrossRef]
99. Hadian, S.; Gabe, D. Residual stresses in electrodeposits of nickel and nickel–iron alloys. Surf. Coat. Technol.
1999, 122, 118–135. [CrossRef]
100. Dzedzina, R.; Hagarova, M. Effect of Additive on the Internal Stress in Galvanic Coatings. Int. J. Electrochem.
Sci. 2013, 8, 8291–8298.
101. Kubart, T.; Mala, Z.; Novak, R.; Novakova, D. Effect of coated edge geometry on internal stress distribution
in multilayered coatings. Surf. Coat. Technol. 2001, 142, 610–614. [CrossRef]
102. El-Sherik, A.; Shirokoff, J.; Erb, U. Stress measurements in nanocrystalline Ni electrodeposits. J. Alloy. Compd.
2005, 389, 140–143. [CrossRef]
55
Materials 2020, 13, 3475
103. Chen, C.J.; Lin, K.L. Internal stress and adhesion of amorphous Ni–Cu–P alloy on aluminum. Thin Solid Film.
2000, 370, 106–113. [CrossRef]
104. Bai, A.; Hu, C.C. Effects of electroplating variables on the composition and morphology of nickel–cobalt
deposits plated through means of cyclic voltammetry. Electrochim. Acta 2002, 47, 3447–3456. [CrossRef]
105. Go, E.; Ramirez, J.; Valle, E. Electrodeposition of Co–Ni alloys. J. Appl. Electrochem. 1998, 28, 71–79.
106. Rafailović, L.D.; Karnthaler, H.P.; Trisovic, T.; Minić, D.M. Microstructure and mechanical properties of
disperse Ni–Co alloys electrodeposited on Cu substrates. Mater. Chem. Phys. 2010, 120, 409–416. [CrossRef]
107. Barbir, F. Transition to renewable energy systems with hydrogen as an energy carrier. Energy 2009, 34,
308–312. [CrossRef]
108. Veziro, T.N.; Barbir, L.F. Hydrogen: The wonder fuel. Int. J. Hydrog. Energy 1992, 17, 391–404. [CrossRef]
109. Bockris, J.N.; Veziroǧlu, T.M. A solar-hydrogen economy for USA. Int. J. Hydrog. Energy 1983, 8, 323–340.
[CrossRef]
110. Mazloomi, K.; Gomes, C. Hydrogen as an energy carrier: Prospects and challenges. Renew. Sust. Energy Rev.
2012, 16, 3024–3033. [CrossRef]
111. Ganley, J.C. High temperature and pressure alkaline electrolysis. Int. J. Hydrog. Energy 2009, 34, 3604–3611.
[CrossRef]
112. Vijayakumar, J.; Mohan, S.; Kumar, S.A.; Suseendiran, S.R.; Pavithra, S. Electrodeposition of Ni–Co–Sn alloy
from choline chloride-based deep eutectic solvent and characterization as cathode for hydrogen evolution in
alkaline solution. Int. J. Hydrog. Energy 2013, 38, 10208–10214. [CrossRef]
113. Lasia, A.; Rami, A. Kinetics of hydrogen evolution on nickel electrodes. J. Electroanal Chem. 1990, 294,
123–141. [CrossRef]
114. Kannan, M.B.; Wallipa, O. Potentiostatic pulse-deposition of calcium phosphate on magnesium alloy for
temporary implant applications—An in vitro corrosion study. Mat. Sci. Eng. C 2013, 33, 675–679. [CrossRef]
115. Baghal, S.M.L.; Sohi, M.H.; Amadeh, A. A functionally gradient nano-Ni–Co/SiC composite coating on
aluminum and its tribological properties. Surf. Coat. Technol. 2012, 206, 4032–4039. [CrossRef]
116. Dieter, P.P. Mechanical Metallurgy; McGraw-Hill Book Co.: New York, NY, USA, 1986.
117. Jeong, D.H.; Gonzalez, F.; Palumbo, G.; Aust, K.T.; Erb, U. The effect of grain size on the wear properties of
electrodeposited nanocrystalline nickel coatings. Scripta. Mater. 2001, 44, 493–499. [CrossRef]
118. Kim, D.; Park, D.Y.; Yoo, B.Y.; Sumodjo, P.T.A.; Myung, N.V. Magnetic properties of nanocrystalline iron
group thin film alloys electrodeposited from sulfate and chloride baths. Electrochim. Acta 2003, 48, 819–830.
[CrossRef]
119. Kusakabe, S.; Rawls, H.R.; Hotta, M. Relationship between thin-film bond strength as measured by a scratch
test, and indentation hardness for bonding agents. Dent. Mater. 2016, 32, e55–e62. [CrossRef]
120. Jiang, W.; Shen, L.; Xu, M.; Wang, Z.; Tian, Z. Mechanical properties and corrosion resistance of Ni-Co-SiC
composite coatings by magnetic field-induced jet electrodeposition. J. Alloy. Compd. 2019, 791, 847–855.
[CrossRef]
121. Chen, X.H.; Chen, C.S.; Xiao, H.N.; Cheng, F.Q.; Zhang, G.; Yi, J.G. Corrosion behavior of carbon nanotubes–Ni
composite coating. Surf. Coat. Technol. 2005, 191, 351–356. [CrossRef]
122. Rogal, Ł.; Kalita, D.; Tarasek, A.; Bobrowski, P.; Czerwinski, F. Effect of SiC nano-particles on microstructure
and mechanical properties of the CoCrFeMnNi high entropy alloy. J. Alloy. Compd. 2017, 708, 344–352.
[CrossRef]
123. Jasim, K.M.; Rawlings, R.D.; West, D.R.F. Metal-ceramic functionally gradient material produced by laser
processing. J. Mater. Sci. 1993, 28, 2820–2826. [CrossRef]
124. Cramer, S.D.; Covino, B.S. ASM Handbook; ASM international Materials Park: Ohio, OH, USA, 2003.
125. Edward, J. Coating and Surface Treatment Systems for Metals: A Comprehensive Guide to Selection; ASM
International: Michigan, MI, USA, 1997.
126. Bakhit, B.; Akbari, A.; Nasirpouri, F.; Hosseini, M.G. Corrosion resistance of Ni–Co alloy and Ni–Co/SiC
nanocomposite coatings electrodeposited by sediment codeposition technique. Appl. Surf. Sci. 2014, 307,
351–359. [CrossRef]
127. Ramanauskas, R.; Quintana, P.; Maldonado, L.; Pomés, R.; Pech, M.A. Corrosion resistance and microstructure
of electrodeposited Zn and Zn alloy coatings. Surf. Coat. Technol. 1997, 92, 16–21. [CrossRef]
56
Materials 2020, 13, 3475
128. García, I.; Conde, A.; Langelaan, G.; Fransaer, J.; Celis, J.P. Improved corrosion resistance through
microstructural modifications induced by codepositing SiC-particles with electrolytic nickel. Corros. Sci.
2003, 45, 1173–1189. [CrossRef]
129. Dieter, G.E.; Bacon, D. Mechanical Metallurgy; McGraw Hill: New York, NY, USA, 1986.
130. Pavlatou, E.A.; Stroumbouli, P.; Gyftou, P.; Spyrellis, N. Hardening effect induced by incorporation of SiC
particles in nickel electrodeposits. J. Appl. Electrochem. 2006, 36, 385–394. [CrossRef]
131. Zimmerman, A.F.; Palumbo, G.; Aust, K.T.; Erb, U. Mechanical properties of nickel silicon carbide
nanocomposites. Mat. Sci. Eng. A-Struct. 2002, 328, 137–146. [CrossRef]
132. Zhou, Y.B.; Ding, Y.Z. Oxidation resistance of co-deposited Ni-SiC nanocomposite coating. Trans. Nonferr.
Met. Soc. 2007, 17, 925–928. [CrossRef]
133. Shi, L.; Sun, C.; Liu, W. Electrodeposited nickel–cobalt composite coating containing MoS2. Appl. Surf. Sci.
2008, 254, 6880–6885. [CrossRef]
134. Bhatnagar, M.; Baliga, B.J. Comparison of 6H-SiC, 3C-SiC, and Si for power devices. IEEE Trans. Electron.
Dev. 1993, 40, 645–655. [CrossRef]
135. Bakhit, B. The influence of electrolyte composition on the properties of Ni–Co alloy coatings reinforced by
SiC nano-particles. Surf. Coat. Technol. 2015, 275, 324–331. [CrossRef]
136. Yang, G.; Yin, L.; Fang, X.; Fang, M.; Liu, Y.; Huang, Z.; Liu, B. Fabrication and liquid–solid, two-phase
erosion wear behaviour of β-Sialon ceramic from pyrophyllite by carbothermal reduction and nitridation.
Ceram. Int. 2014, 40, 10737–10741. [CrossRef]
137. Basu, B.; Vleugels, J.; Van Der Biest, O. Microstructure–toughness–wear relationship of tetragonal zirconia
ceramics. J. Eur. Ceram. Soc. 2004, 24, 2031–2040. [CrossRef]
138. Beiyue, M.; Jingkun, Y. Phase composition of SiC-ZrO2 composite materials synthesized from zircon doped
with La2 O3 . J. Rare Earths 2009, 27, 806–810.
139. Pu, J.; Tong, Y.; Wang, S.; Sheng, E.; Wang, Z. Nickel–cobalt hydroxide nanosheets arrays on Ni foam for
pseudocapacitor applications. J. Power Sources 2014, 250, 250–256. [CrossRef]
140. Ma, C.; Wang, S.C.; Walsh, F.C. Electrodeposition of nanocrystalline nickel and cobalt coatings. Int. J. Surf.
Eng. Coat. 2015, 93, 8–17. [CrossRef]
141. Walsh, F.C.; Wang, S.; Zhou, N. The electrodeposition of composite coatings: Diversity, applications and
challenges. Curr. Opin. Electrochem. 2020, 20, 8–19. [CrossRef]
142. Kannan, M.B. Improving the packing density of calcium phosphate coating on a magnesium alloy for
enhanced degradation resistance. J. Biomed. Mater. Res. Part. A. 2013, 101, 1248–1254. [CrossRef]
143. Tury, B.; Radnoczi, G.Z.; Radnoczi, G.; Varsányi, M.L. Microstructure properties of pulse plated Ni–Co alloy.
Surf. Coat. Technol. 2007, 202, 331–335. [CrossRef]
144. Darband, G.B.; Aliofkhazraei, M.; Rouhaghdam, A.S.; Kiani, M.A. Three-dimensional Ni-Co alloy hierarchical
nanostructure as efficient non-noble-metal electrocatalyst for hydrogen evolution reaction. Appl. Surf. Sci.
2019, 465, 846–862. [CrossRef]
145. Mohan, S.; Raj, V. The effect of additives on the pulsed electrodeposition of copper. Trans. IMF 2005, 83,
194–198. [CrossRef]
146. Nyambura, S.M.; Kang, M.; Zhu, J.; Liu, Y.; Zhang, Y.; Ndiithi, N.J. Synthesis and Characterization of
Ni–W/Cr2 O3 Nanocomposite Coatings Using Electrochemical Deposition Technique. Coatings 2019, 9, 815.
[CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
57
materials
Review
Aqueous Corrosion of Aluminum-Transition Metal Alloys
Composed of Structurally Complex Phases: A Review
Libor Ďuriška, Ivona Černičková, Pavol Priputen and Marián Palcut *
Abstract: Complex metallic alloys (CMAs) are materials composed of structurally complex intermetal-
lic phases (SCIPs). The SCIPs consist of large unit cells containing hundreds or even thousands of
atoms. Well-defined atomic clusters are found in their structure, typically of icosahedral point group
symmetry. In SCIPs, a long-range order is observed. Aluminum-based CMAs contain approximately
70 at.% Al. In this paper, the corrosion behavior of bulk Al-based CMAs is reviewed. The Al–TM
alloys (TM = transition metal) have been sorted according to their chemical composition. The alloys
tend to passivate because of high Al concentration. The Al–Cr alloys, for example, can form protective
passive layers of considerable thickness in different electrolytes. In halide-containing solutions, how-
ever, the alloys are prone to pitting corrosion. The electrochemical activity of aluminum-transition
metal SCIPs is primarily determined by electrode potential of the alloying element(s). Galvanic
microcells form between different SCIPs which may further accelerate the localized corrosion attack.
The electrochemical nobility of individual SCIPs increases with increasing concentration of noble
elements. The SCIPs with electrochemically active elements tend to dissolve in contact with nobler
Citation: Ďuriška, L.; Černičková, I.; particles. The SCIPs with noble metals are prone to selective de-alloying (de–aluminification) and
Priputen, P.; Palcut, M. Aqueous their electrochemical activity may change over time as a result of de-alloying. The metal composition
Corrosion of Aluminum-Transition
of the SCIPs has a primary influence on their corrosion properties. The structural complexity is
Metal Alloys Composed of
secondary and becomes important when phases with similar chemical composition, but different
Structurally Complex Phases: A
crystal structure, come into close physical contact.
Review. Materials 2021, 14, 5418.
https://doi.org/10.3390/ma14185418
Keywords: aluminum; transition metal; corrosion; quasicrystal; approximant
Academic Editor: Daniel de la Fuente
59
alloy by Shechtman et al. [4]. The discovery of quasicrystals was awarded by Nobel prize
in Chemistry in 2011.
In 1982, D. Shechtman discovered a quasiperiodic arrangement of atoms in an Al–
Mn alloy showing an icosahedral symmetry but no unit cell [4]. Many such compounds
correspond to stable or metastable states in various phase diagrams [5]. The icosahe-
dral phase, as the first quasicrystalline structure discovered in the Al–Mn system, has
a long-distance arrangement without translational symmetry [4–7]. Many elements in
thermodynamically stable QCs observed yet belong to alkali, alkali-earth, transition, or
rare-earth metals (Figure 1). Typical examples of QCs and quasicrystalline approximants
are Al–TM alloys, where the TM is formed by one or more transition metals (TM = Cu,
Co, Ni, Fe, etc.) [9]. Thus, the quasicrystalline materials are often low-cost materials that
are easy to produce in large amounts [5–7,9]. Examples of quasicrystalline phases are
summarized in Table 1 [7,8,10–24].
Figure 1. Chemical elements forming thermodynamically stable QCs. Main forming elements (Al, Ti, Zn, Cd) are marked
by orange color, alloying elements are denoted by green.
60
Materials 2021, 14, 5418
61
Materials 2021, 14, 5418
sits inside an icosidodecahedron with 30 atoms. The fifth shell is a rhombic triacontahedron
(32 atoms + 60 atoms in its edges) [46].
Figure 2. Models of Mackay (a), Bergman (b) and Tsai (c) clusters, reproduced from reference [46].
The sequence (1) has features of order, but it is not periodically ordered. There are
sections SL and SLL, which alternate but do not repeat with regular periodicity. In fact, there
is no periodically repeating segment, a so-called unit cell, in the Fibonacci sequence [6,47].
Based on the Fibonacci sequence, it is possible to draw a one-dimensional quasicrystal
graphically (Figure 3). In Figure 3a, letters S and L correspond to shorter and longer
segments used in the Fibonacci sequence, respectively. Figure 3b shows the arrangement
of S and L segments according to the Fibonacci sequence. If the points dividing the line
into segments are atoms, a lattice with the quasicrystalline arrangement in one dimension
can be obtained. By adding the second and third dimensions, a simple example of a
quasicrystalline lattice can be drawn (Figure 3c).
62
Materials 2021, 14, 5418
Figure 4. Schematic structure of quasicrystal (a) and quasicrystalline approximant (b) with denoted
star-shape corresponding to clusters in real structure. Quasicrystalline arrangement is represented by
Penrose tiling.
63
Materials 2021, 14, 5418
Figure 5. Comparison of electron diffraction patterns in [010] zone axis: decagonal quasicrystal in
Al–Co-Cu system (a), ε16 decagonal quasicrystalline approximant in Al–Pd–Co system (b).
The surface structure of SCIPs is significantly less understood compared to bulk [60,61].
Preliminary results show that the adsorption of small, covalently bonding molecules on
icosahedral quasicrystals is very similar to that of pure Al substrate. This is consistent with
other studies, which indicate that the surface termination of most SCIPs is Al-rich [60].
A scanning tunneling microscopy (STM) has been utilized to study the surfaces of Al–
Pd–Mn quasicrystals [62–64]. The STM permits a visualization of the local atomistic
surface structure. Specific planes of the bulk structure have been observed as surface
terminations [63]. The termination planes are characterized by high atomic density and
include elements with the lowest surface energy. Nevertheless, the interpretation of
individual STM images is challenging and often needs to be accompanied by theoretical
models of the surface [61]. Therefore, ab initio density functional theory (DFT) calculations
have been utilized to model quasicrystalline surfaces [65,66]. To perform the calculations,
a Vienna ab initio simulation package (VASP) has been used [66]. The atomic structure
model of the five-fold Al–Pd–Mn surface is derived from the icosahedral approximant
model. In the model, the surface was cut perpendicular to one of its pseudo-five-fold
axes. The cleavage position was selected to create high density surface layers consistent
with experimental findings. The resulting surface structure is characterized by Penrose
tiling [65]. Most tiling vertices coincide with the center of Bergman clusters.
64
Materials 2021, 14, 5418
are released by metal are subsequently consumed by either dissolved oxygen or hydrogen
cations in the electrolyte. The reduction reaction takes place at cathode (cathodic reaction
site). The relative sizes and locations of cathodic and anodic sites are important variables
influencing the overall corrosion rate. The sizes of cathodic and anodic areas may vary
greatly; from atomic scales to macroscopically large dimensions.
Standard electrode potentials of metals are compared in Table 2. Since the Gibbs energy
is related to electrode potential (Equation (3)), the tendency of a metal to corrode in given
environment may be evaluated using E–pH plots. The diagrams have been calculated for
most metals by Pourbaix and are available in ref. [69]. Figure 7 displays the E–pH diagram
for Al–H2 O system [69,70]. The plot indicates the stability regions of different phases
in aqueous solutions. The E–pH diagram shows four different regions where metallic
aluminum, aluminum cations (Al3+ ), aluminum hydroxide and complex anion [Al(OH)4 ]−
are stable. The region, where the metallic Al is stable, is labelled as immunity region. The
areas with aluminum cations and anions as stable species are marked as corrosive regions.
In these areas the corrosion occurs. The passivity region is where the solid hydroxide exists.
In this region, Al is protected by a passive layer. The E–pH diagram demonstrates that
corrosion takes place in both alkaline and acidic environments. The protective layer is
formed at pH 4–9 [71]. The diagram also shows that the equilibrium electrode potential
between [Al(OH)4 ]− and Al, shifts to less noble values with increasing pH.
65
Materials 2021, 14, 5418
Table 2. Standard potentials, E0 , for metal electrodes, compiled from reference [68].
Although E–pH plots are useful in determining the metal’s tendency to corrode in
the given environment, they do not provide a kinetic information. The rate of corrosion
therefore needs to be determined separately by experimental methods. Corrosion rates
are obtained either by weight loss measurements or electrochemical methods [67,68]. The
weight loss measurement is a simple experiment to determine corrosion rates. In the
experiment, a clean weighed piece of material with well-defined dimensions is exposed to
the corrosive environment for a sufficient period. The corrosion rate (vcorr ) is then calculated
based on the recorded weight loss according to the following equation [19,20]
Δw
vcorr = (5)
St
In this equation, Δw is the weight loss, t is the reaction time and S is the exposed
surface area.
Weight loss measurements, although useful, can be time-consuming and may not
provide a complete information about reaction mechanism. Electrochemical techniques are
therefore widely used to study the corrosion mechanisms of metals in different electrolytes.
A potentiodynamic polarization is an electrochemical technique where the progress of
reaction is controlled by potentiostat [67,68,70]. It brings in a variety of parameters and pro-
vides valuable information about reaction mechanism. In the experiment, three electrodes
are assembled in a corrosion cell [72]. The corrosion cell includes a working electrode
(sample), counter (auxiliary) electrode and reference electrode. During the experiment, the
potential of the working electrode is systematically varied with respect to reference elec-
trode. The resulting current is measured by counter electrode. The potential of reference
electrode is constant and serves as reference value. Silver chloride (Ag/AgCl) and calomel
electrodes (Hg/Hg2 Cl2 ) immersed in a saturated KCl solution are most frequently used
66
Materials 2021, 14, 5418
reference electrodes. Platinum mesh is used as counter electrode as this metal is corrosion
resistant in most environments.
A schematic potential versus current density curve recorded during the polarization
experiment is given in Figure 8. Several different regions can be distinguished on the
curve. The first region is immune region. In this region, observed at low potentials,
the metal is thermodynamically stable. In immune region, cathodic reactions prevail at
the metal surface. The second region is labelled as active region. It is observed once
a corrosion potential, Ecorr , has been reached. In the active region, the metal actively
corrodes according to Equation (2). The active corrosion means that the anodic dissolution
of the metal takes place in the studied solution. Some metals can passivate. Therefore,
a passivation region can be also observed on the polarization curve. The passive region
corresponds to passive layer formation on the metal surface. The passivation is reflected
by rapid current density increase or stabilization at potentials higher than Ep (passivation
potential) on the polarization curve. At very high potentials, the current may start to
abruptly increase. The increase is a result of passive film breakdown and happens at
potential higher than transpassive potential, Etr . The passive film breakdown may be
initiated by aggressive halide anions and lead to localized corrosion (pitting). A given alloy
system may contain either some or all regions shown in Figure 8a.
Figure 8. Schematic polarization curve of a passivating metal: (a) full curve, (b) Tafel extrapolation of cathodic and
anodic regions.
A
Δw = Icorr t (6)
zF
In this equation, A is the atomic weight of the metal, Icorr is the corrosion current, t is
the reaction time, F is Faraday’s constant and z is the number of electrons involved in the
electrochemical reaction. The corrosion rate, vcorr , is calculated from the corrosion current
as
Δw A
vcorr = = jcorr (7)
St zF
In this equation S is the sample surface area and jcorr is the corrosion current density
(jcorr = Icorr /S). Equation (6) is valid for pure metals. For alloys, however, an equivalent
67
Materials 2021, 14, 5418
weight, Ew must be introduced to account for different molar masses of constituent metals
and different valence states. The following equation defines the equivalent weight of an
alloy [73]
1
Ew = z f (8)
∑ Ai ii
In the equation, zi is the valence state, fi is the weight fraction and Ai is the atomic
weight of metal i in the alloy. The corrosion rate than becomes
Ew
vcorr = jcorr (9)
F
The assignment of valence states for TMs is often ambiguous as these elements have
multiple stable valences. An independent experimental technique is therefore required, in
addition to corrosion experiments, to establish the proper valence state. Another approach
is to consult equilibrium Pourbaix diagrams [69]. The equilibrium E–pH diagrams can
be used estimate the stable valence state of TM at the experimental conditions (electrode
potential and pH of the electrolyte during corrosion test).
Metals become anodic and corrode only if their equilibrium half-cell potentials are
smaller than the half-cell potential of the corresponding cathodic reaction [67,68]. When
metals are combined into alloys it is no longer possible to define a unique half-cell potential.
In multiphase alloys, different phases may act as local anodes and cathodes. The physical
condition of the material may also be important. Constitutional variables such as the type
and amount of structural defects (dislocations, grain boundaries) and crystal orientation
are also important factors influencing the overall corrosion behavior.
Aluminum has a low standard electrode potential (Table 2, [68]). Therefore, aluminum
and aluminum alloys are prone to corrosion. Nevertheless, the materials are also easily
passivated. The passivation is related to spontaneous aluminum oxide/hydroxide film
formation at the interface [74]. The passive film protects the material and impedes further
reaction with the environment. Oxide layers grown on aluminum alloys at ambient
temperatures are generally non-crystalline, although short-range cubic ordered structure
has also been observed. In humid atmospheres, hydroxyl-oxides such as AlOOH or
Al(OH)3 may also form on aluminum surface. The passive film is generally self-renewing
and self-healing. Therefore, an accidental loss of the passive film due to, for example,
abrasion is rapidly restored.
Aluminum and its alloys are prone to pitting corrosion [75,76]. This type of local
corrosion is often observed in seawater as it is initiated by chlorides and other halide
anions in the electrolyte. The process may lead to passivity breakdown. Secondary phase
particles are important constitutional variables affecting the corrosion rate. They can be
classified into three different groups based on their electrochemical potential [76]: particles
with active elements, noble elements, and particles with both active and noble elements.
Reactive particles with active metals (such as Li, Mg or Zn) have low electrode potentials.
These particles behave as anodes and subsequently dissolve when embedded in aluminum
matrix. Particles with more noble elements (such as Fe or Cu) have higher electrode
potentials and constitute local cathodes. They initiate anodic dissolution of Al matrix. The
matrix adjacent to local cathode is preferentially attacked due to galvanic microcell created
at the matrix/particle interface (Figure 9, [76]).
68
Materials 2021, 14, 5418
Figure 9. Schematic of the de-alloying and subsequent trenching of Al2 CuMg intermetallic in AA2024 aluminum alloy (a);
microstructure image of corroded alloy surface after 1-h exposure in aerated H2 O at 30 ◦ C (b), redrawn (a) and reproduced
(b) from ref. [76].
If intermetallic particles contain both noble and active elements, their electrochemical
behavior changes over time. The active elements may preferentially dissolve, leaving
behind the noble metals. This process is known as de–alloying. It is schematically shown
in Figure 10 for Al2 CuMg [76]. The galvanic interactions at the matrix-particle interface
change because of de–alloying. The de–alloyed particle becomes nobler over time and
may initiate an anodic dissolution of the surrounding matrix. Experimental conditions
may also influence the particle dissolution behavior. For example, Al20 Cu2 Mn3 is a noble
particle with respect to matrix at room temperature. Nevertheless, it may become anodic at
temperatures higher than 50 ◦ C. At 50 ◦ C a dealloying behavior of Al20 Cu2 Mn3 has been
observed, with de–alloying features much the same as Al2 CuMg [77,78].
Figure 10. Schematic of trenching of Al7 Cu2 Fe intermetallic in AA2024-T3 aluminum alloy (a); microstructure image of
corroded alloy surface after 1-h exposure in aerated H2 O at 30 ◦ C, redrawn (a) and reproduced (b) from ref. [76].
69
Materials 2021, 14, 5418
4. Al–Co Alloys
The Al–Co alloys composed of SCIPs were investigated by Lekatou et al. [79–82] The
authors prepared a series of novel Al–Co alloys with 3.3–10.3 at.% Co by arc-melting. The
microstructures obtained were ranging from fully eutectic to hypereutectic microstruc-
tures with primary precipitation of structurally complex Al9 Co2 . Relatively uniform and
directional microstructures were obtained (Figure 11, [81]). The fraction of directionally
solidified Al9 Co2 was increasing with increasing Co concentration. Microstructures of the
materials before and after corrosion are compared in Figure 11. The alloys displayed a
similar corrosion behavior in 3.5 wt.% NaCl. The corrosion attack resulted in a preferential
dissolution of Al solid solution (ss).
Figure 11. Microstructure of the rapidly solidified Al–Co alloys before and after corrosion in aqueous
NaCl, adapted from reference [81].
Al(ss) < Al9 Co2 < Al13 Co4 < Al5 Co2 < β(AlCo) (10)
70
Materials 2021, 14, 5418
The nobility of IMCs increases with increasing Co concentration. The volume fractions
of the phases and physical contacts between them play an important role in the alloy
corrosion behavior. Results indicate that a galvanic mechanism is involved. Moreover, it
should be mentioned that Al–Co IMCs are brittle [85]. Therefore, a piling of noble but brittle
particles, such as β(AlCo), in pores resulting from massive dissolution of surrounding
less-noble phases may significantly influence the alloy stability [82]. The structural defects
in the alloy may act as rapid diffusion paths leading to a significant material degradation
over time. The galvanic coupling of noble IMCs with more active phases may be critical to
the alloy corrosion stability in halide-containing environments.
The parallel occurrence of SCIPs with similar chemical compositions has a positive
effect on the corrosion susceptibility of the alloy [84,85]. The Al74 Co24 alloy was composed
of three phases with close chemical compositions (Z–Al3 Co, Al5 Co2 , and Al13 Co4 , [84]). The
Al74 Co24 alloy had a higher corrosion potential compared to the remaining alloys which is
an indicator of a superior corrosion resistance. The inspection of the alloy after corrosion
testing revealed a relatively uniform phase dissolution [84]. The potential differences
between constituent phases were probably small enough to initiate galvanic corrosion. The
alloy corrosion could only be initiated at high electrode potentials. A polarization at high
potentials resulted into a massive degradation of this alloy.
To further investigate the corrosion susceptibility of individual SCIPs with close chem-
ical composition, an annealing of the Al74 Co26 alloy at 1000 ◦ C for 330 h has been carried
out [86]. The annealing resulted in equilibrium microstructure of the alloy composed of
Z–Al3 Co and Al5 Co2 . The Z–Al3 Co phase in the as-annealed Al74 Co26 alloy was signifi-
cantly less attacked. Although the bulk of this phase comprises less aluminum, it appears
to be nobler and less susceptible to pitting corrosion compared to Al5 Co2 . The reason
for this behavior could stem in a different structure of the phase surface. The Al5 Co2
surface is terminated in puckered layers [87]. The surface of Z–Al3 Co, on the other hand, is
more densely populated compared to Al5 Co2 [88]. Therefore, Z–Al3 Co was less prone to
corrosion attack.
The corrosion behavior of the as-annealed Al74 Co26 alloy was investigated in neu-
tral (NaCl, 0.6 mol dm−3 ), alkaline (NaOH, 10−2 mol dm−3 ) and acidic electrolytes
(HCl, 10−2 mol dm−3 ) by cyclic potentiodynamic polarization [86]. The potentiodynamic
curves are shown in Figure 12. Anodic parts of the curves measured in HCl and NaCl
solutions displayed a passive region which was followed by an abrupt current density
increase. When the polarization scan was reversed, a positive hysteresis was found. These
features indicate pitting corrosion. The polarization behavior in NaOH, on the contrary,
corresponds to uniform alloy corrosion.
Figure 12. Potentiodynamic cyclic polarization curves of near-equilibrium Al74 Co26 alloy in different
electrolytes, re-plotted from reference [86].
71
Materials 2021, 14, 5418
The forward curves were evaluated by Tafel extrapolation and corrosion currents
and corrosion potentials were obtained [86]. The lowest corrosion potential and highest
corrosion current were found for NaOH. The highest corrosion potential and lowest
corrosion current, on the other hand, were observed for the HCl solution. This behavior is
in accordance with equilibrium E–pH diagram of Al (Figure 7).
5. Al–Cr Alloys
The Al–Cr alloys are expected to demonstrate a good corrosion resistance due to
high concentrations of Al and Cr [89]. Both are passivating elements producing protective
scales. The corrosion resistance of an Al70 Cr20 Fe10 alloy was studied by Li et al. [90].
The authors used commercial gas-atomized Al70 Cr20 Fe10 powders that were consolidated
by spark plasma sintering. The phases present in the sintered Al–Cr–Fe pellets were
the following: an icosahedral phase (i–Al–Cr–Fe), decagonal phase (d–Al–Cr–Fe) and
crystalline Al8 (Cr,Fe)5 and Al9 (Cr,Fe)4 phases. Authors measured an open circuit potential
(OCP) of the alloy in 3.5 wt. % NaCl and found that the OCP was nobler compared to
Al. The OCP of the alloy was stable over time, indicating that an equilibrium has been
rapidly established on the alloy surface. The Al70 Cr20 Fe10 alloy had a nobler corrosion
potential and hence a lower susceptibility to corrosion compared to Al. It passivated in
saline solution spontaneously due to significant amount of Cr. The alloy had a higher
corrosion rate compared to pure Al [90]. Nevertheless, the corrosion rate was close to that
of 316 stainless steel and smaller than AISI 440C stainless steel or AISI H13 tool steel.
The passivation behavior of Al–Cr–Fe alloys was studied by Ott et al. [91] The authors
used a flow microcapillary plasma mass spectrometry. The schematic of the experimental
set up is shown in Figure 13 [91]. In the experiment, a tiny microcapillary was positioned
on the alloy surface and continuously filled with the desired solution. The flow injec-
tion was operated in loops by switching the valve. The loop volume was continuously
transferred to the inductively coupled plasma mass spectrometer (ICP MC) for element
analysis. The microcapillary was refilled with fresh electrolyte from the reservoir. The cir-
culation was ensured by a peristaltic pump. A microscope was included to control precise
positioning of the capillary on the alloy surface. The technique provided time-resolved
information about transient electrochemical processes and element-specific dissolution at
the metal–electrolyte interface.
Figure 13. Schematic of the microcapillary flow ICP MS setup, re-drawn from reference [91].
The authors prepared and studied a polycrystalline γ-phase Al64.2 Cr27.2 Fe8.1 alloy
(composition given in at.%, [91]). The corrosion behavior was studied in two acidic
solutions: H2 SO4 (pH 0) and HCl (pH 2). In sulfuric acid, very low element dissolution
rates were found. Neither Fe nor Al is stable at low pH [69]. Therefore, Cr is an essential
element in the passive film stability. It helps to stabilize the Al cations within the passive
72
Materials 2021, 14, 5418
Figure 14. Schematic of the passive film evolution on γ-phase Al64.2 Cr27.2 Fe8.1 alloy, re-drawn from
reference [91].
Ageing Time Applied Potential Dissolved Passive Film Formed Passive Film
Electrolyte pH
[h] [VSCE ] a Thickness [nm] Thickness [nm]
H2 SO4 0 0.5 OCP b 106
H2 SO4 0 0.5 0.18 55.6 8.4
H2 SO4 0 0.5 0.68 153 6.9
H2 SO4 0 3 OCP 116
HCl 2 0.5 OCP 880
HCl 2 0.5 0.18 147 24.6
a Volts versus saturated calomel electrode, b Open circuit potential.
73
Materials 2021, 14, 5418
phase, composed of cubic γ phase) and single crystalline orthorhombic Al79.0 Cr15.0 Fe6.0 . The
corrosion behavior of the different alloys could be explained considering the passivating
role of Cr combined with Fe oxyhydroxide precipitation. The anticipated reaction mech-
anism is presented in Figure 15 [92]. The Al79.5 Cr12.5 Fe8.0 alloy was found to undergo
an active dissolution in the electrolyte, as proven by the high element concentrations in
solution measured by ICP MS. The chromium concentration (12.5 at.%) was small but
sufficient to stabilize the initially air-formed oxyhydroxide for 22 days, as evidenced by the
constant low pH of the solution and low dissolution compared to Al. The concentration of
Cr was, however, too low to provide a long-term protection. A thick and non-protective
layer has been formed on the surface. With increasing Cr concentration, a protective layer
on the alloys started to form. The Cr concentration of 15.0 at.% was sufficient to stabilize
the passive film up to 78 days. A complete and long-lasting protective scale was finally
achieved at 27.2 at.% Cr [92,93].
Figure 15. Schematic of passivation/dissolution processes on pure Al and Al–Cr–Fe alloys in aqueous
HCl + NaCl mixture with initial pH 2, re-drawn from reference [92].
6. Al–Noble-Metal Alloys
Massiani et al. [35] investigated the corrosion behavior of crystalline and quasicrys-
talline phases in the Al–Cu–Fe(–Cr) alloys by potentiodynamic polarization in strongly
acidic and alkaline solutions. They found that the corrosion resistance was determined by
the alloy chemical composition. The complex crystal structure had only a minor influence.
Rüdiger and Köster [94] found that the corrosion behavior of quasicrystals and their ap-
proximants in the Al–Cu–Fe alloy system could be explained based on the electrochemical
behavior of the component elements. The surface of the icosahedral Al63 Cu25 Fe12 was
covered by a thick non-protective layer composed of Cu2 O, Al(OH)3 and metallic Cu.
The scale chemical composition was comparable to crystalline Al7 Cu2 Fe. The complex
crystal structure thus did not have a substantial influence on the corrosion resistance [94].
Furthermore, the authors observed a formation of porous Cu layer in i–Al63 Cu25 Fe12 phase
at pH 0.
While Rüdiger and Köster studied single phase quasicrystalline alloys, Huttunen et al.
investigated Al–Cu–Fe alloys composed of several different phases (Table 4, [95]). The
corrosion behavior was determined by anodic polarization. The microstructural features
and phase constitution of the alloys before and after the polarization were studied by
scanning electron microscopy and X-ray diffraction.
74
Materials 2021, 14, 5418
Table 4. Chemical composition of SCIPs in Al–Cu–Fe alloys (in at.%) studied in reference [95].
The study was focused on four different Al–Cu–Fe alloys: Al67.5 Cu20 Fe12.5 , Al65 Cu20 Fe15 ,
Al62.5 Cu25 Fe12.5 and Al60 Cu27.5 Fe12.5 [95]. The authors found that the presence of structurally
complex phases did not improve the alloys corrosion resistance [95,96]. The chemical
composition of the phases, however, was of great importance. The corrosion potentials of
Al–Cu–Fe alloys with Cu-rich phases were nobler and had lower corrosion rates compared
to Cu-lean alloys [95]. Relative amounts of the phases and their electrical contacts were
also significant factors influencing the overall corrosion behavior. Phases with high Cu
concentration remained virtually unaffected by corrosion. The phases with low Cu atomic
fractions were susceptible to corrosion attack. This behavior could be explained by the
higher electrode potential of Cu compared to Al and Fe (Table 2). The corrosion was
found to occur by galvanic mechanism near phase boundaries. The corrosion behavior
of Al–Cu–Fe alloys was studied in alkaline, neutral, and acidic solutions. In alkaline and
neutral electrolytes, an oxidation of Al and Cr occurred on the surface of the alloys. The
oxidation was accompanied with Cu deposition on the alloy surface. The Cu deposition
interfered with passive layer formation and introduced pores into the oxide film. The
icosahedral ψ–Al65 Cu20 Fe15 was the only phase capable of forming a stable passive layer
on the surface [95].
The Al–noble-metal alloys are interesting materials from electrochemical point of
view. The alloys are prone to selective dissolution of less noble elements (leaching) because
of markedly different electrode potentials of the constituent metals [97]. The less noble
elements tend to dissolve in the electrolyte, leaving behind their vacant positions. As
a result of leaching, a porous de-alloyed structure forms on the alloy surface [97]. The
leaching can be either uniform or localized. Examples of leaching include preferential
dissolution of Zn from brass (de-zincification) or Fe removal from gray cast iron (a so-called
graphitic corrosion) [98]. Other examples include de-aluminification, de-nickelification
and de-cobaltification [99].
Mishra et al. studied a chemical leaching of Al–Cu–Co decagonal quasicrystals [100].
The authors prepared two alloys with Al65 Cu15 Co20 and Al65 Cu20 Co15 chemical compo-
sitions and studied their corrosion behavior in aqueous NaOH (10 mol L−1 ). The alloys
were immersed in the alkaline solution at room temperature for 8 h. Most Al atoms were
removed (Figure 16, [100]). A nearly uniformly distributed metallic Cu, Co and Co3 O4
nanoparticles were found on the alloy surface after leaching. The crystallite size was
calculated from the XRD reflections’ broadening and further confirmed by TEM [100]. The
nanostructure formation of the leached layer was controlled by Al dissolution rate during
leaching. The dispersed Cu and Co nanoparticles were stable in the leached layer and Cu
agglomeration was suppressed.
75
Materials 2021, 14, 5418
Figure 16. Chemical composition (in at.%) of as-cast and as-leached Al–Cu–Co alloys, plotted from
data in reference [100].
Porous nanostructures composed of noble metals are important catalysts. The formation
of Cu-rich porous nanostructure from decagonal Al65CoxCu35−x alloys (x = 12.5, 15, 17.5 at.%)
was studied by Kalai Vani et al. [101] A selective dissolution of Al and Co was achieved by
combined immersion of the alloys in both NaOH (5 mol L−1 ) and Na2 CO3 (0.5 mol L−1 )
electrolytes. A high specific surface of 30 m2 g−1 of the porous Cu structure was achieved.
The electrochemical de-alloying of binary Al–noble metal alloys was also studied [102].
It has been shown that nano-porous Pd, Ag and Au with various structures can be produced
through electrochemical leaching of the Al–based alloys in NaCl aqueous solution. Galvanic
interactions between coexisting phases dominate during corrosion of double phase alloys.
The level of de-alloying depends on the critical de–alloying potential [103], diffusion of
the noble element and reactivity of the noble element and chloride anion. The porosity
evolution is a dynamic process. It is not a simple excavation of the less noble phase from
two phase material. The formation of the porous nanostructure involves selective leaching
of Al and is accompanied with coarsening of the noble element due to surface diffusion.
The corrosion behavior of Al–Pd alloys composed of SCIPs was studied in refer-
ences [104,105]. The open circuit potentials are given in Figure 17. The OCPs decrease in
the following order:
Al67 Pd33 , Al72 Pd28 (group I) > Al77 Pd23 , Al88 Pd12 (group II) (11)
Figure 17. Open circuit potentials of Al–Pd alloys in aqueous NaCl, re-plotted from data in refer-
ence [105].
76
Materials 2021, 14, 5418
The OCPs of the alloys decrease with decreasing Pd concentration. This observation
is in accordance with expectations since Al is electrochemically more active compared to
Pd (Table 2). The corrosion resistance of both as-annealed and as-solidified alloys was
comparable. A large difference, however, between OCP and Ecorr has been found for group
I alloys (Al67 Pd33 and Al72 Pd28 ). The OCPs of these alloys were comparable to their pitting
potentials obtained by potentiodynamic polarization. The Al67 Pd33 and Al72 Pd28 alloys
were probably in a pitting corrosion stage during the OCP measurement. This suggestion
was manifested by oscillations of OCP resulting from a possible pitting behavior (Figure 17).
The anodic dissolution of the alloy at pits requires a generation of cathodic current from
the surrounding surface. The electric current bursts are transient and cause a temporary
decrease in the OCP value. The pitting corrosion sites are usually very small. However,
the current densities during transient bursts inside the pits can be up to 1 A/m2 [106]. The
significant corrosion rates of the alloy are due to aggressive environments developed inside
the pits. Although the pits are small, they may affect the electrochemical response of much
larger surface areas. Therefore, the differences in current densities on separated anodic
and cathodic sites are reflected in potential oscillations (so-called electrochemical noise
associated with localized corrosion).
Interactions between phases with different chemical composition play a significant
role in alloy corrosion. The Al67 Pd33 and Al72 Pd28 alloys were found to be composed
of structurally complex εn (Al3 Pd) and δ(Al3 Pd2 ). The electrochemical nobility of Al–Pd
phases in aqueous NaCl (0.6 mol L−1 ) increases in the following order
The δ phase has a higher concentration of Pd. It serves as a cathode, and thereby
further accelerates the anodic dissolution of the surrounding εn phase. The corrosion
mechanism of Al–Pd alloys in aqueous NaCl involves a rapid passivation stage on the alloy
surface [105,106]. However, once a breakdown potential is reached during potentiodynamic
polarization, the passive layer becomes unstable and susceptible to local attack by chloride
anions. Consequently, chloro–aluminum complex cations are formed and released into
the solution. The local disruption of the passive layer reveals a naked alloy surface which
becomes more susceptible to further corrosion attack.
The microstructures of as-annealed and as-solidified Al72 Pd28 and Al67 Pd33 alloys
had similar features after corrosion testing. In the alloys a high number of inter-penetrating
channels have been found [105,106]. Pits were also observed in the inter-connection
between the channels. The formation of channels was driven by pitting and de–alloying.
The pits were probably initiation sites of the channels. A preferential de-alloying of Al
(de-aluminification) has also been observed. The preferential leaching of Al led to initiation
of microcracks. During rapid solidification residual stresses have been accumulated in
the alloys. The stresses were released during leaching, resulting in continuous tunnels
inter-penetrating the surfaces of de-alloyed materials. A similar corrosion behavior was
also found for the Al–Pd–Co alloys (Figure 18, [107]). The de-alloying of Al was more
pronounced in the as-solidified alloys. This is probably a consequence of their higher defect
concentrations compared to as-annealed alloys. The de–alloying behavior was significantly
reduced in as-annealed alloys [105].
77
Materials 2021, 14, 5418
Figure 18. Microstructure of as-corroded Al–Pd–Co alloys: backscatter scanning electron microscopy
images of Al70 Pd25 Co5 (a) and Al74 Pd12 Co14 (b) and confocal laser scanning images of Al70 Pd25 Co5
(c) and Al74 Pd12 Co14 (d). Reproduced from reference [107].
Ecorr
jcorr
Alloy Condition Electrolyte [mV vs. Reference
[A m−2 ]
Ag/AgCl]
Al96.7 Co3.3 Cast Aerated NaCl (0.6 mol dm−3 ) −838 ± 20 0.7 ± 0.1 [79]
Al96.7 Co3.3 Arc-melted Aerated NaCl (0.6 mol dm−3 ) −820 ± 36 0.3 ± 0.1 [79]
Powder-metallurgy
Al96.7 Co3.3 Aerated NaCl (0.6 mol dm−3 ) −890 ± 50 0.9 ± 0.2 [79]
sintered
Al96.7 Co3.3 Cast Aerated NaCl (0.6 mol dm−3 ) −820 ± 36 0.3 ± 0.1 [81]
Al95.1 Co4.9 Cast Aerated NaCl (0.6 mol dm−3 ) −823 ± 23 0.6 ± 0.1 [81]
Al92.5 Co7.5 Cast Aerated NaCl (0.6 mol dm−3 ) −799 ± 23 0.7 ± 0.1 [81]
Al89.7 Co10.3 Cast Aerated NaCl (0.6 mol dm−3 ) −816 ± 23 0.9 ± 0.1 [81]
Al82.3 Co17.7 Cast Aerated NaCl (0.6 mol dm−3 ) −843 ± 16 1.6 ± 0.1 [82]
Al82.3 Co17.7 Arc-melted Aerated NaCl (0.6 mol dm−3 ) −825 ± 18 0.8 ± 0.1 [82]
Powder-metallurgy
Al82.3 Co17.7 Aerated NaCl (0.6 mol dm−3 ) −877 ± 23 5.8 ± 0.6 [82]
sintered
78
Materials 2021, 14, 5418
Table 5. Cont.
Ecorr
jcorr
Alloy Condition Electrolyte [mV vs. Reference
[A m−2 ]
Ag/AgCl]
Al99.1 Co0.9 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −400 ± 7 1.9 ± 0.3 [80]
Al97.6 Co2.4 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −406 ± 2 3.6 ± 0.6 [80]
Al96.7 Co3.3 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −388 ± 10 2.6 ± 0.6 [80]
Al95.1 Co4.9 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −390 ± 5 1.9 ± 0.6 [80]
Al92.5 Co7.5 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −381 ± 18 3.1 ± 0.8 [80]
Al89.3 Co10.3 Arc-melted Aerated H2 SO4 (1 mol dm−3 ) −372 ± 7 2.9 ± 0.4 [80]
Al76 Co24 Cast Aerated NaCl (0.6 mol dm−3 ) −706 0.13 [84]
Al75 Co25 Cast Aerated NaCl (0.6 mol dm−3 ) −729 0.039 [84]
Al74 Co26 Cast Aerated NaCl (0.6 mol dm−3 ) −515 0.58 [84]
Al73 Co27 Cast Aerated NaCl (0.6 mol dm−3 ) −646 0.05 [84]
Al72 Co28 Cast Aerated NaCl (0.6 mol dm−3 ) −672 0.04 [84]
Al71 Co29 Cast Aerated NaCl (0.6 mol dm−3 ) −530 0.10 [83]
Annealed in Ar 1050 ◦ C
Al74 Co26 Aerated NaCl (0.6 mol dm−3 ) −651 0.051 [86]
330 h
Annealed in Ar 1050 ◦ C
Al74 Co26 Aerated HCl (0.01 mol dm−3 ) −314 0.032 [86]
330 h
Annealed in Ar 1050 ◦C
Al74 Co26 Aerated NaOH (0.01 mol dm−3 ) −1026 2.6 [86]
330 h
Al72 Fe15 Ni13 Cast Aerated NaCl (0.87 mol dm−3 ) - 1.4 [108]
Al69 Co21 Ni10 Cast Aerated NaCl (0.87 mol dm−3 ) - 1.2 [108]
Powder metallurgy
Al70 Cr20 Fe10 Aerated NaCl (0.6 mol dm−3 ) −938 0.018 [90]
sintered
NaCl
Al65 Cu20 Fe15 Cast −638 ± 100 0.37 [109]
(0.6 mol dm−3 )
NaCl
Al78 Cu7 Fe15 Cast −586 ± 100 0.056 [109]
(0.6 mol dm−3 )
NaCl
Al80 Cu5 Fe14 Si1 Cast −570 ± 100 0.14 [109]
(0.6 mol dm−3 )
Na2 SO4
Al70 Cu9 Fe10.5 Cr10.5 Cast −556 1.6 × 10−2 [35]
(0.5 mol dm−3 )
Na2 SO4
Al64 Cu24 Fe12 Cast −555 7.3 × 10−2 [35]
0.5 mol dm−3 )
Na2 SO4
Al63 Cu20 Co15 Si2 Cast −635 2.2 × 10−2 [35]
(0.5 mol dm−3 )
Na2 SO4
Al70 Cu9 Fe10.5 Cr10.5 Cast −496 1.4 × 10−2 [35]
(0.5 mol dm−3 ) + H2 SO4 (pH 2)
Na2 SO4
Al64 Cu24 Fe12 Cast −512 0.8 × 10−2 [35]
(0.5 mol dm−3 ) + H2 SO4 (pH 2)
Na2 SO4
Al63 Cu20 Co15 Si2 Cast −186 0.6 × 10−2 [35]
(0.5 mol dm−3 ) + H2 SO4 (pH 2)
79
Materials 2021, 14, 5418
Table 5. Cont.
Ecorr
jcorr
Alloy Condition Electrolyte [mV vs. Reference
[A m−2 ]
Ag/AgCl]
NaOH
Al70 Cu9 Fe10.5 Cr10.5 Cast −921 1.6 × 10−2 [35]
(0.1 mol dm−3 )
NaOH
Al64 Cu24 Fe12 Cast −1508 336 × 10−2 [35]
(0.1 mol dm−3 )
NaOH
Al63 Cu20 Co15 Si2 Cast −1441 462 × 10−2 [35]
(0.1 mol dm−3 )
Annealed in Ar
Al72 Pd20 Mn8 Deaerated NaCl (0.5 mol dm−3 ) −355 0.5 [36]
800 ◦ C 12 h
Al88 Pd12 Cast Aerated NaCl (0.6 mol dm−3 ) −794 0.89 [105]
Al77 Pd23 Cast Aerated NaCl (0.6 mol dm−3 ) −809 0.82 [105]
Al72 Pd28 Cast Aerated NaCl (0.6 mol dm−3 ) −797 0.63 [105]
Al67 Pd33 Cast Aerated NaCl (0.6 mol dm−3 ) −798 0.62 [105]
Annealed in Ar 700 ◦C
Al77 Pd23 Aerated NaCl (0.6 mol dm−3 ) −763 0.75 [105]
500 h
Annealed in Ar 700 ◦C
Al72 Pd28 Aerated NaCl (0.6 mol dm−3 ) −841 0.68 [105]
500 h
Annealed in Ar 700 ◦ C
Al67 Pd33 Aerated NaCl (0.6 mol dm−3 ) −783 0.72 [105]
500 h
Al88 Pd12 Cast Aerated HCl (0.01 mol dm−3 ) −478 0.26 [104]
Al77 Pd23 Cast Aerated HCl (0.01 mol dm−3 ) −450 0.27 [104]
Al72 Pd28 Cast Aerated HCl (0.01 mol dm−3 ) −253 0.03 [104]
Al67 Pd33 Cast Aerated HCl (0.01 mol dm−3 ) −200 0.17 [104]
Al88 Pd12 Cast Aerated NaOH (0.01 mol dm−3 ) −1019 0.42 [104]
Al77 Pd23 Cast Aerated NaOH (0.01 mol dm−3 ) −1033 0.25 [104]
Al72 Pd28 Cast Aerated NaOH (0.01 mol dm−3 ) −879 0.25 [104]
Al67 Pd33 Cast Aerated NaOH (0.01 mol dm−3 ) −892 0.34 [104]
Al70 Pd25 Co5 Cast Aerated NaCl (0.6 mol dm−3 ) −677 0.10 [107]
Al74 Pd12 Co14 Cast Aerated NaCl (0.6 mol dm−3 ) −758 0.18 [107]
Powder metallurgy
Al93 Co5 Ti2 Aerated NaCl (0.6 mol dm−3 ) −707 1.1 [110]
sintered
Powder metallurgy
Al88 Co10 Ti2 Aerated NaCl (0.6 mol dm−3 ) −676 14 [110]
sintered
Powder metallurgy
Al83 Co15 Ti2 Aerated NaCl (0.6 mol dm−3 ) −490 4.0 × 10−4 [110]
sintered
Powder metallurgy
Al78 Co20 Ti2 Aerated NaCl (0.6 mol dm−3 ) −669 2.39 [110]
sintered
Powder metallurgy
Al73 Co25 Ti2 Aerated NaCl (0.6 mol dm−3 ) −661 0.64 [110]
sintered
Powder metallurgy
Al68 Co30 Ti2 Aerated NaCl (0.6 mol dm−3 ) −649 7.0 [110]
sintered
80
Materials 2021, 14, 5418
Figure 19. Corrosion parameters of Al–Pd, Al–Co and Al–Pd–Co alloys in aqueous NaCl (0.6 mol L−1 ). Data is compiled
from references [81,82,84,105,107].
The corrosion behavior of the Al–Pd–Co alloys is closer to Al–Co alloys (Figure 19).
This observation is unexpected, since Al–Co–Pd and Al–Co alloys have different phase
constitutions. Moreover, the preferentially corroding phase is εn in the Al–Pd–Co alloys.
εn is absent in the Al–Co alloys. It can be noted that Co substitution for Pd significantly
improves the corrosion resistance of εn . The positive influence of Co on the corrosion
resistance of Al–TM alloys has also been observed by Sukhova and Polonskyy [108]. It is
therefore the chemical composition and not the crystal structure of the phase that plays a
dominant role in the corrosion resistance.
To further probe the role of chemical composition, we have compared the corrosion
parameters of the previously discussed Al–TM alloys. The data compilation is plotted in
Figure 20. The parameters are relatively scattered due to large differences in the overall
alloy chemical compositions (Table 5). Nevertheless, some general trends can be noted. The
as-solidified Al–Pd–Co alloys have corrosion current densities comparable to Al–Cu–Fe
alloys. The corrosion potentials of the Al–Pd–Co and Al–Cu–Fe alloys are close to −650 mV
(vs. Ag/AgCl)). The Al–Cr–Fe alloy is also included in Figure 20. This alloy has a lower
corrosion potential compared to the remainder of the alloys. This is related to the absence of
noble metals, such as Pd, in the alloy. Furthermore, the Al–Cr–Fe alloy has a low corrosion
current due to the presence of Cr [90]. This element is responsible for a rapid passive layer
formation on the alloy surface.
Figure 20. Corrosion parameters of ternary Al–TM alloys in aqueous NaCl (0.6 mol L−1 ). Data is compiled from refer-
ences [90,107,109,110].
The corrosion parameters of Al–Co–Ti alloys [110] are also included in the same figure.
The corrosion potentials of these alloys are comparable to Al–Pd–Co alloys (Figure 20).
The concentration of Ti in the Al–Co–Ti alloys was constant (2 at.%). The atomic fraction
of Co was varied between 5–30 at.%. Due to small and constant Ti concentration, the
81
Materials 2021, 14, 5418
microstructural features of the Al–Co–Ti alloys were comparable to Al–Co alloys [81,84].
The corrosion current densities of Al–Pd–Co alloys, however, were smaller compared to
Al–Co–Ti alloys. The Al–15Co–2Ti alloy was an exception as the alloy demonstrated a
lower corrosion current compared to the remainder of the alloys. The difference was related
to different intermetallic particles contained in the alloy (Al13 Co4 , Al9 Co2 , and Al3 Ti). They
had different volume fractions and morphologies compared to the remaining Al–Ti–Co
alloys [110]. The observations show that specific Co atomic fractions may significantly
increase the corrosion resistance of the bulk Al–TM alloys. The εn phase of the Al–Pd–Co
alloys had a high concentration of Co. The Co additions significantly contributed to the
superior corrosion performance of the bi-phasic Al70 Pd25 Co5 alloy.
The corrosion parameters of the structurally complex Al–TM phases are comparable
to previously studied Al–TM intermetallic phases with simpler structures [111]. Therefore,
it is the chemical composition of the SCIP and not the crystal structure that influences
the corrosion behavior. The electrochemical activity of the SCIPs may also vary with
time. Zhu et al. investigated the corrosion performance of Al–TM intermetallic phases
over time [78]. At early stages of exposure, a de–alloying was the primary corrosion
mechanism. The de–alloying led to an ennoblement of intermetallic particles over time
due to preferential Al leaching [78]. The ennoblement speeded up an anodic dissolution of
the adjacent matrix and worsened the corrosion behavior. A long-term annealing may also
influence the corrosion performance of the alloy constituent phases. It reduces internal
stresses generated during casting and contributes to a more uniform element redistribution
in the SCIPs.
8. Conclusions
In this paper the electrochemical corrosion behavior of Al–TM alloys composed of
SCIPs has been reviewed. The following conclusions can be drawn:
1. The Al–TM alloys have a capability of forming passive layers because of high
Al concentration. The Al–Cr alloys, for example, can form protective passive layers of
considerable thickness in different electrolytes.
2. In halide-containing solutions the Al–TM alloys are prone to pitting corrosion. Gal-
vanic microcells between different SCIPs form which may further accelerate the localized
corrosion attack.
3. The electrochemical activity of aluminum–transition-metal SCIPs is primarily
determined by electrode potential of the alloying element(s). The electrochemical nobility
of individual SCIPs increases with increasing concentration of noble elements. The SCIPs
with less noble elements tend to dissolve in contact with nobler particles. The SCIPs with
noble metals are prone to selective de-alloying (de-aluminification). The electrochemical
activity of SCIPs may change over time.
4. The chemical composition of the SCIPs has a primary influence on their corrosion
properties. The structural complexity is secondary. It becomes important when phases
with similar chemical composition come into close physical contact. The phase with higher
structural complexity tends to be cathodic and can be retained during corrosion.
Author Contributions: Conceptualization, M.P. and L.Ď.; methodology, M.P., L.Ď. and I.Č.; formal
analysis, M.P., L.Ď., I.Č. and P.P.; investigation, M.P., L.Ď., I.Č. and P.P.; resources, M.P. and P.P.;
data curation, M.P.; writing—original draft preparation, M.P., L.Ď. and I.Č.; writing—review and
editing, M.P., L.Ď., I.Č. and P.P.; visualization, M.P., L.Ď., I.Č. and P.P.; supervision, M.P.; project
administration, M.P., and P.P.; funding acquisition, M.P. and P.P. All authors have read and agreed to
the published version of the manuscript.
Funding: This research was funded by the Grant Agency VEGA of the Ministry of Education, Science,
Research and Sport of the Slovak Republic and Slovak Academy of Sciences, grant number 1/0330/18,
the Slovak Research and Development Agency project number APVV-20-0124 and the European
Regional Development Fund, project No. ITMS2014+: 313011W085.
Institutional Review Board Statement: Not applicable.
82
Materials 2021, 14, 5418
References
1. Georgantzia, E.; Gkantou, M.; Kamaris, G.S. Aluminium alloys as structural material: A review of research. Eng. Struct. 2021, 227,
111372. [CrossRef]
2. Starke, E.A., Jr.; Staley, J.T. Application of modern aluminium alloys to aircraft. In Fundamentals of Aluminum Metallurgy; Lumley,
R., Ed.; Woodhead Publishing: Cambridge, UK, 2011; pp. 747–783. [CrossRef]
3. Davis, J.R. Aluminum and aluminum alloys. In Alloying: Understanding the Basics; ASM International: Materials Park, OH, USA,
2001; pp. 351–416. [CrossRef]
4. Shechtman, D.; Blech, I.; Gratias, D.; Cahn, J.W. Metallic phase with long-range orientational order and no translational symmetry.
Phys. Rev. Lett. 1984, 53, 1951–1953. [CrossRef]
5. Dubois, J.M. Quasicrystals. J. Phys. Condens. Matter 2001, 13, 7753–7762. [CrossRef]
6. Dubois, J.M. Useful Quasicrystals; World Scientific Publishing: Singapore, 2005; pp. 30–34. ISBN 978-9812561886.
7. Steurer, W.; Deloudi, S. Crystallography of Quasicrystals—Concepts, Methods, and Structures; Springer: Berlin/Heidelberg, Germany,
2009; pp. 258–293. ISBN 978-3-642-01899-2. [CrossRef]
8. Janot, C. Quasicrystals, 2nd ed.; Clarendon Press: Oxford, UK, 1994; ISBN 0-19-851778-5. [CrossRef]
9. Dubois, J.M.; Ferré, E.B.; Feuerbacher, M. Introduction to the Science of Complex Metallic Alloys. In Complex Metallic Alloys:
Fundamentals and Applications; Dubois, J.-M., Belin-Ferré, E., Eds.; Wiley-VCH: Hoboken, NJ, USA, 2011; pp. 1–39. [CrossRef]
10. Priputen, P.; Liu, T.; Černičková, I.; Janičkovič, D.; Kolesár, V.; Janovec, J. Experimental study of Al-Co-Cu phase diagram in
temperature range of 800–1050 ◦ C. J. Phase Equilib. Diff. 2013, 34, 425–429. [CrossRef]
11. Priputen, P.; Černičková, I.; Lejček, P.; Janičkovič, D.; Janovec, J. A partial isothermal section at 1000 ◦ C of Al-Mn-Fe phase
diagram in vicinity of Taylor phase and decagonal quasicrystal. J. Phase Equilib. Diff. 2016, 37, 130–134. [CrossRef]
12. Socolar, J.E.S. Simple octagonal and dodecagonal quasicrystals. Phys. Rev. B 1989, 39, 10519–10551. [CrossRef] [PubMed]
13. Wang, N.; Cen, H.; Kuo, K.H. Two-dimensional quasicrystal with eightfold rotational symmetry. Phys. Rev. Lett. 1987, 59,
1010–1013. [CrossRef]
14. Cao, W.; Ye, H.Q.; Kuo, K.H. A new octagonal quasicrystal and related crystalline phases in rapidly solidified Mn4 Si. Phys. Stat.
Solidi 1988, A107, 511–519. [CrossRef]
15. Wang, N.; Fung, K.K.; Kuo, K.H. Symmetry study of the Mn-Si-Al octagonal quasicrystal by convergent beam electron-diffraction.
Appl. Phys. Lett. 1988, 52, 2120–2121. [CrossRef]
16. Wang, Z.M.; Kuo, K.H. The octagonal quasilattice and electron-diffraction patterns of the octagonal phase. Acta Crystallogr. 1988,
A44, 857–863. [CrossRef]
17. He, L.X.; Zhang, Z.; Wu, Y.K.; Kuo, K.H. Stable decagonal quasicrystals with different periodicities along the tenfold axis in
Al65 Cu20 Co15 . Inst. Phys. Conf. Ser. 1988, 93, 501–502.
18. Tsai, A.P.; Inoue, A.; Masumoto, T. Stable decagonal quasicrystals with a periodicity of 1.6 nm in Al-Pd-(Fe, Ru or Os) alloys.
Philos. Mag. Lett. 1991, 64, 163–167. [CrossRef]
19. Ge, S.P.; Kuo, K.H. Icosahedral and stable decagonal quasicrystals in Ga46 Fe23 Cu23 Si8 , Ga50 Co25 Cu25 and Ga46 V23 Ni23 Si8 . Phil.
Mag. Lett. 1997, 75, 245–253. [CrossRef]
20. Yokoyama, Y.; Yamada, Y.; Fukaura, K.; Sunada, H.; Inoue, A.; Note, R. Stable decagonal quasicrystal in an Al-Mn-Fe-Ge system.
Jpn. J. Appl. Phys. 1997, 36, 6470–6474. [CrossRef]
21. Sato, T.J.; Abe, E.; Tsai, A.P. A novel decagonal quasicrystal in Zn-Mg-Dy system. Jpn. J. Appl. Phys. 1997, 36, L1038–L1039. [CrossRef]
22. Yang, Q.B.; Wei, W.D. Description of the dodecagonal quasicrystal by a projection method. Phys. Rev. Lett. 1987, 58, 1020–1023.
[CrossRef] [PubMed]
23. Chen, H.; Li, D.X.; Kuo, K.H. New type of two-dimensional quasicrystal with twelvefold rotational symmetry. Phys. Rev. Lett.
1988, 60, 1645–1648. [CrossRef] [PubMed]
24. Ďuriška, L. Phase-Constitutional, Thermodynamic, and Electrochemical Studies of Al-Base Complex Metallic Alloys. Ph.D.
Thesis, Slovak University of Technology, Trnava, Slovakia, 2017. Available online: https://opac.crzp.sk/?fn=detailBiblioForm&
sid=7F4700268CB7D7D1FF59AFA53547 (accessed on 6 July 2021).
25. Yokoyama, Y.; Miura, T.; Tsai, A.P.; Inoue, A.; Masumoto, T. Preparation of a large Al70 Pd20 Mn10 single-quasicrystal by the
Czochralski method and its electrical resistivity. Mater. Trans. JIM 1992, 33, 97–101. [CrossRef]
26. Dubois, J.M.; Kang, S.S.; Stebut, J. Quasicrystalline low-friction coatings. J. Mater. Sci. Lett. 1991, 10, 537–541. [CrossRef]
27. Rabson, D.A. Toward theories of friction and adhesion on quasicrystals. Prog. Surf. Sci. 2012, 87, 253–271. [CrossRef]
28. Tsai, A.P.; Suenaga, H.; Ohmori, M.; Yokoyama, Y.; Ioue, A.; Masumoto, T. Temperature dependence of hardness and expansion
in an icosahedral Al-Pd-Mn alloy. Jpn. J. Appl. Phys. 1992, 31, 2530–2531. [CrossRef]
83
Materials 2021, 14, 5418
29. Köster, U.; Liu, W.; Liebertz, H.; Michel, M. Mechanical properties of quasicrystalline and crystalline phases in Al-Cu-Fe alloys. J.
Non-Cryst. Sollids 1993, 153–154, 446–452. [CrossRef]
30. Yokoyama, Y.; Inoue, A.; Masumoto, T. Mechanical properties, fracture mode and deformation behavior of Al70 Pd20 Mn10
single-quasicrystal. Mater. Trans. JIM 1993, 34, 135–145. [CrossRef]
31. Dubois, J.M.; Kang, S.S.; Perrot, A. Towards applications of quasicrystals. Mater. Sci. Eng. A 1994, A179–A180, 122–126. [CrossRef]
32. Masumoto, T.; Inoue, A. YKK Corporation, Honda Giken Kogyo Kabushiki Kaisha. European Patent 94115137.5.
33. Jenks, C.J.; Thiel, P.A. Comments on quasicrystals and their potential use as catalysts. J. Mol. Catal. A—Chem. 1998, 131, 301–306. [CrossRef]
34. Dubois, J.M. New prospects from potential applications of quasicrystalline materials. Mater. Sci. Eng. A 2000, 294–296, 4–9. [CrossRef]
35. Massiani, Y.; Yaazza, S.A.; Croussier, J.P.; Dubois, J.M. Electrochemical behaviour of quasicrystalline alloys in corrosive solutions.
J. Non-Cryst. Solids 1993, 159, 92–100. [CrossRef]
36. Asami, K.; Tsai, A.-P.; Hashimoto, K. Electrochemical behavior of a quasicrystalline Al-Pd-Mn alloy in a chloride-containing
solution. Mater. Sci. Eng. A 1994, 181, 1141. [CrossRef]
37. Rüdiger, A.; Köster, U. Corrosion behavior of Al-Cu-Fe quasicrystals. Mater. Sci. Eng. A 2000, 294–296, 890–893. [CrossRef]
38. Torres, A.; Serna, S.; Patiño, C.; Rosas, G. Corrosion Behavior of W and b Quasicrystalline Al–Cu–Fe Alloy. Acta Metall. Sin. 2015,
28, 1117–1122. [CrossRef]
39. Veys, D.; Rapin, C.; Li, X.; Aranda, L.; Fournee, V.; Dubois, J.-M. Electrochemical behavior of approximant phases in the
Al–(Cu)–Fe–Cr system. J. Non-Cryst. Solids 2004, 347, 1–10. [CrossRef]
40. Wehner, B.I.; Köster, U.; Rüdiger, A.; Pieper, C.; Sordelet, D.J. Oxidation of Al–Cu–Fe and Al–Pd–Mn quasicrystals. Mater. Sci.
Eng. A 2000, 294–296, 830–833. [CrossRef]
41. Rampulla, D.M.; Mancinelli, C.M.; Brunell, I.F.; Gellman, A.J. Oxidative and Tribological Properties of Amorphous and Qua-
sicrystalline Approximant Al-Cu-Fe Thin Films. Langmuir 2005, 21, 4547–4553. [CrossRef]
42. Yamasaki, M.; Pang Tsai, A. Oxidation behavior of quasicrystalline Al63 Cu25 Fe12 alloys with additional elements. Mater. Sci. Eng.
A 2000, 294-296, 890–893. [CrossRef]
43. Pinhero, P.J.; Anderegg, J.W.; Sordelet, D.J.; Besser, M.F.; Thiel, P.A. Surface oxidation of Al-Cu-Fe alloys: A comparison of
quasicrystalline and crystalline phases. Philos. Mag. B 1999, 79, 91–110. [CrossRef]
44. Šulhánek, P.; Drienovský, M.; Černičková, I.; Ďuriška, L.; Skaudžius, R.; Gerhátová, Ž.; Palcut, M. Oxidation of Al-Co alloys at
high temperatures. Materials 2020, 13, 3152. [CrossRef] [PubMed]
45. Graef, M.; McHenry, M. Structure of Materials: An Introduction to Crystallography, Diffraction, and Symmetry, 3rd ed.; Cambridge
University Press: New York, NY, USA, 2008; ISBN 978-0-521-65151-6. [CrossRef]
46. Taylor, J.E.; Teich, E.G.; Damasceno, P.F.; Kallus, Y.; Senechal, M. On the form and growth of complex crystals: The case of
Tsai-type clusters. Symmetry 2017, 9, 188. [CrossRef]
47. Livio, M. Zlatý Řez, 1st ed.; Dokořán: Prague, Czech Republic, 2006; ISBN 80-7363-064-8.
48. Smontara, A.; Smiljanić, I.; Bilušić, A.; Grushko, B.; Balanetskyy, S.; Jagličić, Z.; Vrtnik, S.; Dolinšek, J. Complex ε-phases in the
Al-Pd-transition-metal systems. J. Alloys Compd. 2008, 450, 92–102. [CrossRef]
49. Bancel, P.A.; Heiney, P.A. Icosahedral aluminum-transition-metal alloys. Phys. Rev. B 1986, 33, 7917–7922. [CrossRef]
50. Create and Colour a Penrose Tiling. Available online: https://craftdesignonline.com/penrose/ (accessed on 6 July 2021).
51. Periodic Rhombus Tiling with Penrose Tiles. Available online: https://sk.pinterest.com/pin/268245721535341551/?d=t&mt=login
(accessed on 6 July 2021).
52. Černičková, I.; Švec, P.; Watanabe, S.; Čaplovič, L’.; Mihalkovič, M.; Kolesár, V.; Priputen, P.; Bednarčík, J.; Janičkovič, D.; Janovec,
J. Fine structure of phases of epsilon-family in Al73.8Pd11.9Co14.3 alloy. J. Alloys Compd. 2014, 609, 73–79. [CrossRef]
53. Černičková, I.; Priputen, P.; Liu, T.; Zemanová, A.; Illeková, E.; Janičkovič, D.; Švec, P.; Kusý, M.; Čaplovič, L’.; Janovec, J.
Evolution of phases in Al-Pd-Co alloys. Intermetallics 2011, 19, 1586–1593. [CrossRef]
54. Černičková, I.; Ďuriška, L.; Priputen, P.; Janičkovič, D.; Janovec, J. Isothermal section of the Al-Pd-Co phase diagram at 850 ◦ C
delimited by homogeneity ranges of phases epsilon, U, and F. J. Phase Equilib. Diff. 2016, 37, 301–307. [CrossRef]
55. Černičková, I.; Ďuriška, L.; Drienovský, M.; Janičkovič, D.; Janovec, J. Phase transitions in selected Al-Pd-Co alloys during
continuous cooling. Kovové Mater. 2017, 55, 403–411. [CrossRef]
56. Ďuriška, L.; Černičková, I.; Čička, R.; Janovec, J. Contribution to thermodynamic description of Al-Pd system. J. Phys. Conf. Ser.
2017, 809, 012008. [CrossRef]
57. Adamech, M.; Černičková, I.; Ďuriška, L.; Kolesár, V.; Drienovský, M.; Bednarčík, J.; Svoboda, M.; Janovec, J. Formation of
less-know structurally complex zéta b and orthorhombic quasicrystalline approximant epsilon n on solidification of selected
Al-Pd-Cr alloys. Mater. Charact. 2014, 97, 189–198. [CrossRef]
58. Priputen, P.; Kusý, M.; Drienovský, M.; Janičkovič, D.; Čička, R.; Černičková, I.; Janovec, J. Experimental reinvestigation of Al-Co
phase diagram in vicinity of Al13 Co4 family of phases. J. Alloys Compd. 2015, 647, 486–497. [CrossRef]
59. Kolesár, V.; Priputen, P.; Bednarčík, J.; Černičková, I.; Svoboda, M.; Drienovský, M.; Janovec, J. Evolution of phases in Al55 Ni30 Pd15
alloy at temperatures up to 600 ◦ C. Intermetallics 2014, 46, 141–146. [CrossRef]
60. Jenks, C.J.; Thiel, P.A. Quasicrystals: A Short Review from a Surface Science Perspective. Langmuir 1998, 14, 1392–1397. [CrossRef]
61. Fournée, V.; Ledieu, J.; Park, J.Y. Surface Science of Complex Metallic Alloys. In Complex Metallic Alloys: Fundamentals and
Applications; Dubois, J.-M., Belin-Ferré, E., Eds.; Wiley-VCH: Hoboken, NJ, USA, 2011; pp. 155–206. [CrossRef]
84
Materials 2021, 14, 5418
62. Ebert, P.H.; Feuerbacher, M.; Tamura, N.; Wollgarten, M.; Urban, K. Evidence for a Cluster-Based Structure of AlPdMn Single
Quasicrystals. Phys. Rev. Lett. 1996, 77, 3827–3830. [CrossRef]
63. Papadopolos, Z.; Kasner, G.; Ledieu, J.; Cox, E.J.; Richardson, N.V.; Chen, Q.; Diehl, R.D.; Lograsso, T.A.; Ross, A.R.; McGrath, R.
Bulk termination of the quasicrystalline fivefold surface of Al70 Pd21 Mn9 . Phys. Rev. B 2002, 66, 184207. [CrossRef]
64. Krajčí, M.; Hafner, J. Structure, stability, and electronic properties of the i-AlPdMn quasicrystalline surface. Phys. Rev. B 2005, 71,
054202. [CrossRef]
65. Hafner, J. Ab initio density-functional calculations in materials science: From quasicrystals over microporous catalysts to
spintronics. J. Phys. Condens. Matter 2010, 22, 384205. [CrossRef] [PubMed]
66. Krajčí, M.; Hafner, J. Surfaces of Complex Intermetallic Compounds: Insights from Density Functional Calculations. Acc. Chem.
Res. 2014, 47, 3378–3384. [CrossRef]
67. Stansbury, E.E.; Buchanan, R.A. Introduction and overview of electrochemical corrosion. In Fundamentals of Electrochemical
Corrosion; ASM International: Materials Park, OH, USA, 2000; pp. 1–21. [CrossRef]
68. Perez, N. Electrochemistry. In Electrochemistry and Corrosion Science; Kluwer Academic Publishers: New York, NY, USA, 2004;
pp. 27–70. [CrossRef]
69. Pourbaix, M. Atlas of Electrochemical Equilibria in Aqueous Solution; National Association of Corrosion Engineers: Houston, TX,
USA, 1974.
70. Hasannaeimi, V.; Sadeghilaridjani, M.; Mukherjee, S. Electrochemical and Corrosion Behavior of Metallic Glasses; MDPI: Basel,
Switzerland, 2021. [CrossRef]
71. Burgleigh, T.D. Corrosion of aluminum and its alloys. In Handbook of Aluminum. Alloys Production and Materials Manufacturing;
Totten, G.E., Mackenzie, D.S., Eds.; ASM International: Materials Park, OH, USA, 2003; Volume 2, pp. 421–463. [CrossRef]
72. ISO 17475:2005 Corrosion of Metals and Alloys—Electrochemical Test Methods—Guidelines for Conducting Potentiostatic and
Potentiodynamic Polarization Measurements. Available online: https://www.iso.org/standard/31392.html (accessed on 6
August 2021).
73. ASTM G102–89 Standard Practice for Calculation of Corrosion Rates and Related Information from Electrochemical Measurements; ASTM
International: West Conshohocken, PA, USA, 2015. Available online: https://www.astm.org/Standards/G102.htm (accessed on
6 August 2021).
74. Sukiman, N.L.; Zhou, X.; Birbilis, N.; Hughes, A.E.; Mol, J.M.C.; Garcia, S.J.; Zhou, X.; Thompson, G.E. Durability and Corrosion
of Aluminium and Its Alloys: Overview, Property Space, Techniques and Developments. In Aluminium Alloys—New Trends in
Fabrication and Applications; Ahmad, Z., Ed.; Intechopen: London, UK, 2012; pp. 47–97.
75. Szklarska-Smialowska, Z. Pitting Corrosion of aluminum. Corros. Sci. 1999, 41, 1743–1767. [CrossRef]
76. Li, J.; Dang, J. A Summary of Corrosion Properties of Al-Rich Solid Solution and Secondary Phase Particles in Al Alloys. Metals
2017, 7, 84. [CrossRef]
77. Li, J.; Hurley, B.; Buchheit, R. Effect of temperature on the localized corrosion of AA2024-t3 and the electrochemistry of
intermetallic compounds during exposure to a dilute NaCl solution. Corrosion 2016, 72, 1281–1291. [CrossRef]
78. Zhu, Y.; Sun, K.; Frankel, G.S. Intermetallic Phases in Aluminum Alloys and Their Roles in Localized Corrosion. J. Electrochem.
Soc. 2018, 165, C807–C820. [CrossRef]
79. Lekatou, A.G.; Sfikas, A.K.; Karantzalis, A.E. The influence of the fabrication route on the microstructure and surface degradation
properties of Al reinforced by Al9 Co2 . Mater. Chem. Phys. 2017, 200, 33–49. [CrossRef]
80. Sfikas, A.K.; Lekatou, A.G. Electrochemical Behavior of Al–Al9Co2 Alloys in Sulfuric Acid. Corros. Mater. Degrad. 2020, 1, 249–272. [CrossRef]
81. Lekatou, A.; Sfikas, A.K.; Petsa, C.; Karantzalis, A.E. Al-Co alloys prepared by vacuum arc melting: Correlating microstructure
evolution and aqueous corrosion behavior with Co content. Metals 2016, 6, 46. [CrossRef]
82. Lekatou, A.; Sfikas, A.K.; Karantzalis, A.E.; Sioulas, D. Microstructure and corrosion performance of Al-32%Co alloys. Corros. Sci.
2012, 63, 193–209. [CrossRef]
83. Palcut, M.; Priputen, P.; Kusý, M.; Janovec, J. Corrosion behaviour of Al-29at%Co alloy in aqueous NaCl. Corros. Sci. 2013, 75,
461–466. [CrossRef]
84. Palcut, M.; Priputen, P.; Šalgó, K.; Janovec, J. Phase constitution and corrosion resistance of Al-Co alloys. Mater. Chem. Phys. 2015,
166, 95–104. [CrossRef]
85. Eckert, J.; Scudino, S.; Stoica, M.; Kenzari, S.; Sales, M. Mechanical engineering properties of CMAs. In Complex Metallic Alloys:
Fundamentals and Applications; Dubois, J.-M., Belin-Ferré, E., Eds.; Wiley-VCH: Hoboken, NJ, USA, 2011; pp. 273–315. [CrossRef]
86. Priputen, P.; Palcut, M.; Babinec, M.; Mišík, J.; Černičková, I.; Janovec, J. Correlation between microstructure and corrosion
behavior of near-equilibrium Al-Co alloys in various environments. J. Mater. Eng. Perform. 2017, 26, 3970–3976. [CrossRef]
87. Meier, M.; Ledieu, J.; De Weerd, M.-C.; Huang, Y.-T.; Abreu, G.J.P.; Pussi, K.; Diehl, R.; Mazet, T.; Fournée, V.; Gaudry, E. Interplay
between bulk atomic clusters and surface structure in complex intermetallic compounds: The case study of the Al5 Co2 (001)
surface. Phys. Rev. B 2015, 91, 085414. [CrossRef]
88. Anand, K.; Fournée, V.; Prevot, G.; Ledieu, J.; Gaudry, E. Nonwetting Behavior of Al-Co Quasicrystalline Approximants Owing to
Their Unique Electronic Structures. ACS Appl. Mater. Interfaces 2020, 12, 15793–15801. [CrossRef] [PubMed]
89. Demange, V.; Machizaud, F.; Dubois, J.M.; Anderegg, J.W.; Thiel, P.A.; Sordelet, D.J. New approximants in the Al–Cr–Fe system
and their oxidation resistance. J. Alloys Compd. 2002, 342, 24–29. [CrossRef]
85
Materials 2021, 14, 5418
90. Li, R.T.; Murugan, V.K.; Dong, Z.L.; Khor, K.A. Comparative Study on the Corrosion Resistance of Al–Cr–Fe Alloy Containing
Quasicrystals and Pure Al. J. Mater. Sci. Technol. 2016, 32, 1054–1058. [CrossRef]
91. Ott, N.; Beni, A.; Ulrich, A.; Ludwig, C.; Schmutz, P. Flow microcapillary plasma mass spectrometry-based investigation of new
Al–Cr–Fe complex metallic alloy passivation. Talanta 2014, 120, 230–238. [CrossRef]
92. Beni, A.; Ott, N.; Caporali, S.; Guseva, O.; Schmutz, P. Passivation/precipitation mechanisms of Al-Cr-Fe complex metallic alloys
in acidic chloride containing electrolyte. Electrochim. Acta 2015, 179, 411–422. [CrossRef]
93. Ura-Bińczyk, E.; Homazava, N.; Ulrich, A.; Hauert, R.; Lewandowska, M.; Kurzydlowski, K.J.; Schmutz, P. Passivation of Al-Cr-Fe
and Al-Cu-Fe-Cr complex metallic alloys in 1 M H2 SO4 and 1 M NaOH solutions. Corros. Sci. 2011, 53, 1825–1837. [CrossRef]
94. Rüdiger, A.; Köster, U. Corrosion of Al-Cu-Fe quasicrystals and related crystalline phases. J. Non-Cryst. Solids 1999, 250–252,
898–902. [CrossRef]
95. Huttunen-Saarivirta, E.; Tiainen, T. Corrosion behaviour of Al-Cu-Fe alloys containing a quasicrystalline phase. Mater. Chem.
Phys. 2004, 85, 383–395. [CrossRef]
96. Huttunen-Saarivirta, E. Microstructure, fabrication and properties of quasicrystalline Al–Cu–Fe alloys: A review. J. Alloys Compd.
2004, 363, 150–174. [CrossRef]
97. Erlebacher, J.; Aziz, M.J.; Karma, A.; Dimitrov, N.; Sieradzki, K. Evolution of nanoporosity in dealloying. Nature 2001, 410,
450–453. [CrossRef] [PubMed]
98. Battezzati, L.; Scaglione, F. De-alloying of rapidly solidified amorphous and crystalline alloys. J. Alloys Compd. 2011, 509S, S8–S12. [CrossRef]
99. Yu, J.; Ding, Y.; Xu, C.; Inoue, A.; Sakurai, T.; Chen, M. Nanoporous Metals by Dealloying Multicomponent Metallic Glasses.
Chem. Mater. 2008, 20, 4548–4550. [CrossRef]
100. Mishra, S.S.; Pandey, S.K.; Yadav, T.P.; Srivastava, O.N. Influence of chemical leaching on Al-Cu-Co decagonal quasicrystals.
Mater. Chem. Phys. 2017, 200, 23–32. [CrossRef]
101. Kalai Vani, V.; Kwon, O.J.; Hong, S.M.; Fleury, S.M. Synthesis of porous Cu from Al-Cu-Co decagonal quasicrystalline alloys.
Philos. Mag. 2011, 91, 2920–2928. [CrossRef]
102. Zhang, Q.; Zhang, Z. On the electrochemical dealloying of Al-based alloys in a NaCl aqueous solution. Phys. Chem. Chem. Phys.
2010, 12, 1453–1472. [CrossRef]
103. Sieradzki, K.; Dimitrov, N.; Movrin, D.; McCall, C.; Vasiljevic, N.; Erlebacher, J. The Dealloying Critical Potential. J. Electrochem.
Soc. 2002, 149, B370–B377. [CrossRef]
104. Palcut, M.; Ďuriška, L.; Špoták, M.; Vrbovský, M.; Gerhátová, Ž.; Černičková, I.; Janovec, J. Electrochemical corrosion of Al-Pd
alloys in HCl and NaOH solutions. J. Min. Metall. B 2017, 53, 333–340. [CrossRef]
105. Ďuriška, L.; Palcut, M.; Špoták, M.; Černičková, I.; Gondek, J.; Priputen, P.; Čička, R.; Janičkovič, D.; Janovec, J. Microstructure,
phase occurrence and corrosion behavior of as-solidified and as-annealed Al–Pd alloys. J. Mater. Eng. Perform. 2018, 27,
1601–1613. [CrossRef]
106. Kelly, R.G.; Scully, J.R.; Shoesmith, D.W.; Buchheit, R.G. Electrochemical Techniques in Corrosion Science and Engineering; Marcel
Dekker Inc.: New York, NY, USA, 2003. [CrossRef]
107. Palcut, M.; Ďuriška, L.; Černičková, I.; Brunovská, S.; Gerhátová, Ž.; Sahul, M.; Čaplovič, L’.; Janovec, J. Relationship between
Phase Occurrence, Chemical Composition, and Corrosion Behavior of as-Solidified Al–Pd–Co Alloys. Materials 2019, 12, 1661.
[CrossRef]
108. Sukhova, O.V.; Polonskyy, V.A. Structure and corrosion of quasicrystalline cast Al-Co-Ni and Al-Fe-Ni alloys in aqueous NaCl
solution. East Eur. J. Phys. 2020, 3, 5–10. [CrossRef]
109. Babilas, R.; Bajorek, A.; Spilka, M.; Radoń, A.; Łoński, W. Structure and corrosion resistance of Al–Cu–Fe alloys. Prog. Nat. Sci.
Mater. 2020, 30, 393–401. [CrossRef]
110. Debili, M.Y.; Sassane, N.; Boukhris, N. Structure and corrosion behavior of Al-Co-Ti alloy system. Anti Corros. Methods Mater.
2017, 64, 443–451. [CrossRef]
111. Birbilis, N.; Buchheit, R.G. Electrochemical Characteristics of Intermetallic Phases in Aluminum Alloys: An experimental survey
and discussion. J. Electrochem. Soc. 2005, 152, B140–B151. [CrossRef]
86
materials
Article
Intergranular Corrosion and Microstructural Evolution in a
Newly Designed Al-6Mg Alloy
Kweon-Hoon Choi 1,2 , Bong-Hwan Kim 2, *, Da-Bin Lee 2 , Seung-Yoon Yang 2 , Nam-Seok Kim 2 , Seong-Ho Ha 2 ,
Young-Ok Yoon 2 , Hyun-Kyu Lim 2 and Shae-Kwang Kim 2
1 Department of Industrial Materials and Smart Manufacturing Engineering, Korea University of Science and
Technology, Daejeon 34113, Korea; [email protected]
2 Advanced Process and Materials R&BD Group, Korea Institute of Industrial Technology (KITECH),
Cheonan 31056, Korea; [email protected] (D.-B.L.); [email protected] (S.-Y.Y.);
[email protected] (N.-S.K.); [email protected] (S.-H.H.); [email protected] (Y.-O.Y.);
[email protected] (H.-K.L.); [email protected] (S.-K.K.)
* Correspondence: [email protected]; Tel.: +82-32-850-0440
Abstract: In this work, the microstructure and corrosion behavior of a novel Al-6Mg alloy were
investigated. The alloy was prepared by casting from pure Al and Mg+Al2 Ca master alloy. The
ingots were homogenized at 420 ◦ C for 8 h, hot-extruded and cold-rolled with 20% reduction (CR20
alloy) and 50% reduction (CR50 alloy). The CR50 alloy exhibited a higher value of intergranular
misorientation due to a higher cold rolling reduction ratio. The average grain sizes were 19 ± 7 μm
and 17 ± 9 μm for the CR20 and CR50 alloys, respectively. An intergranular corrosion (IGC) behavior
was investigated after sensitization by a nitric acid mass-loss test (ASTM G67). The mass losses of
both the CR20 and CR50 alloys were similar at early periods of sensitization, however, the CR20
alloy became more susceptible to IGC as the sensitization time increased. Grain size and β phase
Citation: Choi, K.-H.; Kim, B.-H.;
precipitation were two critical factors influencing the IGC behavior of this alloy system.
Lee, D.-B.; Yang, S.-Y.; Kim, N.-S.; Ha,
S.-H.; Yoon, Y.-O.; Lim, H.-K.; Kim,
S.-K. Intergranular Corrosion and
Keywords: light alloys; Al-Mg; high-strength; mechanical properties; intergranular corrosion; precipitation
Microstructural Evolution in a Newly
Designed Al-6Mg Alloy. Materials
2021, 14, 3314. https://doi.org/
10.3390/ma14123314 1. Introduction
The 5xxx series Al-Mg alloys are widely used materials in the automotive industry
Academic Editor: Marián Palcut due to their high strength-to-weight ratios, weldability, and good corrosion resistance [1–3].
Strength and ductility can be improved by adding solute Mg to the alloys because of
Received: 11 May 2021 the solid solution strengthening mechanism [4,5]. Nevertheless, as the amount of Mg
Accepted: 9 June 2021
increases, a selective oxidation prevails and causes a formation of oxide inclusions at
Published: 15 June 2021
elevated temperatures [6]. The rapid oxidation of magnesium alloys can be suppressed
by additions of Ca [7]. With this background, a new alloying strategy has been developed.
Publisher’s Note: MDPI stays neutral
The strategy uses an Mg + Al2 Ca master alloy instead of pure Mg during casting. The
with regard to jurisdictional claims in
Mg + Al2 Ca master alloy improves the oxidation resistance of Al-Mg alloys by forming a
published maps and institutional affil-
protective CaO/MgO mixed layer on the surface [8,9]
iations.
Previous studies have shown that anodic β-phase (β-Mg2 Al3 ) precipitation along grain
boundaries is an important factor affecting the intergranular corrosion (IGC) susceptibility
of Al-Mg alloy [6–8]. The Mg segregation leads to anodic β-phase formation at 50–200 ◦ C.
The β-phase precipitation is referred to as sensitization [10–13]. The β-phase formation is
Copyright: © 2021 by the authors.
usually observed at grain boundaries (GB), sub-grains, intermetallics, dislocations and other
Licensee MDPI, Basel, Switzerland.
defects. Generally, the sequences of β precipitation have been reported as follows [14–16]:
This article is an open access article
Solid solution α → Guinier–Peston zones → β” → β’ → β precipitation
distributed under the terms and
The β phases are electrochemically active compared to the Al matrix which leads to
conditions of the Creative Commons
galvanic corrosion. The distribution and morphology of β precipitates affect the suscep-
Attribution (CC BY) license (https://
tibility of the alloy to IGC. To evaluate the IGC susceptibility, a nitric acid mass-loss test
creativecommons.org/licenses/by/
4.0/).
(NAMLT, ASTM G67) can be used to measure the corrosion rate.
87
In this research, the IGC susceptibility of sensitized Al-6Mg alloy with two different
cold rolling conditions was studied. It is not yet clearly understood how the cold rolling
process affects the corrosion behavior [17]. D’Antuono et al. observed that an increased
rolling reduction increased the growth rate of β precipitation due to lowering nucleation
temperature [18]. Additionally, an initial β precipitation was observed preferentially at
low-angle grain boundaries rather than high-angle grain boundaries [19]. On the contrary,
Wang et al. found that the maximum corrosion depth decreases with increasing cold rolling
reduction ratio. They also stated that larger thickness reductions are attributable to an
increased number of small-sized grains formed at the grain boundary, which can eventually
break off the continuity of corrosion.
This paper aims to figure out the effect of cold rolling and sensitization treatment on
the IGC susceptibility of a newly designed Al-6Mg alloy. In this study, a microstructure
evolution of the alloy was analyzed by electron backscattered diffraction (EBSD). The
continuity of β-phase precipitation at grain boundary was studied by scanning electron mi-
croscopy (SEM) and transmission electron microscopy (TEM). The effects of the cold rolling
reduction ratio and sensitization heat treatment on the IGC susceptibility are discussed.
Alloy Si Fe Cu Mg Mn Ca Al
Al-6Mg 0.05 0.07 <0.01 6.04 0.03 ~0.02 Bal.
88
Materials 2021, 14, 3314
3. Results
3.1. Microstructure
Inverse pole figure (IPF) maps of newly developed Al-6Mg with two different cold
rolling conditions are shown in Figure 1. The results reveal that the CR20 alloy has more
equiaxed grains compared to the CR50 alloy. The CR50 alloy has a deformed microstructure
due to a higher cold rolling reduction ratio. Figure 2 shows the orientation distribution
functions (ODF) of the CR20 and the CR50 alloy. Typical deformation textures of the
cold-rolled Al-6Mg alloys are also shown in Figure 2. As the cold rolling reduction ratio
increased, more deformed textures tended to be obtained.
Figure 1. Inverse pole figure (IPF) maps of the new Al-6Mg CR20 (a) and CR50 alloys (b).
Figure 2. ODF sections of the new Al-6Mg CR20 (a) and CR50 alloys (b). The triangle represents Brass texture, circle is
Copper texture, and the square is used for S texture.
Miller indices of the texture components for rolled samples are listed in Table 2. In the
previous studies it was reported that the fraction of deformation texture components (Brass
{110} <112>, Copper {112} <111> and S {123} <634>) increased while the recrystallization
texture (Cube {100} <010> and Goss {011} <100>) did not change with increasing cold
89
Materials 2021, 14, 3314
rolling reduction ratio in the Al-Mg alloy [22,23]. Figure 2a indicates that the CR20 alloy
has an evolution of copper texture in Φ2 = 45 section. The CR50 alloy shows a strong
copper texture (Figure 2b). In addition, the CR50 alloy shows a stronger brass and S texture
compared to the CR20 alloy. The results are in accordance with previous research [22,23].
However, the recrystallization texture was not found in the newly designed Al-6Mg alloy,
which shows a partial disagreement with previous studies [22,23]. Therefore, the texture of
the newly developed Al-6Mg alloy is yet to be completely understood.
Figure 3. KAM maps of the CR20 (a) and CR50 alloys (b).
90
Materials 2021, 14, 3314
Figure 4. Grain size area histograms of the CR20 (a) and CR50 alloys (b).
Figure 5. NAMLT results of the CR20 alloy and CR50 alloy with different aging times at 100 ◦ C.
91
Materials 2021, 14, 3314
Figure 6. The surface of the CR20 alloy after NAMLT at different sensitization times: (a) 0 h, (b) 7 h, (c) 48 h, (d) 144 h, and
(e) 207 h.
Figure 8 shows the maximum corrosion depth of both CR20 and CR50 alloys. The
values are similar at the maximum sensitization time. On the other hand, the CR20 alloy
shows a higher mass-loss rate compared to the CR50 alloy (Figure 5).
The continuity of β precipitation at grain boundary is critical for the IGC depth. The
grain size, on the other hand, is a more important factor in the mass-loss rate at the early
periods of sensitization. While there is not a significant difference between the alloys in
maximum IGC depth, the NAMLT results show a higher mass-loss rate in the CR20 alloy
after the long heat treatment time. This means that the CR20 alloy is more susceptible
to IGC.
92
Materials 2021, 14, 3314
Figure 7. The surface of the CR50 alloy after NAMLT at different sensitization times: (a) 0 h, (b) 7 h, (c) 48 h, (d) 144 h, and
(e) 207 h.
93
Materials 2021, 14, 3314
94
Materials 2021, 14, 3314
Figure 10. Close inspection of β precipitation covered grain boundary: (a) CR20 alloy, Sensi. 48 h, (b) CR50 alloy, Sensi.
48 h.
Figure 11. Close inspection of β precipitation covered grain boundary: (a) CR20 alloy, Sensi. 207 h, (b) CR50 alloy, Sensi.
207 h.
Previous studies found that the precipitate growth rates increased with rolling reduc-
tion [18,19]. A high density of dislocations can lower the activation energy, which most
likely initiates the precipitation in the rolled specimen [18]. Increasing the dislocation
95
Materials 2021, 14, 3314
Figure 12. Thickness of β-phase precipitates in the CR20, and CR50 alloys sensitized for 48 and
207 h, respectively.
96
Materials 2021, 14, 3314
Figure 13. TEM of β precipitates in: (a) CR20 alloy, 48 h; (b) CR50 alloy, 48 h; (c) CR20 alloy, 207 h; (d) CR50 alloy, 207 h;
and (e) TEM-EDS (Mg element) CR20 alloy, 48 h.
4. Discussion
In this research, we found that both grain size and continuity of β precipitation at
grain boundaries are important factors affecting the Al-Mg IGC susceptibility.
The TEM image in Figure 13 shows that the β precipitates are much thicker in the
CR50 alloy at an early period of sensitization. The β precipitates thickness, however, is al-
most the same at long-term sensitization. Some researchers suggested that grain boundary
misorientation is a crucial factor for the growth rate and the final size of β precipitation.
These factors affect the continuity of β precipitation at grain boundaries [10,12,26,29–34].
Wang et al. concluded that some grain boundaries, e.g., low-angle grain boundaries gener-
ated by plastic deformation, are not susceptible to IGC [28]. On the other hand, D’Antuono
reported that although the β precipitation was preferentially formed at low-angle grain
boundary, the final size of precipitation was larger at high-angle grain boundary [18].
The influence of grain boundary plane orientation was reported to affect the continuity
of precipitation. It was found that grain boundary (GB) planes close to {110} direction
facilitate the β precipitation while the GB plane near {100} direction may be resistant to
β precipitation [31,32].
Previous studies showed that the rolled specimen had a high resistance to IGC coming
from the confluence of refined grain size and the fraction of low-angle grain bound-
aries [17,25,26]. In this study, it was revealed that the effect of grain size on IGC needs to
be considered depending on the sensitization heat treatment which affects the formation of
anodic β-Mg2 Al3 precipitation at the grain boundary. High dislocation density induced by
cold rolling facilitates the precipitate growth rates. The formation of anodic β-Mg2 Al3 is
affected by temperature and the presence of prior strain [35,36]. The increased dislocation
density tends to lower the nucleation temperature and reduce Mg diffusion at a lower
temperature [18,19]. These factors are reflected in increasing the susceptibility of the CR50
alloy in the early period of sensitization. In this situation, the dislocation density and
grain boundary type affect the IGC susceptibility more significantly compared to the grain
size. On the other hand, the grain size affects the IGC susceptibility of cold-rolled Al-6Mg
alloy more dramatically than the grain boundary type. It was found that the large-grained
material tends to be more susceptible to IGC when the precipitation is continuously formed
at the grain boundary due to sufficient sensitization time.
97
Materials 2021, 14, 3314
5. Conclusions
This study has explored the IGC behavior of a newly designed Al-6Mg alloy with
two different cold rolling conditions. It was revealed that the grain size and the continuity
of β precipitation play an important role in IGC. The precipitation growth rate and final
size of precipitates affect how the grain boundary is covered by β precipitation. At the
early period of sensitization, the precipitation growth rate is a crucial factor in IGC. The
dislocation density and grain boundary orientation affect the precipitation growth rate.
The CR50 alloy has a slightly higher precipitation growth compared to the CR20 alloy
because of the high density of dislocations. This results in a higher maximum IGC depth.
However, the grain size effect is more dominant when the sensitization time is long enough
to cover the grain boundary by anodic β precipitation.
References
1. Polmear, I.J. Light Alloys, 4th ed.; Butterworth-Heinemann: Oxford, UK, 2005.
2. Sanders, R.E., Jr.; Hollinshead, P.A.; Simielli, E.A. Industrial Development of Non-Heat Treatable Aluminum Alloys. In
Proceedings of the 9th International Conference on Aluminium Alloys, Brisbane, Australia, 2–5 August 2004.
3. Kubota, M.; Nie, J.F.; Muddle, B.C. Characterisation of Precipitation Hardening Response and As-Quenched Microstructures in
Al-Mg(-Ag) Alloys. Mater. Trans. 2004, 45, 3256–3263. [CrossRef]
4. Lee, B.-H.; Kim, S.-H.; Park, J.-H.; Kim, H.-W.; Lee, J.-C. Role of Mg in simultaneously improving the strength and ductility of
Al–Mg alloys. Mater. Sci. Eng. A 2016, 657, 115–122. [CrossRef]
5. Huskins, E.L.; Cao, B.; Ramesh, K.T. Strengthening mechanisms in an Al–Mg alloy. Mater. Sci. Eng. A 2010, 527, 1292–1298.
[CrossRef]
6. Czerwinski, F. The oxidation behaviour of an AZ91D magnesium alloy at high temperatures. Acta Mater. 2002, 50, 2639–2654.
[CrossRef]
7. You, B.-S.; Park, W.-W.; Chung, I.-S. The effect of calcium additions on the oxidation behavior in magnesium alloys. Scr. Mater.
2000, 42, 1089–1094. [CrossRef]
8. Kim, B.-H.; Ha, S.-H.; Yoon, Y.-O.; Lim, H.-K.; Kim, S.K.; Kim, D.-H. Effect of Ca addition on selective oxidation of Al3Mg2 phase
in Al-5 mass% Mg alloy. Mater. Lett. 2018, 228, 108–111. [CrossRef]
9. Ha, S.-H.; Yoon, Y.-O.; Kim, B.-H.; Lim, H.-K.; Lee, T.-W.; Lim, S.-H.; Kim, S.K. Pilling-Bedworth Ratio Approach to Surface
Oxidation of Al–Mg Alloys Containing Al 2 Ca and Its Experimental Verification. Sci. Adv. Mater. 2018, 10, 697–700. [CrossRef]
10. Steiner, M.A.; Agnew, S.R. Modeling sensitization of Al–Mg alloys via β-phase precipitation kinetics. Scr. Mater. 2015, 102, 55–58.
[CrossRef]
11. Tan, L.; Allen, T.R. Effect of thermomechanical treatment on the corrosion of AA5083. Corros. Sci. 2010, 52, 548–554. [CrossRef]
12. Zhang, R.; Zhang, Y.; Yan, Y.; Thomas, S.; Davies, C.H.J.; Birbilis, N. The effect of reversion heat treatment on the degree of
sensitisation for aluminium alloy AA5083. Corros. Sci. 2017, 126, 324–333. [CrossRef]
13. Jain, S.; Hudson, J.L.; Scully, J.R. Effects of constituent particles and sensitization on surface spreading of intergranular corrosion
on a sensitized AA5083 alloy. Electrochim. Acta 2013, 108, 253–264. [CrossRef]
14. Jain, S.; Lim, M.L.C.; Hudson, J.L.; Scully, J.R. Spreading of intergranular corrosion on the surface of sensitized Al-4.4Mg alloys: A
general finding. Corros. Sci. 2012, 59, 136–147. [CrossRef]
15. Yan, J.; Hodge, A.M. Study of β precipitation and layer structure formation in Al 5083: The role of dispersoids and grain
boundaries. J. Alloy. Compd. 2017, 703, 242–250. [CrossRef]
98
Materials 2021, 14, 3314
16. Lim, M.L.C.; Matthews, R.; Oja, M.; Tryon, R.; Kelly, R.G.; Scully, J.R. Model to predict intergranular corrosion propagation in
three dimensions in AA5083-H131. Mater. Des. 2016, 96, 131–142. [CrossRef]
17. Lin, L.; Liu, Z.; Li, Y.; Han, X.; Chen, X. Effects of Severe Cold Rolling on Exfoliation Corrosion Behavior of Al-Zn-Mg-Cu-Cr
Alloy. J. Mater. Eng. Perform. 2012, 21, 1070–1075. [CrossRef]
18. Scotto D’Antuono, D.; Gaies, J.; Golumbfskie, W.; Taheri, M.L. Direct measurement of the effect of cold rolling on β phase
precipitation kinetics in 5xxx series aluminum alloys. Acta Mater. 2017, 123, 264–271. [CrossRef]
19. Scotto D’Antuono, D.; Gaies, J.; Golumbfskie, W.; Taheri, M.L. Grain boundary misorientation dependence of β phase precipitation
in an Al–Mg alloy. Scr. Mater. 2014, 76, 81–84. [CrossRef]
20. ASTM G67-04. Standard Test Method for Determining the Susceptibility to Intergranular Corrosion of 5XXX Series Aluminum Alloys by
Mass Loss After Exposure to Nitric Acid (NAMLT Test); ASTM International: West Conshohocken, PA, USA, 2004.
21. Bachmann, F.; Hielscher, R.; Schaeben, H. Grain detection from 2d and 3d EBSD data—Specification of the MTEX algorithm.
Ultramicroscopy 2011, 111, 1720–1733. [CrossRef] [PubMed]
22. Duan, X.; Jiang, H.; Mi, Z.; Cheng, L.; Wang, J. Reduce the Planar Anisotropy of AA6016 Aluminum Sheets by Texture and
Microstructure Control. Crystals 2020, 10, 1027. [CrossRef]
23. Choi, S.-H.; Choi, J.-K.; Kim, H.-W.; Kang, S.-B. Effect of reduction ratio on annealing texture and r-value directionality for a
cold-rolled Al–5% Mg alloy. Mater. Sci. Eng. A 2009, 519, 77–87. [CrossRef]
24. Takayama, Y.; Szpunar, J.A. Stored Energy and Taylor Factor Relation in an Al-Mg-Mn Alloy Sheet Worked by Continuous Cyclic
Bending. Mater. Trans. 2004, 45, 2316–2325. [CrossRef]
25. Wang, Z.; Zhu, F.; Zheng, K.; Jia, J.; Wei, Y.; Li, H.; Huang, L.; Zheng, Z. Effect of the thickness reduction on intergranular
corrosion in an under–aged Al–Mg–Si–Cu alloy during cold–rolling. Corros. Sci. 2018, 142, 201–212. [CrossRef]
26. Zhao, Y.; Polyakov, M.N.; Mecklenburg, M.; Kassner, M.E.; Hodge, A.M. The role of grain boundary plane orientation in the β
phase precipitation of an Al–Mg alloy. Scr. Mater. 2014, 89, 49–52. [CrossRef]
27. Zhang, R.; Knight, S.P.; Holtz, R.L.; Goswami, R.; Davies, C.H.J.; Birbilis, N. A Survey of Sensitization in 5xxx Series Aluminum
Alloys. Corrosion 2016, 72, 144–159. [CrossRef]
28. Gupta, R.K.; Zhang, R.; Davies, C.H.J.; Birbilis, N. Influence of Mg Content on the Sensitization and Corrosion of Al-xMg(-Mn)
Alloys. Corrosion 2013, 69, 1081–1087. [CrossRef]
29. Zhang, R.; Qiu, Y.; Qi, Y.; Birbilis, N. A closer inspection of a grain boundary immune to intergranular corrosion in a sensitised
Al-Mg alloy. Corros. Sci. 2018, 133, 1–5. [CrossRef]
30. Ding, Q.; Zhang, D.; Zuo, J.; Hou, S.; Zhuang, L.; Zhang, J. The effect of grain boundary character evolution on the intergranular
corrosion behavior of advanced Al-Mg-3wt%Zn alloy with Mg variation. Mater. Charact. 2018, 146, 47–54. [CrossRef]
31. Guo, C.; Zhang, H.; Wu, Z.; Wang, D.; Li, B.; Cui, J. Effects of Ag on the age hardening response and intergranular corrosion
resistance of Al-Mg alloys. Mater. Charact. 2019, 147, 84–92. [CrossRef]
32. Ramachandran, D.C.; Murugan, S.P.; Kim, Y.-M.; Kim, D.; Kim, G.-G.; Nam, D.-G.; Jeong, C.; Do Park, Y. Effect of Microstructural
Constituents on Fusion Zone Corrosion Properties of GMA Welded AA 5083 with Novel Al–Mg Welding Wires of High Mg
Contents. Met. Mater. Int. 2020, 26, 1341–1353. [CrossRef]
33. Yan, J.; Heckman, N.M.; Velasco, L.; Hodge, A.M. Improve sensitization and corrosion resistance of an Al-Mg alloy by optimization
of grain boundaries. Sci. Rep. 2016, 6, 26870. [CrossRef]
34. Holroyd NJ, H.; Burnett, T.L.; Seifi, M.; Lewandowski, J.J. Improved understanding of environment-induced cracking (EIC) of
sensitized 5XXX series aluminium alloys. Mater. Sci. Eng. A 2017, 682, 613–621. [CrossRef]
35. Nebtp, S.; Hamanai, D.; Cizeron, G. Calorimetric study of pre-precipitation and precipitation in Al-Mg alloy. Acta Metall. Mater.
1994, 43, 3583–3588. [CrossRef]
36. Starink, M.J.; Zahra, A.-M. β’ and β precipitation in an Al–Mg alloy studied by DSC and TEM. Acta Mater. 1998, 46, 3381–3397.
[CrossRef]
99
materials
Article
Effect of Natural Aging on the Stress Corrosion
Cracking Behavior of A201-T7 Aluminum Alloy
Mien-Chung Chen 1 , Ming-Che Wen 2 , Yang-Chun Chiu 1 , Tse-An Pan 1 , Yu-Chih Tzeng 3
and Sheng-Long Lee 1, *
1 Institute of Material Science and Engineering, National Central University, Taoyuan 320, Taiwan;
[email protected] (M.-C.C.); [email protected] (Y.-C.C.);
[email protected] (T.-A.P.)
2 Department of Mechanical Engineering, National Central University, Taoyuan 320, Taiwan;
[email protected]
3 Department of Power Vehicle and Systems Engineering, Chung-Cheng Institute of Technology,
National Defense University, Taoyuan 334, Taiwan; [email protected]
* Correspondence: [email protected]; Tel.: +886-3-4267-325
Abstract: The effect of natural aging on the stress corrosion cracking (SCC) of A201-T7 alloy
was investigated by the slow strain rate testing (SSRT), transmission electron microscopy (TEM),
scanning electron microscopy (SEM), differential scanning calorimetry (DSC), conductivity,
and polarization testing. The results indicated that natural aging could significantly improve
the resistance of the alloys to SCC. The ductility loss rate of the unaged alloy was 28%, while the
rates for the 24 h and 96 h aged alloys were both 5%. The conductivity of the as-quenched alloy was
30.54 (%IACS), and the conductivity of the 24 h and 96 h aged alloys were decreased to 28.85 and
28.65. After T7 tempering, the conductivity of the unaged, 24 h, and 96 h aged alloys were increased
to 32.54 (%IACS), 32.52 and 32.45. Besides, the enthalpy change of the 24 h and 96 h aged alloys
increased by 36% and 37% compared to the unaged alloy. The clustering of the solute atoms would
evidently be enhanced with the increasing time of natural aging. Natural aging after quenching is
essential to improve the alloy’s resistance to SCC. It might be due to the prevention of the formation
of the precipitation free zone (PFZ) after T7 tempering.
Keywords: Al-Cu-Mg-Ag alloy; natural aging; stress corrosion cracking; SSRT; PFZ
1. Introduction
A201 (Al-4.5Cu-0.3Mg-0.7Ag) is a heat treatable aluminum alloy, which has the highest strength
among the casting aluminum alloys, so it has been used in the aerospace and military industries
for many years [1] The primary strengthening phases of A201 are θ’ and Ω, both having a similar
composition to that of CuAl2 [2,3]. The crystal structure of θ’ is tetragonal and with a = b = 0.414 nm
and c = 0.580 nm, forming large rectangular or octagonal plates parallel to the {100}α plane of the matrix
α phase [4]. The Ω phase has a face-centered orthorhombic structure, with a = 0.496 nm, b = 0.859 nm
and c = 0.848 nm, which forms hexagonal plate-like precipitates on the {111}α plane of the matrix
α phase [5–8].
To enhance the mechanical properties, especially the tensile strength, a T6 temper treatment
(solution heat treated then artificially aged) is usually applied in heat treatable alloys [1]. However,
for high strength Al-Cu-Mg (2XX series) or Al-Zn-Mg-Cu (7XXX series) alloys, T6 temper treatment is
not recommended because it will increase the susceptibility of the alloy to stress corrosion cracking
(SCC) [9–14]. SSC can occur when alloys are simultaneously subjected to stress and corrosive
environments. Burleigh [15] specified three SCC mechanisms for aluminum alloys, including anodic
dissolution, hydrogen-induced cracking, and the brittle passive film’s rupturing. Speidel [16] have
indicated that anodic dissolution is the primary mechanism of SCC in Al-Cu alloys. Misra [17] showed
that Al-Cu alloys formed a precipitate free zone (PFZ) along the grain boundary following artificial
aging, and this zone acts as an anode relative to the base of alloy. Eventually, under a corrosive
environment, the grain boundary corrodes quickly and resulting in grain boundary cracking of the alloy.
T7 tempering (solution heat treatment then overaging) is recommended to lower the susceptibility
of high strength Al-Cu-Mg (2XX series) or Al-Zn-Mg-Cu (7XXX series) alloys to SCC [1]. In the AA7075
(Al-Zn-Mg-Cu) alloy, for example, the primary strengthening phase is η (MgZn2 ), which has the lowest
potential compared to the α matrix and PFZ [18]. The T7 temper coarsens the precipitation, resulting in
a discontinuous structure along the grain boundary, thereby decreasing the alloy’s susceptibility to
SCC [19]. However, for a high strength Al-Cu-Mg alloy, PFZ has the lowest potential than the CuAl2
and α matrix. The inhibition of the formation of PFZ is the primary way to improve the resistance of
the Al-Cu-Mg alloy to SCC [20].
Although the effect of aging on the SCC behavior of high strength aluminum alloys had been
investigated for decades [13–21], and these works mainly focused on the effect of single artificial aging
on the SCC, such as T6 (peak aging) or T7 (over aging) treatment. However, the lack of research on
multiple heat treatments (combined natural aging with artificial aging) is the primary purpose of this
work. Hence, the influence of natural aging on the SCC behavior, microstructure, and mechanical
properties of A201-T7 alloy were investigated in this work to find the feasibility of multiple heat
treatments. The results can provide a reference for the development of high-performance alloys with
lower SCC suspicious while the mechanical properties could be maintained.
Alloy Cu Mg Ag Ti Fe Si Al
A201 4.5 (0.1) * 0.3 (0.05) 0.7 (0.05) 0.3 (0.05) 0.05 (0.01) 0.03 (0.01) Balance
* Standard deviations are listed in parentheses.
102
Materials 2020, 13, 5631
103
Materials 2020, 13, 5631
Figure 1. Differential scanning calorimetry (DSC) profile of different naturally aged A201 alloys.
After T7 tempering, with the precipitation of the strengthening phases θ’ and Ω, there was an
increase in the hardness of all the alloys, NA0d, NA1d, and NA4d, to 71HRB. Evidenced that the
clustering of the solute atoms remained at the same level regardless of whether natural aging was
adopted or not, and that lattice distortion would be eliminated after T7 tempering. The percentage
changes [(HT7 −HNA )/HNA × 100%] of the alloys, NA0d, NA1d, and NA4d, were 112%, 53%, and 30%.
Obviously, the hardness did not increase as much with the extension of the natural aging time from
24 h to 96 h. In addition, there was no difference in the yielding stress (YS), ultimate tensile stress
(UTS), or elongation (EL) whether natural aging was adopted or not. The YS, UTS, and EL of the three
alloys were approximately 320 MPa, 397 MPa, and 3.5%, respectively, after T7 tempering.
104
Materials 2020, 13, 5631
Figure 2. Polarization curves of different naturally aged A201-T7 alloys in a 3.5% NaCl solution.
Table 5. Parameters of polarization test of different naturally aged A201-T7 alloys in a 3.5%
NaCl solution.
105
Materials 2020, 13, 5631
respectively, while the loss in the unaged alloy could be as much as to 27.8%. Moreover, the loss of
strength also showed the same tendency. The loss of strength of the unaged alloy was 15%, while the
24 h and 96 h aged alloys showed losses of 2.5% and 3.0%, respectively.
Alloy EL in Air EL in Salt UTS in Air UTS in Salt Water Escc −Eair UTSscc −UTSair
Eair × 100% UTSair × 100%
Notation Eair (%) Water Escc (%) UTSair (MPa) UTSscc (MPa)
NA0d 3.6 (0.1) * 2.6 (0.3) 399 (2.9) 339 (2.1) −27.8 −15.0
NA1d 3.7 (0.2) 3.5 (0.2) 395 (2.6) 385 (2.8) −5.4 −2.5
NA4d 3.5 (0.1) 3.3 (0.1) 397 (2.8) 385 (2.3) −5.7 −3.0
* Standard deviations are listed in parentheses.
The SSRT results indicated that natural aging before T7 tempering was essential, for it had the
great benefit of increased resistance of the A201-T7 alloy to SCC. It also showed that aging for 24 h was
sufficient. Extending the aging time further had no additional benefit. We would also like to remind
the reader that the slow strain rate testing (SSRT) might not be suitable to determine the SCC behavior
of alloys in the latest research due to the sub-critical cracking and to invalidate the SSRT results [25,26].
An examination of Figure 3 shows the fracture surface of the A201-T7 alloys after slow strain rate
testing. During SSRT in air, the fracture surfaces of the alloys, NA0d, NA1d, and NA4d, were similar
with many dimples of different shapes and sizes observed, implying that the fracture mechanism
was ductile fracturing. As a result, natural aging did not affect the elongation in the air. However,
the fracture surfaces of the alloys were quite different when SSRT was conducted in the 3.5% NaCl
solution. For the unaged alloy, nearly no dimples were observed, and the fracture mechanism was
brittle fracturing. For the 24 h and 96 h aged alloys, cleavages and dimples were observed in the
sample, and the fracture mechanism was a combination of ductile and brittle fracturing. The results
were consistent with the mechanical properties presented in Table 4.
106
Materials 2020, 13, 5631
Figure 3. Fracture surface of A201-T7 alloys after slow strain rate testing: (a) NA0d sample in air;
(b) NA1d sample in air; (c) NA4d sample in air; (d) NA0d sample in salt water; (e) NA1d sample in salt
water; (f) NA4d sample in salt water.
Figure 4. Schematic diagram of the diffraction pattern of θ’ and Ω strengthening phases along [011]α
zone in A201-T7 alloy.
107
Materials 2020, 13, 5631
Figure 5. TEM bright field images and the selected area electron diffraction pattern (SAED) of:
(a,d) Alloy NA0d; (b,e) Alloy NA1d; (c,f) Alloy NA4d after T7 tempering.
It is worth noting that the PFZ has the lowest potential compared to CuAl2 phases (θ’ and Ω) and
α matrix in the 2XX series and 2XXX series Al-Cu-Mg alloys [29]. To lower the galvanic corrosion effect,
the inhibition of the formation of PFZ can help prevent SCC [30]. The discussion above is supported by
the polarization test and slow strain rate test results, indicating that naturally aged alloys have better
SCC resistance than unaged alloys.
108
Materials 2020, 13, 5631
4. Conclusions
The effects of natural aging on stress corrosion cracking in A201-T7 alloy were investigated in this
study and the following conclusions can be drawn:
(1) For the as-quenched alloy, the conductivity decrease and the hardness increase during natural
aging. However, the conductivity and mechanical properties (hardness, strength, and elongation)
were unaffected by natural aging after T7 tempering.
(2) Natural aging improved the resistance of A201-T7 alloys to SCC. 24 h aging was sufficient.
Extending the aging time provided no additional benefit.
(3) In the unaged alloys, PFZ existed and brittle fractures could be found on the SCC fracture surface;
for the aged alloys, no PFZ existed, but a combination of fracture types with cleavages and
dimples could be observed on the fracture surface.
References
1. Davis, J.R. ASM Specialty Handbook: Aluminum and Aluminum Alloys; ASM International: Materials Park, OH,
USA, 1994.
2. Chester, R.J.; Polmear, I.J. TEM Investigation of Precipitates in Al-Cu-Mg-Ag and Al-Cu-Mg Alloys. Micron
1980, 11, 311–312. [CrossRef]
3. Kim, K.D.; Zhou, B.C.; Wolverton, C. Interfacial Stability of θ’/Al in Al-Cu Alloys. Scr. Mater. 2019, 159,
99–103. [CrossRef]
4. Phillips, V.A. High Resolution Electron Microscope Observations on Precipitations in Al-3.0% Cu Alloy.
Acta Mater. 1975, 23, 751–767. [CrossRef]
5. Purnendu, K.M. Influence of Micro-alloying with Silver on Microstructure and Mechanical Properties of
Al-Cu Alloy. Mater. Sci. Eng. A 2018, 722, 99–111.
6. Knowles, K.M.; Stobbs, W.M. The Structure of {111} Age-Hardening Precipitates in Al-Cu-Mg-Ag Alloys.
Acta Cryst. 1988, 44, 207–227. [CrossRef]
7. Ivan, Z.; Rustam, K. Aging Behavior of an Al-Cu-Mg Alloy. J. Alloys Compd. 2018, 759, 108–119.
8. Saeed, K.M.; Mohsen, K.; Roland, L. Mechanical Behavior and Texture Development of Over-aged and
Solution Treated Al-Cu-Mg Alloy during Multi-directional Forging. Mater. Charact. 2018, 135, 221–227.
9. Muddle, B.C.; Polmear, I.J. The Precipctate Ω Phase in Al-Cu-Mg-Ag Alloys. Acta Metal. 1989, 37, 777–789.
[CrossRef]
10. Garg, A.; Chang, Y.C.; Howe, J.M. Precipitation of the Ω Phase in an Al-4.0Cu-0.5Mg Alloy. Scr. Mater. 1990,
24, 677–680. [CrossRef]
11. Hu, Y.C.; Liu, Z.Y.; Zhao, Q.; Bai, S.; Liu, F. P-Texture Effect on the Fatigue Crack Propagation Resistance in
an Al-Cu-Mg Alloy Bearing a Small Amount of Silver. Materials 2018, 11, 2481. [CrossRef]
12. ASTM B917. Standard Practice for Heat Treatment of Aluminum-Alloy Castings form All Processes; ASTM
International: West Conshohocken, PA, USA, 2020.
13. Rajan, K.; Wallace, W.; Beddoes, J.C. Microstructure Study of a High Strength Stress-Corrosion Resistant 7075
Aluminum Alloy. J. Mater. Sci. 1982, 17, 2817–2848. [CrossRef]
14. Islam, M.U.; Wallace, W. Retrogression and Reaging Response of 7475 Aluminum Alloy. Met. Technol. 1983,
10, 386–392. [CrossRef]
109
Materials 2020, 13, 5631
15. Burleigh, T.D. The Postulated Mechanisms for Stress Corrosion Cracking of Aluminum Alloys. Corrosion
1991, 47, 89–98. [CrossRef]
16. Speidel, M.O.; Hyatt, M.V. Advances in Corrosion Science and Technology; Springer: Boston, MA, USA, 1972.
17. Misra, M.S.; Oswalt, K.J. Corrosion Behavior of Al-Cu-Mg-Ag (201) Alloy. Met. Eng. Q. 1976, 16, 39–44.
18. Alexander, I.I.; Zhang, B.; Wang, J.Q.; Han, E.H.; Ke, W.; Peter, C.O. SVET and SIET Study of Galvanic
Corrosion of Al/MgZn2 in Aqueous Solutions at Different pH. J. Electrochem. Soc. 2018, 165, 180–194.
19. Shi, Y.J.; Pan, Q.L. Influence of Alloyed Sc and Zr, and Heat Treatment on Microstructures and Stress
Corrosion Cracking of Al-Zn-Mg-Cu Alloys. Mater. Sci. Eng. A 2015, 621, 173–181. [CrossRef]
20. Lee, H.J.; Kim, Y.J.; Jeong, Y.; Kim, S.S. Effects of Testing Variables on Stress Corrosion Cracking Susceptibility
of Al 2024-T351. Corros. Sci. 2012, 55, 10–19. [CrossRef]
21. Cabrini, M.; Bocchi, S.; D’Urso, G.; Giardini, C.; Lorenzi, S.; Testa, C.; Pastore, T. Stress Corrosion Cracking of
Friction Stir-Welded AA-2024 T3 Alloy. Materials 2020, 13, 2610. [CrossRef]
22. ASTM B557. Standard Test Methods for Tension Testing Wrought and Cast Aluminum- and Magnesium-Alloy
Products; ASTM International: West Conshohocken, PA, USA, 2020.
23. Ivanov, R.; Deschamps, A.; Geuser, F.D. Clustering Kinetics during Natural Ageing of Al-Cu Based Alloys
with (Mg, Li) Additions. Acta Mater. 2018, 157, 186–195. [CrossRef]
24. Chang, C.H.; Lee, S.L.; Lin, J.C.; Yeh, M.S.; Jeng, R.R. Effect of Ag Content and Heat Treatment on The Stress
Corrosion Cracking of Al-4.6Cu-0.3Mg alloy. Mater. Chem. Phys. 2005, 91, 454–462. [CrossRef]
25. Martínez-Pañeda, E.; Harris, Z.D.; Fuentes-Alonso, S.; Scully, J.R.; Burns, J.T. On the suitability of slow
strain rate tensile testing for assessing hydrogen embrittlement susceptibility. Corros. Sci. 2019, 163, 108291.
[CrossRef]
26. Nikolaos, D.A.; Christina, C.; Panagiotis, S.; Stavros, K.K. Synergy of Corrosion-induced Micro-cracking and
Hydrogen Embrittlement on The Structural Integrity of Aluminum Alloy (Al-Cu-Mg) 2024. Corros. Sci. 2017,
121, 32–42.
27. Beffort, O.; Solenthaler, C.; Uggowitzer, P.J.; Speidel, M.O. High toughness and high strength spray-deposited
AlCuMgAg-base alloys for use at moderately elevated temperatures. Mater. Sci. Eng. A 1995, 191, 121–134.
[CrossRef]
28. Wang, H.S.; Jiang, B.; Zhang, J.Y.; Wang, N.H.; Yi, D.Q.; Wang, B.; Liu, H.Q. The Precipitation Behavior and
Mechanical Properties of Cast Al-4.5Cu-3.5Zn-0.5Mg Alloy. J. Alloys Compd. 2018, 768, 707–713. [CrossRef]
29. Qi, H.; Liu, X.Y.; Liang, S.X.; Zhang, X.L.; Cui, H.X.; Zheng, L.Y.; Gao, F.; Chen, Q.H. Mechanical Properties and
Corrosion Resistance of Al-Cu-Mg-Ag Heat-resistant Alloy Modified by Interrupted Aging. J. Alloys Compd.
2016, 657, 318–324. [CrossRef]
30. Liu, X.Y.; Li, M.J.; Gao, F.; Liang, S.H.; Zhang, X.L.; Cui, H.X. Effects of Aging Treatment on The Intergranular
Corrosion Behavior of Al-Cu-Mg-Ag alloy. J. Alloys Compd. 2015, 639, 263–267. [CrossRef]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional
affiliations.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
110
materials
Article
Corrosion Behavior of Cold-Formed AA5754 Alloy Sheets
Anna Dobkowska *, Agata Sotniczuk, Piotr Bazarnik, Jarosław Mizera and Halina Garbacz
Faculty of Materials Science and Engineering, Warsaw University of Technology, 02-507 Warsaw, Poland;
[email protected] (A.S.); [email protected] (P.B.); [email protected] (J.M.);
[email protected] (H.G.)
* Correspondence: [email protected]; Tel.: +48-22-234-8399
Abstract: In this work, the influence of bending an AA5457 alloy sheet and the resulting microstruc-
tural changes on its corrosion behavior was investigated. Scanning electron microscopy (SEM) and
transmission electron microscopy (TEM) were used to perform detailed microstructural analyses
of the alloy in its original form and after bending. After immersion in naturally-aged NaCl under
open-circuit conditions (0.5 M, adjusted to 3 by HCl), post-corrosion observations were made, and
electrochemical polarization measurements were performed to investigate the corrosion mechanisms
occurring on both surfaces. The results showed that the corrosion of AA5457 is a complex process
that mainly involves trenching around coarse Si-rich particles, crystallographically-grown large pits,
and the formation of multiple tiny pits around Si-rich nanoparticles. The experimental data showed
that bending AA5457 changed the shape and distribution of Si-rich coarse particles, cumulated a
higher dislocation density in the material, especially around Si-rich nanoparticles, and all of these
factors caused that corrosion behavior of the AA5754 in the bending area was lowered.
1. Introduction
Citation: Dobkowska, A.; Sotniczuk, Due to their high mechanical and anti-corrosion properties, low density, and good
A.; Bazarnik, P.; Mizera, J.; Garbacz, formability, the 5xxx series of Al alloys is widely used in many branches of industry,
H. Corrosion Behavior of particularly in the automotive industry [1–6]. Their chemical composition mainly includes
Cold-Formed AA5754 Alloy Sheets. Al and Mg (up to 7 wt.% Mg), along with secondary additions of Mn and trace amounts
Materials 2021, 14, 394. https:// of Cr [7]. Their corrosion resistance arises from a complex combination of factors, but it
doi.org/10.3390/ma14020394 strongly depends on the microstructural features of the alloy [8–10]. Several studies have
reported the role of intermetallic particles in the corrosion resistance of Al alloys, and these
Received: 29 December 2020
particles can be organized by their size and electrochemical behavior with respect to the
Accepted: 11 January 2021
Al matrix [11–16]. If the particles are relatively large (a few μm), they are classified as
Published: 14 January 2021
coarse particles, but nanoprecipitates also often form in Al-Mg alloys. Due to their various
electrochemical properties (anodic/cathodic), they can either suppress matrix dissolution
Publisher’s Note: MDPI stays neu-
tral with regard to jurisdictional clai-
or have only a small influence on the corrosion of Al alloys [17,18]. Particles rich in Zn or
ms in published maps and institutio-
Si (such as MgZn2 or Mg2 Si) are prone to dissolution [19]. Generally, Fe-rich particles, i.e.,
nal affiliations.
Aln (Fe, Mn), are nobler with respect to the Al matrix and decrease the localized corrosion
resistance of Al alloys depending on the content of other impurities [20–22]. The local
corrosion attack around intermetallic particles has been explained in terms of various
particle behaviors in corrosive media: particle fall-out, selective particle dissolution or
Copyright: © 2021 by the authors. Li- particle dealloying [19,23]. Recent works on the microstructure of alloys have revealed
censee MDPI, Basel, Switzerland. different types of coarse particles in Al-Mg alloys, depending on the additional alloying
This article is an open access article elements.
distributed under the terms and con-
The second type of local attack also occurs at small precipitates formed in 5xxx series
ditions of the Creative Commons At-
Al alloys. When the Mg content exceeds 3.0 wt.% and alloys are exposed to temperatures
tribution (CC BY) license (https://
of 50–225 ◦ C, Mg atoms preferentially diffuse from the supersaturated solid solution (α)
creativecommons.org/licenses/by/
to grain boundaries (GBs). Then, the variety of Mg-rich phases changes from metastable
4.0/).
111
(β and β ) to equilibrium β (Al3 Mg2 ) precipitates at GBs (sensitization) [24–26]. Al3 Mg2 ,
which behaves as an anode with respect to the Al matrix, plays a predominant role in the
corrosion resistance of Al-Mg alloys [19,27–29]; however, due to its small size (~100 nm or
smaller), it is difficult to characterize this phase in situ [30,31]. The quantitative detection
of the dissolution of β precipitates regarding the corrosion behavior of 5xxx alloys has been
thoroughly studied by Guan et al. [25]. Vignesh et al. [32] analyzed the susceptibility to
intergranular corrosion (IGC) of AA5083 and showed that the alloy’s susceptibility to IGC
may be reduced by grain structure refinement, dispersion, and the partial dissolution of
secondary Mg2 Al3 in the matrix. This has been confirmed in another study [33], which
showed that the pitting, IGC and exfoliation all depended on the precipitate distribution.
The study suggested that grain refinement can effectively enhance the IGC resistance of the
commercial AA5083-H111. Moreover, the selective dissolution of Al3 Mg2 phase in AA5083
also determines its resistance to stress corrosion cracking [34]. In 5xxx Al alloys, pitting is
explained in terms of the formation of a defective oxide film [35] and is strongly dependent
on the NaCl concentration and pH of the solution [36–38].
Since the 5xxx alloys do not respond to age hardening, previous reports regarding the
electrochemical behavior of 5xxx alloys have also focused on post-processing heat-treatment
and/or plastic deformation [39–41]. As stated by Neetu et al. [42], the subsequent annealing
of the AA5080 produced by multi-axial forging at room temperature drastically improved
its corrosion resistance. The homogenization heat treatment of AA5083-O plates after
casting had the optimal effect on the overall corrosion resistance of the material [40]. Various
plastically-deformed (i.e., friction stir processing, accumulative roll-bonding) Al-Mg alloys
have shown better corrosion resistance compared with their parent material [43,44]. In
contrast, extruded AA5083 showed similar overall corrosion resistance compared with
conventional Al-Mg alloys [45].
Clearly, the corrosion behavior of Al-Mg alloys is complex and requires more studies
to distinguish the dependence between the formation of various kinds of localized corro-
sion and their mutual interactions. Although Al-Mg alloys display fairly good corrosion
resistance [46,47], subjecting them to additional cold-forming to produce semi-products or
final elements may change the material’s properties and compromise their electrochemical
behavior. Therefore, the optimization of the corrosion performance of commercially appli-
cable Al-Mg alloys is of critical importance in terms of using these alloys for structurally
lightweight devices and when plastic deformation is necessary to obtain the desired shape
of a design element (i.e., bending). For the needs of this project, we produced part of a
construct using AA5457, which has an optimal combination of mechanical properties and
is lightweight, commercially available, and prone to plastic deformation. In this work,
Al-Mg type AA5754 alloy was used to produce the outer components of an experimental
prototype of a capsule that allows for teleperformance assessment tests and telerehabilita-
tion of sensory diseases (i.e., hearing, sight, smell, balance). The main goal of this study
was to analyze how cold-working can change corrosion resistance of AA5754 (O/H111
according to [48]).
2. Materials
A schematic of the construction design is given in Figure 1a. To prepare a final element
in the form of an adjustable angle plate, a 4 mm-thick sheet made of AA5754 (chemical
composition given in Table 1) was cut into strips and subsequently bent to 90◦ (Figure 1). The
main focus was to determine the influence of microstructural changes induced by bending
on the corrosion behavior of the original AA5754 sheet. To accomplish this, the corrosion
behavior in two specified areas of the final element was measured: in the undeformed area
located at the arm of the bent Sheet (A) and at the bent area (B), as shown in Figure 1b.
112
Materials 2021, 14, 394
Figure 1. The outer design of the transferable capsule used for teleperformance assessment tests and
telerehabilitation: (a), schematic of the outer elements, and (b), locations of corrosion tests.
Si Fe Cu Mn Mg Cr Zn Ti Al
max max max max max max max -
0.40 0.40 0.10 0.50 2.6–3.6 0.30 0.20 0.15 Bal.
3. Methodology
3.1. Microstructure Characterization
Microstructure characterization was performed in two areas (undeformed and bent
areas) using scanning electron microscopy (SEM, Hitachi SU8000, Tokyo, Japan) in back-
scattered electron mode (BSE) with an accelerating voltage of 5 kV. The microstructure was
observed on the cross-section of the Sheet in areas A and B shown in Figure 1b. Specimens
for SEM observations were prepared by grinding samples with up to 1200-grit SiC paper
and ion polishing with a Hitachi IM-4000 ion milling system. The ion milling is a damage-
free process that does not cause surface deformation, eliminates stresses and oxide layers;
therefore, the surface quality is good enough to observe the structure under channeling
contrast using SEM. Structural investigations were combined with an energy dispersive X-
ray (EDX) analysis of precipitates and intermetallic particles. To more thoroughly describe
the role of intermetallic particles, the density of coarse particles was calculated by dividing
the number of counted particles (N) by the studied area (A) of the specimen. Moreover,
their size (A—area) and standard deviation (SD ) were calculated using computer software
as per references [49,50].
Detailed microstructural observations of selected areas were performed using a trans-
mission electron microscope (TEM, JEOL 1200, Tokyo, Japan) operating at an accelerat-
ing voltage of 120 kV, and the EDX spectra of the nanoparticles were obtained using a
high-resolution scanning transmission electron microscope (STEM equipped with an EDX
spectrometer, Hitachi S5500). The TEM samples were prepared using twin-jet electropol-
ishing at 3 ◦ C/20 V with an electrolyte composed of perchloric acid (HClO4 ), ethanol and
2-butyxoethanol (type A2 purchased from Struers, Ltd., Ballerup, Denmark) Based on the
observations, the dislocation density (ρ) was calculated using the formula [51]:
N
ρ= (1)
Lr h
where N is the number of dislocation-line intersections, Lr is the total length of all lines,
and h is TEM sample thickness (h = 150 nm).
from analytical-grade reagents and distilled water. A FAS1 Gamry potentiostat (Gamry
Instruments, Warminster, PA, USA) was equipped with three electrodes: Pt as the counter
electrode, Ag/AgCl as the reference electrode (Ag/AgCl wire immersed in a saturated KCl
solution), and the measured sample as the working electrode. The corrosion potential (Ecorr )
was recorded during 3 or 6 h of immersion under open-circuit conditions. After immersion,
the corrosion rate of samples was determined using linear polarization resistance (LPR),
and potentiodynamic tests were conducted. LPR data were recorded over a potential range
of Ecorr ± 10 mV at a scan rate of 5 mV/min. After LPR, the samples were conditioned
for 3 min and potentiodynamic tests were performed at a scan rate of 5 mV/s, starting
from −0.5 V/OCP to 2 V/REF at room temperature. To prevent noise, all measurements
were conducted in a Faraday cage. The polarization curves were fitted using Gamry
Elchem software (Gamry Instruments, Warminster, PA, USA) in Tafel mode. To ensure the
reproducibility of the results, all measurements were repeated at least three times on the
same samples. The surfaces of the samples after immersion under open-circuit potential
were observed using a Hitachi SU8000 SEM.
4. Results
4.1. Microstructure Characterization
Figures 2 and 3 show BSE images of the AA5754 alloy in an undeformed area and a
deformed area, respectively. The BSE image shown in Figure 2a at a lower magnification
shows a typical granular microstructure with many large, irregularly-shaped intermetallic
particles randomly distributed in the alloy. The SEM-EDS analysis reveals that there are two
kinds of particles: the white particles marked as P_1 are Al, Mg, Mn, and Fe-rich particles
(Figure 2a,c), and the dark grey precipitates P_2 are composed of Mg, Al, and Si (Figure 2b,c).
In the deformed region (Figure 3), the subsequent bending altered the microstructure of
the material, and the distribution of the intermetallic particles also changed; more coarse
white precipitates were observed in the bending area (Figure 3a). Additionally, during
deformation, a significant strain was introduced into the material and resulting from this,
higher dislocation density occurred [42]. Therefore, a granular structure is not clearly
distinguishable in the BSE contrast image; however, many dislocations, especially multiple
slip bands, are visible (Figure 3b). Similar to the undeformed area, the coarse intermetallic
particles shown in Figure 3c marked as P_1 are mainly enriched in Mg, Al, Mn, and Fe
(Si was detected in the bending area), while those marked as P_2 are composed of Al, Mg,
and Si (Figure 3d).
Figure 2. SEM images of the microstructure of the AA5457 sheet in the undeformed area (A): (a), an
image showing the overall distribution of intermetallic particles, (b), image showing the morphology
and surrounding area of intermetallic particles, and (c), chemical analysis of the intermetallic particles
P_1 and P_2 marked in panels a and b.
114
Materials 2021, 14, 394
Figure 3. SEM images of the microstructure of the AA5457 sheet in the deformed area (B): (a) and (b),
images showing the overall distribution of intermetallic particles, (c), image showing the morphology
and area surrounding the intermetallic particles P_1 and P_2, and (d), the chemical analysis of the
intermetallic particles P_1 and P_2 marked in panel c.
The density of P_1 particles in the undeformed and deformed areas was comparable,
NA = 0.01 and NA = 0.02 (given in 1 × μm−2 ), for the undeformed and deformed areas,
respectively; however, after deformation, they showed smaller sizes (Table 2). The SEM
images indicate that there was an abrupt change in the density of P_2 particles before
and after deformation (compare Figures 2a and 3a). The calculations demonstrate that the
density of the P_2 particles changed to NA = 0.001 and NA = 0.01 (given in 1 × μm−2 ) in
the undeformed and deformed areas, respectively (Table 2). As a result of bending, the
observed particles were slightly smaller in size (Table 2).
Table 2. The mean surface area (A), standard deviation (SD ), and number per 1 μm2 (NA ) of coarse intermetallic particles
P_1 and P_2.
115
Materials 2021, 14, 394
Figure 4. TEM images of the microstructure of the AA5457 sheet in the undeformed area (A): (a,b), images showing the
overall grain boundaries and dislocation structure, and (c), image showing the overall distribution of nanosized intermetallic
particles. TEM images of the microstructure of the AA5457 sheet in the deformed area (B): (d,e), images showing the increase
in the dislocation density and the formation of cellular dislocation structures and slip bands, and (f), image showing an
increase in the dislocation density around nanosized intermetallic particles. The representative chemical composition of a
nanoparticle is presented below panel (f).
The results from the deformed region clearly show an increase in the number of
dislocations compared with the undeformed section. The dislocation density increased by
nearly an order of magnitude, from 4.5 × 1012 m−2 in the undeformed region (Figure 4b)
to 4.4 × 1013 m−2 in the deformed region (Figure 4d,e). This is a typical phenomenon for
deformed metallic materials. It should be noted that dislocations are distributed rather
uniformly; however, in some regions, the formation of characteristic cellular structures
(Figure 4d) and slip bands (Figure 4e) was observed. Moreover, an increase in the dislo-
cation density was also observed around Si and Mn-rich nanoparticles (Figure 4f). Hard
intermetallic particles pin dislocations during deformation, resulting in the formation of
many dislocations around such particles.
116
Materials 2021, 14, 394
Figure 5. The Ecorr measurements recorded during 6 h of immersion in 0.5 M NaCl (pH 3) in
the undeformed and deformed areas of AA5754. SEM observations were made after 3 and 6 h of
immersion, as marked by the dashed lines.
The potentiodynamic polarization curves recorded for the undeformed and deformed
samples of AA554 after 3 and 6 h of immersion under open-circuit conditions are given in
Figure 6. All curves exhibit similar trends with passive regions, which suggests an anodic
control of the corrosion processes. All curves exhibit a wide passive region when the current
density plateaued, extended by 0.04 V in the undeformed area (both 3 and 6 h immersion),
to around 0.05 V for the deformed area after 3 h of immersion, and 0.07 V for the deformed
area after 6 h of immersion [57] Simultaneously, there is a shift in the pitting current, ip ,
which is the lowest for the undeformed area (after both periods of immersion). The pitting
current, ip , shifted to higher values for the deformed areas after 3 h of immersion and
continued to increase over time. Afterwards, an abrupt change in the current density was
observed, which indicates the position of the pitting potential (Epit ) [25,42,58]. The pitting
potential for both undeformed samples has the same value of Epit = −0.68 V/REF (Table 3).
Slightly more positive pitting potentials were recorded for the deformed samples, with Epit
= −0.64 V/REF after 3 h of immersion. Increasing the immersion time to 6 h decreased the
pitting potential of the deformed area to Epit = −0.66 V/REF (after 6 h of immersion). The
numerical values of pitting corrosion resistance (Rpit ) may be described by the difference
between pitting potential and corrosion potential, Epit and Ecorr [59,60]
Figure 6. The electrochemical curves obtained after 3 and 6 h of immersion in 0.5 M NaCl (pH 3) for undeformed and
deformed areas of the AA5754: (a), potentiodynamic polarization curves, and (b), pitting susceptibility of analyzed materials
calculated using Formula (1).
117
Materials 2021, 14, 394
Table 3. The corrosion potential (Ecorr ), pitting potential (Epit ), corrosion current density (icorr ), and
corrosion rate calculated after 3 and 6 h of immersion in naturally-aerated 0.5 M NaCl (pH 3) for the
undeformed and deformed areas of AA5754.
Rpit = Ecorr − Epit (2)
Considering the data presented in Figure 6b, the smallest difference between Ecorr
and Epit was calculated for the undeformed area of the sample for both immersion periods
(Rpit = 0.16 after 3 h of immersion, and Rpit = 0.14 after 6 h of immersion), suggesting that
the undeformed area is less-resistant to localized corrosion.
Regardless of immersion time, the deformed area of the alloy showed the greatest Rpit
= 0.22V. The higher R and ip values of the deformed areas were attributed to the weakened
passivation properties [59,60].
The extrapolated from Tafel plot data show corrosion potentials of Ecorr = −0.84 V/REF
and Ecorr = −0.82 V/REF for the undeformed areas immersed for 3 and 6 h, respectively.
Corrosion potentials registered for the deformed areas had slightly lower values than
the corrosion potentials of the undeformed areas, with values of Ecorr = −0.88 V/REF
and Ecorr = −0.86 V/REF after 3 and 6 h of immersion, respectively (Table 3). To further
view the overall corrosion behavior of the analyzed materials, LPR data were recorded
after 3 and 6 h of immersion. The corrosion rates after 3 h of immersion were 1.8 mpy
for the undeformed region, and almost twice as high (2.6 mpy) for the deformed region
of the material (Table 3). After 6 h of immersion, the corrosion rate in the undeformed
area increased slightly to 2.6 mpy, and the corrosion rate of the deformed area reached a
significantly higher value of 9.5 mpy.
118
Materials 2021, 14, 394
Figure 7. Post-corrosion observations of AA5754 in the undeformed area (A) after immersion under open-circuit conditions.
Images taken after 3 h of immersion showing: (a), the propagation of crystallographically-grown pits, (b), trenching around
Si-rich (P_2) particles, and (c) high-magnification image of tiny pits. Images taken after 6 h of immersion showing: (d),
the propagation of crystallographically-grown pits, (e), trenching around Si-rich (P_2) particles, and (f) the formation of
tiny pits.
Figure 8. Post-corrosion observations of AA5754 in the deformed area (B) after immersion under open-circuit conditions.
Images taken after 3 h of immersion showing: (a), the propagation of crystallographically-grown pits, (b), the trenching
around Si-rich (P_2) particles, and (c), a high-magnification image of tiny pits. Images taken after 6 h of immersion showing:
(d), trenching around Si-rich (P_2) particles, (e), corrosion products around Fe-rich (P_1) particles, and (f), the formation of
tiny pits.
SEM observations confirmed that after immersion in 0.5 M NaCl (pH 3) trenching
occurred around Si-rich particles over the entire analyzed sample surface. Besides intense
local corrosion around Si-rich particles, after 3 h of immersion, highly-pronounced local
pitting corrosion presented in the form of several superficial crystallographically-grown
pits in the original and deformed areas of AA5754 (Figures 7b and 8a) was formed. As
the experiment duration increased, more crystallographically-grown pits were observed
on the original material compared with the deformed area. Crystallographically-grown
119
Materials 2021, 14, 394
pits limited by perpendicular standing walls of the crystallographic lattice have been
previously observed on Al and Al alloys, and the pit growth kinetics have been previously
explained [62–66]. It was also demonstrated that in acidic chloride-containing solutions, the
general mode of pit propagation is corrosion tunneling [62]. Considering the deformation
effect on corrosion mechanisms on AA5754 after 3 h of immersion, minor differences were
observed regarding the shape of crystallographically-grown pits (compare Figures 7a and
8b), but not their amount. As the experiment duration increased, more crystallographically-
grown pits were observed on the surface of the original material, and some underwent
further lateral spreading, while on the deformed area after 6 h of immersion, only a few
crystallographically-grown pits were observed, suggesting that no new pits were grown on
the deformed area. The locally-formed crystallographic pits on the undeformed material
spread laterally on the surface, whilst the pits on the deformed sample seemed to propagate
into the depth of the material. Some pits were created around Si-rich particles (type P_2),
as shown in Figure 7e and the inset in Figure 8c. In many areas of the observed samples,
these kinds of pits were created after the trenches formed around P_2 particles.
Apart from the microgalvanic coupling between Si-rich particles and the matrix, or
the large crystallographically-grown pits, very tiny metastable ellipsoidal pits formed on
the surface of both areas (Figure 7c,f—undeformed areas after 3 and 6 h of immersion,
respectively; Figure 8c,f—deformed areas after 3 and 6 h of immersion, respectively).
These pits grew larger with the experiment duration (Figures 7f and 8f); however, when
comparing Figures 7c and 8c, it is visible that more and deeper pits were created on the
deformed area of the sample. Moreover, after 6 h of immersion, loosely-adhered corrosion
products were formed on the deformed area of the sample, especially in the locations where
two Fe-rich particles (P_1) were located relatively close to each other (Figure 8e).
5. Discussion
The results of this work show that the observed trenching around Si-rich coarse
particles depends on their number and distribution, and the increased amount of these
particles in the deformed area resulted in more localized corrosion initiation, which was
one of the reasons for its lower corrosion resistance.
As demonstrated in this work, the corrosion that occurred around Si-rich particles
in the deformed area of the material was more severe than the corrosion around the
same type of particles in the undeformed area of AA5457. This phenomenon is related
to the increased number of particles in the material after deformation and due to their
mutual interaction with the matrix, particularly particle-dislocation interactions. In our
opinion, the cumulative stress occurring during deformation was sufficient to cause particle
fragmentation. The increased number of hard coarse particles provided more locations
where trenching could occur, and the intensity of such trenching is related to particle-
dislocation interactions. As is commonly known, the hard coarse particles are the places
where deformation-induced dislocation movement is blocked, leading to the formation of
a high concentration of dislocations around them [67].
The chemical composition of the coarse particles is also important, considering their
corrosion behavior in chloride-containing solution. The Si-rich coarse intermetallics ob-
served in AA5754 also contained Mg and Al. Previous reports have clearly indicated that
the rapid dissolution of Mg from intermetallics occurs during exposure, which leads to the
rapid dealloying of these particles [21,24,61,68,69]. The results of previous studies [68,69]
explain the mechanism of Mg and Al dissolution from the coarse particles composed mainly
of Mg and Si in Al-Mg-Si alloys. The results of this research show that at the beginning
of immersion, the preferential and selective dissolution of Mg occurs, and thus, the Volta
potential of the particle changes, which also changes the nature of Mg-containing particles
from anodic to cathodic, forming a galvanic couple with the Al matrix. The results of this
work show that the increased trenching around Si-rich particles in the deformed area was
a result of the cracking of these particles induced by the cumulative stresses generated
during deformation. During deformation, the coarse hard particles blocked the dislocation
120
Materials 2021, 14, 394
movement; therefore, a higher dislocation density was formed near such particles. As the
locally-formed galvanic cells approach equilibrium, the cathodic particle forces the galvanic
dissolution of the anodic matrix, leading to matrix corrosion. The increased dislocation
density near the coarse particles promotes the corrosion reactions, leading to more intense
trenching around the particles rich in Mg and Si. The reason for this is a higher dislocation
density formed in the bent area of the sample, which promoted pit propagation in the
depth due to their less-ordered structure [58,70]. This hypothesis is indisputably confirmed
by the results presented in this work. It is also worth noting that Fe-rich particles remained
unreactive throughout the entire experiment; however, their presence, especially when
located relatively close to each other, may enhance corrosion product formation.
Another discrepancy that needs to be explained is that Aballe et al. [71] did not observe
any relationship between the formation of crystallographic pitting and the existence of
intermetallic particles, but Neetu et al. [42] claimed that pit initiation occurred at particles
rich in Fe and Mn. We observed spontaneously-formed pits on the undeformed samples;
however, in contrast to both works, our observations indicate that crystallographic pit
growth is often initiated in the areas occluded by secondary particles enriched with Si.
Moreover, this observation is supported by the fact that corrosion attack around Si-rich
particles occurred before crystallographically-grown pits were formed. In a separate set of
experiments, we observed trenching around Si-rich particles, but we did not observe any
crystallographically-grown pits on the samples immersed for one hour in the same solution
(data not published). Since pitting attack is considered to be an autocatalytic process in the
active areas, it is reasonable to assume that the higher number of Si-rich coarse particles
provided more sites were localized corrosion may be initiated. These places are prone to
coalescence and form laterally-spreading local corrosion.
The second type of corrosion that occurred on the AA5754 alloy is microgalvanic
corrosion between Si-rich nanoparticles and the alloy matrix. The observations showed that
tiny pits were created around Si-rich nanoparticles, and the corrosion damage around them
was more severe in the deformed area. Although the deformation-induced stress was not
sufficient to change the size or distribution of the observed nanoparticles, our experiments
clearly show that the nanoparticles in the deformed area with a high dislocation-stacking
fault density around them promoted the formation of many deep tiny pits.
6. Conclusions
Based on the results of this work, the following conclusions can be drawn:
• The AA5754 alloy undergoes microstructure-dependent corrosion attack.
• The bending that was applied to obtain the desired shape of the design elements made
from AA5754 affected the corrosion resistance of the alloy—the bent areas were more
susceptible to corrosion than the original material.
• The lower corrosion resistance of the bent areas was related to the bending-induced
microstructural changes, such as the increased density of Si-rich coarse particles and
nanoparticles, the increased dislocation density around them, and their mutual interactions.
121
Materials 2021, 14, 394
References
1. Ding, L.; Jia, Z.; Zhang, Z.; Sanders, R.E.; Liu, Q.; Yang, G. The natural aging and precipitation hardening behaviour of
Al-Mg-Si-Cu alloys with different Mg/Si ratios and Cu additions. Mater. Sci. Eng. A 2015, 627, 119–126. [CrossRef]
2. Miller, W.S.; Zhuang, L.; Bottema, J.; Wittebrood, A.J.; de Smet, P.; Haszler, A.; Vieregge, A. Recent development in aluminium
alloys for the automotive industry. Mater. Sci. Eng. 2000, A280, 37–49. Available online: www.elsevier.com/locate/msea (accessed
on 20 October 2020). [CrossRef]
3. Abid, T.; Boubertakh, A.; Hamamda, S. Effect of pre-aging and maturing on the precipitation hardening of an Al-Mg-Si alloy. J.
Alloys Compd. 2010, 490, 166–169. [CrossRef]
4. Lathabai, S.; Lloyd, P.G. The effect of scandium on the microstructure, mechanical properties and weldability of a cast Al-Mg
alloy. Acta Mater. 2002, 50, 4275–4292. [CrossRef]
5. Hirsch, J.; Al-Samman, T. Superior light metals by texture engineering: Optimized aluminum and magnesium alloys for
automotive applications. Acta Mater. 2013. [CrossRef]
6. Chehreh, A.B.; Grätzel, M.; Bergmann, J.P.; Walther, F. Effect of corrosion and surface finishing on fatigue behavior of friction stir
welded EN AW-5754 aluminum alloy using various tool configurations. Materials (Basel) 2020, 13, 3121. [CrossRef]
7. Conserva, M.; Leoni, M. Effect of thermal and thermo-mechanical processing on the properties of Al-Mg alloys. Metall. Trans. A
1975, 6, 189–195. [CrossRef]
8. Cavanaugh, M.K.; Birbilis, N.; Buchheit, R.G. Modeling pit initiation rate as a function of environment for Aluminum alloy
7075-T651. Electrochim. Acta 2012, 59, 336–345. [CrossRef]
9. Arrabal, R.; Mingo, B.; Pardo, A.; Mohedano, M.; Matykina, E.; Rodríguez, I. Pitting corrosion of rheocast A356 aluminium alloy
in 3.5wt.% NaCl solution. Corros. Sci. 2013, 73, 342–355. [CrossRef]
10. Włodarczyk, P.P.; Włodarczyk, B. Effect of hydrogen and absence of passive layer on corrosive properties of aluminum alloys.
Materials (Basel)) 2020, 13, 1580. [CrossRef]
11. Yasakau, K.A.; Zheludkevich, M.L.; Ferreira, M.G.S. Role of intermetallics in corrosion of aluminum alloys. Smart corrosion
protection. In Intermetallic Matrix Composites; Mitra, R.B.T.-I.M.C., Ed.; Woodhead Publishing: Cambridge, UK, 2018; pp. 425–462.
[CrossRef]
12. Buchheit, R.G. A Compilation of Corrosion Potentials Reported for Intermetallic Phases in Aluminum Alloys. J. Electrochem. Soc.
1995, 142, 3994–3996. [CrossRef]
13. Li, S.M.; Li, Y.D.; Zhang, Y.; Liu, J.H.; Yu, M. Effect of intermetallic phases on the anodic oxidation and corrosion of 5A06
aluminum alloy. Int. J. Miner. Metall. Mater. 2015, 22, 167–174. [CrossRef]
14. Ma, Y.; Wu, H.; Zhou, X.; Li, K.; Liao, Y.; Liang, Z.; Liu, L. Corrosion behavior of anodized Al-Cu-Li alloy: The role of intermetallic
particle-introduced film defects. Corros. Sci. 2019, 158. [CrossRef]
15. Yuan, D.; Chen, K.; Chen, S.; Zhou, L.; Chang, J.; Huang, L.; Yi, Y. Enhancing stress corrosion cracking resistance of low
Cu-containing Al-Zn-Mg-Cu alloys by slow quench rate. Mater. Des. 2019, 164, 107558. [CrossRef]
16. Revilla, R.I.; Verkens, D.; Rubben, T.; De Graeve, I. Corrosion and corrosion protection of additively manufactured aluminium
alloys—A critical review. Materials (Basel) 2020, 13, 4804. [CrossRef] [PubMed]
17. Davoodi, A.; Pan, J.; Leygraf, C.; Norgren, S. The Role of Intermetallic Particles in Localized Corrosion of an Aluminum Alloy
Studied by SKPFM and Integrated AFM/SECM. J. Electrochem. Soc. 2008, 155, C211. [CrossRef]
18. Buchheit, R.G.; Boger, R.K.; Carroll, M.C.; Leard, R.M.; Paglia, C.; Searles, J.L. The electrochemistry of intermetallic particles and
localized corrosion in Al alloys. JOM 2001, 53, 29–33. [CrossRef]
19. Birbilis, N.; Buchheit, R.G. Electrochemical Characteristics of Intermetallic Phases in Aluminum Alloys. J. Electrochem. Soc. 2005,
152, B140. [CrossRef]
20. Birol, F.; Birol, Y. Corrosion Behavior of Twin-Roll Cast Al-Mg and Al-Mg-Si Alloys. In Proceedings of the 9th International
Conference on Aluminium Alloys, Brisbane, Australia, 2–5 August 2004; pp. 338–344.
21. Yasakau, K.A.; Zheludkevich, M.L.; Lamaka, S.V.; Ferreira, M.G.S. Role of intermetallic phases in localized corrosion of AA5083.
Electrochim. Acta 2007, 52, 7651–7659. [CrossRef]
22. Serdechnova, M.; Volovitch, P.; Brisset, F.; Ogle, K. On the cathodic dissolution of Al and Al alloys. Electrochim. Acta 2014, 124,
9–16. [CrossRef]
23. Ma, Y.; Zhou, X.; Huang, W.; Thompson, G.E.; Zhang, X.; Luo, C.; Sun, Z. Localized corrosion in AA2099-T83 aluminum-lithium
alloy: The role of intermetallic particles. Mater. Chem. Phys. 2015, 161, 201–210. [CrossRef]
24. Gao, W.; Wang, D.; Seifi, M.; Lewandowski, J.J. Anisotropy of corrosion and environmental cracking in AA5083-H128 Al-Mg
alloy. Mater. Sci. Eng. A 2018, 730, 367–379. [CrossRef]
25. Guan, L.; Zhou, Y.; Zhang, B.; Wang, J.Q.; Han, E.-H.; Ke, W. Influence of aging treatment on the pitting behavior associated with
the dissolution of active nanoscale β -phase precipitates for an Al–Mg alloy. Corros. Sci. 2016, 103, 255–267. [CrossRef]
26. Yang, W.; Shen, W.; Zhang, R.; Cao, K.; Zhang, J.; Liu, L. Enhanced age-hardening by synergistic strengthening from Mg–Si and
Mg–Zn precipitates in Al-Mg-Si alloy with Zn addition. Mater. Charact. 2020, 169, 110579. [CrossRef]
27. Yi, G.; Zeng, W.; Poplawsky, J.D.; Cullen, D.A.; Wang, Z.; Free, M.L. Characterizing and modeling the precipitation of Mg-rich
phases in Al 5xxx alloys aged at low temperatures. J. Mater. Sci. Technol. 2017, 33, 991–1003. [CrossRef]
122
Materials 2021, 14, 394
28. Li, J.; Dang, J. A summary of corrosion properties of Al-Rich solid solution and secondary phase particles in al alloys. Metals
(Basel) 2017, 7, 84. [CrossRef]
29. Da Ren, W.; LI, J.-F.; Zheng, Z.-Q.; Chen, W.-J. Localized corrosion mechanism associated with precipitates containing Mg in Al
alloys. Trans. Nonferrous Met. Soc. China 2007, 17, 727–732. [CrossRef]
30. Lyndon, J.A.; Gupta, R.K.; Gibson, M.A.; Birbilis, N. Electrochemical behaviour of the β-phase intermetallic (Mg2Al3) as a
function of pH as relevant to corrosion of aluminium-magnesium alloys. Corros. Sci. 2013, 70, 290–293. [CrossRef]
31. Yang, Y.K.; Allen, T. Direct visualization of β phase causing intergranular forms of corrosion in Al-Mg alloys. Mater. Charact.
2013, 80, 76–85. [CrossRef]
32. Vignesh, R.V.; Padmanaban, R. Intergranular corrosion susceptibility of friction stir processed aluminium alloy 5083. Mater. Today
Proc. 2018, 5, 16443–16452. [CrossRef]
33. Fan, L.; Ma, J.; Zou, C.; Gao, J.; Wang, H.; Sun, J.; Guan, Q.; Wang, J.; Tang, B.; Li, J.; et al. Revealing foundations of the
intergranular corrosion of 5XXX and 6XXX Al alloys. Mater. Lett. 2020, 271, 127767. [CrossRef]
34. Searles, J.L.; Gouma, P.I.; Buchheit, R.G. Stress corrosion cracking of sensitized AA5083 (Al-4.5Mg-1.0Mn). Mater. Sci. Forum.
2002, 396–402, 1437–1442. [CrossRef]
35. Brillas, E.; Cabot, P.L.; Centellas, F.; Garrido, J.A.; Pérez, E.; Rodríguez, R.M. Electrochemical oxidation of high-purity and
homogeneous Al-Mg alloys with low Mg contents. Electrochim. Acta 1997, 43, 799–812. [CrossRef]
36. Sriyono, F.; Hastuti, E.P.; Sunaryo, G.R. Study on Pitting Corrosion of AlMg2 in Solution Containing Chloride. J. Phys. Conf. Ser.
2019, 1198. [CrossRef]
37. Guan, L.; Zhang, B.; Wang, J.Q.; Han, E.H.; Ke, W. The reliability of electrochemical noise and current transients characterizing
metastable pitting of Al-Mg-Si microelectrodes. Corros. Sci. 2014, 80, 1–6. [CrossRef]
38. Laycock, N.J.; Newman, R.C. Localised dissolution kinetics, salt films and pitting potentials. Corros. Sci. 1997, 39, 1771–1790.
[CrossRef]
39. Yi, G.; Cullen, D.A.; Derrick, A.T.; Zhu, Y.; Free, M.L. Effects of Different Temper and Aging Temperature on the Precipitation
Behavior of Al 5xxx Alloy. Light Met. 2015, 359–365. [CrossRef]
40. Li, Y.; Hung, Y.; Du, Z.; Xiao, Z.; Jia, G. The Effect of Homogenization on the Corrosion Behavior of Al–Mg Alloy. Phys. Met.
Metallogr. 2018, 119, 339–346. [CrossRef]
41. Abo Zeid, E.F. Mechanical and electrochemical characteristics of solutionized AA 6061, AA6013 and AA 5086 aluminum alloys. J.
Mater. Res. Technol. 2019, 8, 1870–1877. [CrossRef]
42. Singh, N.S.; Rao, P.N.; Jayaganathan, R.; Midathada, A.; Verma, K.; Ravella, U.K. Elevated corrosion in strain hardened Al–Mg
alloy. Vacuum 2018, 157, 402–413. [CrossRef]
43. Abdi behnagh, R.; Besharati Givi, M.K.; Akbari, M. Mechanical properties, corrosion resistance, and microstructural changes
during friction stir processing of 5083 aluminum rolled plates. Mater. Manuf. Process. 2012, 27, 636–640. [CrossRef]
44. Naeini, M.F.; Shariat, M.H.; Eizadjou, M. On the chloride-induced pitting of ultra fine grains 5052 aluminum alloy produced by
accumulative roll bonding process. J. Alloys Compd. 2011, 509, 4696–4700. [CrossRef]
45. Kus, E.; Lee, Z.; Nutt, S.; Mansfeld, F. A comparison of the corrosion behavior of nanocrystalline and conventional Al 5083
samples. Corrosion 2006, 62, 152–161. [CrossRef]
46. Ezuber, H.; El-Houd, A.; El-Shawesh, F. A study on the corrosion behavior of aluminum alloys in seawater. Mater. Des. 2008, 29,
801–805. [CrossRef]
47. He, C.; Luo, B.; Zheng, Y.; Yin, Y.; Bai, Z.; Ren, Z. Effect of Sn on microstructure and corrosion behaviors of Al-Mg-Si alloys. Mater.
Charact. 2019, 156, 109836. [CrossRef]
48. Products, M.; Plate, T.; Products, M.; Sheet, A.; Environments, S.; Alloys, A. Standard Specification for Aluminum and Aluminum-Alloy
Sheet and Plate; ASTM International: West Conshohocken, PA, USA, 2009; Volume 1, pp. 1–29. [CrossRef]
49. Wejrzanowski, T.; Kurzydlowski, K.J. Stereology of grains in nano-crystals. Diffus. Defect Data Pt. B Solid State Phenom. 2003, 94,
221–228. [CrossRef]
50. Wejrzanowski, T.; Spychalski, W.L.; Rózniatowski, K.; Kurzydłowski, K.J. Image based analysis of complex microstructures of
engineering materials. Int. J. Appl. Math. Comput. Sci. 2008, 18, 33–39. [CrossRef]
51. Norfleet, D.M.; Dimiduk, D.M.; Polasik, S.J.; Uchic, M.D.; Mills, M.J. Dislocation structures and their relationship to strength in
deformed nickel microcrystals. Acta Mater. 2008, 56, 2988–3001. [CrossRef]
52. Guan, L.; Zhou, Y.; Lin, H.Q.; Zhang, B.; Wang, J.Q.; Han, E.H.; Ke, W. Detection and analysis of anodic current transients
associated with nanoscale β-phase precipitates on an Al-Mg microelectrode. Corros. Sci. 2015, 95, 6–10. [CrossRef]
53. Abady, G.M.; Hilal, N.H.; El-Rabiee, M.; Badawy, W.A. Effect of Al content on the corrosion behavior of Mg-Al alloys in aqueous
solutions of different pH, Electrochim. Acta 2010, 55, 6651–6658. [CrossRef]
54. Wang, Z.; Chen, P.; Li, H.; Fang, B.; Song, R.; Zheng, Z. The intergranular corrosion susceptibility of 2024 Al alloy during re–ageing
after solution treating and cold–rolling. Corros. Sci. 2017, 114, 156–168. [CrossRef]
55. Meng, G.; Wei, L.; Zhang, T.; Shao, Y.; Wang, F.; Dong, C.; Li, X. Effect of microcrystallization on pitting corrosion of pure
aluminium. Corros. Sci. 2009, 51, 2151–2157. [CrossRef]
56. Jilani, O.; Njah, N.; Ponthiaux, P. Transition from intergranular to pitting corrosion in fine grained aluminum processed by equal
channel angular pressing. Corros. Sci. 2014, 87, 259–264. [CrossRef]
123
Materials 2021, 14, 394
57. Mance, A.; Cerović, D.; Mihajlović, A. The effect of small additions of indium and thallium on the corrosion behaviour of
aluminium in sea water. J. Appl. Electrochem. 1984, 14, 459–466. [CrossRef]
58. Brunner, J.G.; May, J.; Höppel, H.W.; Göken, M.; Virtanen, S. Localized corrosion of ultrafine-grained Al-Mg model alloys.
Electrochim. Acta 2010, 55, 1966–1970. [CrossRef]
59. Benedetti, A.; Cirisano, F.; Delucchi, M.; Faimali, M.; Ferrari, M. Potentiodynamic study of Al-Mg alloy with superhydrophobic
coating in photobiologically active/not active natural seawater. Colloids Surfaces B Biointerfaces 2016, 137, 167–175. [CrossRef]
[PubMed]
60. Esmailzadeh, S.; Aliofkhazraei, M.; Sarlak, H. Interpretation of Cyclic Potentiodynamic Polarization Test Results for Study of
Corrosion Behavior of Metals: A Review. Prot. Met. Phys. Chem. Surfaces 2018, 54, 976–989. [CrossRef]
61. Eckermann, F.; Suter, T.; Uggowitzer, P.J.; Afseth, A.; Schmutz, P. The influence of MgSi particle reactivity and dissolution
processes on corrosion in Al-Mg-Si alloys. Electrochim. Acta 2008, 54, 844–855. [CrossRef]
62. Baumgärtner, M.; Kaesche, H. Aluminum pitting in chloride solutions: Morphology and pit growth kinetics. Corros. Sci. 1990, 31,
231–236. [CrossRef]
63. Xiao, R.G.; Yan, K.P.; Yan, J.X.; Wang, J.Z. Electrochemical etching model in aluminum foil for capacitor. Corros. Sci. 2008, 50,
1576–1583. [CrossRef]
64. Newman, R.C. Local chemistry considerations in the tunneling corrosion of aluminium. Corros. Sci. 1995, 37, 527–533. [CrossRef]
65. Zaid, B.; Saidi, D.; Benzaid, A.; Hadji, S. Effects of pH and chloride concentration on pitting corrosion of AA6061 aluminum alloy.
Corros. Sci. 2008, 50, 1841–1847. [CrossRef]
66. Towarek, A.; Dobkowska, A.; Zdunek, J.; Jurczak, W.; Mizera, J. The influence of Mg addition and hydrostatic extrusion HE on
the repassivation ability of pure Al, AlMg1 and AlMg3 model alloys in 3.5 wt% NaCl. Corros. Eng. Sci. Technol. 2019, 54, 666–672.
[CrossRef]
67. Dar, S.M.; Liao, H. Creep behavior of heat resistant Al–Cu–Mn alloys strengthened by fine (θ ) and coarse (Al20Cu2Mn3) second
phase particles. Mater. Sci. Eng. A 2019, 763, 138062. [CrossRef]
68. Zheng, Y.Y.; Luo, B.H.; He, C.; Gao, Y.; Bai, Z.H. Corrosion evolution and behaviour of Al–2.1Mg–1.6Si alloy in chloride media.
Rare Met. 2020. [CrossRef]
69. Zhu, Y.; Sun, K.; Frankel, G.S. Intermetallic Phases in Aluminum Alloys and Their Roles in Localized Corrosion. J. Electrochem.
Soc. 2018, 165, C807–C820. [CrossRef]
70. Pouraliakbar, H.; Jandaghi, M.R.; Khalaj, G. Constrained groove pressing and subsequent annealing of Al-Mn-Si alloy: Mi-
crostructure evolutions, crystallographic transformations, mechanical properties, electrical conductivity and corrosion resistance.
Mater. Des. 2017, 124, 34–46. [CrossRef]
71. Aballe, A.; Bethencourt, M.; Botana, F.J.; Cano, M.J.; Marcos, M. Localized alkaline corrosion of alloy AA5083 in neutral 3.5%
NaCl solution. Corros. Sci. 2001, 43, 1657–1674. [CrossRef]
124
materials
Article
Microstructure and Corrosion of Cast Magnesium
Alloy ZK60 in NaCl Solution
Zhen Li 1 , Zeyin Peng 1 , Kai Qi 1,2 , Hui Li 3 , Yubing Qiu 1,2, * and Xingpeng Guo 4,5
1 School of Chemistry and Chemical Engineering, Huazhong University of Science and Technology,
Wuhan 430074, China; [email protected] (Z.L.); [email protected] (Z.P.);
[email protected] (K.Q.)
2 Key Laboratory of Material Chemistry for Energy Conversion and Storage, Huazhong University of Science
and Technology, Ministry of Education, Wuhan 430074, China
3 Changqing Oil and Gas Technology Institute, Changqing Oil Field Company, Xi’an 710021, China;
[email protected]
4 Hubei Key Laboratory of Materials Chemistry and Service Failure, Wuhan 430074, China;
[email protected]
5 School of Chemistry and Chemical Engineering, Guangzhou University, Guangzhou 510006, China
* Correspondence: [email protected]; Tel.: +86-13-545-266-622
Abstract: In this work, the effects of the microstructure and phase constitution of cast magnesium
alloy ZK60 (Mg-5.8Zn-0.57Zr, element concentration in wt.%) on the corrosion behavior in aqueous
NaCl (0.1 mol dm−3 ) were investigated by weight-loss measurements, hydrogen evolution tests,
and electrochemical techniques. The alloy was found to be composed of α-Mg matrix, with large
second-phase particles of MgZn2 deposited along grain boundaries and a Zr-rich region in the
central area of the grains. The large second-phase particles and the Zr-rich regions were more
stable than the Mg matrix, resulting in a strong micro-galvanic effect. A filiform corrosion was
found. It originated from the second-phase particles in the grain boundary regions in the early
corrosion period. The filaments gradually occupied most areas of the alloy surface, and the general
corrosion rate decreased significantly. Corrosion pits were developed under filaments. The pit growth
rate decreased over time; however, it was about eight times larger than the general corrosion rate.
A schematic model is presented to illustrate the corrosion mechanism.
1. Introduction
Magnesium (Mg) alloys have been widely applied as lightweight engineering materials due to
their unique properties [1–8]. As commercial Mg-Zn-based alloys, ZK60 (Mg–Zn–Zr) alloys [9] have
attracted great interest from researchers due to their high strength [10–12]. The microstructure [13,14],
mechanical properties [15–17], and biological applications [18,19] of ZK60 alloys have been studied
widely in recent decades. It has been verified that microstructure evolution is essential for the
mechanical properties of ZK60 alloys, grain refinement, and stable precipitates, having vital effects on
improving the mechanical properties [20–23]. Nevertheless, the weak corrosion resistance of ZK60
alloys limits their further applications.
The microstructure of Mg alloys, especially their second phases, has an evident impact on their
corrosion behavior [3,5]. The second phases of Mg alloys may have a dual role in their corrosion,
i.e., a galvanic acceleration effect or a corrosion blocking effect, depending on their quantities and
distribution [24–26]. In Mg-Al alloys, when the amount of aluminum (Al) is low (e.g., Mg-5Al),
the β-phase (Mg12 Al17 ) is relatively discontinuous in the Mg matrix and mainly acts as a cathode
phase to accelerate the dissolution of the matrix. As the content of the Al element increases (e.g.,
Mg-10Al), the β-phase precipitates are tiny and continuously distributed along the grain boundaries,
producing a barrier to prevent corrosion [27]. The similar effect of the second phase was reported in
other Mg alloys [28,29]. However, the effect of the second phase in ZK60 alloys on their corrosion
behaviors is barely reported. Some researchers tried to enhance the corrosion resistance of ZK60
alloys by modifying their microstructure through heat treatment [30,31], deformation processing [32],
and alloying [33–36]. Even so, the relationship between the microstructure and corrosion behavior of
ZK60 alloys needs more studies for it to be investigated.
Some studies reported the corrosion behavior of ZK60 alloys on different occasions.
Cheng et al. [37] pointed out that the Zr element in ZK60 alloys refined the grain and purified
the alloy composition, which could improve its corrosion resistance in 1 M NaCl. Zeng et al. [38]
investigated the effects of the microstructure and concentration of NaCl (3.5 and 5.0 wt.%) on the
corrosion behavior of an extruded ZK60 alloy. They found that an increase in the grain size of the ZK60
alloy accelerated its corrosion rate. The alloy microstructure played a crucial role in the pitting and
intergranular corrosion. Xu et al. [39] reported that the corrosion rate of a cast ZK60 alloy decreased
with the immersion time in solutions containing 3.5 wt.% NaCl, NaBr, and NaI, while it displayed
passivation in 3.5 wt.% NaF solution. Apart from the above reports, some studies focused on the
biodegradation behavior of ZK60 alloys in Hank’s solution, Ringer’s solution, simulated body fluid,
and artificial urine for biomedical applications [40–43]. The biodegradable property of the ZK60 alloys
is the interesting issue in these investigations. In general, the above research usually concentrated
on the uniform corrosion of ZK60 alloys, and little attention was paid to the development of their
local corrosion. The influence of microstructure on the local corrosion of ZK60 alloys is still not clearly
understood, especially the effect of the second phase.
In this study, a commercial cast ZK60 alloy was selected as the test material. Its microstructure
was characterized by X-ray diffraction (XRD), scanning electron microscopy (SEM), energy-dispersive
X-ray spectroscopy (EDX), and scanning Kelvin probe force microscopy (SKPFM) analysis. The general
and local corrosion of the alloy in 0.1 M NaCl was investigated using weight loss tests, hydrogen
evolution tests, and electrochemical measurements, as well as corrosion morphology monitoring with
an optical microscope and SEM. The effect of the microstructure of the cast ZK60 alloy, especially the
second phase and the distribution of alloying elements, on the corrosion initiation and developmental
features of the alloy were investigated, and the mechanisms involved were studied. This work will
help to verify the relationship between the microstructure and the corrosion behavior of ZK60 alloys.
Moreover, it may also provide a theoretical basis for improving the corrosion resistance of ZK60 alloys
by adjusting the microstructure in future research.
Zn Al Fe Ni Cu Zr Mg
5.80 <0.01 <0.01 <0.01 <0.01 0.57 Bal.
126
Materials 2020, 13, 3833
where the corrosion current density icorr is estimated by the Tafel extrapolation of the cathodic branch
of the polarization curves, and Pi is related to the average corrosion rate.
All the electrochemical tests were repeated at least three times in this study.
Pw = 3.6524ΔW/ρ (2)
where ρ is the metal density (g cm−3 ). For the cast ZK60 alloy, ρ = 1.8 g cm−3 ; thus, Equation (2) becomes:
Pw = 48.67ΔW (3)
127
Materials 2020, 13, 3833
PH = 2.2 vH (4)
Figure 1. Schematic diagram of the test system for the hydrogen evolution measurement.
128
Materials 2020, 13, 3833
Figure 2. (a) Optical and (b) BSE-SEM micrographs of the cast ZK60 alloy.
The BSE-SEM micrograph in Figure 2b indicates the nonuniform distribution of the alloying
elements in the cast ZK60 because the brighter areas contain more elements of higher atomic weight
than the darker regions [50]. Figure 5 presents the area distribution of the alloying elements in the
cast ZK60, proving that the center area of the grains with light color was richer in Zr and Zn (Zr-rich
region) than the neighboring darker zones (grain boundary region). Here, the “grain boundary region”
of the as-cast ZK60 is denoted as the areas between the Zr-rich region, i.e., the dark areas in Figure 2b.
This uneven distribution of the alloying elements in the cast ZK60 alloy may cause the inhomogeneous
electrochemical activity resulting in the micro-galvanic corrosion [51].
Figure 6 presents the SKPFM maps of the cast ZK60 sample, which clearly show that the second
phase had the highest potential. Furthermore, the Volta potential profiles (Figure 6c) along the line A
and line B indicated that the central region of the grains (i.e., Zr-rich region) exhibited higher potential
than those of the grain boundary regions, owing to the higher Zr and Zn contents in the center of grains
and higher Mg content in the grain boundary regions (Figure 5). Thus, the second-phase particles and
the central region of the grains should be more stable than the Mg matrix in grain boundary regions
and more likely to become cathodes in the micro-galvanic cells. A second-phase particle occurred in
the grain area (Figure 6a), which is consistent with Figure 2a, so the micro-galvanic corrosion may also
have been initiated in the grains.
129
Materials 2020, 13, 3833
Figure 4. TEM micrograph of the cast ZK60 alloy and the corresponding selected area electron
diffraction (SAED) patterns of the second-phase particles.
Figure 5. BSE-SEM image of the cast ZK60 alloy and the corresponding area distributions of Mg, Zn,
and Zr.
130
Materials 2020, 13, 3833
Figure 6. (a) Scanning Kelvin probe force microscopy (SKPFM) topography map; (b) Potential map of
the same area; (c) Volta potential profiles along lines A and B in the SKPFM image.
Figure 7. Corrosion rates (ΔW and PW ) of the cast ZK60 alloy after immersion in 0.1M NaCl for different
times measured by weight loss tests.
131
Materials 2020, 13, 3833
Figure 8. Changes in (a) V H and (b) PH of the cast ZK60 alloy as a function of time measured by the
hydrogen evolution test in 0.1M NaCl.
Figure 9. (a) Polarization curves of the cast ZK60 alloy immersed in 0.1 M NaCl for different times;
(b) Changes in the icorr , Pi , and Ebreak with time (25 ◦ C).
132
Materials 2020, 13, 3833
elements (CPE) for the surface film and the electrical double layer, respectively. Rf and Rct represent
the surface film resistance and charge-transfer resistance, respectively. RL and L represent equivalent
resistance and inductance to describe the low-frequency inductance. It should be noted that Rct is the
parallel of the charge-transfer resistance of the anodic process and the cathodic process (Rct,a and Rct,c )
at Ecorr [63]. The fitting curves are also displayed in Figure 10a, and the fitting parameters are listed in
Table 2.
Figure 10. (a) Nyquist plots measured at Ecorr for the cast ZK60 alloy immersed in 0.1 M NaCl for
different times; (b) Rp –t curves at 25 ◦ C.
Figure 11. Equivalent circuit used to model the electrochemical impedance spectroscopy (EIS) response
of the cast ZK60 alloy in 0.1 M NaCl.
According to [59,66,67] and the results in Table 2, the fitting curves can be extrapolated to the
zero-frequency limit to obtain the polarization resistance (Rp ) at different times. Therefore, the Rp –t
curve was obtained as shown in Figure 10b. Rp decreased with time over 0–2 h, indicating corrosion
acceleration, and then, Rp clearly increased between 2 and 24 h before decreasing slightly over 24–72 h.
In this case, the change in Rp with t is consistent with that of Pi (Figure 9b). The Rt and Rf values in
Table 2 show the same change tendencies as those for Rp . The change in the CPEdl , CPEf , RL , and L
with t should be closely related to the surface layer of the cast ZK60 alloy [59]. The specific changes in
these EIS results are discussed later.
133
Materials 2020, 13, 3833
Figure 12. Optical in situ corrosion images of the cast ZK60 etched sample immersed in 0.1 M NaCl for
different times: (a) 20 min; (b) 40 min; (c,d) 60 min.
Figure 13 presents the BSE-SEM micrographs of the etched samples after immersion in 0.1 M
NaCl for 1 and 2 h. The black corrosion filaments (Figure 13a) covered by Mg(OH)2 [70] mainly
occurred on the “darker zones” in grain boundary regions (see Figures 2b and 5) and encompassed
the second-phase particles. Some second-phase particles in the boundary regions (Figure 13b) were
surrounded by a small number of black corrosion products, implying that the corrosion filaments
may have started from the surrounding areas of these second-phase particles. Most of the central
regions of the grains, i.e., the Zr-rich areas with a light color (Figures 2b and 5), were still uncorroded.
When t = 2 h (Figure 13c), only a small number of second-phase particles occurred in the corroded
areas, implying that most of them were removed due to the dissolution of their surrounding Mg
matrix. Some broad corrosion areas in Figure 13 show that the corrosion could develop to the central
area of grains with increasing corrosion time. Moreover, the corrosion filaments in Figure 13c,d show
134
Materials 2020, 13, 3833
an apparent corrosion depth, implying that the corrosion also developed under the black corrosion
product layer.
Figure 13. BSE-SEM micrographs of the cast ZK60 etched samples immersed in 0.1 M NaCl for different
times: (a,b) 1 h (with corrosion products); (c,d) 2 h (without corrosion products).
As shown in Figures 12 and 13, the etched samples were used to observe the origination and
developmental features of the corrosion filaments on the cast ZK60. To avoid the influence of the
etching treatment on the corrosion process, Figure 14 presents the secondary electron (SE) SEM images
of the raw alloy sample after immersion in 0.1 M NaCl for 0.5 and 2 h. The corrosion morphologies
in Figure 14a,b also exhibit the characteristics of filiform corrosion, similar to those in Figure 12b.
The polishing scratches on the surface of the test samples do not appear to influence the initiation and
extension of the corrosion filaments, suggesting that they are controlled by the microstructure of the
cast ZK60 alloy. When t = 2 h, the corrosion images in Figure 14c,d display features similar to those
in Figure 13c and some small corrosion pits occur in the corrosion areas, which may have been the
result of the loss of the second phase-particles. The raw cast ZK60 sample also displayed the filiform
corrosion features in the early corrosion period (0–2 h).
Figure 15 presents the secondary electron (SE) SEM images of the cast ZK60 alloy after immersion
in 0.1 M NaCl for different times to observe the corrosion development over a long period (12–72 h).
The corrosion filaments gradually extended to the whole surface of the test samples after 24 h, and the
number and depth of the corrosion pits increased with time. These results further prove that with
increasing corrosion time, the corrosion of the cast ZK60 alloy can develop toward the central area of
grains and toward the depth of the alloy matrix, which is discussed below.
135
Materials 2020, 13, 3833
Figure 14. SE-SEM micrographs of the cast ZK60 sample immersed in 0.1 M NaCl for different times:
(a,b) 30 min; (c,d) 2 h (without corrosion products).
Figure 15. SE-SEM micrographs of the cast ZK60 alloy immersed in 0.1 M NaCl for different times:
(a) 12 h; (b) 24 h; (c) 48 h; (d) 72 h (with the corrosion products removed).
136
Materials 2020, 13, 3833
to the corrosion morphologies in Figure 15. The corresponding pit depth at different corrosion times
was employed to calculate the average penetration rate (Pdepth , mm y−1 ), which is also presented in
Figure 17. The pit depth increased with time, but the fastest Pdepth occurred in the period of 12–24 h
(~24 mm y−1 ) before gradually decreasing to ~16.6 mm y−1 in the period of 24–72 h.
Figure 16. 3D corrosion images and depth profiles of the cast ZK60 alloy immersed in 0.1 M NaCl for
different times: (a) 12 h; (b) 24 h; (c) 48 h; (d) 72 h (without corrosion products).
Figure 17. Depth of the deepest corrosion pit and the corresponding Pdepth of the cast ZK60 alloy
immersed in 0.1 M NaCl for different times.
137
Materials 2020, 13, 3833
Figure 18 is the cross-sectional EPMA pattern of a corrosion pit on the cast ZK60 sample immersed
in 0.1 M NaCl for 72 h. The corrosion pit was covered by the corrosion product, and the second-phase
particles were encompassed in it. Only Mg, O, and Cl elements were found in the corrosion product
layer, which should result from Mg(OH)2 and Cl− in the test solution. Moreover, a large number of
Cl− through the whole corrosion product layer suggests that Cl− could penetrate the product layer
easily and may become enriched in the bottom of the corrosion pit to propagate the pit corrosion [71].
The second-phase particles in the corrosion pit imply that they may be related to the formation of
corrosion pits, which is discussed later.
Figure 18. Electron probe micro-analyzer (EPMA) maps of the elemental distribution in the cross-section
of a corrosion pit on the cast ZK60 alloy immersed in 0.1 M NaCl for 72 h.
4. Discussion
Because MgO and Mg(OH)2 are both relatively soluble in water according to Reactions (7) and
(8) [3,74], where Ksp is the solubility product constant, MgO will be gradually dissolved and converted
to Mg(OH)2 when they are immersed in NaCl solution.
In this case, the MgO/Mg(OH)2 oxide film formed on the cast ZK60 alloy was partly protective
in a neutral NaCl solution. The anodic and cathodic partial reactions of the corrosion process can be
written as Reactions (9) and (10), respectively, and the corrosion product is formed as Reaction (11) [75].
138
Materials 2020, 13, 3833
Figure 19. Schematic model for the initiation and development of the corrosion on the cast ZK60
alloy in NaCl solution. (a) original state; (b) corrosion initiation; (c) filaments growth; (d) filaments
propagation; (e) pit formation and (f) pit development.
Compared to the Mg matrix in the grain boundary regions, those with stable oxide film act as
micro-galvanic cathodes. Under the attack of Cl− , the film around the second phase could be easily
broken due to the higher numbers of imperfections [78]. Therefore, the micro-galvanic corrosion was
initiated in the areas around the large discontinuous second-phase particles in the grain boundaries,
owing to the strong galvanic effect between these second-phase particles and the Mg matrix in the grain
boundaries (Figure 13b). It should be noted that there were some second-phase particles deposited in
the grains (Figures 2a and 6), and therefore, the corrosion may also be initiated in the grains. However,
we did not observe this phenomenon, which may be due to the stable oxide film in the grains (Zr-rich
region) and the light galvanic effect between them.
After the corrosion initiation (Figure 19b), the oxide film near the second-phase particles in the
grain boundaries cracked first. However, there were no black Mg(OH)2 products observed in this
beginning period (Figure 12a), because they were dissolved in the NaCl solution (Equation (5)). After the
concentration of Mg(OH)2 reached saturation in the NaCl solution, black Mg(OH)2 precipitation
occurred (Figure 12b).
139
Materials 2020, 13, 3833
140
Materials 2020, 13, 3833
5. Conclusions
1. The microstructure of the cast ZK60 alloy was composed of an α-Mg phase and large second-phase
particles (MgZn2 ), which mainly deposited along the grain boundaries, and a Zr-rich region
existed in the central area of the grains. The grain boundaries and their adjacent regions (noted as
grain boundary regions) had relatively higher Mg contents.
141
Materials 2020, 13, 3833
2. The second-phase particles and the central area of the grains (Zr-rich region) were more
electrochemically stable than the grain boundary regions, resulting in a strong micro-galvanic
effect between them.
3. The corrosion of the cast ZK60 alloy in 0.1 M NaCl solution originated from the areas around
the second-phase particles in grain boundaries and firstly developed in the grain boundary
regions, showing filiform-like corrosion characteristics owing to the strong micro-galvanic effect
in its microstructure.
4. The general corrosion rate increased in the early corrosion period (about 0–2 h). Then, the black
corrosion filaments covered with Mg(OH)2 gradually occupied most of the alloy surface to inhibit
the corrosion process and decrease its general corrosion rate. However, corrosion pits occurred
under the corrosion filaments and had a high growth rate (Pdepth ) in this period. After 24 h,
the number of corrosion pits increased, and Pdepth decreased with time (17–24 mm y−1 ) but
was still eight times larger than the general corrosion rates (Pw , PH , and Pi ), which should be
paid more attention. The microstructure of the cast ZK60 alloy has an essential influence on the
initiation and development of its corrosion.
Author Contributions: Conceptualization, Z.L. and Y.Q.; methodology, Z.L.; software, Z.L., Z.P. and K.Q.;
validation, Y.Q., K.Q. and X.G.; formal analysis, Z.L.; investigation, Z.L.; resources, Z.L., Z.P. and H.L.; data
curation, Z.L., Z.P., and H.L.; writing—original draft preparation, Z.L.; writing—review and editing, K.Q., Y.Q.,
and X.G.; visualization, Z.L.; supervision, Y.Q.; funding acquisition, Y.Q. All authors have read and agreed to the
published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: The authors are thankful for the analysis support of the Key Laboratory of Material Chemistry
for Energy Conversion and Storage and the Analytical and Testing Center, Huazhong University of Science
and Technology.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Xu, T.C.; Yang, Y.; Peng, X.D.; Song, J.F.; Pan, F.S. Overview of advancement and development trend on
magnesium alloy. J. Magnesium. Alloy 2019, 7, 536–544. [CrossRef]
2. Zhang, L.L.; Zhang, Y.T.; Zhang, J.S.; Zhao, R.; Zhang, J.X.; Xu, C.X. Effect of alloyed mo on mechanical
properties, biocorrosion and cytocompatibility of as-cast Mg-Zn-Y-Mn alloys. Acta Metall. Sin. 2020, 33,
500–513. [CrossRef]
3. Esmaily, M.; Svensson, J.E.; Fajardo, S. Fundamentals and advances in magnesium alloy corrosion.
Prog. Mater. Sci. 2017, 89, 92–193. [CrossRef]
4. You, S.H.; Huang, Y.D.; Kainer, K.U.; Hort, N. Recent research and developments on wrought magnesium
alloys. J. Magnesium. Alloy 2017, 5, 239–253. [CrossRef]
5. Atrens, A.; Song, G.L. Review of recent developments in the field of magnesium corrosion. Adv. Eng. Mater.
2015, 17, 400–453. [CrossRef]
6. Song, G.L.; Atrens, A. Recent insights into the mechanism of magnesium corrosion and research suggestions.
Adv. Eng. Mater. 2007, 9, 177–183. [CrossRef]
7. Mordike, B.L.; Ebert, T. Magnesium properties-applications-potential. Mater. Sci. Eng. A 2001, 302, 37–45.
[CrossRef]
8. Sezer, N.; Evis, Z.; Kayhan, S.M.; Tahmasebifar, A.; Koç, M. Review of magnesium-based biomaterials and
their applications. J. Magnesium. Alloy 2018, 6, 23–43. [CrossRef]
9. Bettles, C.; Gibson, M.A. Current wrought magnesium alloys: Strengths and weaknesses. JOM 2005, 57,
46–49. [CrossRef]
10. Dharmendra, C.; Rao, K.P.; Jain, M.K.; Prasad, Y.V.R.K. Role of loading direction on compressive deformation
behavior of extruded ZK60 alloy plate in a wide range of temperature. J. Alloys Compd. 2018, 744, 289–300.
[CrossRef]
142
Materials 2020, 13, 3833
11. Torbati-Sarraf, S.A.; Sabbaghianrad, S.; Figueiredo, R.B.; Langdon, T.G. Orientation imaging microscopy and
microhardness in a ZK60 magnesium alloy processed by high-pressure torsion. J. Alloys Compd. 2017, 712,
185–193. [CrossRef]
12. He, Y.B.; Pan, Q.L.; Qin, Y.J. Microstructure and mechanical properties of ZK60 alloy processed by two-step
equal channel angular pressing. J. Alloys Compd. 2010, 492, 605–610. [CrossRef]
13. Yang, Y.; Wang, Z.; Jiang, L.H. Evolution of precipitates in ZK60 magnesium alloy during high strain rate
deformation. J. Alloys Compd. 2017, 705, 566–571. [CrossRef]
14. Wang, C.; Luo, T.J.; Zhou, J.X.; Yang, Y.S. Effects of solution and quenching treatment on the residual stress in
extruded ZK60 magnesium alloy. Mater. Sci. Eng. A 2018, 722, 14–19. [CrossRef]
15. Wang, W.; Zhang, W.C.; Chen, W.Z.; Cui, G.R.; Wang, E.D. Effect of initial texture on the bending behavior,
microstructure and texture evolution of ZK60 magnesium alloy during the bending process. J. Alloys Compd.
2018, 737, 505–514. [CrossRef]
16. Liu, W.C.; Dong, J.; Zhang, P.; Yao, Z.Y.; Zhai, C.Q.; Ding, W.J. High cycle fatigue behavior of as-extruded
ZK60 magnesium alloy. J. Mater. Sci. 2009, 44, 2916–2924. [CrossRef]
17. Lin, J.B.; Wang, Q.D.; Peng, L.M.; Roven, H.J. Microstructure and high tensile ductility of ZK60 magnesium
alloy processed by cyclic extrusion and compression. J. Alloys Compd. 2009, 476, 441–445. [CrossRef]
18. Xia, K.D.; Pan, H.; Wang, T.L. Effect of Ca/P ratio on the structural and corrosion properties of biomimetic
Ca–P coatings on ZK60 magnesium alloy. Mater. Sci. Eng. C 2017, 72, 676–681. [CrossRef]
19. Chen, J.X.; Tan, L.L.; Yang, K. Effect of heat treatment on mechanical and biodegradable properties of an
extruded ZK60 alloy. Biol. Mater. 2017, 2, 19–26. [CrossRef]
20. Cho, J.H.; Han, S.H.; Jeong, H.T.; Choi, S.H. The effect of aging on mechanical properties and texture evolution
of ZK60 alloys during warm compression. J. Alloys Compd. 2018, 743, 553–563. [CrossRef]
21. Torbati–Sarraf, S.A.; Langdon, T.G. Properties of a ZK60 magnesium alloy processed by high–pressure
torsion. J. Alloys Compd. 2014, 613, 357–363. [CrossRef]
22. He, Y.B.; Pan, Q.L.; Qin, Y.J.; Liu, X.Y.; Li, W.B. Microstructure and mechanical properties of ultrafine grain
ZK60 alloy processed by equal channel angular pressing. J. Mater. Sci. 2010, 45, 1655–1662. [CrossRef]
23. Liu, Z.; Xin, R.L.; Wu, X.; Liu, D.J.; Liu, Q. Improvement in the strength of friction-stir-welded ZK60 alloys
via post-weld compression and aging treatment. Mater. Sci. Eng. A 2018, 712, 493–501. [CrossRef]
24. Song, G.L.; Atrens, A. Understanding magnesium corrosion–A framework for improved alloy performance.
Adv. Eng. Mater. 2010, 5, 837–858. [CrossRef]
25. Song, G.L.; Atrens, A.; Dargusch, M. Influence of microstructure on the corrosion of diecast AZ91D. Corros. Sci.
1998, 41, 249–273. [CrossRef]
26. Song, G.L.; Atrens, A.; St John, D.; Li, Z. Magnesium Alloys and Their Applications; Wiley–VCH: Weinheinm,
Germany, 2000; pp. 426–431.
27. Lunder, O.; Lein, J.E.; Aune, T.K.; Nisancioglu, K. The role of Mg17 Al12 phase in the corrosion of Mg alloy
AZ91. Corrosion 1989, 45, 741–748. [CrossRef]
28. Lafront, A.M.; Zhang, W.; Jin, S.; Tremblay, R.; Dube, D.; Ghali, E. Pitting corrosion of AZ91D and AJ62x
magnesium alloys in alkaline chloride medium using electrochemical techniques. Electrochim. Acta 2015, 51,
489–501. [CrossRef]
29. Hara, N.; Kobayashi, Y.; Kagaya, D.; Akao, N. Formation and breakdown of surface films on magnesium and
its alloys in aqueous solutions. Corros. Sci. 2007, 49, 166–175. [CrossRef]
30. Choi, H.Y.; Kim, W.J. Effect of thermal treatment on the bio-corrosion and mechanical properties of
ultrafine-grained ZK60 magnesium alloy. J. Mech. Behav. Biomed. Mater. 2014, 37, 307–322. [CrossRef]
31. Li, X.; Jiang, J.H.; Zhao, Y.H.; Ma, A.B.; Wen, D.J.; Zhu, Y.T. Effect of equal-channel angular pressing and aging
on corrosion behavior of ZK60 Mg alloy. Trans. Nonferrous. Met. Soc. China 2015, 25, 3909–3920. [CrossRef]
32. Orlov, D.; Ralston, K.D.; Birbilis, N.; Estrin, Y. Enhanced corrosion resistance of Mg alloy ZK60 after processing
by integrated extrusion and equal channel angular pressing. Acta Mater. 2011, 59, 6176–6186. [CrossRef]
33. Zengin, H.; Turen, Y.; Ahlatci, H.; Sun, Y. mechanical properties and corrosion behavior of as-cast
Mg-Zn-Zr-(La) magnesium alloys. J. Mater. Eng. Perform. 2018, 27, 389–397. [CrossRef]
34. Baek, S.M.; Kim, B.; Park, S.S. Influence of intermetallic particles on the corrosion properties of extruded
ZK60 Mg alloy containing Cu. Metals 2018, 8, 323. [CrossRef]
143
Materials 2020, 13, 3833
35. Biancardi, O.V.; Rosa, V.L.; Abreu, L.B.; Tavares, A.P.R.; Corrêa, R.G.; Cavalcanti, P.H.; Napoleão, B.I.;
Pereira, S.E. Corrosion behavior of as-cast ZK60 alloy modified with rare earth addition in sodium sulfate
medium. Corros. Sci. 2019, 158. [CrossRef]
36. Zhang, T.T.; Cui, H.W.; Cui, X.L.; Zhao, E.T.; Feng, R.; Pan, Y.K. Investigations on microstructures, mechanical
and corrosion properties of Mg-5.5Zn-0.8Zr alloys with Sm addition. Mater. Res. Express 2019, 6. [CrossRef]
37. Cheng, Y.L.; Qin, T.W.; Wang, H.M.; Zhang, Z. Comparison of corrosion behaviors of AZ31, AZ91, AM60
and ZK60 magnesium alloys. Trans. Nonferrous. Met. Soc. China 2009, 19, 517–524. [CrossRef]
38. Zeng, R.C.; Kainer, K.U.; Blawer, C.; Dietzel, W. Corrosion of an extruded magnesium alloy ZK60
component—The role of microstructural features. J. Alloys Compd. 2011, 509, 4462–4469. [CrossRef]
39. Xu, H.Y.; Diwu, J.T.; Liu, X.; Yang, Y.Q. Corrosion behavior of ZK60 magnesium alloy in sodium halide
solutions. J. Chin. Soc. Corros. Prot. 2015, 35, 245–251.
40. Gu, X.N.; Li, N.; Zheng, Y.F.; Ruan, L.Q. In vitro degradation performance and biological response of a
Mg–Zn–Zr alloy. Mater. Sci. Eng. B 2011, 176, 1778–1784. [CrossRef]
41. Huan, Z.G.; Leeflang, M.A.; Zhou, J.; Fratila-Apachitei, L.E.; Duszczyk, J. In vitro degradation behavior and
cytocompatibility of Mg–Zn–Zr alloys. J. Mater. Sci. Mater. Med. 2010, 21, 2623–2635. [CrossRef]
42. Zhang, S.Y.; Bi, Y.Z.; Li, J.Y.; Wang, Z.G.; Yan, J.M.; Song, J.W.; Sheng, H.B.; Guo, H.Q.; Li, Y. Biodegradation
behavior of magnesium and ZK60 alloy in artificial urine and rat models. Biol. Mater. 2017, 2, 53–62.
[CrossRef] [PubMed]
43. Jamesh, M.I.; Wu, G.S.; Zhao, Y.; McKenzie, D.R.; Bilek, M.M.M.; Chu, P.K. Electrochemical corrosion behavior
of biodegradable Mg–Y–RE and Mg–Zn–Zr alloys in ringer’s solution and simulated body fluid. Corros. Sci.
2015, 91, 160–184. [CrossRef]
44. Shi, Z.M.; Liu, M.; Atrens, A. Measurement of the corrosion rate of magnesium alloys using tafel extrapolation.
Corros. Sci. 2010, 52, 579–588. [CrossRef]
45. Zhao, M.C.; Schmutz, P.; Brunner, S.; Liu, M.; Song, G.L.; Atrens, A. An exploratory study of the corrosion of
Mg alloys during interrupted salt spray testing. Corros. Sci. 2009, 51, 1277–1292. [CrossRef]
46. Shi, Z.M.; Atrens, A. An innovative specimen configuration for the study of Mg corrosion. Corros. Sci. 2011,
53, 226–246. [CrossRef]
47. Wei, L.Y.; Dunlop, G.L.; Westengen, H. The intergranular microstructure of cast Mg–Zn and Mg–Zn–Rare
earth alloys. Metall. Mater. Trans. 1995, 26, 1947–1955. [CrossRef]
48. Pan, F.S.; Mao, J.J.; Chen, X.H.; Peng, J.; Wang, J.F. Influence of impurities on microstructure and mechanical
properties of ZK60 magnesium alloy. Trans. Nonferrous. Met. Soc. China 2010, 20, 1299–1304. [CrossRef]
49. He, Y.B.; Pan, Q.L.; Qin, Y.J. Effect of heat treatment on microstructure and mechanical properties of wrought
ZK60 magnesium alloy. Heat Treat. Met. 2011, 36, 52–57.
50. Song, Y.W.; Shan, D.Y.; Chen, R.S.; Han, E.H. Effect of second phases on the corrosion behaviour of wrought
Mg–Zn–Y–Zr alloy. Corros. Sci. 2010, 52, 1830–1837. [CrossRef]
51. Song, G.L. Corrosion Protection of Magnesium Alloys; Chemical Industry Press: Beijing, China, 2006.
52. Merson, D.; Vasiliev, E.; Markushev, M.; Vinogradov, A. On the corrosion of ZK60 magnesium alloy after
severe plastic deformation. Lett. Mater. 2017, 7, 421–427. [CrossRef]
53. Cao, F.Y.; Shi, Z.M.; Hofstetter, J.; Uggowitzer, P.J.; Song, G.L.; Liu, M.; Atrens, A. Corrosion of ultra-high-purity
Mg in 3.5% NaCl solution saturated with Mg(OH)2 . Corros. Sci. 2013, 75, 78–99. [CrossRef]
54. Sarraf, H.T.; Sarraf, S.A.T.; Poursaee, A.; Langdon, T.G. Electrochemical behavior of a magnesium ZK60 alloy
processed by high-pressure torsion. Corros. Sci. 2019, 154, 90–100. [CrossRef]
55. Song, G.L. Recent progress in corrosion and protection of magnesium alloys. Adv. Eng. Mater. 2005, 7,
563–587. [CrossRef]
56. Song, G.L.; Unocic, K.A. The anodic surface film and hydrogen evolution on Mg. Corros. Sci. 2015, 9, 758–765.
[CrossRef]
57. Song, Y.W.; Shan, D.Y.; Chen, R.S.; Han, E.H. Investigation of surface oxide film on magnesium lithium alloy.
J. Alloys Compd. 2009, 29, 1039–1045. [CrossRef]
58. Gandel, D.S.; Easton, M.A.; Gibson, M.A.; Abbott, T.; Birbilis, N. The influence of zirconium additions on the
corrosion of magnesium. Corros. Sci. 2014, 81, 27–35. [CrossRef]
59. King, A.D.; Birbilisa, N.; Scully, J.R. Accurate electrochemical measurement of magnesium corrosion rates;
A combined impedance, mass-loss and hydrogen collection study. Electrochim. Acta 2014, 121, 394–406.
[CrossRef]
144
Materials 2020, 13, 3833
60. Liu, W.J.; Cao, F.H.; Chen, A.N.; Chang, L.R.; Zhang, J.Q.; Cao, C.N. Corrosion behaviour of AM60
magnesium alloys containing Ce or La under thin electrolyte layers. Part 1: Microstructural characterization
and electrochemical behaviour. Corros. Sci. 2010, 52, 627–638. [CrossRef]
61. Cao, C.N.; Zhang, J.Q. An Introduction of Electrochemical Impedance Spectroscopy; Science Press: Beijing, China, 2002.
62. Song, G.L.; Atrens, A.; John, D.S.; Wu, X.; Nairn, J. The anodic dissolution of magnesium in chloride and
sulphate solutions. Corros. Sci. 1997, 39, 1981–2004. [CrossRef]
63. Song, Y.W.; Han, E.H. The effect of Zn concentration on the corrosion behavior of Mg–xZn alloys. Corros. Sci.
2012, 65, 322–330. [CrossRef]
64. Baril, G.; Galicia, G.; Deslouis, C.; Pebere, N.; Tribollet, B.; Vivier, V. An impedance investigation of the
mechanism of pure magnesium corrosion in sodium sulfate solutions. J. Electrochem. Soc. 2007, 154, 108–113.
[CrossRef]
65. Song, Y.W.; Han, E.H.; Shan, D.Y. Corrosion characterization of Mg–8Li alloy in NaCl solution. Corros. Sci.
2009, 51, 1087–1094. [CrossRef]
66. Shi, Z.M.; Cao, F.Y.; Song, G.L.; Atrens, A. Low apparent valence of Mg during corrosion. Corros. Sci. 2014,
88, 434–443. [CrossRef]
67. Gomes, M.P.; Costa, I.; Pebere, N.; Rossi, J.L.; Tribollet, B.; Vivier, V. On the corrosion mechanism of Mg
investigated by electrochemical impedance spectroscopy. Electrochim. Acta 2019, 306, 61–70. [CrossRef]
68. Jönsson, M.; Persson, D.; Thierry, D. Corrosion product formation during NaCl induced atmospheric
corrosion of magnesium alloy AZ91D. Corros. Sci. 2007, 49, 1540–1558. [CrossRef]
69. Williams, G.; Dafydd, H.L.; Grace, R. The localised corrosion of Mg alloy AZ31 in chloride containing
electrolyte studied by a scanning vibrating electrode technique. Electrochim. Acta 2013, 109, 489–501.
[CrossRef]
70. Williams, G.; Grace, R. Chloride–induced filiform corrosion of organic–coated magnesium. Electrochim. Acta
2011, 56, 1894–1903. [CrossRef]
71. Wang, H.X.; Song, Y.W.; Yu, J.; Shan, D.Y.; Han, E.H. Characterization of filiform corrosion of Mg–3Zn Mg
Alloy. J. Electrochem. Soc. 2017, 1649, 574–580. [CrossRef]
72. Atrens, A.; Song, G.L.; Shi, Z.M.; Soltan, A.; Johnston, S.; Dargusch, M.S. Understanding the corrosion of Mg
and Mg alloys. Encycl. Interfacial Chem. 2018, 515–534.
73. Song, Y.W.; Han, E.H.; Shan, D.Y.; Yim, C.D.; You, B.S. Microstructure and protection characteristics of the
naturally formed oxide films on Mg–x Zn alloys. Corros. Sci. 2013, 72, 133–143. [CrossRef]
74. Nordlien, J.H.; Ono, S.; Masuko, N.; Nisancioglu, K. Morphology and structure of oxide films formed on
magnesium by exposure to air and water. J. Electrochem. Soc. 1995, 142, 3320–3322. [CrossRef]
75. Dinodi, N.; Shetty, A.N. Electrochemical investigations on the corrosion behaviour of magnesium alloy ZE41
in a combined medium of chloride and sulphate. J. Magnesium. Alloys 2013, 1, 201–209. [CrossRef]
76. Kaya, A.; Ben–Hamu, G.; Eliezer, D.; Shin, K.S.; Cohen, S. Corrosion and oxidation of alloys of the
Mg–Y–Zr–REM system. Met. Sci. Heat. Treat. 2006, 48, 518–523. [CrossRef]
77. Ben–Hamu, G.; Eliezer, D.; Shin, K.S.; Cohen, S. The relation between microstructure and corrosion behavior
of Mg–Y–RE–Zr alloys. J. Alloys Compd. 2007, 431, 269–276. [CrossRef]
78. Lindström, R.; Johansson, L.G.; Thompson, G.E.; Skeldon, P.; Svensson, J.E. Corrosion of magnesium in
humid air. Corros. Sci. 2004, 46, 1141–1158. [CrossRef]
79. Song, G.L.; Atrens, A. Corrosion mechanisms of magnesium alloys. Adv. Eng. Mater. 1999, 1, 11–33.
[CrossRef]
80. Zeng, R.C.; Zhang, J.; Huang, W.J.; Kainer, K.U.; Blawer, C.; Dietzel, W.; Ke, W. Review of studies on corrosion
of magnesium alloys. Trans. Nonferrous. Met. Soc. China 2006, 16, 763–771. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
145
materials
Article
Microstructure, Micro-Mechanical and Tribocorrosion Behavior
of Oxygen Hardened Ti–13Nb–13Zr Alloy
Alicja Łukaszczyk 1, *, Sławomir Zimowski 2, *, Wojciech Pawlak 3 , Beata Dubiel 4 and Tomasz Moskalewicz 4
1 Faculty of Foundry Engineering, AGH University of Science and Technology, Reymonta 23,
30-059 Kraków, Poland
2 Faculty of Mechanical Engineering and Robotics, AGH University of Science and Technology,
Mickiewicza Av. 30, 30-059 Kraków, Poland
3 Institute of Materials Science and Technology, Lodz University of Technology, Stefanowskiego 1/15,
90-924 Łódź, Poland; [email protected]
4 Faculty of Metals Engineering and Industrial Computer Science, AGH University of Science and Technology,
Czarnowiejska 66, 30-054 Kraków, Poland; [email protected] (B.D.); [email protected] (T.M.)
* Correspondence: [email protected] (A.Ł.); [email protected] (S.Z.); Tel.: +48-(12)-617-2763 (A.Ł.);
+48-(12)-617-3060 (S.Z.)
Abstract: In the present work, an oxygen hardening of near-β phase Ti–13Nb–13Zr alloy in plasma
glow discharge at 700–1000 ◦ C was studied. The influence of the surface treatment on the alloy
microstructure, tribological and micromechanical properties, and corrosion resistance is presented.
A strong influence of the treatment on the hardened zone thickness, refinement of the α’ laths and
grain size of the bulk alloy were found. The outer hardened zone contained mainly an oxygen-rich
Ti α’ (O) solid solution. The microhardness and elastic modulus of the hardened zone decreased
with increasing hardening temperature. The hardened zone thickness, size of the α’ laths, and grain
size of the bulk alloy increased with increasing treatment temperature. The wear resistance of the
Citation: Łukaszczyk, A.; Zimowski,
S.; Pawlak, W.; Dubiel, B.;
alloy oxygen-hardened at 1000 ◦ C was about two hundred times, and at 700 ◦ C, even five hundred
Moskalewicz, T. Microstructure, times greater than that of the base alloy. Oxygen hardening also slightly improved the corrosion
Micro-Mechanical and Tribocorrosion resistance. Tribocorrosion tests revealed that the alloy hardened at 700 ◦ C was wear-resistant in a
Behavior of Oxygen Hardened corrosive environment, and when the friction process was completed, the passive film was quickly
Ti–13Nb–13Zr Alloy. Materials 2021, restored. The results show that glow discharge plasma oxidation is a simple and effective method to
14, 2088. https://doi.org/10.3390/ enhance the micromechanical and tribological performance of the Ti–13Nb–13Zr alloy.
ma14082088
147
thereby confirming that the higher percentage of β phase is the main reason for the lower
elastic modulus observed. However, the main drawback of all titanium alloys is their
poor tribological performance, as demonstrated by the high coefficient of friction (COF)
combined with a tendency for scuffing and adhesive wear [2,3]. Moreover, the friction
of titanium alloys in a corrosive environment increases their wear rate [16,17], which
is disadvantageous in their application as orthopedic bioimplants. Surface treatment
is usually applied to enhance the properties of titanium alloys as bioimplants. Several
techniques, including oxidizing [18], nitriding [19], ion implantation [20] and physical
vapour deposition (PVD) coating [21], selective laser melting [22], Ar arc melting [12],
spark plasma sintering [23], anodization [14], mechanical blending process [15] have
been investigated. It is known that oxygen hardening can increase the hardness and
tribocorrosion resistance [24–30].
The characterization of the oxygen diffusion zone and the impact of this zone on
macroscopic properties are still subjects of intensive investigation. Dong and Li [28]
reported that the boost diffusion oxidation (BDO) process was an effective way to harden a
titanium surface. The BDO process involved two steps. In the first step, titanium samples
were thermally oxidized in the air. During the second step, the air was removed from the
reaction chamber, and pre-oxidized samples were further diffusion treated in a vacuum.
The method was successfully employed by Zabler and coworkers [30] for the formation
of oxygen diffusion hardening (ODH) zones in the commercially pure Ti (grade 2) and
Ti–6Al–4V alloy. A well-controlled hardness and hardened zone depth were achieved. A
disadvantage of the process is the formation of an oxide scale on the alloy surface. Thicker
titanium oxide layers tend to deteriorate the alloy properties. Jamesh et al. [31] reported
that the thermal oxidation process of commercially pure Ti above 850 ◦ C led to oxide
scale spallation. This behavior is dangerous in frictional contact, where the detached hard
titanium oxide particles accelerate the abrasion.
Januszewicz and Siniarski [32] showed that the assistance of plasma glow discharge
during oxidation is of great benefit as it prevents the formation of brittle rutile phase on
the treated alloy surface by mechanically damaging the oxide scale by high-energy ions. In
our previous work [33], we showed that diffusion hardening of the near-surface zone by
interstitial oxygen atoms with using glow discharge plasma in Ar + O2 atmosphere is a
very effective method to improve the mechanical and tribological properties of two-phase
(α + β) Ti–6Al–4V titanium alloy. It was found that the treatment increased the surface
hardness about three times, from 3.4 GPa to 10.6 GPa. The wear resistance in dry sliding
contact with the Al2 O3 ball was also significantly improved.
The oxygen diffusion hardening of the near-β Ti–13Nb–13Zr was performed by ther-
mal oxidation in an oxygen-containing atmosphere [34]. The results suggest that zirconium
plays a key role in the effective oxygen diffusion hardening at 500 ◦ C for alloys of the
Ti–Nb-Zr system. It was found that a sufficient Zr concentration (>6 wt.%) permits oxy-
gen diffusion into the alloys. The near-surface hardness appears to increase further as
Nb content increases. The high hardness of oxygen hardened Ti–13Nb–13Zr alloy is a
result of an outer oxide layer composed of mixed metal (i.e., Ti, Nb, Zr) oxides on top of
interstitial oxygen hardened alloy. This dense oxide layer is not only highly passive from a
chemical/corrosion point of view but also resistant to abrasive wear, which is attributed to
the mechanical stability of the oxide/substrate interface.
To our best knowledge, there are no data on the influence of oxygen plasma glow-
discharge on the hardening of the Ti–13Nb–13Zr alloy and the influence of the surface
treatment on microstructure and tribological properties in corrosive environments. There-
fore, detailed studies on the effect of oxygen plasma glow discharge on the alloy usage
properties are necessary. The friction wear corrosion results in metal degradation due
to the simultaneous action of mechanical wear and (electro)chemical oxidation. The two
mechanisms do not proceed separately and depend on each other in a complex way. In
many cases, corrosion is accelerated by wear, and similarly, wear may be affected by corro-
sion phenomena. The nature of the passive film and the electrochemical conditions play a
148
Materials 2021, 14, 2088
significant role in friction, wear and corrosion mechanism [35]. When the implants are to
be placed in bone tissue, bone ingrowth into porous implant surfaces starts and osteoblast
cell adhesion on the substrate increases, which may improve osseointegration. Improved
osseointegration provides mechanical stability by interlocking the surrounding bone tissue
with the implant. Tribocorrosion is a material degradation process and involves chemical
and/or electrochemical interactions between material and its environment during a tribolog-
ical process. Friction wear corrosion is a degradation process resulting from the combined
action of small movements between contacting parts and the environment’s corrosivity [36].
The main purpose of this work is to investigate the effect of oxygen plasma glow
discharge temperature on the microstructure development and micromechanical proper-
ties of the near-β Ti–13Nb–13Zr alloy. An influence of the applied surface treatment on
tribocorrosion behavior of the alloy in Ringer’s solution is also studied.
Figure 1. Microstructure of the as-received alloy.
The samples were placed in a quartz tube furnace. After air evacuation, argon of
99.999% purity was purged into the tube. The pressure was adjusted to 10 Pa by flowing
argon (flow rate of 30 standard cm3 /min). When the samples reached the desired tempera-
ture of oxidation, i.e., 700 ◦ C or 850 ◦ C or 1000 ◦ C, a glow discharge was started. The glow
discharge was generated at 1250 V and an electric current of 45 mA. During discharge, high
purity oxygen (99.999%) was mixed with argon and pumped to the furnace in 20 cycles of
1 min duration (flow rate of 6 standard cm3 /min). During oxygen pumping, the pressure
was increased to 12 Pa. After oxidation, cooling to room temperature with working Ar
glow discharge was provided.
The microstructure of the oxygen hardened alloy was characterized by light mi-
croscopy (LM, ZEISS Axio Imager M1m microscope, Oberkochen, Germany), scanning
electron microscopy (SEM, FEI Nova NanoSEM 450 microscope, (Eindhoven, The Nether-
149
Materials 2021, 14, 2088
lands) and transmission electron microscopy (TEM, JEOL JEM-2010 ARP microscope,
(Tokyo, Japan) techniques. The phase constitution was studied by selected area electron
diffraction (SAED). The Java Electron Microscopy Software (JEMS, version 4.4131U2016,
Pierre Stadelmann, Switzerland) was used to interpret the SAED patterns. The lamella from
the cross-section of the coated alloy was prepared for the TEM investigation. The samples
were thinned using FEI Quanta 3D 200i scanning electron microscope equipped with a
Ga+ ion gun, Pt precursor gas injection systems (GIS) and OmniProbe micromanipulator
for in situ lift-out [38]. Ion beam accelerating voltage of 30 kV and ion currents in the
range of 15 nA–30 pA were applied. The lamella was transferred via a micromanipulator
to a TEM half ring, where a focused ion beam (FIB) milling to electron transparency was
performed (ion currents of 500–30 pA). FIB deposition process from Pt precursor was used
to attach the manipulator probe to the lamella and attach it to the grid. The phase identifi-
cation was supplemented by energy-dispersive X-ray spectroscopy (SEM-EDS, TEM-EDS)
microanalysis and STEM-EDS line analysis.
The open-circuit potential (OCP), linear sweep voltamperometry and electrochemical
impedance spectroscopy (EIS) were carried out using a potentiostat Autolab PGSTAT302N
(Metrohm Autolab B.V., Utrecht, The Nederlands). The reference electrode was a saturated
calomel electrode (SCE), and a platinum plate was used as the counter electrode. Ringer’s
solution was used as the electrolyte for the corrosion study. The chemical composition of
the Ringer’s solution was: 8.6 g/L NaCl, 0.3 g/L KCl, 0.25 g/L CaCl2 . Measurements were
performed at pH 7.4 in deaerated solutions at 37 ◦ C. The polarization test was performed
at a scan rate of 1 mV/s from −1.3 V to +2.2 V vs. SCE. For the EIS measurements, the
amplitude was 10 mV; the frequency was 105 Hz to 10−3 Hz. The EIS measurements were
performed at the OCP potential. The EIS data were fitted using ZView software.
The hardness and elastic modulus of the as-received and oxygen hardened Ti–13Nb–
13Zr alloys were determined by micro-combitester (CSM Instruments). The indentation
tests were performed using a Vickers diamond indenter. The indenter was loaded with
a force of 200 mN and kept for 5 s before un-loading. An analysis of the mechanical
properties of the alloy was carried out using the loading/unloading curve by Oliver and
Pharr method [39]. It allowed for the determination of the elastic modulus (EIT ) and
hardness (HIT ). For each sample, ten consecutive experiments at randomly selected places
were performed, and the average of 10 measurements was calculated.
The abrasive wear resistance and the friction coefficient of the as-received and treated
alloy were determined by friction tests based on the ISO 20808 standard [40]. The tests were
carried out in dry sliding contact with an Al2 O3 ball (6 mm diameter) using a ball-on-disc
tribotester (ITeE Radom, Radom, Poland). In a typical ball-on-disc arrangement, a counter-
element in the form of a ball was pressed against a rotating disc (sample) made of titanium
alloy. The tribological tests were repeated three times, with the same parameters: ball load
Fn = 5 N, sliding speed v = 0.07 m/s, sliding distance s = 1000 m, room temperature 23 ◦ C
and relative humidity 55%. The wear rate Wv = V/Fn × s was determined as the ratio of
the volume of material removed during friction (V) to the load (Fn ) and the sliding distance
(s). The volume was calculated based on the size of the cross-sectional area of the wear
groove. The groove profile was measured with a stylus profilometer in six places around
the wear track.
Friction-wear tests with simultaneous measurement of corrosion potential during
friction in Ringer’s solution were also performed. To record the changes in the corrosion
potential, a 2-electrode system with the working electrode (titanium alloy) and reference
electrode (calomel electrode in 3 M KCl) was designed. The schematic of the tribocorrosion
device is shown in Figure 2. A specially designed system ensuring an efficient and stable
electric contact with the alloy sample was used. In this system, the ball with the holder
was moving, whereas the titanium alloy sample was stationary. To fix the ball and sample
positions, special polymer holders were used. The same measurement parameters were
applied as in the case of dry friction tests. The friction process began after a stable corrosion
150
Materials 2021, 14, 2088
potential for the titanium alloy was reached. When the friction was activated, a change in
the potential was recorded as a result of the wear of the alloys’ surface layer.
Figure 2. Schematic of the tribological device with parallel corrosion potential measurement during wear tests of the
titanium alloy in sliding contact with a ball.
Before each test, the alumina ball and titanium alloy were ultrasonically cleaned in
ethanol and left to dry. Additionally, the ball used in the tribocorrosion test was also
washed in Ringer’s solution. The holders were subjected to an identical cleaning procedure.
The whole measurement took place in a plastic vessel filled with Ringer’s solution.
3. Results
3.1. Microstructure of the Oxygen Hardened Alloy
LM and SEM images of the alloy treated at 700 ◦ C, 850 ◦ C and 1000 ◦ C are shown
in Figure 3. The alloy microstructure was dependent on the hardening temperature. The
differences in the thickness of the oxygen-rich hardened zone and the grain size of the bulk
of the material were observed. The highest thickness of the hardened zone, up to 470 μm,
was found in the sample treated at 1000 ◦ C (Figure 3e,f). In the case of alloy treated at
850 ◦ C, the hardened zone thickness was about 160 μm (Figure 3c,d), while in the alloy
treated at 700 ◦ C it had the lowest thickness, up to 120 μm (Figure 3a,b). A coarsening
of the α’ laths in the oxygen-hardened zone with increasing treatment temperature was
also observed. The voids in the hardened zone of the sample treated at 1000 ◦ C were
formed at a distance of ~20–30 μm from the surface. It should be noted that the presence of
pores usually contributes to the formation of microcracks, which were also sporadically
observed in this sample. A typical microcrack is marked with an arrow in Figure 3f.
The observed microstructural defects exclude the sample hardened at 1000 ◦ C from the
intended biological applications.
151
Materials 2021, 14, 2088
Investigation of the cross-section showed a significant grain growth resulting from the
hardening process. The grain size estimated from LM and SEM images was in the range
of 40–100 μm, 100–250 μm and up to 700 μm for the alloys treated at 700 ◦ C, 850 ◦ C and
1000 ◦ C, respectively.
The sample treated at 700 ◦ C was selected for a detailed microstructure characteri-
zation by TEM and STEM. The TEM and STEM images are shown in Figures 4 and 5. In
the near-surface region (depth up to 10 μm from the surface), a high fraction of the Ti α’
152
Materials 2021, 14, 2088
(O) solid solution and low amount of fine laths of the Ti α” (O) solid solution in β phase
were found (Figure 4). In the SAED pattern no. 2, the diffraction spots from crystal planes
belonging to particular three [100] Tiα“ zone axes corresponding to three sets of the Ti α”
laths inclined by the angle of 60◦ are marked with black, green and red color, respectively.
Figure 4. TEM image of the near-surface region in the Ti–13Nb–13Zr alloy after oxygen hardening at 700 ◦ C. SAED patterns
of α’ (hcp) and α” (orthorhombic, Cmcm) were taken from areas marked with 1 and 2, respectively. In the SAED pattern
no. 2, the spots belonging to the three [100] α“ zone axes are marked with black, green and red color. Indices of green spots
are given.
Figure 5. STEM image of the microstructure of the near-surface region in the Ti–13Nb–13Zr alloy cross-section after oxygen
hardening at a temperature of 700 ◦ C (a) and concentration profile of oxygen obtained by TEM-EDS microanalysis performed
in points 1–16 (b). The exemplary grains of the α‘ and β phases, as well as the areas 1 and 2 given in Figure 4, are marked
in Figure 5a.
153
Materials 2021, 14, 2088
6 at.%. At greater depths in the sample, the oxygen concentration in the α‘ phase remained
constant at ~6 at.%. The obtained results of the oxygen distribution should be treated as
approximate values since EDS microanalysis gives only a rough indication of light elements
concentration. Nevertheless, the result is an indication that the Ti α’ (O) solid solution is
enriched in oxygen in the near-surface zone at a depth of up to 6 μm.
In our previous study [33], we showed that the near-surface region of the oxygen
diffusion hardened two-phase (α + β) Ti–6Al–4V alloy consisted of Ti α (O) solid solution
enriched with oxygen mainly. Oxygen is a strong interstitial solid solution strengthening
element of titanium [36]. It is an α phase stabilizer and has a high solubility in the hcp α
phase, up to 31.9 at.%. The solubility of oxygen in the β phase is much lower, maximum
8 at.% [41]. Therefore, the presence of the Ti α’ (O) phase in the near-surface zone is
preferred due to the diffusion of interstitial oxygen atoms.
According to [39], the plasma glow discharge strengthens the oxygen diffusion into
the metallic substrate. It is likely due to an increase in the number of point defects formed
during the first stage of the process. In addition, the plasma glow discharge inhibits the
formation of the rutile layer on the titanium alloy surface. The oxides formed on the alloy
surface act as limited reservoirs of oxygen atoms, which are then forced to diffuse into the
alloy matrix and form a solid solution [32]. Therefore, the surface of the alloy investigated
in this work was not covered by titanium oxide.
Table 1. Microhardness (HIT ), elastic modulus (EIT ), the penetration depth of the indenter (hmax ),
and wear rate (Wv ) of the Ti–13Nb–13Zr alloy.
The hardening improved the mechanical properties of the Ti–13Nb–13Zr alloy, and both
the hardness and elastic modulus increased about three times compared to the baseline alloy.
Interstitial oxygen diffusion hardening of the alloy carried out at 700 ◦ C allowed to achieve
154
Materials 2021, 14, 2088
a hardness comparable to the hardness of the alloy treated by plasma electrolytic oxidation
(PEO) in an electrolyte with and without the addition of zirconia nanoparticles [43].
The wear resistance of the titanium alloy was tested in dry sliding contact. Figure 6
shows the average COF of the as-received and hardened alloy. The COF of the hardened
alloy samples in dry sliding contact with the Al2 O3 ball was in the range of 0.63–0.78. A
significantly lower COF = 0.50 occurred during the friction of the baseline titanium alloy,
and the cooperation with the Al2 O3 counterpart was more stable.
Figure 6. COF of the alloy oxygen hardened at 700 ◦ C (a), 850 ◦ C (b) and 1000 ◦ C (c) compared to as-received alloy (d) in
dry friction condition as well as the alloy hardened at 700 ◦ C in Ringer’s solution (e) against alumina ball.
155
Materials 2021, 14, 2088
typical of dry friction in contact with a hard counterpart and has already been analyzed in
detail elsewhere [47].
Figure 7. Cross-section profile of wear track of as-received Ti–13Nb–13Zr alloy (a) and hardened Ti–13Nb–13Zr alloy
(b) after dry friction.
The best mechanical and tribological properties were found for the Ti–13Nb–13Zr
alloy hardened at 700 ◦ C. Therefore, this alloy was selected for further corrosion resistance
tests. Figure 8a shows the evolution of the OCP for the as-received and the alloy hardened
at 700 ◦ C. The results of Eocp show that the as-received alloy has a less noble potential than
the alloy heat-treated at 700 ◦ C, indicating that the as-received alloy is more susceptible to
corrosion. The OCP slightly increased for the hardened alloy and reached a stable value
after about 2000s.
Figure 8. Electrochemical measurements of as-received and oxygen hardened alloy (Ti–13Nb–13Zr/700 ◦ C) in Ringer’s
solution at 37 ◦ C, (a) evolution of the corrosion potential vs. time and (b) polarization curves at 1 mV/s.
156
Materials 2021, 14, 2088
anodic transition increased from about −0.43 V up to −0.25 V. These results indicate that
the as-received alloy has a slightly smaller corrosion rate than the treated one.
Figure 9 shows the EIS graphs presented as a Bode plot (Figure 9a) and a Nyquist plot
(Figure 9b) of the as-received and hardened alloy in the Ringer’s solution. From Figure 9a,
the Z modulus at a lower frequency in the Bode impedance plot indicated a comparable
corrosion resistance of the investigated samples. The as-received and treated alloys showed
a highly capacitive behavior from medium to low frequencies. The equivalent circuit, as
shown in Figure 10, was used to fit the EIS data. According to the double-layer model
for the oxygen hardened alloy, the equivalent circuit consisted of the electrolyte resistance
(R1), the treated resistance (R2) and the constant phase elements (CPE). A good fitting
between the experimental and simulated results was achieved, and the parameters are
listed in Table 2.
Figure 9. Electrochemical impedance curves of the as-received and oxygen hardened alloy in Ringer’s solution. (a) Bode
impedance and phase angle plot, (b) Nyquist impedance plot.
Table 2. Chi-squared (χ2 ) values obtained by fitting equivalent electrical circuit with Z View software and electrochemical
parameters for the as-received and treated titanium alloy.
To investigate the passivation kinetics of the hardened alloy in a condition where the
oxygen-rich hardened layer of the alloy is abraded, tribological tests were performed in the
presence of Ringer’s solution. The tests were coupled with simultaneous measurement of
the OCP (Figure 2). Figure 11 shows the change in the corrosion potential over time. In the
diagram, we can distinguish 3 characteristic stages:
i) Stage 1—increase and stabilization of the stationary potential;
ii) Stage 2—a drop of the corrosion potential, resulting from the wear of the alloy’s
surface, with the assumed sliding distance equaling 1000 m;
iii) Stage 3—an increase of the stationary potential after the interruption of the friction
process of the alloy’s surface.
157
Materials 2021, 14, 2088
Figure 11. Change in the corrosion potential during the tribological test in Ringer’s solution at 25 ◦ C
of the alloy hardened at 700 ◦ C.
The friction process was activated after 1 h 20 min of stabilization in Ringer’s solu-
tion. A significant drop in the potential corrosion value was observed, resulting from the
change in the measurement conditions. As a result of friction, a groove was formed, with
significant surface roughness and less passivating, thereby showing a considerable drop
of the corrosion potential. The potential reached about −846 ± 50 mV vs. SCE and was
maintained at this value during the whole friction process. When the friction process was
terminated (Stage 3), an increase in the corrosion potential was observed. The increase
in the potential took place abruptly. The strengthened outer layer of the alloy treated at
700 ◦ C provided good protection against tribological wear in a corrosive environment, and
the passive film was quickly restored.
4. Conclusions
In this work, the oxygen hardening of Ti–13Nb–13Zr alloy by plasma glow discharge
at 700–1000 ◦ C was studied.
1. The hardening temperature had a significant influence on the alloy microstructure
and thickness of the hardened zone. A refinement of the α’ laths of the near β-phase
Ti alloy was observed. The thickness of the hardened zone and grain size of the bulk
alloy both increased with increasing temperature. In addition, voids and microcracks
were observed in the near-surface zone of the alloy treated at 1000 ◦ C;
2. The outer hardened zone consisted mainly of the Ti α’ (O) solid solution with small
amounts of fine laths of the Ti α” (O) solid solution in the β phase. Oxygen enrichment
in a depth of up to 6 μm was found;
3. The oxidation of the Ti–13Nb–13Zr alloy under glow discharge conditions resulted
in a significant increase of hardness and elastic modulus compared with the base
alloy. The best results were found for the alloy hardened at 700 ◦ C. With an increasing
temperature, a decrease in both hardness and modulus of elasticity in the hardened
zone were observed;
4. The hardened titanium alloy zone significantly reduces abrasive wear, and the wear
resistance is proportional to the hardness of the alloy;
5. Oxygen hardened alloy does not adversely affect the corrosion resistance;
6. The friction reduces the corrosion resistance of the oxygen-hardened Ti–13Nb–13Zr
alloy. However, when the friction process was stopped, the corrosion potential was
158
Materials 2021, 14, 2088
quickly restored. The strengthened outer layer of the alloy treated at 700 ◦ C provides
good protection against tribological wear in a corrosive environment.
Author Contributions: Conceptualization, A.Ł., S.Z. and T.M.; formal analysis, A.Ł., S.Z., W.P.,
B.D. and T.M.; funding acquisition, T.M. and S.Z.; investigation, A.Ł., S.Z., W.P., B.D. and T.M.;
methodology, A.Ł., S.Z., W.P., B.D. and T.M., Project administration, T.M. and S.Z.; resources, T.M.;
supervision, T.M. and S.Z.; validation, A.Ł., S.Z., B.D. and T.M.; visualization, A.Ł., S.Z., B.D. and
T.M; writing—original draft, A.Ł., S.Z., W.P., B.D. and T.M.; writing—review and editing, A.Ł., S.Z.
and T.M. All authors have read and agreed to the published version of the manuscript.
Funding: The study was supported by the National Science Centre, Poland (decision no. DEC-
2016/21/B/ST8/00238). Part of this work concerning the mechanical and tribocorrosion investigation
was supported by AGH University of Science and Technology, Faculty of Mechanical Engineering
and Robotics, project no. 16.16.130.942/2021.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: The authors appreciate the valuable contributions of M. Gajewska (ACMiN
AGH) for FIB lamella preparation and Ł. Cieniek, for SEM investigation.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or
in the decision to publish the results.
References
1. Slokar, L.; Matković, T.; Matković, P. Alloy design and property evaluation of new Ti–Cr–Nb alloys. Mater. Des. 2012, 33, 26–30.
[CrossRef]
2. Cojocaru, V.D.; Raducanu, D.; Cinca, I.; Dan, I.; Ivanescu, S.; Jalba, E. The mechanical properties evaluation for an as-cast Ti-Ta-Zr
alloy. Metal. Int. 2011, 16, 35–38. [CrossRef]
3. Raducanu, D.; Vasilescu, E.; Cojocaru, V.D.; Cinca, I.; Drob, P.; Vasilescu, C.; Drob, S.I. Mechanical and corrosion resistance of a
new nanostructured Ti-Zr-Ta-Nb alloy. J. Mech. Behav. Biomed. Mater. 2011, 4, 1421–1430. [CrossRef]
4. Tong, Y.X.; Guo, B.; Zheng, Y.F.; Chung, C.Y.; Ma, L.W. Effects of Sn and Zr on the Microstructure and Mechanical Properties of
Ti-Ta-Based Shape Memory Alloys. J. Mater. Eng. Perform. 2011, 20, 762–766. [CrossRef]
5. Niinomi, M. Mechanical properties of biomedical titanium alloys. Mater. Sci. Eng. A 1998, 243, 231–236. [CrossRef]
6. Niinomi, M. Mechanical biocompatibilities of titanium alloys for biomedical applications. J. Mech. Behav. Biomed. Mater. 2008, 1,
30–42. [CrossRef]
7. Cremasco, A.; Messias, A.D.; Esposito, A.R.; de RezendeDuek, E.A.; Caram, R. Effects of alloying elements on the cytotoxic
response of titanium alloys. Mater. Sci. Eng. C 2011, 31, 833–839. [CrossRef]
8. Robin, A.; Carvalho, O.A.S.; Schneider, S.G.; Schneider, S. Corrosion behavior of Ti-xNb-13Zr alloys in Ringer’s solution. Mater.
Corros. 2008, 59, 929–933. [CrossRef]
9. Khan, M.A.; Williams, R.L.; Williams, D.F. The corrosion behaviour of Ti-6Al-4V, Ti-6Al-7Nb and Ti-13Nb-13Zr in protein
solutions. Biomaterials 1999, 20, 631–637. [CrossRef]
10. Zhao, C.; Zhang, X.; Cao, P. Mechanical and electrochemical characterization of Ti–12Mo–5Zr alloy for biomedical application. J.
Alloys Compd. 2011, 509, 8235–8238. [CrossRef]
11. Geetha, M.; Singh, A.K.; Asokamani, R.; Gogia, A.K. Ti based biomaterials, the ultimate choice for orthopaedic implants—A
review. Prog. Mater. Sci. 2009, 54, 397–425. [CrossRef]
12. Ke, Z.; Yi, C.; Zhang, L.; Yuan, Z.; He, Z.; Tan, J.; Jiang, Y. Characterization of a new Ti-13Nb-13Zr-10Cu alloy with enhanced
antibacterial activity for biomedical applications. Mater. Lett. 2019, 253, 335–338. [CrossRef]
13. Mohammed, M.; Khan, Z.; Siddiquee, A. Beta Titanium Alloys: The Lowest Elastic Modulus for Biomedical Applications: A
review. Int. J. Chem. Mol. Nucl. Mater. Metall. Eng. 2014, 8, 822–827. [CrossRef]
14. Barjaktarević, D.; Medjo, B.; Gubeljak, N.; Cvijović-Alagić, I.; Štefane, P.; Djokic, V.; Rakin, M. Experimental and numerical
analysis of tensile properties of Ti-13Nb-13Zr alloy and determination of influence of anodization process. Procedia Struct. Integr.
2020, 28, 2187–2194. [CrossRef]
15. Mohan, P.; Rajak, D.K.; Pruncu, C.I.; Behera, A.; Amigó-Borrás, V.; Abou Bakr Elshalakany, A. Influence of β-phase stability in
elemental blended Ti-Mo and Ti-Mo-Zr alloys. Micron 2021, 142, 102992. [CrossRef]
159
Materials 2021, 14, 2088
16. Dong, H. Tribological properties of titanium-based alloys. In Woodhead Publishing Series in Metals and Surface Engineering. Surface
Engineering of Light Alloys. Aluminium, Magnesium and Titanium Alloys; Dong, H., Ed.; Woodhead Publishing: Cambridge, UK,
2010; pp. 58–80. [CrossRef]
17. Garbacz, H.; Wieciński, P.; Ossowski, M.; Ortore, G.; Wierzchoń, T.; Kurzydłowski, K.J. Surface engineering techniques used for
improving the mechanical and tribological properties of the Ti6A14V alloy. Surf. Coat. Technol. 2008, 202, 2453–2457. [CrossRef]
18. Dong, H.; Bell, T. Enhanced wear resistance of titanium surfaces by a new thermal oxidation treatment. Wear 2000, 238, 131–137.
[CrossRef]
19. Muraleedharan, T.M.; Meletis, E.I. Surface modification of pure titanium and Ti-6A1-4V by intensified plasma ion nitriding. Thin
Solid Films 1992, 221, 104–113. [CrossRef]
20. Hutchings, R.; Oliver, W.C. A study of the improved wear performance of nitrogen-implanted Ti-6Al-4V. Wear 1983, 92, 143–153.
[CrossRef]
21. Wilson, A.D.; Leyland, A.; Matthews, A. A comparative study of the influence of plasma treatments, PVD coatings and ion
implantation on the tribological performance of Ti–6Al–4V. Surf. Coat. Technol. 1999, 114, 70–80. [CrossRef]
22. Zhou, L.; Tiechui Yuan, T.; Ruidi Li, R.; Lanbo Li, L. Two ways of evaluating the wear property of Ti-13Nb-13Zr fabricated by
selective laser melting. Mater. Lett. 2019, 242, 9–12. [CrossRef]
23. Kong, Q.; Lai, X.; An, X.; Feng, W.; Lu, C.; Wu, J.; Wu, C.; Wu, L.; Wang, Q. Characterization and corrosion behaviour of
Ti-13Nb-13Zr alloy prepared by mechanical alloying and spark plasma sintering. Mater. Today Commun. 2020, 23, 101130.
[CrossRef]
24. Jugowiec, D.; Kot, M.; Moskalewicz, T. Electrophoretic deposition and characterization of chitosan coating on near-beta titanium
alloy. Arch. Metall. Mater. 2016, 61, 657–664. [CrossRef]
25. Sak, A.; Moskalewicz, T.; Zimowski, S.; Cieniek, Ł.; Dubiel, B.; Radziszewska, A.; Kot, M.; Łukaszczyk, A. Influence of
polyetheretherketone coatings on the Ti-13Nb-13Zr titanium alloy’s bio-tribological properties and corrosion resistance. Mater.
Sci. Eng. C 2016, 63, 52–61. [CrossRef] [PubMed]
26. Hornberger, H.; Randow, C.; Fleck, C. Fatigue and surface structure of titanium after oxygen diffusion hardening. Mater. Sci. Eng.
A 2015, 630, 51–57. [CrossRef]
27. Grabarczyk, J.; Batory, D.; Kaczorowski, W.; Pazik, ˛ B.; Januszewicz, B.; Burnat, B.; Czerniak-Reczulska, M.; Makówka, M.;
Niedzielski, P. Comparison of Different Thermo-Chemical Treatments Methods of Ti-6Al-4V Alloy in Terms of Tribological and
Corrosion Properties. Materials 2020, 13, 5192. [CrossRef]
28. Dong, H.; Li, X.Y. Oxygen boost diffusion for the deep-case hardening of titanium alloys. Mater. Sci. Eng. A 2000, 280, 303–310.
[CrossRef]
29. Zhang, Z.X.; Dong, H.; Bell, T.; Xu, B.S. The effect of deep-case oxygen hardening on the tribological behaviour of a-C:H DLC
coatings on Ti6Al4V alloy. J. Alloys Compd. 2008, 464, 519–525. [CrossRef]
30. Zabler, S. Interstitial Oxygen diffusion hardening—A practical route for the surface protection of titanium. Mater. Charact. 2011,
62, 1205–1213. [CrossRef]
31. Jamesha, M.; Sankara Narayanan, T.S.N.; Chu, P.K. Thermal oxidation of titanium: Evaluation of corrosion resistance as a function
of cooling rate. Mater. Chem. Phys. 2013, 138, 565–572. [CrossRef]
32. Januszewicz, B.; Siniarski, D. The glow discharge plasma influence on the oxide layer and diffusion zone formation during
process of thermal oxidation of titanium and its alloys. Vacuum 2006, 81, 215–220. [CrossRef]
33. Moskalewicz, T.; Wendler, B.; Zimowski, S.; Dubiel, B.; Czyrska-Filemonowicz, A. Microstructure, micro-mechanical and
tribological properties of the nc-WC/a-C nanocomposite coatings magnetron sputtered on non-hardened and oxygen hardened
Ti–6Al–4V alloy. Surf. Coat. Technol. 2010, 205, 2668–2677. [CrossRef]
34. Poggie, R.A.; Kovacs, P.; Davidson, J.A. Oxygen Diffusion Hardening of Ti-Nb-Zr Alloys. Mater. Manuf. Process. 1996, 11, 185–197.
[CrossRef]
35. Barril, S.; Mischler, S.; Landolt, D. Electrochemical effects on the fretting corrosion behaviour of Ti6Al4V in 0.9% sodium chloride
solution. Wear 2005, 259, 282–291. [CrossRef]
36. Albayrak, Ç.; Hacısalihoğlu, İ.; Yenalvangölü, S.; Alsaran, A. Tribocorrosion behavior of duplex treated pure titanium in Simulated
Body Fluid. Wear 2013, 302, 1642–1648. [CrossRef]
37. Davidson, J.A.; Mishra, A.K.; Kovacs, P.; Poggie, R.A. New surface-hardened, low-modulus, corrosion-resistant Ti-13Nb-13Zr
alloy for total hip arthroplasty. Biomed. Mater. Eng. 1994, 4, 231–243. [CrossRef]
38. Langford, R.M.; Clinton, C. In situ lift-out using a FIB-SEM system. Micron 2004, 35, 607–611. [CrossRef] [PubMed]
39. Oliver, W.C.; Pharr, G.M. Measurement of hardness and elastic modulus by instrumented indentation: Advances in understanding
and refinements to methodology. J. Mater. Res. 2004, 19, 3–20. [CrossRef]
40. ISO 20808:2016 Fine Ceramics (Advanced Ceramics, Advanced Technical Ceramics)—Determination of Friction and Wear
Characteristics of Monolithic Ceramics by Ball-on-Disc Method. Available online: https://www.iso.org/standard/65415.html
(accessed on 10 April 2021).
41. Massalski, T.B.; Okamoto, H.; Subramanian, P.R.; Kacprzak, L. Binary Alloy Phase Diagrams; ASM International: Materials Park,
OH, USA, 1990; Volume 3. [CrossRef]
42. Lee, T. Variation in Mechanical Properties of Ti-13Nb-13Zr Depending on Annealing Temperature. Appl. Sci. 2020, 10, 7896.
[CrossRef]
160
Materials 2021, 14, 2088
43. Lederer, S.; Lutz, P.; Fürbeth, W. Surface modification of Ti 13Nb 13Zr by plasma electrolytic oxidation. Surf. Coat. Technol. 2018,
335, 62–71. [CrossRef]
44. Krishna, D.S.R.; Brama, Y.L.; Sun, Y. Thick rutile layer on titanium for tribological applications. Tribol. Int. 2007, 40, 329–334.
[CrossRef]
45. Kolubaev, A.V.; Tarasov, S.Y. Studies on formation and destruction of surface layers under severe friction. Facta Univ. 1997, 1,
429–432.
46. Majumdar, P.; Singh, S.B.; Chakraborty, M. Wear response of heat-treated Ti–13Zr–13Nb alloy in dry condition and simulated
body fluid. Wear 2008, 264, 1015–1025. [CrossRef]
47. Mohan, L.; Anandan, C. Wear and corrosion behavior of oxygen implanted biomedical titanium alloy Ti–13Nb–13Zr. Appl. Surf.
Sci. 2013, 282, 281–290. [CrossRef]
48. Mc Cafferty, E.; Hubler, G.K. Electrochemical behavior of palladium implanted titanium. J. Electrochem. Soc. 1978, 125, 1892–1893.
[CrossRef]
161
materials
Article
The Effect of Sn Addition on Zn-Al-Mg Alloy; Part I:
Microstructure and Phase Composition
Peter Gogola *, Zuzana Gabalcová, Martin Kusý and Henrich Suchánek
Institute of Materials Science, Faculty of Materials Science and Technology in Trnava, Slovak University of
Technology in Bratislava, Ulica Jána Bottu 25, 917 24 Trnava, Slovakia; [email protected] (Z.G.);
[email protected] (M.K.); [email protected] (H.S.)
* Correspondence: [email protected]
Abstract: In this study, the addition of Sn on the microstructure of Zn 1.6 wt.% Al 1.6 wt.% Mg alloy
was studied. Currently, the addition of Sn into Zn-Al-Mg based systems has not been investigated in
detail. Both as-cast and annealed states were investigated. Phase transformation temperatures and
phase composition was investigated via DSC, SEM and XRD techniques. The main phases identified
in the studied alloys were η(Zn) and α(Al) solid solutions as well as Mg2 Zn11 , MgZn2 and Mg2 Sn
intermetallic phases. Addition of Sn enabled the formation of Mg2 Sn phase at the expense of Mgx Zny
phases, while the overall volume content of intermetallic phases is decreasing. Annealing did not
change the phase composition in a significant way, but higher Sn content allowed more effective
spheroidization and agglomeration of individual phase particles.
Keywords: Zn-based alloy; phase composition; XRD; DSC; microstructure formation; Sn-addition;
intermetallic phases
200 ◦ C, the pure Zn coating starts to react with the steel [1,3,4] substrate and continues to
Received: 27 July 2021
Accepted: 14 September 2021
form ZnFe intermetallics. Such a reaction can reduce the actual steels cross section area
Published: 18 September 2021
and thus reduce the cables’ mechanical properties. This limitation can be overcome by
alloying [5,6].
Publisher’s Note: MDPI stays neutral
Nowadays, various Zn-Al-Mg alloy system coatings are available, such as the commer-
with regard to jurisdictional claims in
cially well-known Magizinc (MZ) with a chemical composition of Zn 1.6 wt.% Al 1.6 wt.%
published maps and institutional affil- Mg [7–14].
iations. Mg is added to Zn-based coatings to further increase their corrosion protection ca-
pabilities. Mg is useful especially for increasing the galvanic protection offered by a
coating at cut-edges and mechanically damaged spots. Additional alloying may further
improve the properties of Zn-Al-Mg alloys as reported by several literature sources [15–25].
Unfortunately, there is still lack of information about Sn addition into these systems [26].
Copyright: © 2021 by the authors.
Licensee MDPI, Basel, Switzerland.
In recent years, the effects of Sn addition on the microstructure of Mg- and Al-based
This article is an open access article
alloys has been studied. This includes microstructural stability upon thermal exposure. Sn
distributed under the terms and is known to have a high affinity to Mg, creating mainly the Mg2 Sn intermetallic phase. All
conditions of the Creative Commons in all, a positive influence of Sn on Mg and Al-based alloys has been reported depending
Attribution (CC BY) license (https:// on the amount of Sn added to these alloys. Alloying enabled the formation of Mg2 Sn
creativecommons.org/licenses/by/ in all these alloys [27–33]. The aim of this research is to confirm if Mg2 Sn phase is also
4.0/). preferred compared to Mgx Zny phases in the current alloy system as suggested by studied
163
literature sources [32,34]. Phase composition and overall microstructure character will be
investigated on as-cast samples.
As one of the potential applications involves a long-term thermal exposure, microstruc-
ture and phase composition will be investigated after a representative annealing treatment
of each alloy as well. Experimental annealing temperature will be set to 310 ◦ C to be clearly
above potential exposure temperatures, but below the melting temperatures (~340 ◦ C) of
the investigated alloys.
Mg2 Zn11 and MgZn2 are the most common intermetallic phases in the currently in-
vestigated Zn-based systems. MgZn2 is reported to be less ductile, while both are reported
to be slightly less ductile compared to Mg2 Sn [18–20,29,35,36]. The possibility to replace
Mgx Zny phases by Mg2 Sn and thus reduce the overall volume content of intermetallic
phases in a Zn 1.6 wt.% Al 1.6 wt.% Mg based alloy would indicate an interesting re-
search path for further extensive investigation of corrosion and mechanical properties of
such alloys.
Alloy Al Mg Sn Zn
MZ + 0.0Sn 1.56 ± 0.07 1.40 ± 0.01 0.07 ± 0.02 bal.
MZ + 0.5Sn 1.64 ± 0.02 1.41 ± 0.01 0.52 ± 0.01 bal.
MZ + 1.0Sn 1.62 ± 0.03 1.45 ± 0.02 1.06 ± 0.02 bal.
MZ + 2.0Sn 1.57 ± 0.01 1.44 ± 0.01 1.95 ± 0.01 bal.
MZ + 3.0Sn 1.57 ± 0.12 1.43 ± 0.05 2.69 ± 0.06 bal.
Casting was done from 470 ◦ C of melt temperature into a water-cooled copper mold
with a diameter of 30 mm and depth of 20 mm. During casting, the sample temperature
was continuously measured at a sampling frequency of 25 Hz with K-type thermocouples
attached to the mold surface. The cooling rate of 60 ◦ C/s was established.
Two types of cylindrical samples were prepared for each alloy: (i) as-cast samples
and (ii) cast and subsequently solution annealed at 310 ◦ C for 1 h. The annealing step was
finished by ice-water quenching. A cooling rate of 75 ◦ C/s was recorded. The selected
solution annealing temperature corresponds to the γ + η region of the Zn-Al system [37],
while it is clearly below the melting point of all chosen alloys. Cooling in cold water
ensured a very good control of the annealing time.
Vickers hardness tests were carried out in line with ISO 6507-1 [38] on polished
surfaces of as-cast and annealed samples via a BUEHLER Indentamet 1105 (Buehler Ltd.,
Lake Bluff, IL, USA) hardness tester at an applied load of 9.8 N, holding time at the point
of load application was 10 s.
DSC measurements were carried out by the Perkin Elmer Diamond DSC (Perkin Elmer
Inc., Billerica, MA, USA) device. The DSC samples were cut from as-cast samples to a
target weight of 5 mg. The samples were heated to the temperature of 500 ◦ C at a heating
rate of 10 ◦ C/min and then cooled to ambient temperature at a cooling rate of 10 ◦ C/min
under the protective argon atmosphere.
The XRD analysis was carried out on metallic filings of the as-cast and annealed
samples by the PANalytical Empyrean X-ray diffractometer (XRD) (Malvern Panalytical
164
Materials 2021, 14, 5404
Ltd., Malvern, UK). The procedure to measure on metallic filings instead of bulk castings
was chosen to limit the influence of casting texture on the recorded XRD pattern. The
casting texture added additional complexity to the XRD measurements by influencing the
theoretical relative intensities for the individual crystallographic planes. This made the
quantitative analysis very unreliable due to the complex texture corrections needed. The
measurements were performed in Bragg–Brentano geometry. Theta-2Theta angle range
between 10◦ and 148◦ 2Theta was chosen. The XRD source was set to 40 kV and 40 mA. The
incident beam was modified by 0.04 rad soller slit, 1/4◦ divergence slit and 1/2◦ anti-scatter
slit. The diffracted beam path was equipped with a 1/2◦ anti-scatter slit, 0.04 rad soller slit,
Ni beta filter and PIXcel3D position sensitive detector operated in 1D scanning mode. The
phase quality was analyzed using PANalytical Xpert High Score program (HighScore Plus
version 3.0.5) with the ICSD FIZ Karlsruhe database. Quantitative results were determined
from XRD patterns using the Rietveld refinement-based program MAUD version 2.84 [39].
The program uses an asymmetric pseudo-Voight function to describe experimental peaks.
Instrument broadening was determined by measuring the NIST660c LaB6 (The National
Institute of Standards and Technology, Gaithersburg, MD, USA) line position and line
broadening standard and introduced to the Rietveld refinement program via the Caglioti
equation. Anisotropy size-strain model was applied to Zn solid solution while other phases
were treated by isotropic models. A minor discrepancy between the nominal and measured
peak intensities was corrected using the spherical harmonic functions with fibre symmetry.
The quality of the fit was in all analyzed samples below 10% Rwp .
The metallographic preparation of DSC, as-cast and annealed samples consisted of
standard grinding using abrasive papers and polishing on diamond pastes with various
grain sizes of down to 0.25 μm.
The microstructure evaluation was performed by the JEOL JSM 7600F scanning elec-
tron microscopy (SEM, Jeol Ltd., Tokyo, Japan) with a Schottky field emission electron
source operating at 20 kV and 90 μA. The samples were placed at a working distance of
15 mm and documented using a backscattered electron detector. The chemical element
analysis was performed via the Oxford Instruments X-Max silicon drift detector, energy
dispersive X-ray spectrometer (EDS, Oxford Instruments plc, Abingdon, UK).
Image analysis was performed on at least 15 sites for each sample by ImageJ FIJI
1.53c software [40]. Area ratio of η(Zn) based areas and other microstructure components
were established.
All results are listed as the average values of multiple measurements with ±standard
deviation error bars.
3. Results
The DSC curves of MZ + xSn alloys system in Figure 1a–e show charts of DSC heating
runs. Figure 1 shows the most relevant section of the recorded data, while measurements
were done from 20 to 500 ◦ C at a heating rate of 10 ◦ C/min. The first heating runs recorded
are presented to observe the reactions in the as-cast samples during heating and subsequent
melting, including eutectoid reactions (not visible in DSC cooling runs).
165
Materials 2021, 14, 5404
Hence, the individual microstructure features in the as-cast samples cannot be clearly
distinguished; the DSC samples after the cooling run were investigated via SEM. These
observations enabled us to identify the individual reactions for each recorded peak.
Peaks were recorded describe the reactions of the following phases: η(Zn)–hcp Zn-
based solid solution; γ(Al)–fcc Al-based solid solution present above the eutectoid reaction
temperature reported at 275 ◦ C in the Zn-Al system [37] up to melting temperature; α(Al)–
fcc Al-based solid solution present below the Zn-Al system eutectoid reaction; Mg2 Zn11 ,
MgZn2 and Mg2 Sn intermetallic phases of respective systems.
The curve in Figure 1a corresponds to the MZ + 0.0Sn alloy. The first peak observed at
about 285.0 ◦ C corresponds to the eutectoid transformation α(Al) + η(Zn) → γ(Al). This
peak is repeated for all alloys observed. Melting of pure MZ + 0.0Sn starts at 344.0 ◦ C
(peak maximum at 347.5 ◦ C) with melting of the ternary eutectic consisting of η(Zn), γ(Al)
and Mg2 Zn11 phases. An example of such areas is provided in Figure 2a. This is followed
by the melting of the Zn/Mgx Zny binary eutectic (peak maximum at 357.0 ◦ C). A clear
example can be observed in Figure 2b. Mgx Zny corresponds to a mixture of Mg2 Zn11 and
MgZn2 phases as documented in Figure 2c. A closer detail of this area is given in Figure 2d)
showing also the α(Al) + η(Zn) eutectoid particles in detail. Chemical composition of the
individual phases is documented in Table 2. The last peak corresponds with the melting of
the Zn rich dendrites (peak maximum at 378.9 ◦ C) [41].
166
Materials 2021, 14, 5404
Figure 2. Selected features of MZ + 0.0Sn alloy microstructure after DSC measurement: (a) example of ternary eutectic;
(b) example of binary eutectic; (c) example of Mg2 Zn11 and MgZn2 mixture; (d) area with EDS measurement points 1–4
listed in Table 2.
Site No.
Chemical Element (at.%) 1 2 3 4 5
Zn 99.20 54.50 83.70 66.30 66.10
Al 0.80 45.50 - - -
Mg - - 16.30 33.70 -
Sn - - - - 33.90
Phase/Region η(Zn) α(Al) + η(Zn) eutectoid Mg2 Zn11 MgZn2 Mg2 Sn
As indicated in Figure 2a, η(Zn) dendrites are always decorated by a needle like
α(Al) particles. These are formed as a result of the decreasing solubility of Al in Zn in
the temperature range below ~285 ◦ C (see Figure 1). These particles are observed for all
alloys investigated.
As observed in Figure 1b, the addition of 0.5 wt.% of Sn enables the formation of a peak
at 335 ◦ C. This effect corresponds to newly emerging quaternary eutectic areas consisting
of η(Zn), γ(Al), Mg2 Zn11 and Mg2 Sn phases. A typical such area is shown in Figure 3a)
with a selected detail in Figure 3b. The chemical composition of the Mg2 Sn phase was
measured and listed in Table 2. These areas start to melt at about 334 ◦ C (peak maximum
at 335.0 ◦ C). The next peak corresponds to the melting of the ternary eutectic (342.0 ◦ C)
composed of η(Zn), γ(Al) and Mg2 Zn11 . The adjacent peak indicates the melting of the
η(Zn) + Mg2 Zn11 binary eutectic (353.9 ◦ C). The peak at 378.3 ◦ C indicates the melting of
the remaining η(Zn) dendrites.
167
Materials 2021, 14, 5404
Figure 3. Quaternary eutectic area in MZ + 0.5Sn alloy microstructure after DSC measurement: (a) example of quaternary
eutectic; (b) closer detail of such area including EDS measurement point 5 listed in Table 2.
The alloy with 1 wt.% of Sn (Figure 1c) has a minor peak left corresponding to
the ternary eutectic (339.1 ◦ C), while the peak corresponding to the quaternary eutectic
(335.3 ◦ C) increased further in peak area. Both other peaks correspond to the same reactions
as described above.
With 2 wt.% of Sn (Figure 1d), the melting starts at 333.5 ◦ C. This reaction melts the
complex eutectic area shown in Figure 4. The peak corresponding to the ternary eutectic
reaction is not resolved separately anymore. The peak at 349.6 ◦ C in this alloy represents
the melting of two binary eutectics: η(Zn) + Mg2 Zn11 as well as η(Zn) + Mg2 Sn. The last
peak in this DSC curve at 373.4 ◦ C corresponds again to η(Zn) dendrites.
Figure 4. Binary and quaternary eutectic areas in MZ + 2.0Sn alloy microstructure after DSC measurement.
Adding 3 wt.% of Sn (Figure 1e) changes the peak of the quaternary eutectic reaction
only slightly (peak maximum at 336.8 ◦ C). The peak observed at 346.3 ◦ C corresponds
according to microstructure observations solely to the melting of the binary η(Zn) + Mg2 Sn
eutectic. Melting of the remaining (η)Zn is indicated by the peak at 372.6 ◦ C.
XRD measurements were performed on the metallic filings prepared from the bulk
samples. Figures 5 and 6 shows the XRD patterns for selected alloys in the as-cast and
annealed state, respectively. A quantitative analysis using the Rietveld method was per-
formed considering the phases listed in Table 3 characterized in the ICSD FIZ Karlsruhe
database. These phases enabled the identification of all significant peaks in the measured
XRD patterns.
168
Materials 2021, 14, 5404
Figure 5. XRD diffraction patterns recorded on powder samples of selected alloys in as-cast state
(a) MZ + 0.0Sn, (b) MZ + 0.5Sn, (c) MZ + 3.0Sn.
Figure 6. XRD diffraction patterns recorded on powder samples of selected alloys in annealed state
(a) MZ + 0.0Sn, (b) MZ + 0.5Sn, (c) MZ + 3.0Sn.
169
Materials 2021, 14, 5404
As the cooling speed in all experiments was rather high at 60–75 ◦ C/s for both as-cast
and annealed samples, it was assumed that the solubility changes below the eutectoid
transformation [γ(Al) → αAl + η(Zn) at 275 ◦ C] will be significantly limited. Such a
phenomenon was reported for α(Al) as well as η(Zn) phases by Gogola et al. [42] based on
XRD measurements of Zn-Al based samples. To enable the correct quantitative analysis of
α(Al) as well as η(Zn) phases, their chemical composition had to be changed by adding
14 at.% of Zn and 2 at.% of Al, respectively, as suggested in this publication. The chemical
composition of the phases was changed in MAUD software before the quantitative analysis
of each XRD pattern. The soundness of this approach was double checked comparing the
GDOES chemical composition data with the chemical composition calculated from XRD
quantitative analysis for each sample.
The volume content of individual phases in the as-cast samples evolved as shown in
Figure 7.
Figure 7. Phase composition in vol.% in metallic bulk samples as measured on metallic filings from
as-cast samples.
In all alloys, both Mg2 Zn11 as well as the non-equilibrium MgZn2 , the intermetallic
phases can be detected. The overall content of Mgx Zny intermetallic phases was reduced
from 23.5 vol.% to about 1.5 vol.% by adding 3 wt.% of Sn into the basic Zn 1.6 wt.% Al
1.6 wt.% Mg (MZ) alloy. At the same time, the Mg2 Sn phase occupied about 8 vol.% of the
as-cast MZ + 3.0Sn alloy.
The content of MgZn2 was reduced from 3.5 vol.% to ~1.5 vol.% by adding 0.5 wt.%
of Sn. The further addition of Sn did not change the content of this phase significantly. Its
content was gradually further reduced to ~1 vol.% by adding up to 3 wt.% of Sn into the
alloy. However, at ~1 vol.% of MgZn2 , the detectability limit of MgZn2 was likely reached
in the current alloy with the applied measurement setup.
Mg2 Zn11 phase was detected in all alloys. Its content was gradually reduced from
~20 vol.% down to below 1 vol.% by adding up to 3 wt.% of Sn.
170
Materials 2021, 14, 5404
Peaks corresponding to Mg2 Sn can be already clearly identified in the as-cast MZ + 0.5Sn
sample representing as low as 1.5 vol.% of this phase. Its volume content clearly gradually
increased up to ~8 vol.% when 3 wt.% of Sn was added.
The content of α(Al) is calculated to be 2.5 vol.% on average across all alloys investi-
gated. Addition of Sn did not change the content of α(Al) in a significant way.
Annealing the investigated alloys clearly influenced their phase composition (Figure 8)
as calculated from XRD measurements (Figure 6).
Figure 8. Phase fractions in vol.% in metallic bulk samples as measured on metallic filings of
annealed samples.
171
Materials 2021, 14, 5404
Figure 9. Microstructure images for selected samples, longitudinal section, near sample surface: (a) MZ + 0.0Sn as-cast,
(b) MZ + 0.0Sn annealed, (c) MZ + 0.5Sn as-cast, (d) MZ + 0.5Sn annealed, (e) MZ + 3.0Sn as-cast, (f) MZ + 3.0Sn annealed.
Figure 10 summarizes the vol.% of all other microstructure components apart from η(Zn).
For as-cast samples, this represents the interdendritic spaces which are formed mainly
by various eutectics including a certain portion of η(Zn) phase solidified within them as
well as α(Al) + η(Zn) eutectoid particles.
For the annealed samples, it was possible to clearly distinguish between η(Zn) matrix
and all intermetallic phase particles along with α(Al) + η(Zn) eutectoid particles.
The difference between data for as-cast and annealed samples is mainly caused by
the fact that η(Zn) solidified in the interdendritic spaces of the as-cast samples cannot be
separately identified, while during annealing, these small η(Zn) particles are allowed to
connect to the larger primary η(Zn) areas forming a uniform matrix.
172
Materials 2021, 14, 5404
Figure 10. Vol.% of microstructural components as determined from SEM image analysis.
4. Discussion
Sn was chosen to supplement the composition of ZnAlMg based alloys. Similar Zn-
based alloy compositions have not been reported in the literature so far, probably due to
the concerns related to the corrosion properties of such alloys. These properties will be
investigated in adjacent research.
Different aspects of the alloy’s microstructure were investigated. The main phases
identified were in the general agreement with the published data as follows: η(Zn), α(Al),
Mg2 Zn11 , MgZn2 [7–18,20–24] and Mg2 Sn [26,27,32,33].
DSC curves can be clearly described only by investigating the microstructure of DSC
samples after cooling in the DSC equipment (Figures 2–4). Copper mold as-cast microstruc-
ture is an order of magnitude finer and hence less likely to be clearly identified. A clear
comparison can be for example given by comparing the size Mg2 Sn/η(Zn) eutectic particles
in Figure 4 (DSC sample of MZ + 3.0Sn) and Figure 9e (as-cast sample of MZ + 3.0Sn). Fur-
thermore, the kinetic of solidification may also affect the order in which the phase or phase
mixtures are formed. DSC curves show that the ternary η(Zn)/α(Al)/Mgx Zny eutectics
of MZ + 0.0Sn alloy are replaced by quaternary η(Zn)/α(Al)/Mgx Zny /Mg2 Sn eutectics
by adding 1 wt.% of Sn. By gradually adding Sn from 0.0 to 1.0 wt.%, the peak of the
quaternary eutectic areas is formed at ~335 ◦ C, while the peak of the ternary eutectic areas,
found at temperatures in the range from 347.5 to 339.1 ◦ C, is being gradually suppressed.
Further addition of Sn (2.0 and 3.0 wt.%) causes the ternary reaction peak to shift towards
even lower temperatures and being completely overlapped by the quaternary reaction
peak. This is in line with available assessment of liquidus projection for the Zn-Mg-Sn
ternary system [32,45]. These systems also predict a decrease in liquidus temperature for
less complex eutectics when Sn concentration is approaching a more complex eutectic point
near the Zn-rich corner of this system.
Peaks at 357–346.3 ◦ C represent the binary eutectics. While DSC curves suggest
only a gradual peak shift of binary eutectic reaction, the microstructure investigation
showed that Mg2 Zn11 /η(Zn) binary eutectics is being replaced by Mg2 Sn/η(Zn) in case
of the MZ + 3.0Sn alloy. Based on available ternary Zn-Mg-Sn assessments [32,45], it is
173
Materials 2021, 14, 5404
hypothesized, that Sn supports the preferential formation of Mg2 Sn/η(Zn) binary eutectic
instead of Mg2 Zn11 /η(Zn) eutectic mixture. This is observed in the currently investigated
system as well, despite the presence of Al as an additional alloying element. It is also
worth mentioning that the temperature difference between the binary and ternary eutectic
points calculated [32,45] is only 1 ◦ C. This may cause difficulties to reveal the real order
of solidification reactions since even a slight local chemical difference or temperature
heterogeneity may cause local fluctuation and a competitive formation of Mg2 Sn/η(Zn)
and Mg2 Zn11 /η(Zn) binary eutectic areas as indicated for the MZ + 2.0Sn alloy.
Heating curves shown in Figure 1 depicted also the peaks corresponding to the
α(Al) + η(Zn) → γ(Al) eutectoid transformation at ~285 ◦ C. The corresponding reaction
cannot be observed during a cooling DSC run since this eutectoid transformation is rather
sluggish; therefore, it is without a detectable heat release. The microstructure investigation
shows that the γ(Al) → α(Al) + η(Zn) reaction clearly occurs; however, probably over a
much broader temperature range compared to the heating curves. This reaction might be
finished even at an ambient temperature [46].
Adding Sn reduced the volume content of Mgx Zny intermetallic phases, and these
were replaced by Mg2 Sn particles. This behavior is in line with the literature findings on
similar systems [27,32]. The addition of Sn mainly affects the volume content of Mg2 Zn11
(Figure 7).
For similar Zn-Al-Mg alloys, the sources report the same two Mgx Zny phases to
be present [21,22,25,47]. Vlot et al. [47] identified only MgZn2 in similar alloys, while
other literature sources confirmed the presence of both phases mentioned [21,22,25]. In
our samples, Mg2 Zn11 is the primary phase; however, MgZn2 was also clearly identified
by both XRD and even SEM/EDX. At 3.5 vol.%, its content was highest in the as-cast
MZ + 0.0Sn sample. As MgZn2 is a non-equilibrium phase in the current system, its
content is significantly reduced by annealing the basic MZ + 0.0Sn alloy at 310 ◦ C for 1 h.
Mg2 Zn11 and MgZn2 are competing phases and their final ratio is complex to predict
and control even in a simple Mg-Zn alloy as reported by several sources [23,24]. All in
all, their presence will depend on several factors like exact alloy composition or cooling
rate [48].
The amount and distribution of intermetallic particles appears to have a direct influ-
ence on the microhardness of the studied alloys. Overall volume content of intermetallic
phases is decreasing with the increasing wt.% of Sn as measured by XRD (Figures 7 and 8)
as well as the SEM image analysis (Figure 10). This is reflected in the decreasing alloy
hardness summarized in Figure 11.
Formation of Mg2 Sn particles and the gradual increase of their vol.% is reported to
cause an increase in hardness for specific Mg-based alloys with similar Sn content [29–31].
174
Materials 2021, 14, 5404
The same mechanism does not apply to our Zn-based alloys, as in our samples, the overall
vol.% of intermetallics is decreasing.
The hardness of as-cast samples is marginally higher compared to annealed samples
(Figure 11). This is most probably caused mainly by the change in shape and distribution
of the intermetallic particles. This can be observed when comparing the images of as-cast
vs annealed conditions in Figure 9. The annealing allows η(Zn) to diffuse from eutectics in
the interdendritic areas towards the primary η(Zn) dendrites, hence changing the eutectic
nature of the interdendritic areas. For the MZ + 0.0Sn and MZ + 0.5Sn alloys, the original
dendritic character of the microstructure can still be recognized even after annealing
Figure 9a vs. Figure 9b,c vs. Figure 9d. For higher Sn content, this is not possible. The
addition of 1 to 3 wt.% of Sn into this alloy system, enabled a more effective spheroidization
and agglomeration of individual phase particles. Hence, the annealing had a more apparent
influence on the microstructure of these alloys (MZ + 1.0Sn, MZ + 2.0Sn, MZ + 3.0Sn).
For as-cast and annealed samples a different ratio of microstructural components was
established for the same alloys by SEM. Nevertheless, both as-cast and annealed samples
show the same trend compared to the XRD quantitative analysis. Additionally, for the
annealed samples, the SEM image analysis and XRD analysis are in even better agreement.
5. Conclusions
Melting of MZ + 0.0Sn starts at 344 ◦ C, while with the addition of 0.5–3.0 wt.% of Sn,
melting starts already at 334 ◦ C. Melting is finished at 382 ◦ C for the MZ + 0.0Sn and this
temperature is being continuously decreased to 376 ◦ C by the addition of up to 3 wt.%
of Sn.
Main phases identified in the MZ + 0.0Sn alloy were η(Zn) and α(Al) solid solutions
as well as Mg2 Zn11 and MgZn2 intermetallic phases. Addition of Sn enabled the formation
of Mg2 Sn intermetallic phase at the expense of Mgx Zny phases, while mainly affecting the
vol.% of Mg2 Zn11 .
The microstructure is dendritic for all as-cast alloys. The interdendritic areas are
formed by the binary, ternary and quaternary eutectics specific for each alloy. Alloying
with Sn causes the following changes of microstructural components: ternary eutectics
consisting of η(Zn), α(Al) and Mgx Zny phases are gradually replaced by quaternary η(Zn),
α(Al), Mgx Zny and Mg2 Sn eutectics. Binary η(Zn) + Mgx Zny eutectics are gradually
replaced by binary η(Zn) + Mg2 Sn eutectics.
For the MZ + 0.0Sn and MZ + 0.5Sn alloys, the original dendritic character of the
microstructure can still be recognized even after annealing. At the same time, the individual
phases from the eutectics are connected to discrete particles, and thus the original eutectics
are not recognizable anymore. Introducing 1 to 3 wt.% of Sn into this alloy system enabled a
more effective spheroidization and agglomeration of individual phase particles significantly
changing even the shape of the primary η(Zn) dendrites.
Annealing causes slight changes in the phase composition. For MZ + 0.0Sn mainly
MgZn2 is transformed to Mg2 Zn11 . For the alloys with Sn, the volume content of Mg2 Zn11
is partially increased mainly at the expense of Mg2 Sn.
The microhardness is decreasing with the increasing of Sn content. The annealing
changes the microhardness only slightly.
Based on microstructure observation, these alloys are overall suitable for coatings
exposed to extended high temperature exposure. As coatings of steel substrates, their
corrosion properties will be at least maintained as reported in part two of this research:
The effect of Sn addition on Zn-Al-Mg alloy-Part II.
Author Contributions: Conceptualization, P.G. and M.K.; methodology, P.G., Z.G. and M.K.; valida-
tion, P.G., Z.G., M.K. and H.S.; formal analysis, P.G., Z.G.; investigation, P.G., Z.G., M.K. and H.S.;
resources, P.G., Z.G. and H.S.; data curation, P.G., Z.G. and M.K.; writing—original draft preparation,
P.G. and Z.G.; writing—review and editing, P.G. and M.K.; visualization, P.G. and Z.G.; supervision,
P.G. and Z.G.; project administration, P.G.; funding acquisition, M.K. All authors have read and
agreed to the published version of the manuscript.
175
Materials 2021, 14, 5404
Funding: This research was supported by the Grant Agency VEGA of the Slovak Ministry of
Education, Research, Science and Sport, Project No. 1/0490/18: “The effect of microstructure
and phase composition on corrosion resistance of hot dip alloys” and by the Slovak Research and
Development Agency under the Contract no. APVV-20-0124.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data sharing is not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Coffin, C.; Depamelaere, H.; King, D.; Van Raemdonck, W. Evaluation of High Temperature Behavior of Zn and ZnAl Coatings on
Core Wires and Strands for ACSR, ACSS and Alike Overhead Power Conductors, WJI 2010, Monterrey ITC Preview. pp. 68–75.
Available online: https://issuu.com/wirejournal/docs/de-aug10-reduced (accessed on 24 February 2021).
2. Kiessling, F.; Nefzger, P.; Nolasco, J.F.; Kaintzyk, U. Overhead Power Lines—Planning, Design, Construction, 1st ed.; Springer:
Berlin/Heidelberg, Germany, 2003; p. 250. [CrossRef]
3. Mingyuan, G.U.; Notis, M.R.; Marder, A.R. The Effect of Continuous Heating on the Phase Transformations in Zinc-Iron
Electrodeposited Coatings. Metall. Trans. A 1991, 22, 1737–1743. [CrossRef]
4. Onishi, M.; Wakamatsu, Y.; Miura, H. Formation and Growth Kinetics of Intermediate Phases in Fe-Zn Diffusion Couples. Trans.
JIM 1974, 15, 331–337. [CrossRef]
5. Rico, Y.; Carrasquero, E.J. Microstructural Evaluation of Double-Dip Galvanized Coatings on Carbon Steel. MRS Adv. 2017, 2,
3917–3923. [CrossRef]
6. Wright, R.N. Wire coatings. In Wire Technology—Process Engineering and Metallurgy, 2nd ed.; Elsevier—Butterworth Heinemann
Books: Oxforfd, UK, 2011; pp. 245–256, ISBN 9780128026786.
7. Tanaka, S.; Honda, K.; Takahashi, A.; Morimoto, Y.; Kurosaki, M.; Shindo, H.; Nishimura, K.; Sugiyama, M. The performance of a
Zn-Al-Mg-Si hot-dip galvanized steel sheet. In Proceedings of the 5th International Conference on Zinc and Zinc Alloy Coated
Steel Sheet (GALVATECH 2001), Brussels, Belgium, 26–28 June 2001; Lamberights, M., Ed.; Verl. Stahleisen: Düsseldorf, Germany,
2001; p. 153, ISBN 3-514-00673-3.
8. Morimoto, Y.; Honda, K.; Nishimura, K.; Tanaka, S.; Takahashi, A.; Shindo, H.; Kurosaki, M. Excellent Corrosion-resistant
Zn-Al-Mg-Si Alloy Hot-dip Galvanized Steel Sheet “SUPER DYMA”. Nippon Steel Tech. Rep. 2003, 87, 24–26.
9. Nishimura, K.; Kato, K.; Shindo, H. Highly Corrosion-resistant Zn-Mg Alloy Galvanized Steel Sheet for Building Construction
Materials. Nippon. Steel Tech. Rep. 2000, 81, 85–88. [CrossRef]
10. Shindo, H.; Nishimura, K.; Okado, T.; Nishimura, N.; Asai, K. Developments and Properties of Zn-Mg Galvanized Steel Sheet
“DYMAZING” Having Excellent Corrosion Resistance. Nippon. Steel Tech. Rep. 1999, 79, 63–67.
11. Nishimura, K.; Shindo, H.; Kato, K.; Morimoto, Y.; Funaki, S.O. Microstructure and corrosion behaviour of Zn–Mg–Al hot-
dip galvanized steel sheet. In Proceedings of the 4th International Conference on Zinc and Zinc Alloy Coated Steel Sheet
(GALVATECH ’98), Chiba, Japan, 20–23 September 1998; Masuko, N., Ed.; ISIJ: Tokyo, Japan, 1998; pp. 437–442.
12. Shindo, H.; Nishimura, K.; Kato, K. Anti-Corrosion in Atmospheric Exposure of Zn–Mg–Al Hot-Dip Galvanized Steel Sheet. In
Proceedings of the 4th International Conference on Zinc and Zinc Alloy Coated Steel Sheet (GALVATECH ’98), Chiba, Japan,
20–23 September 1998; Masuko, N., Ed.; ISIJ: Tokyo, Japan, 1998; pp. 433–4363.
13. Tsujimura, T.; Komatsu, A.; Andoh, A. Influence of Mg content in coating layer and coating structure on corrosion resistance of
hot-dip Zn–Al–Mg–Si alloy coated steel shee. In Proceedings of the 5th International Conference on Zinc and Zinc Alloy Coated
Steel Sheet (GALVATECH 2001), Brussels, Belgium, 26–28 June 2001; Lamberights, M., Ed.; Verl. Stahleisen: Düsseldorf, Germany,
2001; pp. 145–152, ISBN 3-514-00673-3.
14. Kittaka, T.; Andoh, A.; Komatsu, A.; Tsujimura, T.; Yamaki, N.; Watanabe, K. Hot-Dip Zn-Al-Mg Coated Steel Sheet Excellent in
Corrosion Resistance and Surface Appearance and Process for the Production Thereof. U.S. Patent 6,235,410, 22 May 2001.
15. Tokuda, S.; Muto, I.; Sugawara, Y.; Takahashi, M.; Matsumoto, M.; Hara, N. Micro-electrochemical investigation on the role of Mg
in sacrificial corrosion protection of 55mass%Al-Zn-Mg coated steel. Corros. Sci. 2017, 129, 126–135. [CrossRef]
16. Vida, T.A.; Brito, C.; Lima, T.S.; Spinelli, J.E.; Cheung, N. Near-eutectic Zn-Mg alloys: Interrelations of solidification thermal
parameters, microstructure length scale and tensile/corrosion properties. Curr. Appl. Phys. 2009, 51, 2355–2363. [CrossRef]
17. Krystýnová, M.; Doležal, P.; Fintová, S.; Zapletal, J.; Marada, T.; Wasserbauer, J. Characterization of Brittle Phase in Magnesium
Based Materials Prepared by Powder Metallurgy. Key Eng. Mater. 2018, 784, 61–66. [CrossRef]
18. Vida, T.A.; Soares, T.; Septimio, R.S.; Brito, C.C.; Cheung, N.; Garcia, A. Effects of Macrosegregation and Microstructure on the
Corrosion Resistance and Hardness of a Directionally Solidified Zn-5.0wt.%Mg Alloy. Mater. Res. 2019, 22, 1–13. [CrossRef]
19. Pinc, J.; Čapek, J.; Kubásek, J.; Veřtát, P.; Hosová, K. Microstructure and mechanical properties of the potentially biodegradable
ternary system Zn-Mg0.8-Ca0.2. Procedia Struct. Integr. 2019, 23, 21–26. [CrossRef]
20. De Bruycker, E.; Zermout, Z.; De Cooman, B.C. Zn-Al-Mg Coatings—Thermodynamic Analysis and Microstructure Related
Properties. Mater. Sci. Forum 2007, 539–543, 1276–1281. [CrossRef]
176
Materials 2021, 14, 5404
21. De Bruycker, E.; De Cooman, B.C.; De Meyer, M. Experimental study and microstructure simulation of Zn-Al-Mg coatings. Rev.
Metall—CIT 2005, 102, 543–550. [CrossRef]
22. Akdeniz, V.M.; Wood, J.V. Microstructures and phase selection in rapidly solidified Zn-Mg alloys. J. Mater. Sci. 1996, 31, 545–550.
[CrossRef]
23. Liu, H.Y.; Jones, H. Solidification Microstructure Selection and Characteristics in the Zinc-Based Zn-Mg System. Acta Metall.
Mater. 1992, 40, 229–239. [CrossRef]
24. Prosek, T.; Nazarov, A.; Goodwin, F.; Šerák, J.; Thierry, D. Improving corrosion stability of Zn-Al-Mg by alloying for protection of
car bodies. Surf. Coat. Technol. 2016, 306, 439–447. [CrossRef]
25. Farahany, S.; Tat, L.H.; Hamzah, E.; Bakhsheshi-Rad, H.R.; Cho, M.H. Microstructure development, phase reaction characteristics
and properties of quaternary Zn-0.5Al-0.5Mg-xBi hot dipped coating alloy under slow and fast cooling rates. Surf. Coat. Technol.
2017, 315, 112–122. [CrossRef]
26. Gondek, J.; Babinec, M.; Kusý, M. The corrosion performance of Zn-Al-Mg based alloys with tin addition in neutral salt spray
environment. J. Achiev. Mater. Manuf. Eng. 2015, 70, 70–77.
27. Chen, J.; Chen, Z.; Yan, H.; Zhang, F.; Liao, K. Effects of Sn addition on microstructure and mechanical properties of Mg–Zn–Al
alloys. J. Alloys Compd. 2008, 461, 209–215. [CrossRef]
28. Chen, L.; Yan, A.; Liu, H.; Li, X. Strength and fatigue fracture behaviour of Al-Zn-Mg-Cu-Zr(-Sn) alloys. Trans. Nonferrous Met.
Soc. China 2013, 223, 2817–2825. [CrossRef]
29. Kim, B.; Do, J.S.; Lee, H.; Park, I. In situ fracture observation and fracture toughness analysis of squeeze cast AZ51–xSn magnesium
alloys. Mater. Sci. Eng. A 2010, 527, 6745–6757. [CrossRef]
30. Turen, Y. Effect of Sn addition on microstructure, mechanical and casting properties of AZ91 alloy. Mater. Des. 2013, 49, 1009–1015.
[CrossRef]
31. Wang, X.-Y.; Wang, Y.-F.; Wang, C.; Xu, S.; Rong, J.; Yang, Z.-Z.; Wang, J.-G.; Wang, H.-Y. A simultaneous improvement of both
strength and ductility by Sn addition in as-extruded Mg-6Al-4Zn alloy. J. Mater. Sci. Technol. 2020, 49, 117–125. [CrossRef]
32. Ghosh, P.; Mezbahul-Islam, M.; Medraj, M. Critical assessment and thermodynamic modeling of Mg-Zn, Mg-Sn, Sn-Zn and
Mg-Sn-Zn systems. Calphad 2012, 36, 28–43. [CrossRef]
33. Guangyin, Y.; Yangshan, S.; Wenjiang, D. Effects of Sn addition on the microstructure and mechanical properties of AZ91
magnesium alloy. Scripta Mater. 2001, 308, 34–38. [CrossRef]
34. Mezbahul-Islam, M.; Mostafa, A.O.; Medraj, M. Essential Magnesium Alloys Binary Phase Diagrams and Their Thermochemical
Data. J. Mater. Hindawi 2014, 2014, 704283. [CrossRef]
35. Agarwal, G. (RWTH Aachen, Germany). Personal communication, 2014.
36. Kevorkijan, V.; Škapin, S.D. Preparation and Study of Mg2Sn-based Composites with Different Compositions. Mater. Tehnol. 2010,
44, 251–259.
37. Durmus, Y.E.; Montiel Guerrero, S.S.; Tempel, H.; Hausen, F.; Kungl, H.; Eichel, R.-A. Influence of Al alloying on the electro-
chemical behavior of Zn electrodes for Zn–Air batteries with neutral sodium chloride electrolyte. Front. Chem. 2019, 7, 800.
[CrossRef]
38. ISO. Metallic Materials—Vickers Hardness Test; 6507-1:2018; International Organization for Standardization: Geneva, Switzer-
land, 2018.
39. Lutterotti, L.; Matthies, S.; Wenk, H.R. MAUD (Material Analysis Using Diffraction): A user friendly Java program for rietveld
texture analysis and more. In Proceedings of the 12th International Conference on Textures of Materials (ICOTOM-12), Mon-
treal, QC, Canada, 9–13 August 1999; Volume 1, p. 1599. Available online: http://hdl.handle.net/11572/57067 (accessed on
25 March 2021).
40. Schindelin, J.; Arganda-Carreras, I.; Frise, E.; Kaynig, V.; Longair, M.; Pietzsch, T.; Preibisch, S.; Rueden, C.; Saalfeld, S.; Schmid,
B.; et al. Fiji: An open-source platform for biological-image analysis. Nat. Methods 2012, 9, 676–682. [CrossRef]
41. De Bruycker, E. Zn-Al-Mg coatings: Thermodynamic Analysis and Microstructure-Related Properties. Ph.D. Thesis, Ghent
University, Ghent, Belgium, 2006.
42. Gogola, P.; Gabalcová, Z.; Suchánek, H.; Babinec, M.; Bonek, M.; Kusý, M. Quantitative x-ray diffraction analysis of Zn-Al based
alloys. Arch. Metall. Mater. 2020, 65, 959–966. [CrossRef]
43. Prosek, T.; Persson, D.; Stoulil, J.; Thierry, D. Composition of corrosion products formed on Zn–Mg, Zn–Al and Zn–Al–Mg
coatings in model atmospheric conditions. Corros. Sci. 2014, 86, 231–238. [CrossRef]
44. Raghavan, V. Al-Mg-Zn (Aluminum-Magnesium-Zinc). JPED 2007, 28, 203–208. [CrossRef]
45. Meng, F.G.; Wang, J.; Liu, L.B.; Jin, Z.P. Thermodynamic modeling of the Mg–Sn–Zn ternary system. J. Alloys Compd. 2010, 508,
570–581. [CrossRef]
46. Larsson, L.E. Pre-precipitation and precipitation phenomena in the Al-Zn system. Acta Metall. 1967, 15, 35–45. [CrossRef]
47. Vlot, M.; Zuijderwijk, M.; Toose, M.; Elliot, L.; Bleeker, R.; Maalman, T. Hot dip ZnAlMg coatings: Microstructure and forming
properties. In Proceedings of the 7th International Conference on Zinc and Zinc Alloy Coated Steel Sheet (Galvatech ’07), Osaka,
Japan, 19–21 November 2007; Tsuru, T., Ed.; ISIJ: Tokyo, Japan, 2007; pp. 574–579.
48. Kim, J.N.; Lee, C.S.; Jin, Y.S. Structure and Stoichiometry of MgxZny in Hot-Dipped Zn–Mg–Al Coating Layer on Interstitial-Free
Steel. Met. Mater. Int. 2018, 24, 1090–1098. [CrossRef]
177
materials
Article
The Effect of Sn Addition on Zn-Al-Mg Alloy; Part II:
Corrosion Behaviour
Zuzana Gabalcová *, Peter Gogola, Martin Kusý and Henrich Suchánek
Faculty of Materials Science and Technology in Trnava, Institute of Materials Science, Slovak University of
Technology in Bratislava, Ulica Jána Bottu 25, 917 24 Trnava, Slovakia; [email protected] (P.G.);
[email protected] (M.K.); [email protected] (H.S.)
* Correspondence: [email protected]
Abstract: Corrosion behaviour of Sn (0.0, 0.5, 1.0, 2.0 and 3.0 wt.%)-doped Zn 1.6 wt.% Al 1.6 wt.%
Mg alloys exposed to salt spray testing was investigated. Intergranular corrosion was observed for
all alloys in both as-cast and annealed states. However, due to microstructure spheroidisation in
the annealed samples, potential intergranular corrosion paths are significantly reduced. Samples
with 0.5 wt.% of Sn showed the best corrosion properties. The main corrosion products identified by
XRD analysis for all samples were simonkolleite and hydrozincite. Occasionally, ZnO and AlO were
identified in limited amounts.
Keywords: Zn-based alloy; Sn-addition; corrosion products; salt spray test; intergranular corrosion;
corrosion penetration depth; weight loss
1. Introduction
Citation: Gabalcová, Z.; Gogola, P.; A wide range of commercial Zn-based hot-dip coatings are used for corrosion protec-
Kusý, M.; Suchánek, H. The Effect of tion. These also include Zn-Al-Mg-based coatings such as Magizinc (MZ) with Zn 1.6 wt.% Al
Sn Addition on Zn-Al-Mg Alloy; and 1.6 wt.% Mg. It is widely used in the coating industry including steel sheet production
Part II: Corrosion Behaviour. Materials for building, energetics, and the automotive industry [1–12].
2021, 14, 5290. https://doi.org/ Neutral salt spray testing (NSST) is used as an industry standard for corrosion resis-
10.3390/ma14185290
tance testing. Zn-Al-Mg coatings perform notably better compared to conventional hot-dip
Zn coatings. The presence of Mg in the Zn-Al-Mg coatings enables the stabilisation of
Academic Editor: Daoguang He
protective corrosion products like simonkolleite and hydrozincite [13–16]. Regarding the
microstructure, Mg addition to binary Zn-Al alloys results in the formation of intermetallic
Received: 2 August 2021
phases such as Zn2 Mg and Zn11 Mg2 . These phases are more corrosion active even com-
Accepted: 11 September 2021
pared to the η(Zn) phase, hence enabling the more effective cathodic protection of steel
Published: 14 September 2021
substrates [17]. They are formed within eutectics in the interdendritic areas of primary
η(Zn) dendrites. Unfortunately, these phases are also enabling the cathodic protection
Publisher’s Note: MDPI stays neutral
with regard to jurisdictional claims in
of this Zn-based matrix, hence overall corrosion attack starts as the intergranular (IG)
published maps and institutional affil-
corrosion. Sources have reported this phenomenon, however only on the coatings with a
iations. limited thickness of up to 50 μm. In all these corrosion test results, substantial parts of the
coatings were affected by IG corrosion locally, even across the entire coating [18–20].
The potentials of additional alloying of Zn-Al-Mg systems by Cr, Zr, Ti Mo, Mn, Si,
etc. have been already studied in the literature [10]. Sn is also an interesting candidate
due to its high affinity to Mg [21]. The preliminary research [12] into the development
Copyright: © 2021 by the authors.
of microstructure and corrosion resistance of the Zn-Al-Mg + Sn alloy system has shown
Licensee MDPI, Basel, Switzerland.
This article is an open access article
that Sn can affect the phase composition, and consequently the corrosion properties of
distributed under the terms and
MZ. In the follow-up to these results, this system is being investigated with an extended
conditions of the Creative Commons
experimental scope in Parts I and II of the current articles. The main aim of these additional
Attribution (CC BY) license (https:// experiments is to observe if long time exposure to rather high temperatures (1 h at 310 ◦ C)
creativecommons.org/licenses/by/ have a significant influence on the corrosion properties of these alloys. Based on these
4.0/). inputs, bulk samples were chosen for our research. This enabled to investigate the IG
179
corrosion phenomena for these alloys in both as-cast and annealed states without the limit
of a coating’s thickness.
Alloy Al Mg Sn Zn
MZ + 0.0Sn 1.56 ± 0.07 1.40 ± 0.01 0.07 ± 0.02 bal.
MZ + 0.5Sn 1.64 ± 0.02 1.41 ± 0.01 0.52 ± 0.01 bal.
MZ + 1.0Sn 1.62 ± 0.03 1.45 ± 0.02 1.06 ± 0.02 bal.
MZ + 2.0Sn 1.57 ± 0.01 1.44 ± 0.01 1.95 ± 0.01 bal.
MZ + 3.0Sn 1.57 ± 0.12 1.43 ± 0.05 2.69 ± 0.06 bal.
As a reference material for the corrosion test, 4N5 purity Zn-samples were cast. Two
types of cylindrical samples were prepared for each alloy: (i) as-cast samples and (ii) cast
and subsequently solution annealed at 310 ◦ C for 1 h.
Casting was done from 470 ◦ C of melt temperature into a water-cooled copper mould
with a diameter of 30 mm and depth of 20 mm. During casting, the sample temperature
was continuously measured and an average cooling rate of 60 ◦ C/s was established. The
annealing step was finished by quenching it in a water bath below 10 ◦ C at an average
cooling rate of 75 ◦ C/s.
The investigated surface of the as-cast and annealed samples was subjected to grinding
using up to 4000 grit abrasive papers. The surface topography was determined using a
ZEISS LSM700 laser scanning confocal microscope (LSCM, Carl Zeiss AG, Oberkochen,
Germany). The 405 nm light source was used, which in combination with a Epiplan-
Apochromat 50×/0.95 objective enabled to reach step sizes of 250 nm on the X and Y axes
as well as 200 nm on the Z axis. These surfaces were subjected to the corrosion in the
salt chamber.
The investigated samples were coated with Lacomit Varnish to prevent the corrosion
of the entire sample and limit the exposed area. The exposed surface was digitally scanned
to double check the exposed area. These data, together with the surface topography data,
made it possible to calculate the real surface area exposed to the corrosion on each sample.
The neutral salt spray corrosion test (NSST) was performed in a Co.Fo.Me.Gra 400E
(CO.FO.ME.GRA. Srl, Milano, Italy) corrosion chamber according to the ISO 9227:2017
Standard [23]. The NSST samples were immediately exposed in the cabinet to a 5 wt.%
NaCl solution. The air pressure of the atomized saline solution was maintained in the
range of 95–105 kPa, and the temperature inside the cabinet was 35 ± 2 ◦ C, pH level was
6.6–7.1, and the salt solution deposition rate 125–200 mL/h/m2 . Custom holders were
used to keep the prescribed sample orientation of 15◦ from the vertical position.
Exposure times for all types of samples were 250, 500, 750 and 1000 h. Three samples
were prepared for all as-cast and annealed conditions for all exposure times. All in all,
144 individual samples were exposed at the same time. After the salt spray testing, the
samples were dried at room temperature for 24 h at minimum before being further pro-
cessed. After drying, loose corrosion products were removed and collected separately. It
was of upmost importance to prevent any kind of a mechanical damage to the metallic
sample surface. The bulk samples were cleaned by acetone and dried on air. The initial
180
Materials 2021, 14, 5290
weight of the specimen was measured (w0 ) by using the Mettler Toledo XPR205 weighing
balance (Mettler-Toledo International Inc., Columbus, OH, USA). According to the ASTM
G31 Standard [24], the specimens were immersed in the chromate acid (CrO3 ) to ensure
that the corrosion products were removed. Samples were cleaned in 60 s intervals. After
each cleaning interval, the samples were repeatedly weighed. This process was considered
finished when less than 5 mg of weight was lost after a cleaning cycle for all three repeats
of a condition [23–25]. The final weight for each sample was recorded (wn ). The recorded
weight difference was normalized by the exposed area of each sample (An ) corrected by
the sample topography coefficient (k). The topography coefficient is retrieved from LSCM
software as the ratio between real surface area, incorporating surface topography, and
the ideal surface. This value was 1.09 on average. These data enabled the calculation the
average weight change (w ) for each condition in mg/mm2 according to equation:
w0 − w n
w = (1)
An k
The metallographic preparation on the longitudinal section (along the cylinder axis)
of corroded as-cast and annealed samples consisted of standard grinding using abrasive
papers, polishing on diamond pastes with various grain sizes down to 0.25 μm.
The microstructure evaluation was performed by the JEOL JSM 7600F scanning elec-
tron microscopy (SEM, Jeol Ltd., Tokyo, Japan) with a Schottky field emission electron
source operating at 20kV and 90 μA. The samples were placed at a working distance of
15 mm and documented using a backscattered electron detector.
The quantitative analysis of IG corrosion depth was performed by ImageJ 1.53c
software [26] along the longitudinal section for each condition. At least 150 individual
values were recorded for each data point.
The weight measurements are displayed with +/− standard deviation error bars and
the depth of IG corrosion measurements are given with +/− standard error.
The X-ray diffraction (XRD) analysis was carried out by the PANalytical Empyrean
X-ray diffractometer (Malvern Panalytical Ltd., Malvern, UK) with configurations as de-
tailed in Table 2. The measurements were performed on the samples after 1000 h of NSST
with Ni filtered Cu-radiation. X-ray diffraction data were further analysed qualitatively
using the PANalytical Xpert High Score program (HighScore Plus 3.0.5 version) with ICSD
FIZ Karlsruhe database. These findings were confirmed and enhanced using the Rietveld
refinement-based program, MAUD version 2.84 [27]. The program uses an asymmetric
pseudo-Voight function to describe the experimental peaks. The instrument broadening
was determined by measuring the NIST660c LaB6 (The National Institute of Standards and
Technology, Gaithersburg, MD, USA) line position and line broadening standard and intro-
duced to the Rietveld refinement program (MAUD version 2.84) via the Caglioti equation.
An anisotropic size-strain model was applied to the majority of corrosion products, while
the other phases were treated by isotropic models. A minor discrepancy between nominal
and measured peak intensities was corrected using the spherical harmonic functions with
fibre symmetry. The quality of the fit was in all analysed patterns achieved below 10% Rwp .
181
Materials 2021, 14, 5290
3. Results
As mentioned before, the weight changes for each sample were calculated according
to Equation (1) and the obtained data are plotted in Figures 1 and 2. Reference Zn samples
showed a gradual weight loss for both as-cast and annealed conditions as expected. It can
be observed that for several as-cast samples, a weight gain rather than a weight loss was
recorded. The annealed samples showed the weight loss for all conditions as expected.
MZ + 3.0Sn showed the best results at even 40% lower values compared to MZ + 0.0Sn.
Since the weight gain instead of the weight loss was recorded for several as-cast
conditions, it was decided to prepare longitudinal cuts of the samples and investigate
potential reasons of this phenomena. The intergranular corrosion was present in most
samples to a significant extent. Most phases present in the interdendritic spaces were
corroded. Such corrosion products could not be cleaned by CrO3 acid solution [24]. These
corrosion products, anchored among the still mainly intact η(Zn) dendrites, were increasing
the total weight of the samples even after the cleaning process (Figure 3a). Their presence
is visualised by chemical element distribution maps in Figure 3b.
Backscattered-electron scanning electron microscopy (BSEM) images of the longitudi-
nal sections for representative as-cast samples with 0.0, 0.5 and 3.0 wt.% of Sn after 1000 h of
NSST are given in Figure 4. Corresponding quantitative analysis results of the intergranular
corrosion penetration depth are summarized in Figure 5. The same is available for the
annealed samples in Figures 6 and 7.
182
Materials 2021, 14, 5290
Figure 3. Anchoring effect of η(Zn) dendrites with corroded interdendritic spaces (MZ + 2.0Sn, as-cast): (a) overview BSEM
image; (b) chemical element distribution maps of Zn, Sn, O, Mg, Al and Cl.
Figure 4. BSEM images indicating the extend of IG corrosion observed for the as-cast samples after 1000 h of NSST:
(a) MZ + 0.0Sn (b) MZ + 0.5Sn (c) MZ + 3.0Sn.
183
Materials 2021, 14, 5290
Figure 6. BSEM images indicating the extent of IG corrosion observed for the annealed samples: (a) MZ + 0.0Sn
(b) MZ + 0.5Sn (c) MZ + 3.0Sn.
184
Materials 2021, 14, 5290
Figure 8. MZ + 2.0Sn as-cast microstructure after 1000 h in NSST affected by IG corrosion: (a) overview, (b) detail.
The depth of IG corrosion is significantly lower for the annealed samples with maxi-
mums reaching only about 70 μm even after 1000 h of NSST. For 250 and 500 h, all alloys
behaved rather similar with IG corrosion depths of 10 and 22 μm, respectively. MZ + 1.0Sn
and MZ + 2.0Sn seem to be more susceptible to the IG corrosion when comparing the
samples after the full 1000 h test. On the contrary, annealed MZ + 3.0Sn samples showed
values comparable even to MZ + 0.0Sn, or MZ + 0.5Sn reaching a maximum depth of about
45 μm.
The examples of areas affected by the intergranular corrosion are given in Figures 8 and 9
for the as-cast and annealed samples, respectively. The EDS chemical analysis of the
microstructure in Table 3 confirms the intergranular corrosion attack. Figure 8a shows
the η(Zn) dendritic microstructure affected by the corrosion along the interdendritic areas.
185
Materials 2021, 14, 5290
η(Zn) primary dendrites also showed the signs of corrosion in the form of fine cracks. These
can be attributed to the presence of fine, sub-micron Al-rich particles observed within
the η(Zn) primary dendrites. In a more detailed image (Figure 8b) it can be seen that
α(Al) and Mgx Zny particles were corroded. Mg2 Sn particles were subject to the process of
dealloying [28–31], leaving thus pure metallic Sn particles behind.
Figure 9. MZ + 2.0Sn annealed microstructure after 1000 h in NSST affected by IG corrosion: (a) overview, (b) detail.
As reported in the first part of this research [22], the basic dendritic character of
the microstructure was still rather well maintained for the MZ + 0.0Sn and MZ + 0.5Sn
alloys even after annealing. Hence, the IG corrosion is observed to propagate preferably
along the interdendritic areas. For the annealed samples with 1 and more wt.% of Sn the
microstructure is more spheroidized. The individual intermetallic phases were coalesced
into coarse, discrete particles, while η(Zn) dendrites were reshaped and new grains are
formed within the microstructure. The boundaries of these grains contained a significant
portion of intermetallic phase particles. As such, they were more susceptible to the IG
corrosion. The propagation of the IG corrosion is documented in Figure 9a and the grain
boundaries decorated by intermetallic particles are shown in closer detail in Figure 9b.
The XRD analysis was performed on all samples after NSST. As described, loose
corrosion products were gathered and investigated. The XRD was used to determine
the phase composition of the corrosion products formed on the samples during NSST.
The XRD patterns for all corrosion products retrieved from the as-cast and annealed
samples are summarized in Figures 10 and 11, respectively. The presence of the identified
phases was also confirmed using the Rietveld method (Table 4). Despite the differences
between the microstructure of the as-cast and annealed samples, their corrosion products
showed an identical qualitative phase composition. The semi-quantitative results from
these calculations indicate that the majority of the corrosion products were formed by a
hydrozincite for all samples. About 20 vol.% of simonkolleite was measured for all pure
186
Materials 2021, 14, 5290
Zn samples (Figures 10a and 11a). The corrosion products of MZ-based samples contained
only about 10 vol.% of simonkolleite on average (Figures 10b–d and 11b–d).
ZnO was identified solely in the corrosion products of the pure as-cast Zn sample
(Figure 10a), representing only about 2 vol.%.
NaCl was identified in randomly varying amounts in the corrosion products as a
remainder of the corrosion environment.
Next to hydrozincite and simonkolleite, the sources indicated that other phases might
also be formed [17–20,32–39]. Therefore, the corroded surfaces of the bulk metallic samples
were investigated after the loose corrosion products were removed. The measurement in
grazing incident diffraction mode with 0.5◦ incident angle was chosen to limit the signal
from the substrate (mainly Zn) as much as possible. Additionally, to previously identified
phases, zincite (ZnO) and aluminium (II) oxide (AlO) were identified as present directly
attached to the sample surface. An example of such a pattern is given in Figure 12 for the
MZ + 3.0Sn annealed sample surface after 1000 h of NSST. However, only about 2 and
5 vol.% of ZnO and AlO, respectively, were identified using the Rietveld method.
Reference Crystallography
Phase Chemical Space Group
Phase Name Code—ICSD FIZ Open Database Crystal System Space Group
Formula Number
Karlsruhe Database COD ID
Hydrozincite Zn5 (OH)6 (CO3 )2 01-072-1100 9007481 Monoclinic C2/m 12
Simonkolleite Zn5 (OH)8 Cl2 ·H2 O 98-003-4904 9004683 Hexagonal R3m 166
Zincite ZnO 98-015-4487 9004178 Hexagonal P63 /mc 186
Aluminium (II) Oxide AlO 98-002-8920 - Cubic Fm3m 225
Sodium Chloride NaCl 01-075-0306 1000041 Cubic Fm3m 225
Figure 10. XRD patterns of corrosion product powders retrieved from the as-cast samples after 1000 h
of NSST: (a) pure Zn (b) MZ + 0.0Sn (c) MZ + 0.5Sn (d) MZ + 3.0Sn.
187
Materials 2021, 14, 5290
Figure 11. XRD patterns of corrosion product powders retrieved from the annealed samples after
1000 h of NSST. (a) pure Zn (b) MZ + 0.0Sn (c) MZ + 0.5Sn (d) MZ + 3.0Sn.
Figure 12. XRD pattern on MZ + 3.0Sn annealed sample surface after 1000 h NSST, measured in
grazing incident geometry.
4. Discussion
4.1. SEM vs. Mass Balance after NSST
Both weight loss and weight gain were observed for a significant portion of the
as-cast samples due to several related properties of the ZnAlMg alloy system. During
directional cooling of these alloys, the η(Zn) phase forms the primary dendrites, while
the interdendritic spaces are formed by a fine mixture of various phases including α(Al)
solid solution, MgZn2 , Mg2 Zn11 , and Mg2 Sn intermetallic phases. There is an inherent
difference in the open circuit potential (OCP) of these phases mainly compared to the
η(Zn) phase (Figure 13). Consequently, the interdedritic phases seem to offer the galvanic
188
Materials 2021, 14, 5290
protection to the η(Zn) phase dendrites. Due to this phenomenon, the interdendritic spaces
corrode prior to the η(Zn) phase. The still intact η(Zn) phase dendrites act as anchors
holding these corrosion products in place. These corrosion products cannot be removed by
the environment during the NSST, nor by chemical cleaning done in the preparation for
the sample weighing after the test. Naturally, the total weight of such corrosion products
is greater as the weight of the original metallic phases. This phenomenon will cause
weight gain for several samples even after the corrosion products were removed as much
as possible before weighing. This increase in weight is also followed by an increase in
volume. Following the BSEM images, it is clear that the η(Zn) phase dendrites are cracking
as seen in Figure 8. This could be attributed to volume expansion-induced cracking
(Figures 3a and 8a).
Figure 13. Overview of OCP values for phases present in the investigated system [34,40–44].
189
Materials 2021, 14, 5290
Figure 14. Behaviour of Mg2 Sn intermetallic phase during corrosion (MZ + 2.0Sn, annealed): (a) Influence of Mg2 Sn on
initial stage of corrosion of Mg2 Zn11 intermetallic particle; (b) dealloying of Mg2 Sn particle.
Figure 15. Details of Mg2 Sn particles affected by dealloying (MZ + 3.0Sn, annealed): (a) overview BSEM image; (b) chemical
element distribution maps of Mg, O, Sn and Zn.
190
Materials 2021, 14, 5290
5. Conclusions
Based on the experimental results discussed in part I and part II of this research, the
following conclusions can be drawn:
• Weight change cannot be correlated with alloy composition nor NSST exposure time
due to presence of IG corrosion.
• Increasing the exposure time in NSST from 250 h to 1000 h increases the intergranular
corrosion penetration depth, regardless of the chemical composition and heat treatment.
• As-cast samples were more susceptible to the IG corrosion as interdendritic areas are
forming a connected network of less noble phases. These include α(Al), Mgx Zny and
Mg2 Sn, while dendrites are formed mainly by η(Zn).
• Adding 0.5 wt.% Sn has almost no effect on the weight change of the as-cast samples
after NSST compared to MZ + 0.0Sn, while being significantly less susceptible to the
IG corrosion. As a result, the as-cast MZ + 0.5Sn samples show the most favourable
corrosion behaviour.
• Adding 1 to 3 wt.% of Sn yields in the weight gain instead of the weight loss as well
as a significant increase in the IG corrosion depth.
• Annealed alloys are less susceptible to the IG corrosion as intermetallic phases are
coalesced, spheroidised, and more uniformly distributed within the η(Zn) matrix or at
newly formed grain boundaries of η(Zn).
• Changing the alloy composition of the annealed samples has only a slight effect on
the weight change. Nevertheless, the samples with 3 wt.% of Sn showed the most
favourable results. Meanwhile, the IG corrosion depth is comparable to MZ + 0.0Sn,
resulting in overall best performance of the annealed MZ + 3.0Sn samples.
• Hydrozincite and simonkolleite were identified as the main corrosion products on all
samples. A small portion of ZnO was identified only on pure Zn samples. GI XRD
measurements indicated a small amount of AlO formed on most MZ-based samples.
• Current results show that even the high temperature exposure of up to 310 ◦ C does
not negatively affect the corrosion performance of these alloys. It could be noted that
such exposure even provides a beneficial effect and enhances the corrosion resistance
of the coating by suppressing the IG corrosion.
Author Contributions: Conceptualization, P.G. and Z.G.; methodology, P.G., Z.G. and M.K.; valida-
tion, P.G., Z.G., M.K. and H.S.; formal analysis, P.G. and Z.G.; investigation, P.G., Z.G., M.K. and H.S.;
resources, Z.G. and P.G.; data curation, P.G., Z.G. and M.K.; writing—original draft preparation, P.G.
and Z.G.; writing—review and editing, P.G., Z.G. and M.K.; visualization, Z.G. and P.G.; supervision,
P.G. and Z.G.; project administration, M.K.; funding acquisition, M.K. All authors have read and
agreed to the published version of the manuscript.
Funding: This research was supported by the Grant Agency VEGA of the Slovak Ministry of
Education, Research, Science and Sport Project No. 1/0490/18: “The effect of microstructure and
phase com-position on corrosion resistance of hot dip alloys” and by the Slovak Research and
Development Agency under the Contract No. APVV-20-0124.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data sharing is not applicable.
191
Materials 2021, 14, 5290
References
1. Vida, T.A.; Brito, C.; Lima, T.S.; Spinelli, J.E.; Cheung, N. Near-eutectic Zn-Mg alloys: Interrelations of solidification thermal
parameters, microstructure length scale and tensile/corrosion properties. Curr. Appl Phys. 2009, 51, 2355–2363. [CrossRef]
2. Tokuda, S.; Muto, I.; Sugawara, Y.; Takahashi, M.; Matsumoto, M.; Hara, N. Micro-electrochemical investigation on the role of Mg
in sacrificial corrosion protection of 55mass%Al-Zn-Mg coated steel. Corros. Sci. 2017, 129, 126–135. [CrossRef]
3. Krystýnová, M.; Doležal, P.; Fintová, S.; Zapletal, J.; Marada, T.; Wasserbauer, J. Characterization of Brittle Phase in Magnesium
Based Materials Prepared by Powder Metallurgy. Key Eng. Mater 2018, 784, 61–66. [CrossRef]
4. Vida, T.A.; Soares, T.; Septimio, R.S.; Brito, C.C.; Cheung, N.; Garcia, A. Effects of Macrosegregation and Microstructure on the
Corrosion Resistance and Hardness of a Directionally Solidified Zn-5.0wt.%Mg Alloy. Mater. Res. 2019, 22, 1–13. [CrossRef]
5. Pinc, J.; Čapek, J.; Kubásek, J.; Veřtát, P.; Hosová, K. Microstructure and mechanical properties of the potentially biodegradable
ternary system Zn-Mg0.8-Ca0.2. Procedia Struct. Integr. 2019, 23, 21–26. [CrossRef]
6. De Bruycker, E.; Zermout, Z.; De Cooman, B.C. Zn-Al-Mg Coatings—Thermodynamic Analysis and Microstructure Related
Properties. Mater. Sci. Forum 2007, 539–543, 1276–1281. [CrossRef]
7. De Bruycker, E.; De Cooman, B.C.; De Meyer, M. Experimental study and microstructure simulation of Zn-Al-Mg coatings.
Rev. Metall-CIT 2005, 102, 543–550. [CrossRef]
8. Akdeniz, V.M.; Wood, J.V. Microstructures and phase selection in rapidly solidified Zn-Mg alloys. J. Mater. Sci. 1996, 31, 545–550.
[CrossRef]
9. Liu, H.Y.; Jones, H. Solidification Microstructure Selection and Characteristics in the Zinc-Based Zn-Mg System. Acta Metall.
Mater. 1992, 40, 229–239. [CrossRef]
10. Prosek, T.; Nazarov, A.; Goodwin, F.; Šerák, J.; Thierry, D. Improving corrosion stability of Zn-Al-Mg by alloying for protection of
car bodies. Surf. Coat. Technol. 2016, 306 Pt B, 439–447. [CrossRef]
11. Farahany, S.; Tat, L.H.; Hamzah, E.; Bakhsheshi-Rad, H.R.; Cho, M.H. Microstructure development, phase reaction characteristics
and properties of quaternary Zn-0.5Al-0.5Mg-xBi hot dipped coating alloy under slow and fast cooling rates. Surf. Coat. Technol.
2017, 315, 112–122. [CrossRef]
12. Gondek, J.; Babinec, M.; Kusý, M. The corrosion performance of Zn-Al-Mg based alloys with tin addition in neutral salt spray
environment. J. Achiev. Mater. Manuf. Eng. 2015, 70, 70–77.
13. Vargel, C. Corrosion of Aluminium, 1st ed.; Elsevier Science: Amsterdam, The Netherlands, 2004; pp. 167–182. [CrossRef]
14. Volovitch, P.; Allely, C.; Ogle, K. Understanding corrosion via corrosion product characterization: I Case study of the role of Mg
alloying in Zn–Mg coating on steel. Corros. Sci. 2009, 51, 1251–1262. [CrossRef]
15. Odnevall, W.; Leygraf, C. A Critical Review on Corrosion and Runoff from Zinc and Zinc-Based Alloys in Atmospheric
Environments. Corros. J. Sci. Eng. 2017, 73, 1060–1077. [CrossRef]
16. De la Fuente, D.; Castaño, J.G.; Morcillo, M. Long-term atmospheric corrosion of zinc. Corros. Sci. 2007, 51, 1420–1436. [CrossRef]
17. Volovitch, P.; Vu, T.N.; Allély, C.; Abdel Aal, A.; Ogle, K. Understanding corrosion via corrosion product characterization: II Role
of alloying elements in improving the corrosion resistance of Zn–Al–Mg coatings on steel. Corros. Sci. 2011, 53, 2437–2445. [CrossRef]
18. Thierry, D.; Persson, D.; Luckeneder, G.; Stellnberger, K.-H. Atmospheric corrosion of ZnAlMg coated steel during long term
atmospheric weathering at different worldwide exposure sites. Corros. Sci. 2019, 148, 338–354. [CrossRef]
19. LeBozec, N.; Thierry, D.; Persson, D.; Riener, C.K.; Luckeneder, G. Influence of microstructure of zinc-aluminium-magnesium
alloy coated steel on the corrosion behavior in outdoor marine atmosphere. Surf. Coat. Technol. 2019, 374, 897–909. [CrossRef]
20. Azevedo, M.S.; Allély, C.; Ogle, K.; Volovitch, P. Corrosion mechanisms of Zn(Mg, Al) coated steel in accelerated tests and natural
exposure 1. The role of electrolyte composition in the nature of corrosion products and relative corrosion rate. Corros. Sci. 2015,
90, 472–481. [CrossRef]
21. Ghosh, P.; Mezbahul-Islam, M.; Medraj, M. Critical assessment and thermodynamic modeling of Mg-Zn, Mg-Sn, Sn-Zn and
Mg-Sn-Zn systems. Calphad 2012, 36, 28–43. [CrossRef]
22. Gogola, P.; Gabalcová, Z.; Kusý, M.; Suchánek, H. The effect of Sn addition on Zn-Al-Mg alloy; Part I: Microstructure and phase
composition. Materials 2021, under review.
23. Corrosion Tests in Artificial Atmospheres-Salt Spray Tests; ISO 9227:2017; International Organization for Standardization: Geneva,
Switzerland, 2017.
24. Standard Guide for Laboratory Immersion Corrosion Testing of Metals; ASTM G31-21; ASTM International: West Conshohocken, PA,
USA, 2021.
25. Jokar, M.; Aliofkhazraei, M. Comprehensive Materials Finishing, 1st ed.; Elsevier Science: Amsterdam, The Netherlands, 2017;
pp. 306–335. [CrossRef]
26. Schneider, C.A.; Rasband, W.S.; Eliceiri, K.W. NIH Image to ImageJ: 25 Years of image analysis. Nat. Methods 2012, 9, 671–675.
[CrossRef]
27. Lutterotti, L.; Matthies, S.; Wenk, H.R. MAUD (Material Analysis Using Diffraction): A user friendly Java program for rietveld
texture analysis and more. In Proceedings of the 12th International Conference on Textures of Materials (ICOTOM-12), McGill
University Montreal, Montréal, QC, Canada, 9–13 August 1999; pp. 1599–1604. Available online: http://hdl.handle.net/11572/
57067 (accessed on 28 July 2021).
192
Materials 2021, 14, 5290
28. Nguyen, G.T.H.; Nguyen, D.-T.; Song, S.-W. Unveiling the Roles of Formation Process in Improving Cycling Performance of
Magnesium Stannide Composite Anode for Magnesium-Ion Batteries. Adv. Mater. Interfaces 2018, 5, 1801039. [CrossRef]
29. Nguyen, D.-T.; Song, S.-W. Magnesium stannide as a high-capacity anode for magnesium-ion batteries. J. Power Sources 2017, 368,
11–17. [CrossRef]
30. Singh, N.; Arthur, T.S.; Ling, C.; Matsui, M.; Mizuno, F. A high energy-density tin anode for rechargeable magnesium-ion batteries.
Chem. Commun. 2013, 49, 149–151. [CrossRef]
31. Yaghoobnejad Asl, H.; Fu, J.; Kumar, H.; Welborn, S.S.; Shenoy, V.B. In Situ Dealloying of Bulk Mg2 Sn in Mg-Ion Half Cell as an
Effective Route to Nanostructured Sn for High Performance Mg-Ion Battery Anodes. Chem. Mater. 2018, 30, 1815–1824. [CrossRef]
32. Prosek, T.; Persson, D.; Stoulil, J.; Thierry, D. Composition of corrosion products formed on Zn–Mg, Zn–Al and Zn–Al–Mg
coatings in model atmospheric conditions. Corros. Sci. 2014, 86, 231–238. [CrossRef]
33. Zhang, X.G. Corrosion and Electrochemistry of Zinc, 1st ed.; Springer: Boston, MA, USA, 1996; pp. 157–181. [CrossRef]
34. Buyn, J.M.; Yu, J.M.; Kim, D.K.; Kim, T.-Y.; Jung, W.-S.; Kim, Y.D. Corrosion Behavior of Mg2 Zn11 and MgZn2 Single Phases.
Korean J. Met. Mater. 2012, 51, 413–419. [CrossRef]
35. McMahon, M.E.; Burns, T.J.; Scully, J.R. Development of new criteria for evaluating the effectiveness of Zn-rich primers in
protecting Al-Mg alloys. Prog. Org. Coat. 2019, 135, 392–409. [CrossRef]
36. Persson, D.; Thierry, D.; LeBozec, N.; Prosek, T. In situ infrared reflection spectroscopy studies of the initial atmospheric corrosion
of Zn–Al–Mg coated steel. Corros. Sci. 2013, 72, 54–63. [CrossRef]
37. Li, B.; Dong, A.; Zhu, G.; Chu, S.; Qian, H.; Hu, C.; Sun, B.; Wang, J. Investigation of the corrosion behaviors of continuously
hot-dip galvanizing Zn–Mg coating. Surf. Coat. Technol. 2012, 206, 3989–3999. [CrossRef]
38. Zhu, Z.; Li, A.; Xu, B. Study on corrosion mechanism of arc sprayed Zn-Al-Mg coatings by XRD and EIS. Adv. Mater. Res. 2011,
230–232, 85–88. [CrossRef]
39. Ennadi, A.; Legrouri, A.; De Roy, A.; Besse, J.P. X-ray Diffraction Pattern Simulation for Thermally Treated [Zn-Al-Cl] Layered
Double Hydroxide. J. Solid State Chem. 2000, 152, 568–572. [CrossRef]
40. Singh, I.B.; Singh, M.; Das, S. A comparative corrosion behavior of Mg, AZ31 and AZ91 alloys in 3.5 NaCl solution. J. Magnesium
Alloys 2015, 3, 142–148. [CrossRef]
41. Špoták, M.; Drienovský, M.; Rízeková-Trnková, L.; Palcut, M. Corrosion of Candidate Lead-Free Solder Alloys in Saline
Solution. In Proceedings of the 24th International Conference on Metallurgy and Materials METAL 2015, Brno, Czech Republic,
3–5 June 2015; pp. 1650–1656.
42. Hu, C.-C.; Wang, C.-K. Effects of composition and reflowing on the corrosion behavior of Sn–Zn deposits in brine media.
Electrochim. Acta 2006, 51, 4125–4134. [CrossRef]
43. Calabrese, L.; Bonaccorsi, L.; Capri, A.; Proverbio, E. Electrochemical behavior of hydrophobic silane-zeolite coatings for corrosion
protection of aluminum substrate. J. Coat. Technol. Res. 2014, 11, 883–898. [CrossRef]
44. Gogola, P.; Gabalcová, Z.; Palcut, P. Experimental determination of the corrosion potential for the intermetallic Mg2 Sn phase.
(manuscript in preparation).
45. Mindat.org Hydrozincite. Available online: https://www.mindat.org/min-1993.html (accessed on 28 July 2021).
46. Mindat.org Simonkolleite. Available online: https://www.mindat.org/min-3668.html (accessed on 28 July 2021).
47. Zhitova, E.S.; Krivovichev, S.V.; Pekov, I.; Greenwell, H.C. Crystal chemistry of natural layered double hydroxides. 5. Single-
crystal structure refinement of hydrotalcite, [Mg6 Al2 (OH)16 ](CO3 )(H2 O)4 . Mineral. Mag. 2019, 83, 269–280. [CrossRef]
193
materials
Article
Influence of Degradation Product Thickness on the Elastic
Stiffness of Porous Absorbable Scaffolds Made from an
Bioabsorbable Zn–Mg Alloy
Jannik Bühring 1, *, Maximilian Voshage 2 , Johannes Henrich Schleifenbaum 2 and Holger Jahr 3
and Kai-Uwe Schröder 1
Abstract: For orthopaedic applications, additive manufactured (AM) porous scaffolds made of
absorbable metals such as magnesium, zinc or iron are of particular interest. They do not only offer
the potential to design and fabricate bio-mimetic or rather bone-equivalent mechanical properties,
they also do not need to be removed in further surgery. Located in a physiological environment,
scaffolds made of absorbable metals show a decreasing Young’s modulus over time, due to product
dissolution. For magnesium-based scaffolds during the first days an increase of the smeared Young’s
Citation: Bühring, J.; Voshage, M.;
modulus can be observed, which is mainly attributed to a forming substrate layer of degradation
Schleifenbaum, J.H.; Jahr, H.;
products on the strut surfaces. In this study, the influence of degradation products on the stiffness
Schröder, K.-U. Influence of
Degradation Product Thickness on
properties of metallic scaffolds is investigated. For this, analytical calculations and finite-element
the Elastic Stiffness of Porous simulations are performed to study the influence of the substrate layer thickness and Young’s
Absorbable Scaffolds Made from an modulus for single struts and for a new scaffold geometry with adapted polar cubic face-centered
Bioabsorbable Zn–Mg Alloy. Materials unit cells with vertical struts (f2cc,z). The finite-element model is further validated by compression
2021, 14, 6027. https://doi.org/ tests on AM scaffolds made from Zn1Mg (1 wt% Mg). The results show that even low thicknesses
10.3390/ma14206027 and Young’s moduli of the substrate layer significantly increases the smeared Young’s modulus
under axial compression.
Academic Editor: Marián Palcut
195
properties at the same time is still challenging. Biocompatible materials can be found in a
wide variety of material classes [3]. One example are polymer-based materials, which offer
great advantages in terms of customized biodegradation and design [4]. Further to mention
are ceramic materials, which also exhibit the aforementioned biodegradation and offer
particularly good healing properties for bone defects [5]. However, for fully load-bearing
applications only metals fulfill the needed properties, especially regarding strength and
stiffness [6]. The Laser Powder Bed Fusion (LPBF) process enables the individualized
production of high-resolution lattice structures with very fine struts (<250 μm) [7,8] at
reasonable costs, and is thus ideal for the production of personalized implants [9]. In
particular, the use of zinc (Zn), magnesium (Mg), iron (Fe) and their alloys, are increasingly
coming into focus for orthopaedic applications [10,11]. Although Fe-based implants would
biomechanically, and with respect to their corrosion speed [12,13], gain most from increased
porosity [14], their limited cytocompatibility is a concern [15]. Nevertheless, in comparison
to pure zinc and magnesium, iron has the highest values regarding yield strength and
Young’s modulus (σy,Fe ≈ 200–352 MPa, EFe ≈ 188–215 GPa [16–18]; σy,Zn ≈ 12–32 MPa,
EZn ≈ 43–150 GPa [12,18,19]; σy,Mg ≈ 51 MPa, E Mg ≈ 27–35 GPa [20–22]) and offers a large
margin for introducing a controlled porosity, which directly influences the strength and
stiffness properties of the material. Alloying can further improve the mechanical properties.
Adding Zn to Mg-based alloys increases the yield strength and Young’s modulus of the
material [3,13,23]. Same goes for Zn alloyed with Mg [3,19,24], whereas adding aluminum
to Zn-based alloys leads to a decrease in stiffness and strength [19].
Examples for Mg- and Zn-based studies on porous scaffolds can be found, e.g., in
Witte et al. [25], who show the feasibility of producing AM open-porous, biodegradable
and biocompatible Mg scaffolds. Li et al. [2] produced AM WE43 (Mg alloy with 4 wt%
yttrium and 3 wt% rare earth elements) scaffolds based on diamond unit cells, to show
the in vitro biodegradation behavior, mechanical properties and biocompatibility. Further-
more, Kopp et al. [26] showed that the pore size of Mg scaffolds influences the long-term
stability, while heat treatment especially effects the degradation and mechanical stability.
Cockerill et al. [27] used a casting approach to produce porous structures made of pure Zn
and studied the topology, mechanical properties, biodegradation and biocompatibility. An-
other example is shown by Li et al. [28], who produced scaffolds from Zn with a diamond
lattice structure via LPBF and studied the static and dynamic biodegradation behavior.
In a physiological environment biodegradable metals usually show a decreasing
Young’s modulus during the degradation process, due to the progressive absorption of the
metallic surface, which consequently leads to a reduction of the strut cross section [29–32].
Since the strut thickness is directly related to the stiffness, the latter will also decrease.
Interestingly, during the first days of in vitro corrosion of Mg-based (WE43) scaffolds, an
increase of around 40% in the Young’s modulus was recently reported [2]. This increase
in stiffness is mainly attributed to the formation of a composite cross section, consisting
of the base strut and an adherend layer of degradation products. A brief review of the
literature shows [3,10,29,31] that the compound of degradation products, which adheres
to the surface of the struts, consists for the most parts of hydroxides, phosphates and car-
bonates, for which only insufficient mechanical properties can be found. The phosphates
and carbonates form a compound of usually unspecified chemical composition that further
changes over time. Furthermore, a hydroxide layer is forming on the metallic surface. The
basic biochemical processes, responsible for this, can be summarized as followed [29,31,32]:
196
Materials 2021, 14, 6027
der the release of hydrogen and hydroxide ions, which form together with the metal a
hydroxide layer on the surfaces of the struts. From equivalent reactions, phosphates and
carbonates form on the strut surfaces [29]. These processes are responsible for an increase
in stiffness during the early phases of the corrosion process [2]. Later, chloride ions start
the dissolution of the biodegradable metal to cause a decrease of the cross-sectional strut
diameter of the scaffold.
%RG\IOXLG
Q H
6WUXWVXUIDFH
Figure 1. Schematic process sketch of the degradation process of absorbable metals according to
Han et al. and Li [29,31].
PL
EV = (1)
DS vs Δys
Within the scope of this work, all AM scaffolds were manufactured with a constant
layer thickness of 30 μm and EV was set for all scaffolds to 133 J/mm3 . The scaffolds were
afterwards sandblasted with 2.5 bar, to remove adhering powder particles.
197
Materials 2021, 14, 6027
Table 1. Resulting geometric parameters of the scaffold used for this study (Rm = 1.4 mm); i defines
the actual ring, starting from the middle with i = 1 according to Figure 2.
198
Materials 2021, 14, 6027
Figure 3. Resulting LPBF produced scaffold used for the physical evaluation.
where Es is the base materials Young’s modulus, Esub is the Young’s modulus of the
compound of degradation products in the substrate layer, rs is the inner radius of the
substrate layer, or rather the base strut radius, and tsub is the thickness of the substrate
layer. For the equivalent composite bending stiffness EJ results:
π π
EJ = ∑ Ej Jj = Es rs4 + Esub (rs + tsub )4 − rs4 . (4)
4 4
199
Materials 2021, 14, 6027
Figure 4. Cross section of the idealized corroded strut; in grey: base strut, in orange: compound of
degradation/reaction products.
x2 x2
Cross Secon
x3 x3
x1 x1
x3 x3
Figure 5. Finite-Element Mesh; (a) solid model, (b) beam model with schematic cross section.
an Aramis 4M system by GOM. By this, the real displacement of the specimen can be
measured. Figure 6 shows the used setup for the compression tests.
7HVW6HWXS
',&6\VWHP
3. Results
3.1. Analytical Results
Figure 7 shows the results of the analytical calculations for the Zn1Mg single struts
under axial compression. Shown is resulting composite Young’s modulus E as a function of
the substrate layer thickness tsub for different strut radii rs (50 μm–250 μm). Furthermore,
Figure 7 (a) shows the resulting absolute composite Young’s modulus (left axis) and relative
increase E/EZn1Mg ) (right axis) for a Young’s modulus twice as high, (b) three times as high,
(c) four times as high and (d) five times as high as the base materials Young’s modulus.
It can be noticed that the thinner the struts and the thicker the substrate layer, the higher
the resulting composite axial stiffness of the strut. Especially for smaller strut radii, i.e.,
rs = 50 μm, as well as for small substrate thicknesses, the effect of an increase in axial
stiffness is clearly visible. Same applies for high Young’s moduli of the substrate layer.
Figure 7. Analytical calculation of the axial stiffness of a single composite strut for varying parameters
of the substrate layer.
201
Materials 2021, 14, 6027
Figure 8 shows the results for the analytical observations of the Zn1Mg single struts
under bending. Shown is the resulting composite bending stiffness EJ as a function
of the substrate layer thickness tsub for different substrate Young’s moduli Esub , which
is set to 2–5 times the base materials Young’s modulus EZn1Mg . Furthermore, Figure 8
shows the resulting absolute composite bending stiffness EJ (left axis) and relative in-
crease EJ/( EJ ) Zn1Mg ) (right axis) for (a) a base strut radius rs = 50 μm, (b) rs = 100 μm,
(c) rs = 150 μm and (d) rs = 200 μm. With increasing substrate layer thickness and higher
substrate Young’s modulus, a higher increase in bending stiffness can be observed. Espe-
cially for small strut radii, such as rs = 50 μm, very high increases in bending stiffness can
be achieved. This is not only the case for high moduli of the substrate layer, but also in the
case when the composite of degradation products has the same Young’s modulus.
Figure 8. Analytical calculation of the bending stiffness of a single composite strut for varying
parameters of the substrate layer.
202
Materials 2021, 14, 6027
Figure 9. Reaction force (RF) comparison of modeling approaches for base and corroded strut under
axial compression; (left) beam elements, (right) solid elements.
Figure 10. Reaction force (RF) and reaction moment (RM ) comparison of modeling approaches for
base and corroded strut under bending; (left) beam elements, (right) solid elements.
203
Materials 2021, 14, 6027
Table 2. Percentage increase of the smeared Young’s modulus E for varying substrate Young’s moduli
Esub and layer thicknesses tsub .
rs [mm]
Esub [GPa] tsub [μm]
0.05 0.1 0.15 0.2 0.25
5 22% 11% 8% 6% 4%
19 10 46% 23% 15% 11% 9%
20 102% 49% 35% 23% 18%
5 43% 22% 14% 11% 8%
38 10 91% 44% 29% 22% 17%
20 201% 95% 61% 45% 35%
5 64% 32% 21% 16% 13%
57 10 136% 66% 43% 32% 25%
20 300% 140% 90% 66% 52%
5 85% 42% 28% 21% 17%
76 10 180% 87% 57% 42% 34%
20 353% 164% 106% 78% 61%
Figure 11. Results of the FE simulations of corroded scaffolds for varying parameters.
Figure 3. Furthermore, the modeling using beam elements neglects the accumulation of
material in the nodes of the real scaffold. Furthermore, the used Young’s modulus is based
on literature data and it is well known that Young’s moduli of AM materials tend to show
slight differences (see also Section 1). Nevertheless, the tests show that the FE model based
on beam elements provides sufficiently accurate results in terms of the resulting smeared
axial stiffness and can be used for the parametric study.
Figure 12. Validation of FE model: Resulting load-displacement curve of two tested LPBF produced
scaffolds and equivalent FE model.
4. Discussion
We investigated the influence of degradation products on the elastic stiffness proper-
ties of biodegradable metallic scaffolds. For this, a hypothetical compound of degradation
products was modeled as a thin-walled layer with a homogeneous cross section. The
compound of degradation products consists for the most parts of hydroxides, phosphates
and carbonates [29–32]. Since there is no sufficient database, yet, for the mechanical
properties of the degradation products, hypothetical Young’s moduli were defined using
multiples of the Young’s modulus of the base material, which was obtained from litera-
ture data [3,12,13,16–24]. By this, the influence of the degradation products on the elastic
stiffness properties as a function of the layer thickness and Young’s modulus could be
investigated. This was done using analytical models and finite-element simulations for
single struts, to show the direct influence of the layer of degradation products on the axial
and bending stiffness, as well as for whole scaffold geometries, to show the superposed
influence on the axial smeared Young’s modulus of a specific scaffold geometry. Two
modeling approaches were contrasted for the FE simulations, first a meshing strategy using
a 3D volume mesh and second using beam elements. Both approaches show concurring
results. For this reason, the beam model was used for a parametric study on whole lattice
scaffold geometries, due to the enormous difference regarding the calculation time. To
validate the FE model, scaffolds were produced via LPBF and compression tests on two
scaffolds were done.
From the single strut investigations can be concluded that depending on the substrates
Young’s modulus and the ratio of strut radius to thickness of the substrate layer, significant
increases of the composite axial and bending stiffness is expected. The effect intensifies,
the smaller the base strut radius in the initial state is. This applies as well as for relatively
low Young’s moduli of the substrate layer as for very high Young’s moduli. In comparable
studies [2,14,15,25,26], mentioned in the introduction part, strut diameters of 300–400 μm
were used for orthopaedic scaffolds. Even for low layer thicknesses (i.e., 10 μm) and low
Young’s moduli, for single struts with diameters in this range, depending on the thickness
of the substrate layer and the composite module, an increase of more than 10% for the
Young’s modulus under axial compression and more than 40% in bending stiffness can be
205
Materials 2021, 14, 6027
expected, which is not to be confused with the Young’s modulus in the bending load case.
To validate the base FE model, physical test results were compared to an equivalent FE
simulation, using beam elements for meshing. As presented in the results section, the beam
modeling shows similar results, compared to a much more numerically expensive meshing
strategy with solid elements. The compression tests on LPBF produced scaffolds show
reproducible results and furthermore equivalent smeared Young’s moduli in the FE model
and physical tests. For this reason, a FE parametric study on the tested geometry was done
by varying the substrate layer thickness and the Young’s moduli of the compound of the
degradation products in the substrate layer, to study the influence of the substrate layer
on the smeared Young’s modulus of complex scaffold geometries. Our results show that
an enormous increase in stiffness can be expected even for complex geometries, which
was also observed by Li et al. [2] for diamond lattice structures made from WE43. For the
previously mentioned example of strut diameters of 300–400 μm, the investigations on the
scaffolds show that a much stronger effect can be observed due to the superposition of the
axial and bending stiffness increase. As presented in the results section, the increase of the
smeared axial Young’s modulus under compression can be quantified to approximately
10–40% for a layer thickness of 10 μm and varying Young’s moduli. The effect intensifies to
values of approximately 20–80%, if i.e., a layer thickness of 20 μm is assumed. From this can
be followed that compared to the separated reflection of the influence of the substrate layer
on axial and bending stiffness of single struts, the effect of a stiffness increase is clearly more
pronounced in the case of scaffold geometries. This is mainly attributable to the combined
loading in compression and bending of the struts, which both ultimately have a direct
effect on the smeared Young’s modulus of the scaffold. Nevertheless, the investigations
on single struts give clear indications about the formation of the effect. Furthermore, the
analytical expressions show the direct influence of the thickness and Young’s modulus of
the degradation products.
5. Conclusions
In conclusion, our analytical and numerical modeling approach basically confirmed
earlier assumptions by Li et al. [2] that the increase in stiffness of corrosion product
layer-coated AM WE43 is indeed due to formation of a composite beam of base strut
and substrate layer. As shown in this discussion, even for low thicknesses and Young’s
moduli of the degradation product layer, axial stiffness increases of more than 40% can be
achieved. Even though the geometry of the scaffold is different at the investigations of Li
et al., this study clearly shows the influence on the stiffness. Nevertheless, our results must
be validated by further investigations on corroded single struts or equal, to validate the
formation of an almost homogeneous layer of degradation products and to obtain more
knowledge about the real composite Young’s modulus, or rather the Young’s modulus of
the compound of degradation products.
Author Contributions: J.B. performed most of the analytical analyses, J.B. modeling, physical testing,
and drafted the paper. J.B., M.V. and H.J. contributed to the design of the scaffold, while M.V.
produced them H.J. initiated and supervised the study, including the analyses. J.H.S. and K.-U.S.
contributed their extensive experience, gave advice regarding the content and the manuscript. All
authors have read and agreed to the published version of the manuscript.
Funding: Parts of this work were supported by the Federal Ministry of Education and Research
(BMBF) and the Ministry of Culture and Science of the State of North Rhine-Westphalia (MKW)
under the Excellence Strategy of the Federal Government and the Länder (OPSF597).
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Acknowledgments: The authors thank the BMBF for funding of the work.
206
Materials 2021, 14, 6027
References
1. Böstman, O.; Pihlajamäki, H. Clinical biocompatibility of biodegradable orthopaedic implants for internal fixation: A review.
Biomaterials 2000, 21, 2615–2621. [CrossRef]
2. Li, Y.; Zhou, J.; Pavanram, P.; Leeflang, M.A.; Fockaert, L.I.; Pouran, B.; Tümer, N.; Schröder, K.U.; Mol, J.M.C.; Weinans, H.; et al.
Additively manufactured biodegradable porous magnesium. Acta Biomater. 2018, 67, 378–392. [CrossRef] [PubMed]
3. Jahr, H.; Li, Y.; Zhou, J.; Zadpoor, A.A.; Schröder, K.U. Additively Manufactured Absorbable Porous Metal Implants—Processing,
Alloying and Corrosion Behavior. Front. Mater. 2021, 8, 292. [CrossRef]
4. Liu, X.; Ma, P.X. Polymeric scaffolds for bone tissue engineering. Ann. Biomed. Eng. 2004, 32, 477–486. [CrossRef] [PubMed]
5. Seitz, H.; Rieder, W.; Irsen, S.; Leukers, B.; Tille, C. Three-dimensional printing of porous ceramic scaffolds for bone tissue
engineering. J. Biomed. Mater. Res. Part B—Appl. Biomater. 2005, 74B, 782–788. [CrossRef]
6. Chen, Q.; Thouas, G.A. Metallic implant biomaterials. Mater. Sci. Eng. R Rep. 2015, 87, 1–57. [CrossRef]
7. Cutolo, A.; Engelen, B.; Desmet, W.; van Hooreweder, B. Mechanical properties of diamond lattice Ti–6Al–4V structures produced
by laser powder bed fusion: On the effect of the load direction. J. Mech. Behav. Biomed. Mater. 2020, 104, 103656. [CrossRef]
8. Gümrük, R.; Mines, R.; Karadeniz, S. Static mechanical behaviours of stainless steel micro-lattice structures under different
loading conditions. Mater. Sci. Eng. A 2013, 586, 392–406. [CrossRef]
9. Yan, Q.; Dong, H.; Su, J.; Han, J.; Song, B.; Wei, Q.; Shi, Y. A Review of 3D Printing Technology for Medical Applications.
Engineering 2018, 4, 729–742. [CrossRef]
10. Li, Y.; Jahr, H.; Zhou, J.; Zadpoor, A.A. Additively manufactured biodegradable porous metals. Acta Biomater. 2020, 115, 29–50.
[CrossRef] [PubMed]
11. Wang, J.L.; Xu, J.K.; Hopkins, C.; Chow, D.H.K.; Qin, L. Biodegradable Magnesium-Based Implants in Orthopedics-A General
Review and Perspectives. Adv. Sci. 2020, 7, 1902443. [CrossRef]
12. Wen, P.; Voshage, M.; Jauer, L.; Chen, Y.; Qin, Y.; Poprawe, R.; Schleifenbaum, J.H. Laser additive manufacturing of Zn metal parts
for biodegradable applications: Processing, formation quality and mechanical properties. Mater. Des. 2018, 155, 36–45. [CrossRef]
13. Qin, Y.; Wen, P.; Guo, H.; Xia, D.; Zheng, Y.; Jauer, L.; Poprawe, R.; Voshage, M.; Schleifenbaum, J.H. Additive manufacturing of
biodegradable metals: Current research status and future perspectives. Acta Biomater. 2019, 98, 3–22. [CrossRef] [PubMed]
14. Li, Y.; Jahr, H.; Lietaert, K.; Pavanram, P.; Yilmaz, A.; Fockaert, L.I.; Leeflang, M.A.; Pouran, B.; Gonzalez-Garcia, Y.; Weinans, H.;
et al. Additively manufactured biodegradable porous iron. Acta Biomater. 2018, 77, 380–393. [CrossRef] [PubMed]
15. Li, Y.; Jahr, H.; Pavanram, P.; Bobbert, F.S.L.; Paggi, U.; Zhang, X.Y.; Pouran, B.; Leeflang, M.A.; Weinans, H.; Zhou, J.; et al.
Additively manufactured functionally graded biodegradable porous iron. Acta Biomater. 2019, 96, 646–661. [CrossRef] [PubMed]
16. Song, B.; Dong, S.; Deng, S.; Liao, H.; Coddet, C. Microstructure and tensile properties of iron parts fabricated by selective laser
melting. Opt. Laser Technol. 2014, 56, 451–460. [CrossRef]
17. Song, B.; Dong, S.; Liu, Q.; Liao, H.; Coddet, C. Vacuum heat treatment of iron parts produced by selective laser melting:
Microstructure, residual stress and tensile behavior. Mater. Des. 2014, 54, 727–733. [CrossRef]
18. Montani, M.; Demir, A.; Mostaed, E.; Vedani, M.; Previtali, B. Processability of pure Zn and pure Fe by SLM for biodegradable
metallic implant manufacturing. Rapid Prototyp. J. 2017, 23, 514–523. [CrossRef]
19. Yang, Y.; Yuan, F.; Gao, C.; Feng, P.; Xue, L.; He, S.; Shuai, C. A combined strategy to enhance the properties of Zn by laser rapid
solidification and laser alloying. J. Mech. Behav. Biomed. Mater. 2018, 82, 51–60. [CrossRef]
20. Zhou, Y.; Wu, P.; Yang, Y.; Gao, D.; Feng, P.; Gao, C.; Wu, H.; Liu, Y.; Bian, H.; Shuai, C. The microstructure, mechanical properties
and degradation behavior of laser-melted Mg Sn alloys. J. Alloys Compd. 2016, 687, 109–114. [CrossRef]
21. Ng, C.C.; Savalani, M.M.; Lau, M.L.; Man, H.C. Fabrication of magnesium using selective laser melting technique. Rapid Prototyp.
J. 2011, 17, 479–490. [CrossRef]
22. Ng, C.C.; Savalani, M.M.; Lau, M.L.; Man, H.C. Microstructure and mechanical properties of selective laser melted magnesium.
Appl. Surf. Sci. 2011, 257, 7447–7454. [CrossRef]
23. Wei, K.; Zeng, X.; Wang, Z.; Deng, J.; Liu, M.; Huang, G.; Yuan, X. Selective laser melting of Mg-Zn binary alloys: Effects of Zn
content on densification behavior, microstructure, and mechanical property. Mater. Sci. Eng. A 2019, 756, 226–236. [CrossRef]
24. Kubásek, J.; Dvorský, D.; Čapek, J.; Pinc, J.; Vojtěch, D. Zn-Mg Biodegradable Composite: Novel Material with Tailored
Mechanical and Corrosion Properties. Materials 2019, 12, 3930. [CrossRef] [PubMed]
25. Frank, W.; Lucas, J.; Wolfgang, M.; Zienab, K.; Kristin, S.; Tanja, S. Open-porous biodegradable magnesium scaffolds produced by
selective laser melting for individualized bone replacement. Front. Bioeng. Biotechnol. 2016, 4. [CrossRef]
26. Kopp, A.; Derra, T.; Müther, M.; Jauer, L.; Schleifenbaum, J.H.; Voshage, M.; Jung, O.; Smeets, R.; Kröger, N. Influence of design
and postprocessing parameters on the degradation behavior and mechanical properties of additively manufactured magnesium
scaffolds. Acta Biomater. 2019, 98, 23–35. [CrossRef] [PubMed]
27. Cockerill, I.; Su, Y.; Sinha, S.; Qin, Y.X.; Zheng, Y.; Young, M.L.; Zhu, D. Porous zinc scaffolds for bone tissue engineering
applications: A novel additive manufacturing and casting approach. Mater. Sci. Eng. C Mater. Biol. Appl. 2020, 110, 110738.
[CrossRef] [PubMed]
28. Li, Y.; Pavanram, P.; Zhou, J.; Lietaert, K.; Taheri, P.; Li, W.; San, H.; Leeflang, M.A.; Mol, J.M.C.; Jahr, H.; et al. Additively
manufactured biodegradable porous zinc. Acta Biomater. 2020, 101, 609–623. [CrossRef]
207
Materials 2021, 14, 6027
29. Han, H.S.; Loffredo, S.; Jun, I.; Edwards, J.; Kim, Y.C.; Seok, H.K.; Witte, F.; Mantovani, D.; Glyn-Jones, S. Current status and
outlook on the clinical translation of biodegradable metals. Mater. Today 2019, 23, 57–71. [CrossRef]
30. Zheng, Y.F.; Gu, X.N.; Witte, F. Biodegradable metals. Mater. Sci. Eng. R Rep. 2014, 77, 1–34. [CrossRef]
31. Li, P. Absorbable Zinc-based alloy for craniomaxillofacial osteosynthesis implants. In Faculty of Medicine; Eberhard Karls
University Tübingen: Tübingen, Germany, 2020.
32. Kannan, M.B.; Moore, C.; Saptarshi, S.; Somasundaram, S.; Rahuma, M.; Lopata, A.L. Biocompatibility and biodegradation
studies of a commercial zinc alloy for temporary mini-implant applications. Sci. Rep. 2017, 7, 15605. [CrossRef] [PubMed]
33. Meiners, W. Direktes Selektives Laser-Sintern Einkomponentiger Metallischer Werkstoffe. Ph.D. Thesis, RWTH Aachen
University, Aachen, Germany, 1999.
34. Ulutan, S.; Gilbert, M. Mechanical properties of HDPE/magnesium hydroxide composites. J. Mater. Sci. 2000, 35, 2115–2120.
[CrossRef]
35. Ulian, G.; Valdrè, G. Anisotropy and directional elastic behavior data obtained from the second-order elastic constants of
portlandite Ca(OH)2 and brucite Mg(OH)2 . Data Brief 2018, 21, 1375–1380. [CrossRef] [PubMed]
36. Yao, C.; Wu, Z.; Zou, F.; Sun, W. Thermodynamic and Elastic Properties of Magnesite at Mantle Conditions: First-Principles
Calculations. Geochem. Geophys. Geosyst. 2018, 19, 2719–2731. [CrossRef]
208
materials
Article
Corrosion Susceptibility and Allergy Potential of
Austenitic Stainless Steels
Lucien Reclaru 1 and Lavinia Cosmina Ardelean 2, *
1 Scientific Independent Consultant Biomaterials and Medical Devices, 103 Paul-Vouga,
2074 Marin-Neuchâtel, Switzerland; [email protected]
2 Department of Technology of Materials and Devices in Dental Medicine, “Victor Babes” University of
Medicine and Pharmacy Timisoara, 2 Eftimie Murgu sq, 300041 Timisoara, Romania
* Correspondence: [email protected]
Abstract: Although called stainless steels, austenitic steels are sensitive to localized corrosion,
namely pitting, crevice, and intergranular form. Seventeen grades of steel were tested for localized
corrosion. Steels were also tested in general corrosion and in galvanic couplings (steels–precious
alloys) used in watchmaking applications. The evaluations have been carried out in accordance with
the ASTM standards which specifically concern the forms of corrosion namely, general (B117-97,
salt fog test), pitting (G48-11, FeCl3 ), crevice (F746-87) and intergranular (A262-15, Strauss chemical
test and G108-94, Electrochemical potentiodynamic reactivation test). All tests revealed sensitivity to
corrosion. We have noticed that the transverse face is clearly more sensitive than the longitudinal
face, in the direction of rolling process. The same conclusion has been drawn from the tests of
nickel release. It should be pointed out that, despite the fact that the grade of steel is in conformity
with the classification standards, the behavior is very different from one manufacturer to another,
due to parameters dependent on the production process, such as casting volume, alloying additions,
and deoxidizing agents. The quantities of nickel released are related to the operations involved in the
manufacturing process. Heat treatments reduce the quantities of nickel released. The surface state
has little influence on the release. The hardening procedures increase the quantities of nickel released.
The quantities of released nickel are influenced by the inclusionary state and the existence of the
secondary phases in the steel structure. Another aspect is related to the strong dispersion of results
concerning nickel release and corrosion behavior of raw materials.
Keywords: austenitic steels; general (uniform) corrosion; pitting corrosion; crevice corrosion; intergranular
corrosion; galvanic couplings; nickel release; contact with skin; medical devices; watchmaking
1. Introduction
Corrosion represents an important factor in the design and selection of metals and alloys for
different purposes, as various corrosion mechanisms can lead to failure [1,2]. Corrosion resistance is an
important criterion for selecting materials used, because the cost of their degradation due to corrosion
and the associated environmental impact are quite substantial [3]. Like all metals, stainless steels can
undergo chemical corrosion over time [4–6].
Corrosion manifests in different forms and depends on a multitude of physico-chemical factors
(chemical composition and microstructure of the alloy, temperature, pH, chemical composition of the
environment) and mechanical factors (stresses, friction) [7]. The relationship between corrosion rate
and grain size has been revealed in numerous studies [8–11]. Its importance lies in the fact that this
parameter can be tailored by the producers [8–11].
The austenitic steels belong to the stainless steels family and are being characterized by high
Nieq and Creq [12,13]. Over time, the chemical composition, mechanical properties, resistance to
corrosion, machinability and polish ability of the austenitic steels have evolved considerably, and new
production processes have been developed by the steel manufacturers [14]. Each chemical element
in their composition plays an important role in their properties [15], including corrosion resistance,
and can be substantially modified by adding certain elements as Cu, Ti, Nb, Al, Si and Ca. Generally,
the composition of austenitic stainless steels is adjusted to meet service requirements in various
corrosive environments [8]. The corrosion sensitivity of austenitic steels mainly takes the form of
pitting, crevice and intergranular type [16].
An important aspect which concerns the austenitic steels is the release of nickel in contact with the
skin. The role of nickel in the biological response to alloys is significant with regard to toxicology and
biological performance. The current trend is to eliminate nickel from alloys for medical applications.
However, this needs a careful evaluation since no compromise is acceptable concerning the mechanical
properties, corrosion resistance or any other possible undesirable consequences due to the substitution
of nickel [7,17].
Nickel allergy is the most widespread of all contact allergies. In the European population,
the prevalence of nickel allergy is of 10%–15% of adult females and 1%–3% of adult males [18–22].
Of nickel-sensitive people in the general population, 30% develop hand eczema. Teenagers and young
adults tend to have a higher prevalence due to frequent body piercing.
In Europe, for objects containing nickel, intended for permanent contact with skin, Directive
94/27/EC imposed a ban if the rate of nickel release exceeds 0.5 μg/cm2 ·week. The subsequent Directive
2004/96/EC: “Piercing in the Human Body” specifies that the limit rate of nickel release, for these cases,
is 0.2 μg/cm2 ·week [23–25].
The aim of this study is to evaluate the sensitivity, under the same conditions, of 17 austenitic
steels of the 304, 316 and 904 series, for uniform, pitting, crevice and galvanic corrosion. Our interest
was also to assess the behavior differences of the transverse surface of the samples compared to the
longitudinal one.
Table 1. Chemical composition (wt.%) of the grades of austenitic steels used in corrosion tests.
210
Materials 2020, 13, 4187
According to the classification of Fontana [26], the evaluation of their behavior in uniform, pitting,
crevice, intergranular corrosion was presented, as well as in galvanic couplings.
Test Conditions
NaCl
Electrolyte
Artificial sweat
Temperature 35 ◦ C
Total duration 12 days
NaCl 5% 6 days
Artificial sweat
6 days
Dilution x40
Table 3. Types of electrolytes used and the experimental conditions for testing pitting corrosion [31–33].
211
Materials 2020, 13, 4187
The immersion times were therefore different: 15 days for FeCl3 0.1 M, 26 days for NaCl 0.5 M and
30 days for artificial sweat (Table 3), because of the difference in aggressiveness of the media (chemical
composition, concentration) towards the evaluated materials [34].
Figure 1. Specific assembly for the crevice corrosion test. Crevice corrosion aspect of the test sample.
The sample was mounted in the rotating electrode of the measuring cell (Figure 2).
The test consisted of two stages:
• In the first stage, anodic excitation of the sample to be evaluated was carried out at 800 mV vs.
SCE (saturated calomel electrode) for 10 s;
• In the second stage, the potentiost at imposed the abandon potential value for 15 min. As a result,
the current variation was plotted as a function of time for an imposed potential.
7DFK\SURFHVVRU
3RWHQWLRVWDW
5RWDWLQJ
HOHFWURGH
&RXQWHUHOHFWURGH
5HIHUHQFH
HOHFWURGH
0HDVXUHPHQW
FHOO /XJJLQ
SUREH
:RUNLQJ
HOHFWURGH
If the current recorded stayed within the cathodic domain (negative values), a fresh measurement
cycle was started: excitation for 10 s at 800 mV and current measurement for a potential set at Eabandon
+25 mV (ASTM recommends +50 mV). The cycles were repeated, each time at a higher potential,
until the current measured moved into the anodic domain (positive values). As a result, the crevice
potential was determined, which corresponded to the last-but-one measurement for which the current
212
Materials 2020, 13, 4187
was positive. Once the Teflon ring was removed, a groove of crevice corrosion was noted on the surface
of the metal sample.
Table 4. ASTM A262-15 [36] Standard practice for detecting susceptibility to intergranular attack in
austenitic stainless steels.
Testing Evaluation
Designation Test Temperature Applicability
Time Method
Chromium
Oxalic Acid Etch Microscopic
Practice A Ambient 1.5 min Carbide
Screening Test Examination
sensitization only
Ferric Sulfate and 50% Chromium Weight loss/
B Boiling 120 h
Sulfuric Acid Carbide Corrosion Rate
Chromium Weight loss/
C 65% Nitric Acid Boiling 4h
Carbide Corrosion Rate
10% Nitric-3% Hydro Corrosion Rates of
Chromium
Fluoric Acid (This test “unknown” over
D 70 ◦ C 4h carbide in 316,
has been removed that of solution
316L, 317, 317L
from A 262-15) annealed specimen
6% Copper Sulfate 16% Examination for
Chromium
E Sulfuric Acid with Boiling 24 h fissures after
Carbide
metallic copper bending
Copper Sulfate 50% Chromium
Weight loss/
F Sulfuric Acid with Boiling 120 h Carbide in 316
Corrosion rate
metallic copper and 316L
Evaluation of intergranular corrosion according to standard ASTM A262-15-the Strauss test [36].
The test medium was a solution of copper sulphate-50% sulfuric acid in the presence of metallic
copper, brought to the boiling point (125 ◦ C) for 120 h. This test investigated the intergranular corrosion
behavior of steel in the potential range between 110–350 mV. The test setup was in accordance with
ASTM A262-15 [36]. The samples to be tested, tubes of #3, 1.4435/316L steel, were placed in specific
glass cradle.
213
Materials 2020, 13, 4187
a solution of 0.5 M H2 SO4 + 0.01 M KSCN. Subsequently, the potential was decreased, at a rate of
6 V/h, to the corrosion potential, Ecorr . This decrease resulted in reactivation of the specimen, involving
breakdown of the passive film covering chromium depleted areas of material. The area under the large
loop generated in the curve of potential vs. current (Figure 3) was proportional to the electric charge Q,
that depends on surface area and grain size. In non-sensitized material, the passive film was intact and
the loop size was small.
3DVVLYH
PYYV6&(
1RQ6HQVLWL]HG 6HQVLWL]HG
ORJ,
Figure 3. Procedures of single loop EPR test method according to ASTM G108–94 (2015) [48].
The samples tested were tubes of 304L AISI, used for manufacturing medical endoscopes (Table 5).
The tubes #Test 1 and #Test 2 were suspected to be sensitized in intergranular corrosion. For the
evaluation test, reference samples #Brut 1 and #Brut 2 and three samples from the supplier stock,
subjected to heat treatments at 500, 620 and 750 ◦ C, respectively, were used (Table 5).
Code Description
#Test 1 Tube returned from user
#Test 2 Tube returned from user
#Brut 1 Tube from supplier stock (reference)
#Brut 2 Tube from supplier stock (second reference)
#500 Tube from supplier stock+ 500 ◦ C heat treatment, 1 h
#620 Tube from supplier stock+ 621 ◦ C heat treatment, 1 h
#750 Tube from supplier stock+ 750 ◦ C heat treatment, 1 h
The sample (Table 5), transversely cut, was incorporated into a resin and “mirror” polished.
The resin was machined to be adapted to the working electrode. The mounting of the electrodes,
the electrochemical cell and the EPR measurement conditions were those of ASTM G108-94 (2015) [48].
The ASTM G108-94 method allows a quantitative evaluation of the intergranular corrosion sensitization
of steels AISI 304 and AISI 304L. The purpose of the test was to evaluate the intergranular corrosion of
the transverse surface by an electrochemical scanning method from +200 mV to −400 mV vs. SCE. In the
mathematical calculation for evaluating the sensitivity to intergranular corrosion, another parameter,
the corresponding value of the grain index, has to be considered, according to ASTM E112-13 [49].
214
Materials 2020, 13, 4187
Steels: 1.4441, 1.4435, 316 L F, 316 L F Cu, 1.4301, 1.4305, Sandvik 1802, 1.4104. 1.4105, 1.4539
and 12/12.
The samples, in form of 10-mm-diameter discs, were “mirror” polished, washed with a mixture of
acetone and ethanol, and rinsed with deionized water 18 MΩ·cm. After drying with hot air, the samples
were introduced into the PTFE sample holder, specially designed for the rotating electrode test.
The electrochemical measurements were made with a potentiostatic assembly of three electrodes:
a working electrode (rotating electrode), a platinum counter-electrode and a reference SCE electrode.
Given that diffusion phenomena play a major role with regard to the changes produced at the
metal/solution interface and consequently to the state and composition of the metal surfaces layers,
readings were made in a laminar system (criterion of Re = 3200) with a limit current iL = 56 mA, and a
rotational velocity of 300 rpm, to control the mass transfer phenomena. The open circuit potentials
(Eoc ) were measured after 24 h of immersion.
In the second stage, our interest was focused on galvanic couplings in the assembly of steel watch
strap links with precious metal alloys (18K gold). The evaluation was indirectly made, by measuring
the quantities of nickel released after 7 days of immersion in artificial sweat, according to standard EN
1811-2011+A1:2015 [50]. The tests were carried out on 27 gold-steel links (Figure 4), 5N18 (18K gold
alloy)-1.4441 (316L) and 5N18 (18K alloy)-1.4539 (904L).
215
Materials 2020, 13, 4187
Figure 5. Uniform corrosion salt fog test, according to ASTM B 117-97. Samples #2-1.4427 So,
#3-1.4435/316L, #4-316LUgim and #5-316L Val.
According to Zanotto et al., the test has limitations, it becomes non-discriminating in case of steels
with excellent corrosion resistance [51].
In case of steels #8, #10, #6 and #17, the test, carried out under the same experimental conditions,
showed no signs of corrosion of the transversal or longitudinal surfaces, similar to sample #3 (Figure 6).
Figure 6. Pitting test results of the transverse and longitudinal surfaces for various grades of steel
alloys (0.5 M FeCl3 test medium at 50◦ for 2 h).
The results obtained are presented in Table 6. In general, the transverse surfaces were corroded
with the exception of steel #3. Examination of steel #3 did not reveal any traces of corrosion either in
the cold-worked state.
Table 6. Observations after the salt fog test (samples tested in annealed, cold-worked state).
216
Materials 2020, 13, 4187
'HQVLW\SLWWLQJ QUFPA
7UDQVYHUVDO /RQJLWXGLQDO
Figure 7. Pitting test results of the transverse and longitudinal surfaces for various grades of steels
(0.1 M FeCl3 test medium at 37 ◦ C for 15 days).
'HQVLW\SLWWLQJ QUFPA
7UDQVYHUVDO /RQJLWXGLQDO
Figure 8. Pitting test results of the transverse and longitudinal surfaces for various grades of steels
(0.5 M NaCl test medium at 37 ◦ C for 26 days).
'HQVLW\SLWWLQJ QUFP
7UDQVYHUVDO /RQJLWXGLQDO
Figure 9. Pitting test results of the transverse and longitudinal surfaces for various grades of steels
(artificial sweat test medium at 37 ◦ C for 30 days).
217
Materials 2020, 13, 4187
In all test environments, the transverse surfaces showed a higher pitting density compared to the
longitudinal ones.
Using a Kontron KS 300 Version 1.2 image analysis program, the cross-sectional area of alloys
#1, #2, #3, #4, #5, #7, #8 and #10, tested in 0.5 M FeCl3 at 50 ◦ C for 2 h was analyzed statistically in
relation to the area size of the pits. The following classes were established accordingly: <20, 20–50,
50–150, 150–500, 500–1000 and >1000 μm2 . The densities (number of pits/cm2 ) according to the above
mentioned classes are presented in Figure 10.
The image analysis revealed a density of pits which can be significant (more than 10,000 for
sample #2, Figure 10). Under these conditions, it was necessary to define criteria which enable easier
identification of pitting corrosion. Examination of the surfaces revealed numerous cavities which were
not necessarily pitting.
Figure 10. Number of pits counted by class, according to the size of the pit surface area.
At this point, it is essential to clarify the definition of pitting corrosion. According to the ASTM,
a pitting corrosion is electrochemically active if an anodic dissolution of the alloy occurs within the
cavity. Thus, by definition, a pitting corrosion releases cations in the electrolyte. It is therefore sufficient
to carry out the corrosion test in an electrolyte containing traces of an analytical reagent which forms,
with one of the cations released by the active corrosion pits, an insoluble colored compound which will
deposit near the pit. The common element for all the alloys in question is iron, the iron dissolution
mechanism being based on ferrous ions (Fe2+ ). Tests carried out on several reagents showed that
potassium ferricyanide (K3 [Fe (CN6 ]) is suitable to form an insoluble colored complex. Corrosion pits,
revealed as blue discs encircling the pits (Figure 11), were used to determine the number per unit area.
μP
Figure 11. Corrosion pits revealed as blue discs (cross section sample #2).
In conclusion, the FeCl3 solution, according to the ASTM G48-11 [31] standard, was aggressive
with respect to the pitting corrosion behavior of the alloys studied; within two hours, most steel
218
Materials 2020, 13, 4187
grades show readily identifiable macroscopic corrosion pits. On the other hand, in case of NaCl
0.5 M or artificial sweat ISO 3160-2 [30], the evaluation of the pits density was problematic given the
difficulty of identifying the effectively electrochemically active pits. The use of potassium ferricyanide,
which precipitates in the form of Turnbull blue in the presence of Fe2+ ions, greatly facilitated the
evaluation of the density of pits after immersion in this type of electrolyte.
According to Blackwood [2], pitting corrosion was a common problem with the early 304 stainless
steels. In the case of 316L stainless steels, the addition of 2–3 wt% Mo has greatly reduced the number
of failures due to pitting corrosion [2].
After initiation, pits either keep growing or repassivation may occur. According to Virtanen,
an alloy with a high pitting corrosion resistance should ideally combine low susceptibility to pit
initiation, low pit propagation rate, and fast repassivation [53]. According to Melchers, in case of metals
with electron-conducting passive films such as stainless steels, the number of pits usually correlates
inversely with their average depth, since the cathodic current consumed by the large passive surface
area fosters anodic dissolution inside the pits [54]. The pits may grow at different rates, depending on
the number of active pits [54]. According to Abbasi Aghuy et al., changes in electrochemical behavior of
metal due to grain refinement as a consequence of changing grain boundary densities may occur [55].
P9
&XUUHQWX$FP
&XUUHQWX$FP
P9
P9
P9
VDPSOH VDPSOH
7LPH PLQ 7LPH PLQ
(a) (b)
Figure 12. The crevice test potentiostatic curves for the longitudinal surface of sample #8 (a) and #10 (b).
Figure 13 shows the values of the crevice potentials determined for both longitudinal and
transverse surfaces, according to ASTM F746-87 standard [35].
&UHYLFH3RWHQWLDO( PY6&(
/ 7
Figure 13. Crevice potential values measured for the transverse and longitudinal surfaces of the steels
considered (L: longitudinal, T: transverse).
219
Materials 2020, 13, 4187
The values of the crevice potentials measured did not reveal any difference in susceptibility to
crevice corrosion between the two surfaces. The only difference was that the transverse surface showed
lower values than the longitudinal surface, but these differences remained in the field of experimental
errors. In other words, there was no significant difference in the crevice corrosion behavior between
the two surfaces.
In case of 316L steels (#2, #3, #4, #6, #8, and #9), the values of the crevice potentials were different,
due to the structure type of inclusions and composition in minor chemical elements.
The study of Poyetet et al., involving 18-10 type stainless steels [56], has concluded that the
reactivity of the inclusions, in terms of their contribution to the onset of pitting, is a function of their
association (Table 7).
Mixed oxide-sulfide or silicate-sulfide inclusions are the most susceptible to pitting. The corrosion
susceptibility of inclusions might be ranked, in increasing order: sulfides < alumina-sulfides <
silicate-sulfides < Mg-oxide-sulfides. By themselves, sulfides do not have a particularly detrimental
action on the pitting corrosion resistance of steel, but they become particularly harmful when associated
in the form of mixed inclusions. As far as shape is concerned, globular inclusions (present only in the
as-cast, undeformed material) seem to be less harmful than inclusions deformed during hot working
of the metal [56].
When considering the final values of currents recorded after 15 min for each level and representing
the current as a function of potential, a series of “polarization curves”, specific to the crevice corrosion
process were obtained (Figure 14).
&XUUHQW,DUHD Q$FP
7
7 7 7
7 / / / / /
&XUUHQW Q$FP
/
7
7
3RWHQWLHO( P9 3RWHQWLHO( P9
7 7 7 7 7 7 7 / / / / / / /
(a) (b)
Figure 14. Polarization curves (current value recorded after 15 min vs. preselected potential;
(a) transverse and (b) longitudinal surface.
When comparing the crevice corrosion behavior of the two surfaces, no real difference in
susceptibility to corrosion was noticed. On the other hand, in accordance to Bryant et al. [57], each steel
has a different behavior to crevice corrosion. Some steels do not reveal a “passivation capacity” before
220
Materials 2020, 13, 4187
reaching the value of crevice initiation potential. In case of 316L, respectively #2, #3, #4, #6 and #8 this
difference was noticed. According to Liu et al., in case of 316L stainless steel, widely used as a metallic
biomaterial, crevice corrosion has been a serious concern [58]. In case of #10, a difference was expected,
due to the better corrosion resistance compared to the 316L family.
In conclusion, the evaluation of the crevice corrosion resistance did not reveal marked differences
between the behavior of the transverse surface compared to the longitudinal one. This shows that the
pitting and crevice corrosion mechanisms, although having some similarities, are different. In case
of certain grades of steel, particularly sensitive to crevice corrosion, sometimes crevice corrosion can
interfere with pitting corrosion measurements. Figure 15 shows a situation where crevice corrosion
strongly interfered during the measurement of pitting corrosion by the rotating electrode technique.
The crevice corrosion developed under a defective collar, making the measurements unusable for
the characterization of pitting corrosion. Sometimes, the observation of the corrosion morphology
provides information on the metallographic structure of the alloy. The morphology of crevice corrosion
on the transverse surface (Figure 15) showed a particular structure, the orientation of the corroded
structures suggesting a preferential longitudinal dissolution. This reveals a manifestation of a higher
corrosion sensitivity of the transverse direction compared to longitudinal direction.
Figure 15. Scanning electron microscopy (SEM) of the transverse surface, corroded under a defective
PTFE collar. The columnar morphology suggests preferential longitudinal dissolution due to the texture
of the material.
μP
221
Materials 2020, 13, 4187
μP
Figure 17. Metallographic section of the tube (#3, 1.4435/316L steel) after Strauss’s test.
The Strauss test clearly and unequivocally showed that the 316L steel tubes were sensitized to
intergranular corrosion. The corrosion rate was higher on the interior compared to the exterior.
The energy-dispersive X-ray spectroscopy (EDX) analysis of the corroded areas showed the
presence of elements which did not belong to the alloy: sulfur, chlorine, calcium, sodium, aluminum
and potassium. These elements most likely resulted from lubricants formulated as additives or base oil.
The sensitization to intergranular corrosion was probably due to the pyrolysis of residual oil present
on the tube surface—in other words, poor cleaning during the manufacturing process.
(a) (b)
Figure 18. (a)#Test 1 and (b) #Test 2. Grain index (ASTM E112-13) G = 11.
222
Materials 2020, 13, 4187
According to the EPR method ASTM G108-94 (2015) [48], after the cyclic polarization scans,
the evaluation parameter is the normalized charge (Pa), measured in coulombs/cm2 , calculated with
the formula:
Pa = Q/X (1)
where Q = measured on current integration measuring instrument (coulombs), normalized for both
specimen size and grain size X = As [5.1 × 10−3 e0.35G ], where As = specimen area (cm2 ), G = grain index
at 100× according to ASTM E112-13 [49].
In the derivation of the equation, it was assumed that the Q value was the result of the attack
on the specimen surface that was distributed uniformly over the entire grain boundary region of a
constant width of 2 × (5 × 10−5 ) cm. This may not represent the actual physical processes.
The potentiokinetic electrochemical reactivation results are presented in Table 10.
223
Materials 2020, 13, 4187
B
&XUUHQW Q$FPA
7HVW
B
%UXW
B
3RWHQWLDO( P9
Figure 20. Potentio kinetic reactivation curves recorded for #Brut 1, #Test 1, #500_1, #620_2, and #750_2.
The peak valuesfor Ir , given in Table 10, were specific to the intergranular corrosion degradation
of the tubes. The higher the intensity, the greater the degradation. Thus, according to Figure 20,
it was noted that the highest sensitization of the tubes was generated by the heat treatment at 620 ◦ C.
The minimum sensitization corresponded to the 500 ◦ C heat treatment. The overall results (Table 10) for
the normalized charge (Pa) calculated (Figure 21) for all the samples indicated that the heat treatment
over 500 ◦ C for 304 steels is not indicated, the risks of inducing an intergranular corrosion process
being obvious. Consequently, the 500 ◦ C heat treatment should be used in the manufacturing process.
Type 304 steel was more sensitive to intergranular corrosion compared to other steels. Consequently,
in the manufacturing process, great importance must be given to this type of corrosion morphology.
The temperature of 620 ◦ C was critical for generating the process and therefore used in the ASTM tests
(A262-15 and G108-94) for evaluating intergranular corrosion. The goal is to near the behavior of the
tube in raw state (#Brut 1).
The susceptibility to intergranular corrosion of stainless steels is not always due to heat treatment
with precipitation of chromium carbides. Under certain conditions, the precipitation of intermetallic
compounds of (Fe, Cr)Mo2 or (Cr, Ni, Fe)3 P2 type can occur. According to Stonawská et al., the structural
sensitization of 316 L steel is due to the precipitation of secondary phases along the grain boundaries [59].
The studies of Liu et al. regarding 316 L steels [60] and Fujii et al. [61] regarding 304 steel, also supports
this statement. According to Liu et al. [62], the chromium-depleted zones near grain boundaries
represent the corrosion nucleation sites for austenitic steels. According to Eliaz, since the carbon content
in 316 stainless steel was lowered in the 316L and 316LVM grades, sensitization of this steel is less
problematic as it used to be [7].
224
Materials 2020, 13, 4187
comparison of the alloys nobility in the considered medium and to construct a galvanic series.
Higher potential values mean higher corrosion resistance. The alloys series with higher negative
potentials (anodic) generally tend to undergo greater corrosion in the event of a galvanic coupling,
while other metals (cathodic) will generally undergo a reduced attack. According to Mansfeld and
Kenkel, the corrosion potential of each alloy is a criterion in the analysis of galvanic corrosion behavior,
but it is still insufficient. The electrical potential values can only indicate a trend and state absolutely
nothing about the rate of corrosion and the type of control of the galvanic cell (mixed, cathodic or
anodic) [63,64].
Table 11. Galvanic series established in an EN1811-2011+A1:2015 artificial sweat type environment.
The most frequent cases we encountered are precious metal-austenitic steel assemblies in watch
straps. Thus millions of gold-steel links are produced to assemble straps, this is the ideal case for the
formation of a galvanic cell, a significant difference in electrical potential being involved. In the case of
gold-steel, a difference in electrical potential of around 300 mV can be calculated.
According to Gilbert and Mali, while corrosion per se may not be of great concern, when combined
with mechanical effects, restricted crevice-like geometries or any combination thereof, considerably
amplified corrosion rates might arise [65].
One of the several available techniques to realize the gold-steel assembly is brazed gold caps.
The brazed gold caps reveal a particularity due to brazing. The solder acts as the anode (a small area)
and the steel and gold parts are the cathode (large area). Thus, the corrosion process results in the
dissolution of the solder (Figures 22 and 23). In such a type of assembly it is particularly important to
make the right choice of solder.
225
Materials 2020, 13, 4187
Figure 23. Corrosion of the transverse surface, at the level of the gold-steel interface.
(a) The cathode–anode relationship. The precious metal surfaces will act as the cathode, and the
less noble parts will be the anode. Constructions with large cathode surfaces and small anode
surfaces are very dangerous. The galvanic cell will output a strong anodic current which will
lead to the rapid degradation of the anodic part by mechanisms of crevice or pitting corrosion;
(b) The nickel release in contact with the skin has to be considered as the current legislation tolerates
an amount of 0.5 μg/cm2 ·week.
Table 12 presents a series of tests carried out on the same types of gold-stainless steel links.
The 14441/316 LM steel originated from five different steelmakers from the EU, USA, Japan and China.
# Gold Steel Nickel Release (μg·cm−2 ·Week−1 ) Corroded Parts Rate Corrosion Rate (%)
#1 5N18 1.4441 0.14 0/6 0%
0.25
0.13
316LM 0.04
0.22
0.03
#2 5N18 1.4441 0.44 2/6 33%
0.05
316LM 0.01
0.10
#4 5N18 1.4441 0.03 6/9 67%
0.06
316LM
0.09
#5 5N18 1.4441 0.09 0/3 0%
0.08 - -
316LM
0.06
#6 5N18 1.4539 2.3 0/3 0%
2.6
904L
2.2
In case of #6 (gold-steel 904L), the nickel release was greater than 2 μg·cm−2 ·week−1 , despite
the absence of visible corrosion. This was due to a different behavior compared to a medical 316L
steel. The comparison of the EDX profile between a gold-904L steel (Figure 24) and a gold-316LM
steel (Figure 25) revealed a very different nickel profile with the disappearance of the nickel peak in
226
Materials 2020, 13, 4187
the gold-904L steel solder (Figure 24). In the gold-904L system, the solder was in the anodic position
(gold and 904L steel being cathodic), with a very unfavorable surface report. It revealed a selective
corrosion morphology powered by a galvanic battery; this would explain the significant nickel release
from the gold-904L steel coupling, despite the absence of visible corrosion.
(a) (b)
Figure 24. Sample #6. (a) EDX profiles of the gold-904 L steel solder for iron, nickel, silver, gold, copper
and zinc; (b) EDX Ni profile.
(a) (b)
Figure 25. Sample #4. (a) EDX profiles of the gold-316 LM steel solder for iron, nickel, silver, gold,
copper and zinc; (b) EDX Ni profile.
SEM examination of sample #4 (5N18) showed that the corrosion was localized and did not
develop at the level of the steel, but of the brazing, causing its dissolution (Figure 26). This demonstrates
that the solder represents the weak point of the gold-steel assembly.
The steels which are being used for watch straps are of grades 316 and 904L. The other steel
grades, such as 304, 304L, 316LS, and 316Ti, are not usable; their rate of Ni release does not respect
the limits imposed by the EU directives, or other countries legislations (USA, Japan, China, Korea,
Canada). The difficulty consists in eliminating the corrosion process and achieving a rate of nickel
release which respects the legislation: max 0.5 μg·cm−2 ·week−1 .
The use of a Ni-Cr-P solder involves the risk of increasing the nickel amounts by chemical
dissolution or corrosion of the solder. In the case of a gold base solder (melting range 750–850 ◦ C) the
risk is to initiate corrosion in the steel; with a Ni-Cr-P based solder (melting range 800–950 ◦ C) the risk
is to start corrosion in the solder. To find the best compromise, testing the link assemblies is necessary.
227
Materials 2020, 13, 4187
Because the quality of 316 LM steel is highly dependent on the supplier, most straps manufacturers
use steels they have exclusivity for.
+ 67521*'(&5($6(
+($775($70(17
1 6/,*+7'(&5($6(
528*+
685)$&( 32/,6+(' 6/,*+7,1)/8(1&(
6$7,1<
:25. 675$,1! ,1&5($6(
+$5'(1,1*
6758&785( ,1&/86,216$1' ,1&5($6(
6(&21'3+$6(6
Figure 27. Factors influencing the amounts of nickel released during the manufacturing process.
The heat treatments resulted in a reduction of the nickel release rates. The surface state had little
influence. On the other hand, the hardening processes strongly influenced the quantities of nickel
released. The increase in hardness greatly decreased the corrosion resistance and increased the amount
of nickel released. Another factor which strongly influenced the quantities of nickel released was
the inclusion state and the existence of secondary phases in the structure of steels. Being aware of
these causes, subcontractors demand from the steelmakers a very strict specification respecting in the
manufacturing of steels.
228
Materials 2020, 13, 4187
4. Conclusions
Seventeen grades of stainless steels were assessed for specific types of corrosion: general, pitting,
crevice, intergranular and galvanic. It was noted that there are significant differences between the
grades of the austenitic steels studied.
The conclusions are as follows:
– The intensity of the corrosion was dependent on the production parameters, such as the casting
volume, alloying additions, and deoxidizing agents;
– The amount of nickel release was dependent on the heat treatment, hardening rate, and other
parameters of the manufacturing process;
– The quantity of nickel released is strongly influenced by the inclusion state and the existence of
secondary phases;
– There is a clear difference of corrosion between the transverse surface and the longitudinal surface.
The longitudinal surface (in the rolling direction) reveals a better resistance to corrosion than the
transverse surface.
– The quantities of nickel released are highly dependent on the grade of steel. As a result,
manufacturers can use only steels that meet the current legislative requirements;
– Top range watches manufacturers use steels with exclusivity labels, so the chemical compositions,
structures, inclusive states, mechanical properties, machinability, polishing are very well defined
in their specifications. The rule also applies to medical devices manufacturers;
– Finally, a compromise in choosing a steel over another has to be made, depending on the
application and the legal requirements for the final products on a specific market.
Author Contributions: Conceptualization, L.R.; methodology, L.R.; validation, L.R.; formal analysis, L.R.;
investigation, L.R.; resources, L.R.; data curation, L.R.; writing—original draft preparation, L.R., L.C.A.;
writing—review and editing, L.C.A. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Eliaz, N. Degradation of Implant. Materials; Springer: New York, NY, USA, 2012.
2. Blackwood, D.J. Biomaterials: Past successes and future problems. Corros. Rev. 2003, 21, 97–124. [CrossRef]
3. Boonruang, C.; Thong-On, A.; Kidkhunthod, P. Effect of nanograin-boundary networks generation on
corrosion of carburized martensitic stainless steel. Sci. Rep. 2018, 8, 2289. [CrossRef]
4. Mudali, U.K.; Sridhar, T.M.; Eliaz, N.; Raj, B. Failures of stainless steel orthopaedic devices—Causesand
remedies. Corros. Rev. 2003, 21, 231–267. [CrossRef]
5. Pound, B.G. Corrosion behavior of metallic materials in biomedical applications. I. Ti and its alloys. Corros. Rev.
2014, 32, 1–20. [CrossRef]
6. Revie, R.; Uhlig, H.H. Corrosion and Corrosion Control: An Introduction to Corrosion Science and Engineering,
4th ed.; John Wiley & Sons: Hoboken, NJ, USA, 2008; p. 84.
7. Eliaz, N. Corrosion of metallic biomaterials: A review. Materials 2019, 12, 407. [CrossRef] [PubMed]
8. Tiamiyu, A.A.; Eduok, U.; Szpunar, J.A.; Odeshi, A.G. Corrosion behavior of metastable AISI 321 austenitic
stainless steel: Investigating the effect of grain size and prior plastic deformation on its degradation pattern
in saline media. Sci. Rep. 2019, 9, 12116. [CrossRef]
9. Gupta, R.K.; Birbilis, N. The influence of nanocrystalline structure and processing route on corrosion of
stainless steel: A review. Corros. Sci. 2015, 92, 1–15. [CrossRef]
10. Ralston, K.D.; Birbilis, N.; Davies, C.H.J. Revealing the relationship between grain size and corrosion rate of
metals. Scr. Mater. 2010, 63, 1201–1204. [CrossRef]
11. Schino, A.D.I.; Materiali, C.S.; Brin, V.B. Effects of grain size on the properties of a low nickel austenitic
stainless steel. J. Mater. Sci. 2003, 8, 4725–4733. [CrossRef]
229
Materials 2020, 13, 4187
12. Key to Steel-Stahlschlüssel, 25th ed.; Verlag Stahlschlüssel Wegst GmbH: Marbach, Germany, 2019. Available
online: http://www.online.stahlschluessel.de/BrowseProducts.aspx (accessed on 30 April 2020).
13. Amaro Vicente, T.; Oliveira, L.A.; Correa, E.O.; Barbosa, R.P.; Macanhan, V.B.P.; Alcântara, N.G. Stress
corrosion cracking behaviour of dissimilar welding of AISI 310S austenitic stainless steel to 2304 duplex
stainless steel. Metals 2018, 8, 195. [CrossRef]
14. Lo, K.H.; Shek, C.H.; Lai, J.K.L. Recent developments in stainless steels. Mater. Sci. Eng. R. 2009, 65, 39–104.
[CrossRef]
15. Ha, H.-Y.; Lee, T.-H.; Bae, J.-H.; Chun, D.W. Molybdenum effects on pitting corrosion resistance of
FeCrMnMoNC austenitic stainless steels. Metals 2018, 8, 653. [CrossRef]
16. Ha, H.-Y.; Jang, J.H.; Lee, T.-H.; Won, C.; Lee, C.-H.; Moon, J.; Lee, C.-G. Investigation of the localized
corrosion and passive behavior of type 304 stainless steels with 0.2–1.8 wt% B. Materials 2018, 11, 2097.
[CrossRef] [PubMed]
17. Desestret, A.; Charle, J. Les aciersinoxydablesausténico-ferritiques. In Les AciersInoxydables; Lacombe, P.,
Baroux, B., Beranger, G., Colombier, L., Hochmann, J., Eds.; Les Ulis, Les Editions de Physique: Paris, France,
1990; pp. 631–678.
18. Reclaru, L.; Ziegenhagen, R.; Eschler, P.Y.; Blatter, A.; Lemaître, J. Comparative corrosion study of “Ni-free”
austenitic stainless steels in view of medical applications. Acta. Biomater. 2006, 2, 433–444.
19. Baker, M. European standards developed in support of the european union nickel directive. In Metal Allergy;
Chen, J.K., Thyssen, J.P., Eds.; Springer: Berlin, Germany, 2018; pp. 23–29. [CrossRef]
20. Liden, C. Nickel in jewelry and associated products. Contact Dermat. 1992, 26, 73–75. [CrossRef]
21. Belsito, D.V. The diagnostic evaluation, treatment, and prevention of allergic contact dermatitis in the new
millennium. J. Allergy Clin. Immunol. 2000, 105, 409–420. [CrossRef]
22. Schafer, T.; Bohler, E.; Ruhdorfer, S.; Weigl, L.; Wessner, D.; Filipiak, B.; Wichmann, H.E.; Ring, J. Epidemiology
of contact allergy in adults. Allergy 2001, 56, 1192–1196. [CrossRef] [PubMed]
23. Publication Office of the European Union. Commission communication in the framework of the
implementation of regulation (EC) No 1907/2006 of the european parliament and of the council concerning
the registration, evaluation, authorisation and restriction of chemicals (REACH). Off. J. Eur. Union 2012,
C142, 8.
24. Publication Office of the European Union. European parliament and council directive 94/27/EC of 30 June
1994: Amending for the 12th time directive 76/769/EEC on the approximation of the laws, regulations and
administrative provisions of the member States relating to restrictions on the marketing and use of certain
dangerous substances and preparations. Off. J. Eur. Commun. 1994, L188, 1–2.
25. Heim, K.; Basketter, D. Metal exposure regulations and their effect on allergy prevention. In Metal Allergy;
Chen, J.K., Thyssen, J.P., Eds.; Springer: Berlin, Germany, 2018; pp. 39–54. [CrossRef]
26. Fontana, M. Corrosion Engineering, 5th ed.; McGraw Hill International Edition: New York, NY, USA, 1987;
p. 556.
27. ASTM B117-97. Standard Practice for Operating Salt Spray (Fog) Apparatus; ASTM International: West Conshohocken,
PA, USA, 2010. Available online: www.astm.org (accessed on 30 April 2020).
28. Bech-Nielsen, G. The anodic dissolution of iron–V. Some observations regarding the influence of cold working
and annealing on the two anodic reactions of the metal. Electrochim. Acta 1974, 19, 821–828. [CrossRef]
29. Cigada, A.; Mazza, B.; Pedeferri, P.; Sinigaglia, D. Influence of cold plastic deformation on critical pitting
potential of AISI 316 L and 304 L steels in an artificial physiological solution simulating the aggressiveness of
the human body. J. Biomed. Mater. Res. 1977, 11, 503–512. [CrossRef] [PubMed]
30. ISO 3160-2:2015. Watch-Cases and Accessories—Gold Alloy Coverings—Part 2: Determination of Fineness,
Thickness, Corrosion Resistance and Adhesion. Available online: https://www.iso.org/standard/66162.html
(accessed on 30 April 2020).
31. ASTM G48-11. Standard Test Methods for Pitting and Crevice Corrosion Resistance of Stainless Steels and Related
Alloys by Use of Ferric Chloride Solution; ASTM International: West Conshohocken, PA, USA, 2015. Available
online: www.astm.org (accessed on 30 April 2020).
32. Hoar, T.P.; Mears, D.C. Corrosion-resistant alloys in chloride solutions: Materials for surgical implants.
Proc. R. Soc. Lond. 1966, 294, 486–510.
33. Zabel, D.D.; Brown, S.A.; Merritt, K.; Payer, J.H. AES analysis of stainless steel corroded in saline, in serum
and in vivo. J. Biomed. Mater. Res. 1988, 22, 31–44. [CrossRef] [PubMed]
230
Materials 2020, 13, 4187
34. Turnbull, A. The solution composition and electrode potential in pits, crevices and cracks. Corros. Sci. 1983,
23, 833–870. [CrossRef]
35. ASTM F746-87. Standard Test Method for Pitting or Crevice Corrosion of Metallic Surgical Implant Materials;
ASTM International: West Conshohocken, PA, USA, 1999. Available online: www.astm.org (accessed on
30 April 2020).
36. ASTM A262-15. Standard Practices for Detecting Susceptibility to Intergranular Attack in Austenitic Stainless
Steels; ASTM International: West Conshohocken, PA, USA, 2015. Available online: www.astm.org (accessed
on 2 May 2020).
37. Streicher, M.A. Intergranular. In Corrosion Tests and Standards: Application and Interpretation, 2nd ed.;
Baboian, R., Ed.; ASTM International: West Conshohocken, PA, USA, 2004; pp. 244–266.
38. Henthorne, M. Localized Corrosion. ASTM Tech. Publ. 1972, 518, 108.
39. Cihal, V.; Stefec, R. On the development of the electrochemical potentiokinetic method. Electrochim. Acta
2001, 46, 3867–3877. [CrossRef]
40. Iacoviello, F.; Di Cocco, V.; D’Agostino, L. Analysis of the intergranular corrosion susceptibility in stainless
steel by means of potentiostatic reactivation tests. Procedia Struct. Integr. 2017, 3, 269–275. [CrossRef]
41. Clarke, W.L.; Romero, V.M.; Danko, J.C. Detection of sensitization in stainless steel using electrochemical
techniques. In Corrosion 77, International Corrosion Forum; National Association of Corrosion Engineers:
San Francisco, CA, USA, 1977; p. 180.
42. Novak, P.; Stefec, R.; Franz, F. Testing the susceptibility of stainless steels to intergranular corrosion by a
reactivation method. Corrosion 1975, 31, 344–347. [CrossRef]
43. Desestret, A.; Froment, M.; Guiraldenq, P. Intergranular corrosion of austenitic stainless steels in the
sensitized or solution-annealed condition. In Proceedings of the 23rd Meeting of I.S.E., Stockholm, Sweden,
27 August–2 September 1972; Volume 87.
44. Aydoğdu, G.H.; Aydinol, K. Determination of susceptibility to intergranular corrosion and electrochemical
reactivation behaviour of AISI 316L type stainless steel. Corr. Sci. 2006, 48, 3565–3583. [CrossRef]
45. Cíhal, V.; Lasek, S.; Blahetová, M.; Kalabisová, E.; Krhutová, Z. Trends in the electrochemical polarization
potentiodynamic reactivation method-EPR. Chem. Biochem. Eng. Q. 2007, 21, 47–54.
46. Parvathavarthini, N.; Dayal, R.K.; Seshadri, S.K.; Gnanamoorthy, J.B. Influence of prior deformation on the
sensitisation of AISI type 304 stainless steel and applicability of EPR technique. Br. Corros. J. 1991, 26, 67–76.
[CrossRef]
47. Majidi, A.P.; Streicher, M.A. The double loop reactivation method for detecting sensitization in AISI 304
stainless steels. Corrosion 1984, 40, 584–593. [CrossRef]
48. ASTM G108-94. Standard Test Method for Electrochemical Reactivation (EPR) for Detecting Sensitization of AISI
Type 304 and 304L Stainless Steels; ASTM International: West Conshohocken, PA, USA, 2015; Available online:
www.astm.org (accessed on 2 May 2020).
49. ASTM E112-13. Standard Test Methods for Determining Average Grain Size; ASTM International:
West Conshohocken, PA, USA, 2013. Available online: www.astm.org (accessed on 6 May 2020).
50. EN 1811-2011+A1:2015, Reference Test Method for Release of Nickel From All Post Assemblies which are
Inserted Into Pierced Parts of the Human Body and Articles Intended to Come Into Direct or Prolonged Contact
With the Skin, CEN/TC347. 2011. Available online: https://www.en-standard.eu/din-en-1811 (accessed on
10 October 2018).
51. Zanotto, F.; Grassi, V.; Balbo, A.; Monticelli, C.; Zucchi, F. Resistance of thermally aged DSS 2304 against
localized corrosion attack. Metals 2018, 8, 1022. [CrossRef]
52. ASTM G46-94. Standard Guide for Examination and Evaluation of Pitting Corrosion; ASTM International:
West Conshohocken, PA, USA, 2018. Available online: www.astm.org (accessed on 6 May 2020).
53. Virtanen, S. Degradation of titanium and its alloys. In Degradation of Implant Materials; Eliaz, N., Ed.; Springer:
New York, NY, USA, 2012; pp. 29–55.
54. Melchers, R.E. A review of trends for corrosion loss and pit depth in longer-term exposures. Corros. Mater.
Degrad. 2020, 1, 4. [CrossRef]
55. Abbasi Aghuy, A.; Zakeri, M.; Moayed, M.H.; Mazinani, M. Effect of grain size on pitting corrosion of 304L
austenitic stainless steel. Corros. Sci. 2015, 94, 368–376. [CrossRef]
231
Materials 2020, 13, 4187
56. Poyet, P.; Desestret, A.; Coriou, H.; Grall, L. Contribution à L’étude de la Corrosion par Piqûres des Aciers
Inoxydables 18/10. Influence de Divers Facteurs sur le Processus D’amorçage des Piqûres; Société française de
métallurgie. Journées d’automne: Paris, France, 1973.
57. Bryant, M.; Hu, X.; Farrar, R.; Brummitt, K.; Freeman, R.; Neville, A. Crevicecorrosion of biomedical
alloys: A novel method of assessing the effects of bone cement and its chemistry. J. Biomed. Mater. Res. B
Appl. Biomater. 2013, 101, 792–803. [CrossRef]
58. Liu, Y.; Zhu, D.; Pierre, D.; Gilbert, J.L. Fretting initiated crevicecorrosion of 316LVM stainless steel in
physiological phosphate buffered saline: Potential and cycles to initiation. Acta Biomater. 2019, 97, 565–577.
[CrossRef]
59. Stonawská, Z.; Svoboda, M.; Sozańska, M.; Krístková, M.; Sojka, J.; Dagbert, C.; Hyspecká, L. Structural
analysis and intergranularcorrosion tests of AISI 316L steel. J. Microsc. 2006, 224, 62–64. [CrossRef]
60. Liu, T.; Xia, S.; Bai, Q.; Zhou, B.; Lu, Y.; Shoji, T. Evaluation of grain boundary network and improvement of
intergranular cracking resistance in 316L stainless steel after grain boundary engineering. Materials 2019,
12, 242. [CrossRef]
61. Fujii, T.; Furumoto, T.; Tohgo, K.; Shimamura, Y. Crystallographic evaluation of susceptibility to intergranular
corrosion in austenitic stainless steel with various degrees of sensitization. Materials 2020, 13, 613. [CrossRef]
62. Liu, G.; Liu, Y.; Cheng, Y.; Li, J.; Jiang, Y. The intergranular corrosion susceptibility of metastable austenitic
Cr-Mn-Ni-N-Cu high-strength stainless steel under various heat treatments. Materials 2019, 12, 1385.
[CrossRef] [PubMed]
63. ASTM G82-98. Standard Guide for Development and Use of a Galvanic Series for Predicting Galvanic Corrosion
Performance; ASTM International: West Conshohocken, PA, USA, 2014. Available online: www.astm.org
(accessed on 6 May 2020).
64. Mansfeld, F.; Kenkel, J.V. Laboratory studies of galvanic corrosion of aluminium alloys. In Galvanic and
Pitting Corrosion–Field and Laboratory Studies; Baboian, R., France, W.D., Eds.; ASTM: West Conshohocken,
PA, USA, 1976; pp. 20–47.
65. Gilbert, J.L.; Mali, S. Medical implant corrosion: Electrochemistry at metallic biomaterial surfaces.
In Degradation of Implant Materials; Eliaz, N., Ed.; Springer: New York, NY, USA, 2012; Chapter 1, pp. 1–28.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
232
materials
Article
Effect of Cr on Aqueous and Atmospheric Corrosion of
Automotive Carbon Steel
Sang-won Cho 1 , Sang-Jin Ko 1 , Jin-Seok Yoo 1 , Yun-Ha Yoo 2 , Yon-Kyun Song 2 and Jung-Gu Kim 1, *
Abstract: This study investigated the effect of Cr alloying element on the corrosion properties of
automotive carbon steel (0.1C, 0.5Si, 2.5Mn, Fe Bal., composition given in wt.%) in aqueous and
atmospheric conditions using electrochemical measurement and cyclic corrosion tests. Three steels
with 0, 0.3, and 0.5 wt.% Cr were studied by electrochemical impedance spectroscopy. Polarization
resistance (Rp ) of 0.3 Cr and 0.5 Cr steels was higher than that of 0 Cr steel, and the Rp also increased
as the Cr content increased. Therefore, Cr increases the corrosion resistance of automotive carbon
steel immersed in a chloride ion (Cl− )-containing aqueous solution. In the cyclic corrosion test results,
Cl− was concentrated at the metal/rust interface in all of the steels regardless of Cr content. The
Cl− was uniformly concentrated and distributed on the 0 Cr steel, but locally and non-uniformly
concentrated on the Cr-added steels. The inner rust layer consisted of β-FeOOH containing Cl− and
Cr-goethite, while the outer rust layer was composed of amorphous iron oxyhydroxide mixed with
various types of rust. FeCl2 and CrCl3 are formed from the Cl− nest developed in the early stage,
Citation: Cho, S.-w.; Ko, S.-J.; Yoo, and the pitting at CrCl3 -formed regions are locally accelerated because Cr is strongly hydrolyzed to a
J.-S.; Yoo, Y.-H.; Song, Y.-K.; Kim, J.-G. very low pH.
Effect of Cr on Aqueous and
Atmospheric Corrosion of Keywords: automotive steel; atmospheric corrosion; electrochemical impedance spectroscopy; cyclic
Automotive Carbon Steel. Materials corrosion test; iron oxide
2021, 14, 2444. https://doi.org/
10.3390/ma14092444
233
Steels Cr C Si Mn Fe
0 Cr 0.01 0.10 0.52 2.49 Bal.
0.3 Cr 0.32 0.10 0.52 2.49 Bal.
0.5 Cr 0.50 0.10 0.52 2.49 Bal.
Figure 1. SEM images of the microstructure of (a) 0 Cr, (b) 0.3 Cr, and (c) 0.5 Cr, etched with 2% nital solution.
234
Materials 2021, 14, 2444
The specimens were cut to a size of 1.5 cm × 1.5 cm, polished with 600-grit SiC paper,
and cleaned with distilled water. Additionally, the solution for electrochemical measure-
ments was 3.5 wt.% NaCl. All of the electrochemical measurements were conducted in
a 3-electrode electrochemical cell. The test specimen was used as the working electrode,
a carbon rod was used as the counter electrode, and a saturated calomel electrode was
used as the reference electrode. Potentiodynamic polarization tests were performed with a
potential sweep of 0.166 mV/s according to ASTM G5. To establish a stable potential, the
scan was initiated after the specimen was stabilized in the solution [10]. Electrochemical
impedance spectroscopy (EIS) tests were performed with an amplitude of 10 mV in the
frequency range of 100 kHz to 10 mHz. Electrochemical tests were conducted by a poten-
tiostat (BioLogics, VMP-2, Seyssinet-Pariset, France). After the CCT, the test specimens
were mounted with epoxy and analyzed by an optical microscope (OM), and the compo-
nents of the corrosion product were analyzed by an electron probe micro-analyzer (EPMA;
JEOL, JXA-8530F, Fukuoka, Japan), X-ray diffraction (XRD; Rigaku, D/max-2500V/PC,
Tokyo, Japan), and transmission electron microscopy (TEM; FEI, Tecnai F20 G2, Hillsboro,
OR, USA).
The specific CCT process is shown in Figure 2. The specimens used for the CCT were
cut to a size of 3 cm × 7 cm, and exposed on only one side to the corrosive environment.
The CCT was performed for 10, 20, and 30 cycles, respectively. The salt solution used for
the CCT was 5 wt.% NaCl. The length of a CCT cycle was 24 h, consisting of a wet stage
for 21 h and a dry stage at 30% of relative humidity and 50 ◦ C for 3 h.
235
Materials 2021, 14, 2444
0.0
0 Cr
-0.2 0.3 Cr
0.5 Cr
-0.4
VSCE (V)
-0.6
-0.8
-1.0
-1.2
-8 -7 -6 -5 -4 -3 -2 -1 0 1
10 10 10 10 10 10 10 10 10 10
2
log i (A/cm )
Figure 3. Potentiodynamic polarization curves in 3.5 wt.% NaCl solution.
236
Materials 2021, 14, 2444
Figure 4. Nyquist and Bode impedance plots of EIS data of (a,b) 0 Cr steel, (c,d) 0.3 Cr steel, and (e,f) 0.5 Cr steel in 3.5 wt.%
NaCl solution.
To determine the optimized values for the resistance and capacitance parameters,
the equivalent circuit was used as shown in Figure 5. Rs is the test solution resistance,
Rfilm is the oxide film resistance, Rct is the charge transfer resistance, and Rfilm + Rct is
total resistance or polarization resistance (Rp ), which is proportional to the radius of the
capacitive loop in the Nyquist plot. The constant phase element (CPE) is the capacitive
response of the system. CPE1 is the capacitive response of the oxide film, and CPE2 is
237
Materials 2021, 14, 2444
the capacitive response of the double layer caused by the dissolution of the metal and the
charge separation between the metal/electrolyte interface [16–19]. In the equivalent circuit,
CPE is defined as below:
ZCPE = Q0−1 (jω)−n (1)
where Z is the impedance, Q0 is the coefficient of proportionality, j is the imaginary number,
ω is the angular frequency, and n is the empirical CPE exponent (0 ≤ n ≤ 1) measuring
the deviation from the behavior of an ideal electric capacity [20,21]. CPE can represent
resistance (n = 0), capacitance (n = 1), inductance (n = −1), or Warburg impedance (n = 0.5)
in accordance with n [22].
The EIS data were fitted using the ZSimpWin (Princeton Applied Research, Oak Ridge,
TN, USA) program and the results are shown in Table 3. The Rp of 0.3 Cr and 0.5 Cr steels
was higher than that of 0 Cr steel, and the Rp also increased as the Cr content increased.
This indicates that the corrosion resistance is increased as the Cr content increases. This
is believed to be due to the bigger grain size of the steel with higher Cr content. Metal
with active polarization behavior decreases the corrosion rate with bigger grain size [23].
Therefore, Cr improved the corrosion resistance of the ACS that was immersed in the
Cl- -containing aqueous solution.
CPE1 CPE2
Immersion Rs Qfilm Rfilm Qct Rct Rp
Steel (Ω−1 (Ω−1 ·
Time (Ω·cm−2 ) n1 (Ω·cm−2 ) n1 (Ω·cm−2 ) (Ω·cm−2 )
cm− 2 ·sn ) cm− 2 ·sn )
0h 1.424 7.52 × 10− 4 0.8355 29.9 3.97 × 10− 4 0.9671 636.5 665.7
1h 2.531 6.11 × 10− 4 0.8146 192.6 1.29 × 10− 4 0.9516 951.6 1144.2
2h 2.522 6.36 × 10− 4 0.8037 294 1.48 × 10− 4 0.9883 801.2 1095.2
0 Cr 3h 2.542 6.65 × 10− 4 0.801 239.9 1.55 × 10− 4 0.9724 831.3 1071.2
4h 2.549 6.84 × 10− 4 0.796 267.4 1.61 × 10− 4 0.997 752.1 1019.5
5h 2.574 6.72 × 10− 4 0.7959 217.5 1.73 × 10− 4 0.9593 766.3 983.8
6h 2.589 6.81 × 10− 4 0.7953 203.7 1.85 × 10− 4 0.9509 733.1 936.8
0h 1.773 1.71 × 10− 4 1 3 9.06 × 10− 4 0.7621 1304 1307
1h 1.227 2.05 × 10− 4 0.9653 32.3 3.76 × 10− 4 0.7516 1210 1242.3
2h 1.22 1.92 × 10− 4 0.9714 29.2 3.79 × 10− 4 0.7541 1151 1180.2
0.3 Cr 3h 1.212 1.83 × 10− 4 0.9759 26.5 3.90 × 10− 4 0.7497 1215 1241.5
4h 1.212 1.69 × 10− 4 0.9843 24.3 4.00 × 10− 4 0.7475 1223 1247.3
5h 1.217 1.59 × 10− 4 0.9914 23.7 4.10 × 10− 4 0.7442 1266 1289.7
6h 1.215 1.57 × 10− 4 0.993 23.1 4.08 × 10− 4 0.7424 1285 1308.1
238
Materials 2021, 14, 2444
Table 3. Cont.
CPE1 CPE2
Immersion Rs Qfilm Rfilm Qct Rct Rp
Steel (Ω−1 (Ω−1 ·
Time (Ω·cm−2 ) n1 (Ω·cm−2 ) n1 (Ω·cm−2 ) (Ω·cm−2 )
cm− 2 ·sn ) cm− 2 ·sn )
0h 2.809 4.71 × 10− 4 0.8449 21.6 2.14 × 10− 4 0.8069 1414 1435.6
1h 1.986 5.09 × 10− 4 0.8474 59 1.63 × 10− 4 0.7983 1543 1602
2h 1.978 5.18 × 10− 4 0.8405 63.5 1.67 × 10− 4 0.7818 1419 1482.5
0.5 Cr 3h 1.964 3.86 × 10− 4 0.8637 20.6 2.95 × 10− 4 0.7261 1270 1290.6
4h 1.963 3.24 × 10− 4 0.8782 14.9 3.39 × 10− 4 0.7095 1489 1503.9
5h 2.001 1.18 × 10− 4 0.9798 4.5 5.31 × 10− 4 0.7418 1502 1506.5
6h 2.013 9.42 × 10− 4 0.9999 3.6 5.44 × 10− 4 0.748 1542 1545.6
(a)
(b)
(c)
Figure 6. Cross-section OM images of (a) 0 Cr steel, (b) 0.3 Cr steel, and (c) 0.5 Cr steel after CCT.
To determine the localized corrosion tendency, the pitting factor (PF) with a concept
similar to that given in ASTM G46 was used. A PF value of 1 means perfect uniform
corrosion, and a higher PF means an increased localized corrosion tendency. The PF
239
Materials 2021, 14, 2444
for each CCT cycle was derived by the following equation, and the variation of the PF
according to CCT cycle is shown in Figure 7.
p
PF = (2)
d
where p is the maximum penetration depth, and d is the average penetration depth.
7
0 Cr
0.3 Cr
6
0.5 Cr
5
Pitting factor
1
10 20 30
Cycle
In the case of 0 Cr steel, the PF was approximately 2 regardless of the cycle, while the
PF of 0.3 Cr and 0.5 Cr steels changed depending on the cycle. In all cycles, the PF of the
Cr-added steels was higher than that of the 0 Cr steel, but the PF was not proportional to
Cr content. This indicates that the Cr alloying element can accelerate localized corrosion,
and the presence or absence of Cr greatly affects the localized corrosion, not the Cr content.
The cross-section of the specimen after 10 and 30 cycles was analyzed to determine the
chemical composition using EPMA, and the results are shown in Figure 8. The rust layer of
0 Cr steel was composed entirely of porous iron oxide (e.g., γ-FeOOH, γ-Fe2 O3 , Fe3 O4 ).
In addition, Cl− was accumulated at the metal/rust interface and on the inner layer with
uniform concentration and distribution. The Cr-added steels had a very dense and uniform
Cr-enriched region in the inner rust layer, while the outer rust layer was composed of
porous iron oxide, like 0 Cr. Cl− was detected underneath the Cr-enriched layer and at the
metal/rust interface, but unlike 0 Cr steel, it was localized and non-uniformly concentrated.
The rust layer of the 30-cycle steel was exfoliated from the metal, and the Cl− concentration
in the inner rust layer was increased significantly compared to 10 cycles. Therefore, it is
considered that the corrosion is accelerated because the protective oxide layer loses its
protective property after the rust layer exfoliates.
To summarize the above results, Cl− was concentrated at the metal/rust interface
in all of the specimens regardless of Cr content. Generally, since the localized corrosion
in an atmospheric environment is caused by Cl− enrichment [4,24], localized corrosion
with a PF of approximately 2 or higher occurred in all of the steels, as shown in Figure 7.
However, 0.3 Cr and 0.5 Cr steels had higher PFs than 0 Cr steel because Cl- was localized
and non-uniformly concentrated as compared with 0 Cr steel.
240
Materials 2021, 14, 2444
Figure 8. EPMA analysis of (a) 0 Cr steel, (b) 0.3 Cr steel, and (c) 0.5 Cr steel after CCT.
241
Materials 2021, 14, 2444
The EDS results, TEM images, and diffraction patterns of the inner and outer rust
formed on the 0.5 Cr steel were analyzed, and the results are shown in Figure 10. Chlorine
was observed in the inner rust particle as an acicular single crystal with a size of about
100 nm, as shown in Figure 10a. This is the major characteristic of akaganeite [27]. In
Figure 10b, Cr and Cl− were observed together in the inner rust particle. The particle was
polycrystalline and was a spherical agglomeration with a size of several nanometers. As
the spherical-shaped rust is the main feature of goethite [28], the particle is Cr-containing
nanoscale goethite (Cr-goethite). Since dissolved or enriched Cr suppresses the growth
of goethite crystals [7], the size of the Cr-goethite particles is very small. Cr-goethite is so
small in size that it acts as a protective film that is densely formed in the inner rust layer.
Furthermore, Cr-goethite has cation selectivity so it can inhibit the penetration of aggressive
anions such as Cl− and SO4 2− and improve corrosion resistance [7,25,29–31]. In short,
Cr-goethite was formed in the inner rust layer of Cr-added steels, which blocked the inflow
of additional Cl− from the outside and consequently improved the corrosion resistance.
As shown in Figure 10c, Cl− and Cr were not detected in the outer rust particle. Therefore,
the outer rust layer is composed of various rusts such as lepidocrocite, maghemite, and
magnetite detected from the XRD analysis results. In summary, the inner rust layer consists
of akaganeite containing Cl− and Cr-goethite, while the outer rust layer is composed of
amorphous iron oxyhydroxide mixed with various types of rust.
242
Materials 2021, 14, 2444
Figure 10. TEM images and the diffraction patterns of (a,b) inner rust and (c) outer rust formed on the 0.5 Cr steel.
243
Materials 2021, 14, 2444
Figure 11. Schematic diagram of the mechanism of localized corrosion of Cr-added steel under wet/dry conditions.
Corrosion of steel begins in areas where the inherent oxide film is weak. During the
wet stage, Cl− ions existing in the aqueous adsorption layer move to these weak areas, and
a Cl-concentrated region (nest) is formed [32]. The Cl− is adsorbed on the steel surface,
and then atmospheric corrosion initiates. Thereafter, Fe2+ reacts with H2 O to form Fe(OH)2 ,
and with salt or Cl− in the air to form FeCl2 .
244
Materials 2021, 14, 2444
after the formation of Cr-goethite, the inflow of extra Cl− from the outside is blocked, so
that Cl− is locally accumulated underneath the Cr-enriched layer.
During the wet stage, Cl− ions in the akaganeite formed in the inner rust layer are
dissolved and eluted in water, resulting in the formation of FeCl2 and CrCl3 . The hydrolysis
reactions of these Fe and Cr salts occur.
4. Conclusions
In this study, the effect of Cr alloying element on the corrosion properties of ACS in
aqueous and atmospheric conditions was investigated using electrochemical measurements
and a CCT. The conclusions based on the investigations are as follows:
• In the electrochemical measurement results, the Cr alloying element improves the cor-
rosion resistance of the ACS that was immersed in the Cl-containing aqueous solution.
• Cl is concentrated at the metal/rust interface in all of the specimens regardless of Cr
content after the CCT. The Cl is uniformly concentrated and distributed on the 0 Cr
steel, whereas Cl is localized and non-uniformly concentrated on the Cr-added steels.
The PF of the Cr-added steels is higher than that of the 0 Cr steel during the CCT.
• The inner rust layer consists of Cl-containing akaganeite and Cr-goethite, while the
outer rust layer is composed of amorphous iron oxyhydroxide mixed with various
types of rust.
• FeCl2 and CrCl3 are formed from the Cl nest developed in the early stage, and the
pitting at CrCl3 -formed regions is locally accelerated because Cr is strongly hydrolyzed
to a very low pH.
Author Contributions: Conceptualization, S.-w.C. and Y.-H.Y.; Data curation, S.-w.C.; Formal anal-
ysis, S.-w.C.; Funding acquisition, J.-G.K.; Investigation, S.-w.C., S.-J.K., and J.-G.K.; Methodology,
J.-G.K.; Project administration, Y.-H.Y., Y.-K.S., and J.-G.K.; Resources, Y.-H.Y. and Y.-K.S.; Software,
S.-J.K. and J.-S.Y.; Validation, Y.-H.Y. and J.-G.K.; Visualization, S.-w.C.; Writing—original draft,
S.-w.C.; Writing—review and editing, S.-w.C., S.-J.K., J.-S.Y., Y.-H.Y., Y.-K.S., and J.-G.K. All authors
have read and agreed to the published version of the manuscript.
Funding: This research was funded by POSCO, grant number 2018Z098.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Lee, C.; Kang, B.; Choi, B.-H.; Lee, J.; Lee, K. Observation and characterization of squeak noises of polymeric materials for
automotive interior parts under field-degradation. Trans. KSAE 2017, 25, 257–265. [CrossRef]
2. Adikari, A.; Munasinghe, R.D.S.; Jayatileke, S. Prediction of atmospheric corrosion—A Review. Engineer 2014, 47, 75–83.
[CrossRef]
245
Materials 2021, 14, 2444
3. Alcántara, J.; Chico, B.; Simancas, J.; Díaz, I.; Morcillo, M. Marine atmospheric corrosion of carbon steel: A review. Materials 2017,
10, 406. [CrossRef]
4. Kamimura, T.; Stratmann, M. The influence of chromium on the atmospheric corrosion of steel. Corros. Sci. 2001, 43, 429–447.
[CrossRef]
5. Asami, K.; Kikuchi, M. Characterization of rust layers on weathering steels air-exposed for a long period. Mater. Trans. 2002, 43,
2818–2825. [CrossRef]
6. Yamashita, M.; Shimizu, T.; Konishi, H.; Mizuki, J.; Uchida, H. Structure and protective performance of atmospheric corrosion
product of Fe–Cr alloy film analyzed by Mössbauer spectroscopy and with synchrotron radiation X-rays. Corros. Sci. 2003, 45,
381–394. [CrossRef]
7. Yamashita, M.; Miyuki, H.; Matsuda, Y.; Nagano, H.; Misawa, T. The long term growth of the protective rust layer formed on
weathering steel by atmospheric corrosion during a quarter of a century. Corros. Sci. 1994, 36, 283–299. [CrossRef]
8. Zhao, Q.-H.; Liu, W.; Zhu, Y.-C.; Zhang, B.-L.; Li, S.-Z.; Lu, M.-X. Effect of small content of chromium on wet–dry acid corrosion
behavior of low alloy steel. Acta Metall. Sin-Engl. 2017, 30, 164–175. [CrossRef]
9. Park, S.-A.; Kim, J.-G.; Lee, B.-H.; Yoon, J.-B. Development of sulfuric and hydrochloric acid dew-point corrosion-resistant steels:
1. Effect of alloying elements on the corrosion resistance of low-alloy steels. Korean J. Met. Mater. 2014, 52, 837–855.
10. Kim, S.-H.; Lee, J.-H.; Kim, J.-G.; Kim, W.-C. Effect of the crevice former on the corrosion behavior of 316L stainless steel in
chloride-containing synthetic tap water. Met. Mater. Int. 2018, 24, 516–524. [CrossRef]
11. Cheng, Q.; Chen, Z. The cause analysis of the incomplete semi-circle observed in high frequency region of EIS obtained from
TEL-covered pure copper. Int. J. Electrochem. Sci 2013, 8, 8282–8290.
12. Keddam, M.; Mottos, O.R.; Takenouti, H. Reaction model for iron dissolution studied by electrode impedance: I. Experimental
results and reaction model. J. Electrochem. Soc. 1981, 128, 257–266. [CrossRef]
13. Liu, W.; Dou, J.; Lu, S.; Zhang, P.; Zhao, Q. Effect of silty sand in formation water on CO2 corrosion behavior of carbon steel. Appl.
Surf. Sci. 2016, 367, 438–448. [CrossRef]
14. Zeng, L.; Zhang, G.; Guo, X. Erosion–corrosion at different locations of X65 carbon steel elbow. Corros. Sci. 2014, 85, 318–330.
[CrossRef]
15. Srinivasan, A.; Blawert, C.; Huang, Y.; Mendis, C.; Kainer, K.; Hort, N. Corrosion behavior of Mg–Gd–Zn based alloys in aqueous
NaCl solution. J. Magnes. Alloy 2014, 2, 245–256. [CrossRef]
16. Kim, M.; Jang, S.; Woo, S.; Kim, J.; Kim, Y. Corrosion resistance of ferritic stainless steel in exhaust condensed water containing
aluminum cations. Corrosion 2015, 71, 285–291. [CrossRef]
17. Cho, S.; An, J.-H.; Lee, S.-H.; Kim, J.-G. Effect of pH on the passive film characteristics of lean duplex stainless steel in chloride-
containing synthetic tap water. Int. J. Electrochem. Sci 2020, 15, 4406–4420. [CrossRef]
18. Chen, Y.; Hong, T.; Gopal, M.; Jepson, W. EIS studies of a corrosion inhibitor behavior under multiphase flow conditions. Corros.
Sci. 2000, 42, 979–990. [CrossRef]
19. Hamdy, A.S.; El-Shenawy, E.; El-Bitar, T. Electrochemical impedance spectroscopy study of the corrosion behavior of some
niobium bearing stainless steels in 3.5% NaCl. Int. J. Electrochem. Sci. 2006, 1, 171–180.
20. Bentiss, F.; Lebrini, M.; Vezin, H.; Chai, F.; Traisnel, M.; Lagrené, M. Enhanced corrosion resistance of carbon steel in normal
sulfuric acid medium by some macrocyclic polyether compounds containing a 1,3,4-thiadiazole moiety: AC impedance and
computational studies. Corros. Sci. 2009, 51, 2165–2173. [CrossRef]
21. Lopez, D.A.; Simison, S.; De Sanchez, S. The influence of steel microstructure on CO2 corrosion. EIS studies on the inhibition
efficiency of benzimidazole. Electrochim. Acta 2003, 48, 845–854. [CrossRef]
22. Mansfeld, F. Recording and analysis of AC impedance data for corrosion studies. Corrosion 1981, 37, 301–307. [CrossRef]
23. Ralston, K.D.; Birbilis, N.; Davies, C.H.J. Revealing the relationship between grain size and corrosion rate of metals. Scripta Mater.
2010, 63.12, 1201–1204. [CrossRef]
24. Xiao, H.; Ye, W.; Song, X.; Ma, Y.; Li, Y. Evolution of akaganeite in rust layers formed on steel submitted to wet/dry cyclic tests.
Materials 2017, 10, 1262–1275. [CrossRef] [PubMed]
25. Asami, K.; Kikuchi, M. In-depth distribution of rusts on a plain carbon steel and weathering steels exposed to coastal–industrial
atmosphere for 17 years. Corros. Sci. 2003, 45, 2671–2688. [CrossRef]
26. Alcántara, J.; Chico, B.; Díaz, I.; De la Fuente, D.; Morcillo, M. Airborne chloride deposit and its effect on marine atmospheric
corrosion of mild steel. Corros. Sci. 2015, 97, 74–88. [CrossRef]
27. Senthilnathan, A.; Dissanayake, D.; Chandrakumara, G.; Mantilaka, M.; Rajapakse, R.; Pitawala, H.; Nalin de Silva, K. Akaganeite
nanorices deposited muscovite mica surfaces as sunlight active green photocatalyst. R. Soc. Open Sci. 2019, 6, 1–12. [CrossRef]
[PubMed]
28. Verma, S.; Baig, R.N.; Nadagouda, M.N.; Varma, R.S. Oxidative CH activation of amines using protuberant lychee-like goethite.
Sci. Rep. 2018, 8, 1–7. [CrossRef] [PubMed]
29. Kimura, M.; Kihira, H. Nanoscopic mechanism of protective-rusts formation on weathering steel surfaces. SHINNITTETSU GIHO
2004, 91, 77–81.
30. Xu, Q.; Gao, K.; Lv, W.; Pang, X. Effects of alloyed Cr and Cu on the corrosion behavior of low-alloy steel in a simulated
groundwater solution. Corros. Sci. 2016, 102, 114–124. [CrossRef]
246
Materials 2021, 14, 2444
31. Melchers, R.E. Effect of small compositional changes on marine immersion corrosion of low alloy steels. Corros. Sci. 2004, 46,
1669–1691. [CrossRef]
32. Henriksen, J. The distribution of NaCl on Fe during atmospheric corrosion. Corros. Sci. 1969, 9, 573–576. [CrossRef]
33. Ma, Y.; Li, Y.; Wang, F. Corrosion of low carbon steel in atmospheric environments of different chloride content. Corros. Sci. 2009,
51, 997–1006. [CrossRef]
34. Ma, Y.; Li, Y.; Wang, F. The effect of β-FeOOH on the corrosion behavior of low carbon steel exposed in tropic marine environment.
Mater. Chem. Phys. 2008, 112, 844–852. [CrossRef]
35. Misawa, T.; Hashimoto, K.; Shimodaira, S. The mechanism of formation of iron oxide and oxyhydroxides in aqueous solutions at
room temperature. Corros. Sci. 1974, 14, 131–149. [CrossRef]
36. Yang, W.; Ni, R.-C.; Hua, H.-Z.; Pourbaix, A. The behavior of chromium and molybdenum in the propagation process of localized
corrosion of steels. Corros. Sci. 1984, 24, 691–707. [CrossRef]
37. SA, P.; DP, L. Alloying effect of chromium on the corrosion behavior of low-alloy steels. Meter. Trans. 2013, 54, 1770–1778.
38. Jones, D.A. Principles and Prevention of Corrosion, 2nd ed.; Prentice Hall, Inc.: Upper Saddle River, NJ, USA, 1996; p. 217.
247
materials
Article
Effect of Grain Size on the Corrosion Behavior of
Fe-3wt.%Si-1wt.%Al Electrical Steels in Pure Water Saturated
with CO2
Gaetano Palumbo 1, *, Dawid Dunikowski 1 , Roma Wirecka 2,3 , Tomasz Mazur 2 , Urszula Lelek-Borkowska 1 ,
Kinga Wawer 4 and Jacek Banaś 1
1 Faculty of Foundry Engineering, Department of Chemistry and Corrosion of Metals, AGH University of
Science and Technology, Mickiewicza St. 30, 30-059 Krakow, Poland; [email protected] (D.D.);
[email protected] (U.L.-B.); [email protected] (J.B.)
2 Academic Centre for Materials and Nanotechnology, AGH University of Science and Technology,
Mickiewicza St. 30, 30-059 Kraków, Poland; [email protected] (R.W.); [email protected] (T.M.)
3 Department of Condensed Matter Physics, Faculty of Physics and Applied Computer Science,
AGH University of Science and Technology, Mickiewicza St. 30, 30-059 Krakow, Poland
4 Łukasiewicz Research Network–Institute of Aviation, Al. Krakowska 110/114, 02-256 Warsaw, Poland;
[email protected]
* Correspondence: [email protected]; Tel.: +48-12-888-27-63
Abstract: The corrosion behavior of two silicon steels with the same chemical composition but
different grains sizes (i.e., average grain area of 115.6 and 4265.9 μm2 ) was investigated by metallo-
graphic microscope, gravimetric, electrochemical and surface analysis techniques. The gravimetric
and electrochemical results showed that the corrosion rate increased with decreasing the grain size.
Citation: Palumbo, G.; The scanning electron microscopy/energy dispersive x-ray spectroscopy and X-ray photoelectron
Dunikowski, D.; Wirecka, R.;
spectroscopyanalyses revealed formation of a more homogeneous and compact corrosion product
Mazur, T.; Lelek-Borkowska, U.;
layer on the coarse-grained steel compared to fine-grained material. The Volta potential analysis,
Wawer, K.; Banaś, J. Effect of Grain
carried out on both steels, revealed formation of micro-galvanic sites at the grain boundaries and
Size on the Corrosion Behavior of
triple junctions. The results indicated that the decrease in corrosion resistance in the fine-grained
Fe-3wt.%Si-1wt.%Al Electrical Steels
in Pure Water Saturated with CO2 .
steel could be attributed to the higher density of grain boundaries (e.g., a higher number of active
Materials 2021, 14, 5084. https:// sites and defects) brought by the refinement. The higher density of active sites at grain boundaries
doi.org/10.3390/ma14175084 promote the metal dissolution of the and decreased the stability of the corrosion product layerformed
on the metal surface.
Academic Editor: Marián Palcut
Keywords: silicon steel; electrical steel; grain size; electrochemical corrosion; carbon dioxide;
Received: 27 July 2021 AM-KPFM Volta potential measurements
Accepted: 1 September 2021
Published: 5 September 2021
249
introduction of any liquids, they usually operate in severe working environments, such as
high pressure, high temperature, and presence of aggressive gases (e.g., CO2 , H2 S), which
can easily compromise their mechanical integrity. With time, the steam condenses into
droplets of liquid inside this equipment. The CO2 corrosion in the oil and gas industry is
one of the greatest challenges [5,6]. The gaseous CO2 dissolves in the condensed water,
forming carbonic acid, which successively dissociates into bicarbonate and carbonate
anions [5,6]. The combination of liquid water and CO2 creates aggressive conditions,
which may lead to severe corrosion attack, leading to their performance degradation and
hence, compromising the functionality of the plant over time. The grain size plays an
important role in the design of electrical steel. From the magnetic properties point of
view, large grains are more benefical. Grain size has also a strong effect on the mechanical
and corrosion properties of the steel [7–11]. The relationship between the grain size and
mechanical properties of the steel is well defined by the Hall-Petch relationship. However,
the correlation between the grain size and its corrosion behavior is still an open field for
investigation. Onyeji et al. [8] studied the corrosion behavior of two X65 steels with the same
chemical composition but different grain sizes in aerated and deaerated brine solutions. The
authors reported that the steel with coarser grains showed a higher corrosion resistance in
both solutions. Li et al. [12] observed that the corrosion resistance of nanocrystallized low-
carbon steels in 0.05 M H2 SO4 + 0.05 M Na2 SO4 aqueous solution increased with decreasing
the grain size. Palumbo et al. [13] observed that an increase in grain refinement leads to an
increase in the volume fraction of intercrystalline areas such as grain boundaries and triple
junctions. Many authors argued that the grain boundaries and triple junctions have higher
energies compared to the bulk and, as such, are more chemically active with respect to the
adjacent matrix [7,8,11,12,14–17]. Therefore, the grain refinement enhances the reactivity of
the surface, which may cause a preferential dissolution of the grains [7,8,11,12,14–16,18].
However, it is worth mentioning that there is not an unanimous consensus regarding the
effect of the grain size on the corrosion resistance of ferrous alloys. Some studies showed
that the environment plays a crucial role. Wang et al. [7] reported that the grain refinement
decreased the corrosion resistance of the low alloy steel in a 3.5 wt.% NaCl solution, but the
same steel showed an improvement in corrosion resistance in a 0.1 M NaHCO3 solution.
Similar behavior was observed by Zeiger et al. [16]. The authors found that the grain
refinement led to a decrease in the corrosion resistance of the steel in a Na2 SO4 solution
with pH = 1, but the corrosion resistance increased in a Na2 SO4 solution with pH = 6. The
little consensus reported in the literature is related to the difficulty of isolating the effect of
the grain size from other microstructural changes introduced during the grain refinement
processes such as, for example, rolling or plastic deformation. Consequently, a case-by-case
study is needed to understand the corrosion effect of the grain size of a given metal in a
given environment.
The objective of this work is to study the corrosion behavior of two electrical steel
sheets with similar chemical composition, but different grain sizes in pure water saturated
with CO2 . To this end, the study was carried out using weight loss and electrochemical
measurements. The scanning electron microscopy-energy dispersive x-ray spectroscopy
(SEM-EDS) and x-ray photoelectron spectroscopy (XPS) analyses were employed to charac-
terize the corrosion product layer and to support the gravimetric and electrochemical re-
sults. Moreover, to highlight the micro-galvanic activities occurring at the grain boundaries
and triple junctions on the metal surface, Volta potential measurements were performed.
2. Experimental Procedures
2.1. Materials
The study was performed on two types of silicon steels labeled 200 and 300. The
samples were supplied by Łukasiewicz Research Network according to IEC 60404 8 5
standard. Both samples in as-received conditions were covered with a phosphate protective
coating. Each time prior to a test, the coating was removed by grinding the surface with SiC
250
Materials 2021, 14, 5084
abrasive paper up to 1200 grit and finishing the surface with levigated alumina, cleaned
ultrasonically in absolute ethanol and dried before immersion in the tested solution.
The coating-free sample surface was analyzed on a spark spectrometer to identify the
chemical composition of samples (Table 1).
Table 1. Chemical composition of the tested samples.
Element Sample
wt.% 200 300
C 0.011 ± 0.01 0.007 ± 0.01
Si 2.93 ± 0.17 3.24 ± 0.11
Mn 0.22 ± 0.03 0.18 ± 0.02
P 0.03 ± 0.01 0.026 ± 0.03
S 0.004 ± 0.001 0.003 ± 0.001
Al 0.91 ± 0.13 1.05 ± 0.08
Fe Bal. Bal.
3.27×10−3 icorr EW
CR (mm y −1 ) = (2)
d
where icorr is the corrosion current density, Ew is the equivalent weight of the metal, and d
is the density of the metal. 3.27 × 10−3 is the conversion factor.
251
Materials 2021, 14, 5084
(a) (d)
(b) (e)
(c) (f)
Figure 1. Microstructure and the histogram showing the grain size number distribution of 200 (a–c)
and 300 steels (d–f).
252
Materials 2021, 14, 5084
200 300
Average Average Average Average Average Average
grain size diameter grain area grain size diameter grain area
number (G) (μm) (μm2 ) number (G) (μm) (μm2 )
10.1 ± 0.23 10.8 ± 0.87 115.6 ± 20.9 4.9 ± 0.32 65.3 ± 7.49 4265.9 ± 499.7
1
ZCPE = (3)
Q( jω )n
where Q stands for CPE constant, n is the exponent, j is the imaginary unit, and ω is the
angular frequency at which Z reaches its maximum value.
253
Materials 2021, 14, 5084
(a) (b)
(c) (d)
(e) (f)
Figure 2. EIS plots obtained after different immersion times for the 200 steel before (a,c,e) and after (b,d,f) the IR
drop correction.
254
Materials 2021, 14, 5084
(a) (b)
(c) (d)
(e) (f)
Figure 3. EIS plots obtained after different immersion times for the 300 steel before (a–c) and after (d–f) the IR
drop correction.
255
Materials 2021, 14, 5084
Figure 4. Equivalent circuit used to fit the EIS plots. Here, Rs is the electrolyte resistance, Qdl is
the constant phase element representing the double charge layer capacitance and Rct is the charge
transfer resistance. L and RL represent the inductance and inductance resistance, respectively.
As can be seen from figures (Figure 2b,d,f and Figure 3b,d,f), the fitted results were
similar to those obtained experimentally and the values of χ2 were very low (Table 4),
indicating that the equivalent circuit, employed to simulate the system under investigation,
was the most appropriate one. It follows from the Nyquist diagrams (Figures 2c and 3c) that
the shape of the curve did not change with the immersion time and exhibited a depressed
semicircle in the whole frequency range due to the inherent charge transfer processes
controlling the corrosion reactions. An inductive loop is also visible at low frequencies,
likely due to the relaxation time of the intermediate adsorbed species.
Table 4. Electrochemical impedance parameters of the two steels in the tested solution after the IR drop correction.
Time Rs Q1 Rct L RL χ2
Sample n
(h) (Ω cm2 ) (mΩ−1 sn cm−2 ) (Ω cm2 ) (H cm2 ) (Ω cm2 ) (10−4 )
3 0.96 0.48 ± 0.06 0.57 ± 0.02 0.47 ± 0.03 0.27 ± 0.03 0.14 ± 0.011 1.05
6 0.98 0.38 ± 0.03 0.59 ± 0.04 0.45 ± 0.05 0.15 ± 0.02 0.01 ± 0.005 0.99
200 12 0.98 0.36 ± 0.03 0.58 ± 0.05 0.54 ± 0.06 0.19 ± 0.04 0.01 ± 0.003 0.38
18 0.97 0.26 ± 0.02 0.63 ± 0.03 0.65 ± 0.03 0.12 ± 0.02 0.04 ± 0.001 0.46
24 0.99 0.19 ± 0.01 0.67 ± 0.05 0.64 ± 0.03 0.07 ± 0.01 0.07 ± 0.002 1.6
3 0.95 0.04 ± 0.001 0.61 ± 0.04 2.32 ± 0.16 0.24 ± 0.01 0.70 ± 0.06 2.02
6 0.98 0.04 ± 0.001 0.65 ± 0.03 2.35 ± 0.21 0.25 ± 0.02 0.69 ± 0.02 2.06
300 12 0.89 0.04 ± 0.003 0.69 ± 0.04 2.03 ± 0.22 0.64 ± 0.01 0.49 ± 0.02 1.99
18 0.98 0.05 ± 0.005 0.71 ± 0.03 2.26 ± 0.15 0.36 ± 0.02 0.52 ± 0.03 1.41
24 0.97 0.05 ± 0.003 0.72 ± 0.02 2.61 ± 0.28 0.36 ± 0.01 0.51 ± 0.05 0.94
The CO2 gas dissolves in the solution forming carbonic acid, which successively
dissociates into bicarbonate and carbonate anions, according to the following reactions:
HCO3− ↔ H +
+2CO23− (6)
In the presence of CO2 , the process is controlled by the three cathodic reactions [6,22,23]:
(a) (b)
(c)
Figure 5. EIS plots comparing the two steels after 24 h of immersion and after the IR drop correction in the tested solution.
(a) Nyquist, (b) Bode and (c) Phase angle.
257
Materials 2021, 14, 5084
Figure 6. Potentiodynamic polarization curves obtained after 24 h of immersion in the tested solution.
Sample Ecorr (V vs. SCE) − βc (V dec−1 ) icorr (μA cm−2 ) CR (mm y−1 )
200 −0.696 ± 0.015 0.433 ± 0.087 24.44 ± 1.87 0.28 ± 0.02
300 −0.680 ± 0.011 0.588 ± 0.051 12.98 ± 2.57 0.15 ± 0.03
To confirm the correctness of the results obtained in pure water, the electrochemical
experiments were also carried out in a 3.5 wt.% NaCl aqueous solution saturated with CO2 .
The EIS and PDP results are displayed in Figures 7 and 8 and Tables 6 and 7, respectively.
The results are in agreement with ones observed in pure water saturated with CO2 , which
shows that the coarse-grained steel still displays a better corrosion resistance compared to
the fine grained steel. In particular, it can be seen from the potentiodynamic experiments
(Figure 8) that the 300 steel exhibits a pseudo-passive region, only, extends to a small range
of potential (i.e., −612 to −586 mV vs. SCE). This result confirms that the coarse-grained
steel tends to form a more stable corrosion product layer on the surface, which led to an
increase in its corrosion resistance.
258
Materials 2021, 14, 5084
(a) (b)
(c)
Figure 7. EIS plots comparing the two different samples after 24 h of immersion in a 3.5 wt.% NaCl saturated with CO2 .
(a) Nyquist, (b) Bode, and (c) Phase angle.
Figure 8. Potentiodynamic polarization curves obtained after 24 h of immersion in a 3.5 wt.% NaCl
aqueous solution saturated with CO2 .
259
Materials 2021, 14, 5084
Table 6. Electrochemical impedance parameters of the steels after 24 h of immersion in a 3.5 wt.% NaCl aqueous solution
saturated with CO2 .
Q1
Time Rs Rct L RL χ2
Sample (mΩ−1 sn n
(h) (Ω cm2 ) (Ω cm2 ) (H cm2 ) (Ω cm2 ) (10−4 )
cm−2 )
200 24 10.04 0.46 ± 0.03 0.78 ± 0.03 421.10 ± 28.53 50.11 ± 9.07 49.22 ± 8.59 8.27
300 24 10.26 0.30 ± 0.06 0.80 ± 0.05 968.30 ± 40.89 60.10 ± 10.03 31.81 ± 6.89 5.69
Sample Ecorr (V vs. SCE) − βc (V dec−1 ) icorr (μA cm−2 ) CR (mm y−1 )
200 −0.735 ± 0.019 0.517 ± 0.044 70.42 ± 2.34 0.82 ± 0.03
300 −0.746 ± 0.016 0.471 ± 0.067 29.43 ± 1.37 0.34 ± 0.02
(a) (b)
Figure 9. SEM analysis on the sample (a) 200 and (b) 300, after 24 h of immersion. (The red square
area corresponds to the area of the EDS analysis).
260
Materials 2021, 14, 5084
The precipitation of FeCO3 depends not only on the concentration of Fe2+ and CO23−
ions, but is also affected by other factors such as temperature, CO2 partial pressure, and pH.
Among the abovementioned factors, the pH of the solution can be regarded as one of the
most influential factors [25]. Dugstad [25] reported that an increase in pH of the solution
significantly reduced the concentration of Fe2+ ions required to exceed the FeCO3 solubility
product and therefore promoteing its precipitation. In this study, the pH increased as
the immersion time increased, going from 4.12, at the beginning of the experiment, to
5.61 after 24 h of immersion. As such, making the precipitation of FeCO3 more probable.
Moreover, Dugstad [25] also reported that the concentration of Fe2+ ions is higher at the
surface/solution interface compared to the bulk solution. Consequently, the concentration
of Fe2+ ions required to promote the formation of FeCO3 on the metal surface is lower and
thus, increasing the likelihood of having FeCO3 on surface. However, only small traces
of FeCO3 could be found as confirmed by the XPS analysis (Figure 10). Previous studies
reported that FeCO3 begins to decompose at temperatures below 100 ◦ C according to the
following reaction [6,23,26]:
FeCO3 → FeO + CO2 (14)
In the presence of CO2 or water vapor, FeO transforms into Fe3 O4 [23,26].
the data observed for the O1s spectrum, which shows three peaks at 284.8, ~286.3, and
~288.5 eV. The 284.6 eV peak may correspond to secondary carbon [28], whereas the peaks
at ~286.3 are ~288.5 eV are attributed to the C–O and C=O bonds of FeCO3 [6]. As in the
case of the O1S spectrum, for the 300 steel, a fourth peak was observed at 289.8 eV, which
is characteristic of FeCO3 [29]. The deconvoluted Fe2p3/2 peaks (Figure 10j,k) could be
attributed to α-Fe2 O3 or/and γ- Fe2 O3 oxides [6,30], likely due to the partial decomposition
of iron carbonate (i.e., Equations (14)–(18)).
(a)
(b) (c)
(d) (e)
Figure 10. Cont.
262
Materials 2021, 14, 5084
(f) (g)
(h) (i)
(j) (k)
Figure 10. XPS analysis carried out after 24 h of immersion. Survey (a), 200 steel (b,d,f,h,j) and
300 steel (c,e,g,i,k).
263
Materials 2021, 14, 5084
200 300
Peak
Binding Binding
Assignment
Energy %Area Energy %Area
(eV) (eV)
C1s - 23.0 - 14.5
C–C,
284.8 47.5 284.8 52.4
C–H
C–O,
286.5 35.2 286.3 24.5
C–OH
C=O 288.7 17.3 288.5 14.2
CO23− - - 289.8 8.9
Al2p - 12.0 - 14.5
Al2 O3 74.7 85.8 74.7 72.6
AlOOH 77.5 14.2 77.3 27.4
Si2p - 4.6 - 7.2
SiOx 101.8 66.7 102.1 66.7
splitting 102.5 33.3 102.7 33.3
O1s - 53.7 - 54.0
O–Me 530 36.6 530.1 38.2
O–C 531.6 45.3 531.6 38.9
O=C 532.9 18.1 533.0 17.8
CO23− - - 534.6 5.1
Fe2p - 5.7 - 8.3
Fe3+ 710.5 44.5 710.7 51.0
Fe3+ 713.2 15.2 714.0 15.3
satellite 717.6 13.2 718.6 9.7
- 723.9 22.2 724.0 19.1
- 726.7 4.9 726.5 4.9
The XPS analysis is in agreement with the EDS analysis (Table 8), indicating that the
corrosion products formed on both samples were mainly composed of Al2 O3 and SiO2 ,
with traces of FeCO3 /Fe2 O3 .
To study the micro-galvanic activities occurring at the grain boundaries and triple
junction on the metal surface, AM-KPFM measurements were carried out. AM-KPFM is
a powerful technique for assessing the Volta potential (ΔΨ) of a metal surface. The ΔΨ
is a characteristic property of the metal surface and can provide an insight into the local
electrochemical activities on the metal surface [31]. Figure 11 the topography and the
corresponding Volta potential maps of the two tested steels. The Volta potential maps
show darker color representing the anodic regions, whereas the lighter color representing
cathodic regions. The ΔΨ mapping clearly shows potential differences at grain boundaries
and triple junctions, indicating higher electrochemical activity in these regions. The grain
boundaries and triple junctions are characterized in the maps as lighter zones, thus rep-
resenting the cathodic areas and displaying a relative ΔΨ difference of circa +30 mV with
respect to the adjacent matrix. The results confirmed that these regions are more active com-
pared to the adjacent matrix, which makes them more susceptible to corrosion attack during
the exposure to electrolytes. Therefore, the grain refinement enhances the reactivity of the
surface, which could promote a preferential dissolution of the grains [7,8,11,12,14–16,18].
264
Materials 2021, 14, 5084
(a) (b)
(c) (d)
Figure 11. AM-KPFM analysis of the steels with the topography and the corresponding Volta
potential map, respectively. Steel 200 (a,b); 300 (c,d).
4. Conclusions
The effect of the grain size on the corrosion behavior of two electric steels in pure
water saturated by CO2 can be summarized as follows:
• The gravimetric results indicated that the grain refinement decreases the corrosion
resistance of the steel.
• The EIS measurements showed that the coarse-grained sample displayed a higher
capacitive loop compared to fine-grained steel. This result is related to the formation
of a thicker and more stable protective corrosion product layer.
• The potentiodynamic measurements showed that the corrosion current density of the
coarse-grained steel was much smaller compared fine-grained steel. Both the anodic
and cathodic current densities were found to be lower for the coarse-grained steel.
• The SEM-EDS and XPS analyses confirmed presence of a thicker and more homoge-
nous protective layer on the coarse-grained steel, consisting mainly of Al2 O3 , SiO2 ,
and traces of FeCO3 .
• The Volta potential measurements showed potential differences between the grains
and grain boundaries, indicating a higher electrochemical activity in these regions,
which would couse a preferential dissolution of grains.
Author Contributions: G.P. conceived and designed this study. G.P. performed the electrochemical
experiments, analyzed the experimental data, wrote and edited the manuscript; D.D. performed
the gravimetric and SEM-EDS experiments; R.W. performed the XPS analysis; T.M. performed the
AM-KPFM analysis; U.L.-B. prepared the samples; K.W. supplied materials, supervision, and funding
acquisition; J.B. supervision and funding acquisition. All authors have read and agreed to the
published version of the manuscript.
Funding: Part of this work was supported by AGH University of Science and Technology, Faculty of
Foundry Engineering, Department of Chemistry and Corrosion of Metals, project no. 16.16.170.654.
R.W. has been partly supported by the EU Project POWR.03.02.00-00-I004/16.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the
corresponding author.
Conflicts of Interest: The authors declare no conflict of interest. The funders had no role in the design
of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or
in the decision to publish the results.
265
Materials 2021, 14, 5084
References
1. Xu, Y.; Jiao, H.; Qiu, W.; Misra, R.D.K.; Li, J. Effect of cold rolling process on microstructure, texture and properties of strip cast
Fe-2.6% Si Steel. Materials 2018, 11, 1161. [CrossRef] [PubMed]
2. Xia, C.; Wang, H.; Wu, Y.; Wang, H. Joining of the laminated electrical steels in motor manufacturing: A Review. Materials 2020,
13, 4583. [CrossRef] [PubMed]
3. Petryshynets, I.; Kováč, F.; Petrov, B.; Falat, L.; Puchý, V. Improving the magnetic properties of non-oriented electrical steels by
secondary recrystallization using dynamic heating conditions. Materials 2019, 12, 1914. [CrossRef] [PubMed]
4. Lee, K.; Park, S.; Huh, M.; Kim, J.; Engler, O. Effect of texture and grain size on magnetic flux density and core loss in non-oriented
electrical steel containing 3.15% Si. J. Magn. Magn. Mater. 2014, 354, 324–332. [CrossRef]
5. Palumbo, G.; Górny, M.; Banaś, J. Corrosion inhibition of pipeline carbon steel (N80) in CO2 -saturated chloride (0.5 M of KCl)
solution using gum arabic as a possible environmentally friendly corrosion inhibitor for shale gas industry. J. Mater. Eng. Perform.
2019, 28, 6458–6470. [CrossRef]
6. Palumbo, G.; Kollbek, K.; Wirecka, R.; Bernasik, A.; Górny, M. Effect of CO2 partial pressure on the corrosion inhibition of
N80 carbon steel by gum arabic in a CO2 -water saline environment for shale oil and gas industry. Materials 2020, 13, 4245.
[CrossRef] [PubMed]
7. Wang, P.; Ma, L.; Cheng, X.; Li, X. Effect of grain size and crystallographic orientation on the corrosion behaviors of low alloy
steel. J. Alloys Compd. 2021, 857, 158258. [CrossRef]
8. Onyeji, L.; Kale, G. Preliminary investigation of the corrosion behavior of proprietary micro-alloyed steels in aerated and
deaerated brine solutions. J. Mater. Eng. Perform. 2017, 26, 5741–5752. [CrossRef]
9. Di Schino, A.; Barteri, M.; Kenny, J.M. Grain size dependence of mechanical, corrosion and tribological properties of high nitrogen
stainless steels. J. Mater. Sci. 2003, 38, 3257–3262. [CrossRef]
10. Di Schino, A.; Kenny, J.M. Effects of the grain size on the corrosion behavior of refined AISI 304 austenitic stainless steels. J. Mater.
Sci. Lett. 2002, 21, 1631–1634. [CrossRef]
11. Ralston, K.D.; Birbilis, N. Effect of grain size on corrosion: A review. Corrosion 2010, 66, 075005. [CrossRef]
12. Li, Y.; Wang, F.; Liu, G. Grain size effect on the electrochemical corrosion behavior of surface nanocrystallized low-carbon steel.
Corrosion 2004, 60, 891–896. [CrossRef]
13. Palumbo, G.; Thorpe, S.; Aust, K. On the contribution of triple junctions to the structure and properties of nanocrystalline
materials. Scr. Metall. Mater. 1990, 24, 1347–1350. [CrossRef]
14. Soleimani, M.; Mirzadeh, H.; Dehghanian, C. Effect of grain size on the corrosion resistance of low carbon steel. Mater. Res.
Express 2019, 7, 016522. [CrossRef]
15. Chen, Y.T.; Zhang, K.G. Influence of grain size on corrosion resistance of a HSLA steel. Adv. Mater. Res. 2012, 557–559,
143–146. [CrossRef]
16. Zeiger, W.; Schneider, M.; Scharnweber, D.; Worch, H. Corrosion behaviour of a nanocrystalline FeA18 alloy. Nanostructured Mater.
1995, 6, 1013–1016. [CrossRef]
17. Wang, S.G.; Shen, C.B.; Long, K.; Zhang, T.; Wang, F.H.; Zhang, Z.D. The electrochemical corrosion of bulk nanocrystalline ingot
iron in acidic sulfate solution. J. Phys. Chem. B 2006, 110, 377–382. [CrossRef] [PubMed]
18. Seikh, A.H. Influence of heat treatment on the corrosion of microalloyed steel in sodium chloride solution. J. Chem. 2013,
2013, 587514. [CrossRef]
19. ASTM International. ASTM-E112. Standard Test Methods for Determining Average Grain Size; ASTM-E112: West Conshohocken, PA,
USA, 2013.
20. ASTM International. ASTM-G102. Standard Practice for Calculation of Corrosion Rates and Related Information from Electrochemical
Measurements; ASTM-G102: West Conshohocken, PA, USA, 1994.
21. Palumbo, G.; Berent, K.; Proniewicz, E.; Banaś, J. Guar gum as an eco-friendly corrosion inhibitor for pure aluminium in 1-M HCl
solution. Materials 2019, 12, 2620. [CrossRef]
22. Mustafa, A.H.; Ari-Wahjoedi, B.; Ismail, M.C. Inhibition of CO2 corrosion of X52 steel by imidazoline-based inhibitor in high
pressure CO2 -water environment. J. Mater. Eng. Perform. 2012, 22, 1748–1755. [CrossRef]
23. Islam, M.A.; Farhat, Z.N. Characterization of the corrosion layer on pipeline steel in sweet environment. J. Mater. Eng. Perform.
2015, 24, 3142–3158. [CrossRef]
24. Zhang, G.; Cheng, Y. Electrochemical characterization and computational fluid dynamics simulation of flow-accelerated corrosion
of X65 steel in a CO2 -saturated oilfield formation water. Corros. Sci. 2010, 52, 2716–2724. [CrossRef]
25. Dugstad, A. Fundamental aspects of CO2 metal loss corrosion—Part 1: Mechanism. In Corrosion 2006; NACE International: San
Diego, CA, USA, 2006; p. 18.
26. Heuer, J.; Stubbins, J. An XPS characterization of FeCO3 films from CO2 corrosion. Corros. Sci. 1999, 41, 1231–1243. [CrossRef]
27. Barrera, A.; Tzompantzi, F.; Campa-Molina, J.; Casillas, J.E.; Pérez-Hernández, R.; Ulloa-Godinez, S.; Velásquez, C.;
Arenas-Alatorre, J. Photocatalytic activity of Ag/Al2 O3 -Gd2 O3 photocatalysts prepared by the sol–gel method in the degradation
of 4-chlorophenol. RSC Adv. 2018, 8, 3108–3119. [CrossRef]
28. Dong, B.; Zeng, D.; Yu, Z.; Cai, L.; Shi, S.; Yu, H.; Zhao, H.; Tian, G. Corrosion mechanism and applicability assessment of N80
and 9Cr steels in CO2 auxiliary steam drive. J. Mater. Eng. Perform. 2019, 28, 1030–1039. [CrossRef]
266
Materials 2021, 14, 5084
29. Garcia, S.; Rosenbauer, R.; Palandri, J.; Maroto-Valer, M. Sequestration of non-pure carbon dioxide streams in iron oxyhydroxide-
containing saline repositories. Int. J. Greenh. Gas Control. 2012, 7, 89–97. [CrossRef]
30. Singh, A.; Ansari, K.R.; Quraishi, M.A.; Lgaz, H. Effect of electron donating functional groups on corrosion inhibition of J55 steel
in a sweet corrosive environment: Experimental, density functional theory, and molecular dynamic simulation. Materials 2018,
12, 17. [CrossRef]
31. Örnek, C.; Engelberg, D. SKPFM measured Volta potential correlated with strain localisation in microstructure to understand
corrosion susceptibility of cold-rolled grade 2205 duplex stainless steel. Corros. Sci. 2015, 99, 164–171. [CrossRef]
267
materials
Article
Effect of Sub-Zero Treatments and Tempering on Corrosion
Behaviour of Vanadis 6 Tool Steel
Peter Jurči 1, *, Aneta Bartkowska 2 , Mária Hudáková 1 , Mária Dománková 1 , Mária Čaplovičová 3
and Dariusz Bartkowski 4
269
ii. The martensite of SZT steel is refined compared to that developed by conventional
room-temperature quenching.
iii. Sub-zero treatment leads to acceleration of precipitation of transient carbides during
either the re-heating to room temperature or short-term storage at room temperature.
iv. The last consequence of the SZT is the considerably enhanced number and pop-
ulation density of small globular carbides (SGCs). The formation of the SGCs at
cryogenic temperature is time-dependent and follows the reduction of the retained
austenite amount well.
The corrosion resistance is not commonly considered as a key property for tool ma-
terials. This is due to the fact that the dominant number of technological applications is
conducted in corrosion-friendly environments, or, if not, there exist surfacing techniques
(such as physical vapour deposition, for instance) through which a sufficiently high cor-
rosion resistance of tools can be ensured. Despite this, it might be desirable to keep at
least acceptable corrosion behaviour of tools made of Cr and Cr-V ledeburitic tool steels in
some applications, where, for instance, surface techniques fail for several reasons. These
applications may comprise industrial branches, such as mining, earth-handling, milling,
powder compaction, mineral processing, etc. Here, high processing reliability and sufficient
tools’ durability require materials with not only excellent wear performance but also with
at least acceptable corrosion resistance.
However, the characterisation of the effect of SZT on the corrosion behaviour of Cr
and Cr-V ledeburitic steels is almost completely lacking. There are only a few relevant
studies published on this topic, and the obtained results manifest clear inconsistency. Amini
et al. [20] reported on the worsened corrosion behaviour of 1.2080 grade (AISI D3) steel due
to the SZT in liquid nitrogen for 24–48 h. According to their consideration, the worsening
of corrosion resistance can be ascribed to the increased carbide percentage, which decreases
the number of solutionised chromium atoms in the martensite, as well as to the increased
martensite/carbide interfaces (galvanic cell areas). Hill et al. [21], on the contrary, recorded
an improvement of corrosion resistance of sub-zero-treated (−196 ◦ C/15 min) X190CrVMo
20-4 ledeburitic steel when tested in 0.5 molar sulphuric acid solution.
The present study attempts to overcome the limitation of the lack of data on the
corrosion behaviour of sub-zero-treated ledeburitic tool steels. It describes the effects of
different sub-zero treatment temperatures (−75, −140, −196, and −269 ◦ C) and tempering
regimes on the corrosion behaviour of Cr-V ledeburitic steel Vanadis 6. Conventionally
treated steel has been used as a reference. In addition, the relationships between the
microstructures, heat treatment parameters, and corrosion behaviour are presented and
thoroughly discussed.
Element (Mass %) C Si Mn Cr V Mo Fe
Content 2.1 1.0 0.4 6.8 5.4 1.5 balance
270
Materials 2021, 14, 3759
SZT temperature (4). The SZTs were carried out at temperatures of −75, −140, −196, or
−269 ◦ C, and for 17 h each (5). After that, the material was re-heated to room temperature,
at a heating rate of 1 ◦ C/min (6). Immediately after re-heating, sub-zero-treated specimens
were subjected to tempering. Tempering was carried out at temperatures of 170, 330, 450,
or 530 ◦ C, 2 times each, for 2 h (7, 8).
than 500 nm were then denoted as SGCs. For the determination of the population densities
of ECs, SCs, and SGCs, 25 randomly acquired SEM micrographs, at a magnification of
3000×, were used. Then, the mean values and standard deviations were calculated from
the obtained experimental data.
Semi-quantitative chemical composition of phases of sub-micron size (metallic ele-
ments only) and the microstructure of both the martensite and the austenite were examined
by transmission electron microscopy (TEM). Thin foils for TEM were prepared by section-
ing specimens, with a special saw, to obtain pieces with a thickness of 0.1 mm, which were
then ground mechanically on silicon carbide paper (1200 grit) to a thickness of approxi-
mately 0.02 mm. Final thinning was carried out using an electro-polisher Struers-Tenupol
5. A TEM JEM ARM 200 cF (Jeol Ltd., Tokyo, Japan) equipped with an energy-dispersive
spectroscopy (EDS) detector was used for acquisition of micrographs as well as for the
determination of phase compositions.
The phases in differently heat-treated specimens were identified from the X-ray diffrac-
tion (XRD) profiles by using a Phillips PW 1710 diffractometer. A filtered Coα1,2 charac-
teristic radiation, obtained at the voltage of 40 kV and current of 40 mA, has been used
for acquisition of diffraction line profiles, within a range of 37–130◦ of two-theta angles.
The analysis was coupled with Rietveld refinement of the X-ray line profiles. The retained
austenite amount was determined following the ASTM E975-13 standard [23].
Corrosion resistance studies were carried out by the potentiodynamic polarisation mea-
surements (TAFEL). The TAFEL measurements were completed by using the potentiostat-
galvanostat ATLAS 0531 EU (Atlas-Sollich, R˛ebiechowo, Poland) and IA ATLAS SOLLICH
(Atlas-Sollich, R˛ebiechowo, Poland). A platinum electrode was used as an auxiliary elec-
trode, while a calomel electrode was used as the reference electrode and tested specimens
as the working electrodes. The data were recorded by AtlasCorr (Atlas-Sollich, version 3.19)
and AtlasLab (Atlas-Sollich, version 2.24) computer software. After corrosion tests, the
tested specimens were examined by using SEM coupled with EDS.
Before testing, all the specimens were ground using metallographic emery papers
with a grit size up to 2000, and finally, were polished using 3 μm diamond slurry. Just prior
to measurements, the specimens were degreased in ethanol and warm air-dried. The 3.5 (in
mass %) NaCl water solution was prepared using distilled water and high-purity reagent
grade sodium chloride. Then, the solution was subjected to Ar gas bubbling for 30 min
in order to its deaerate. For the corrosion tests, the temperature of the solution was kept
constant at 22 ◦ C.
The potentiodynamic polarisation measurements were carried out within the range of
potentials of −1.5 to 1.5 V, and at a scan rate of 1 mV/s.
The corrosion rate for each specimen was estimated according to the ASTM G 102-89
standard [24]. The calculation was based on the validity of the Faraday’s Law, and it
followed Equation (1):
icor
C R = K1 · ·E (1)
ρ W
where CR is the corrosion rate (mm/year), Icor is the current density (μA/cm2 ), ρ is the
specific density of the alloy (g/cm3 ), K1 = 3.27·10−3 , (mm × g/μA × cm × year), and EW
is the equivalent weight of the experimental steel.
Equivalent weight, EW , represents the mass of metal, in grams, that will be oxidised by
the passage of one Faraday of electric charge. The value of EW of the experimental material
was calculated upon its known chemical composition, atomic weight of the major elements,
W, the most common valence of a particular element n, and as a reciprocal value of the sum
of the electron equivalents of all the major elements, Q, following Equations (2) and (3):
1
EW = ni · f i
(2)
∑ Wi
272
Materials 2021, 14, 3759
ni · f i
Q= ∑ Wi
(3)
where fi is the mass percentage of the ith element in the alloy, Wi is the atomic weight of
the ith element in the alloy, and ni is the most common valence of the ith element of the
alloy. All the input data for the calculations are collected in Table 2. Additionally, it should
be noted that an assumption of corrosion uniformity was adopted in the present study, in
order to simplify the considerations.
Table 2. The mass percentage, f, atomic weight, W, most common valance, n, electron equivalents of
the main elements in the Vanadis 6 steel, Q, and calculated electron equivalent of the steel, Qtotal , as
well as its equivalent weight, Ew .
In order to analyse the differences in nobility between the carbides and matrix, the
Kelvin probe force microscopy (KPFM) has been adopted. This technique enables to
distinguish between local potentials of different phases, at the sub-micron level. Hence, it
is suitable for analysing fine-grained PM ledeburitic steels. The analyses were performed at
The University of Manchester, Department of Materials Corrosion, by using a Multimode 8
instrument (Bruker), in the amplitude modulated mode. The imaging parameters were
the following: 50 × 50 μm scans, potential maps obtained from a 50 nm lift height, using a
0.3 Hz scan rate, 512 points per line, 256 lines, and height images obtained using 100 nm
peak force amplitude.
3. Results
3.1. Microstructure
SEM images (Figure 2) show the microstructures of the Vanadis 6 steel obtained by the
conventional quenching and by different sub-zero treatments. The steel after conventional
room temperature quenching is composed of martensitic matrix with the presence of a
relatively large retained austenite amount, and of three carbide types (Figure 2a). The
carbides are, according to the classification reported in [14]: eutectic carbides (ECs), sec-
ondary carbides (SCs), and small globular carbides (SGCs). The character of the matrix
microstructure does not change significantly with the application of sub-zero treatment, at
the magnification used for the SEM observations (Figure 2b–e). The formations of retained
austenite become almost invisible after SZT, due to the significant reduction of γR amount
due to this kind of treatment. Alternatively, sub-zero-treated steel differs from that after
CHT in terms of the number and population density of carbide particles (compare the SEM
micrograph in Figure 2a with micrographs in Figure 2b–e). This concerns mainly the SGCs:
the population density of these particles is much higher after SZT, while the SZT does not
alter the population densities of ECs and SCs, as demonstrated recently [4,13,19].
273
Materials 2021, 14, 3759
Figure 2. Scanning electron microscopy (SEM) micrographs showing the microstructure of Vanadis
6 ledeburitic steel after conventional room temperature quenching (a) and after SZT at −75 ◦ C (b),
−140 ◦ C (c), −196 ◦ C (d), and −269 ◦ C (e).
274
Materials 2021, 14, 3759
by diffraction patterns in Figure 4c. The martensite produced by SZT manifests consid-
erable refinement compared to that developed by CHT. The typical width of martensitic
domains ranges between 50 and 150 nm (Figure 4d). It is also shown here that the size
of the martensitic domains manifests a great level of variability—there are some coarser
domains visible in the micrograph, while in some sites, the domains are much smaller.
A much higher amount of dislocations are generated within the martensite, as a result of
plastic deformation that takes place during the isothermal hold at cryotemperatures. The
retained austenite amount is significantly reduced by application of SZT. Additionally, the
size of γR formations is much smaller compared to the state after CHT (Figure 4e). The
presence of retained austenite at the interfaces of martensitic domains is confirmed by
diffraction patterns in Figure 4f.
Figure 5 shows X-ray diffraction spectra obtained on the conventionally quenched
specimen and on specimens with SZT at −140 and −269 ◦ C. All the spectra contain the
peaks of the martensite, retained austenite, and major carbides. In the diffraction profile of
the CHT specimen, there are the peaks of (111)γ, (200)γ, (220)γ, and (311)γ, all well-visible.
On the other hand, the last two peaks disappear almost completely from the XRD spectra
of SZT steel, suggesting a significant reduction of retained austenite amounts after this
kind of treatment.
Figure 3. Population density of small globular carbides (SGCs) for differently sub-zero treated (SZT) specimens as a function
of tempering temperature: (a) conventional heat treatment (CHT), (b) SZT −75 ◦ C, (c) SZT −140 ◦ C, (d) SZT −196 ◦ C,
(e) SZT −269 ◦ C.
275
Materials 2021, 14, 3759
Figure 4. Transmission electron microscopy (TEM) micrographs of the CHT specimen (a–c) and the
specimen that was subjected to SZT at −196 ◦ C (d–f). (a) Bright-field image showing martensitic
needle microstructure with retained austenite at needle interfaces, (b) corresponding dark-field
image showing the retained austenite, (c) diffraction patterns of the retained austenite, (d) bright-
field image showing martensitic microstructure with a small amount of retained austenite at the
interfaces of martensitic domains, (e) corresponding dark-field image, (f) diffraction patterns of the
retained austenite.
Variations in the retained austenite amounts for CHT steel, and for steel samples
subjected to different SZTs, as a function of tempering temperature are summarised in
Figure 6. It is shown that the application of SZT reduces the γR amount to one tenth—
one fourth compared to the state after room-temperature quenching. Tempering at low
temperatures does not influence the γR amounts. Alternatively, tempering in the secondary
hardening temperature range evokes almost complete retained austenite decomposition in
the case of CHT steel. The application of SZT, in addition, accelerates the decomposition of
γR . Therefore, the amounts of metastable retained austenite lie below the detection limit
of XRD after tempering at 450 ◦ C and higher, in the case of the steel SZT in either liquid
nitrogen or in liquid helium.
276
Materials 2021, 14, 3759
Figure 5. X-ray diffraction spectra of CHT steel and steel after SZT at −140 and −269 ◦ C.
Figure 6. Retained austenite amounts for different SZT specimens, as a function of tempering
temperature: (a) CHT, (b) SZT −75 ◦ C, (c) SZT −140 ◦ C, (d) SZT −196 ◦ C, (e) SZT −269 ◦ C.
277
Materials 2021, 14, 3759
Figure 7. Potentiodynamic polarisation curves of CHT and different SZT specimens in un-tempered state.
The potentiodynamic curves of CHT steel in un-tempered state as well as in the states
after different tempering treatments are shown in Figure 8. The un-tempered specimen
manifests the highest corrosion potential (Ecorr ), and the Ecorr decreases with the increasing
tempering temperature. This indicates that the tempering treatment deteriorates the re-
sistance of conventional room-temperature-quenched Vanadis 6 steel against oxidation in
3.5% water solution of NaCl. The variations in the pitting potentials, Epit (see also Table 4),
do not manifest a clear tendency with respect to the level of tempering temperature. The
specimens treated at either low (170 ◦ C) or high (530 ◦ C) tempering temperatures mani-
fest more anodic behaviour as compared with un-tempered steel. Conversely, the pitting
potentials, Epit , of specimens tempered at intermediate temperatures indicate higher sus-
ceptibility to the stable pitting corrosion. The corrosion current, Icorr (see also Table 4), does
not manifest significant changes after tempering at 170 ◦ C. However, it raises rapidly after
tempering at 330 ◦ C and higher, suggesting that the corrosion rate of the material increases
278
Materials 2021, 14, 3759
Table 3. Corrosion current, Icorr , corrosion potential, Ecorr , and pitting potential, Epit , values acquired
from corrosion tests in 3.5 mass % NaCl water solution, for CHT and different SZT specimens in the
prior-to-tempering state.
Figure 8. Potentiodynamic polarisation curves of CHT Vanadis 6 steel in un-tempered state and in
the states after different tempering treatments.
Table 4. Corrosion current, Icorr , corrosion potential, Ecorr , and pitting potential, Epit , values acquired
from corrosion tests in 3.5 mass % NaCl water solution, for CHT and differently tempered specimens.
specimens, indicating almost no effect of tempering on the steel susceptibility to the pitting
corrosion with SZT at −140 ◦ C. The Icorr follows a very similar tendency to what was
recorded for CHT steel. However, it is obvious (see also Table 5) that the values of Icorr
are lower than what was recorded for the CHT steel tempered at the same temperatures.
It should be noted that similar variations in corrosion characteristics were recorded for
other SZTs. One can thus surmise that the SZT makes an overall amelioration of corrosion
behaviour of the Vanadis 6 steel when tested in the 3.5 mass % NaCl water solution.
Table 5. Corrosion current, Icorr , corrosion potential, Ecorr , and pitting potential, Epit , values acquired
from corrosion tests in 3.5 mass % NaCl water solution, for specimens after SZT at −140 ◦ C and
different tempering regimes.
Figure 10 shows the potentiodynamic polarisation curves for the CHT steel specimen
and steel specimens after SZT at −75, −140, −196, and −269 ◦ C, after tempering at 530 ◦ C.
As shown, the corrosion potential of SZT specimens is higher (more anodic) as compared to
the material state after CHT. Further, it is evident that the SZTs at −75, −140, and −269 ◦ C
produced higher Ecorr than the treatment at −196 ◦ C. The pitting potentials, Epit (see also
Table 6), on the other hand, are shifted to the more cathodic values. This suggests that the
application of SZTs provides the examined steel with higher susceptibility to the stable
formation of pitting, in the state after tempering at 530 ◦ C. The variations of corrosion
current well-follow the changes in Ecorr : the samples after CHT have the highest dissolution
rate (the highest Icorr ), and the Icorr decreases in the order of SZT −196 ◦ C, SZT −75 ◦ C,
SZT −140 ◦ C, and SZT −269 ◦ C. Here, a very important finding is that the tendency of the
corrosion behaviour improvement, due to the SZT, is partly maintained after tempering at
temperatures normally used for the secondary hardening.
280
Materials 2021, 14, 3759
Figure 10. Potentiodynamic polarisation curves of CHT Vanadis 6 steel, and Vanadis 6 subjected to
different SZTs, after tempering at 530 ◦ C.
Table 6. Corrosion current, Icorr , corrosion potential, Ecorr , and pitting potential, Epit , values acquired
from corrosion tests in 3.5 mass % NaCl water solution, for CHT and different SZT specimens, after
tempering at 530 ◦ C.
Figure 11 provides an overview of the dependence of the corrosion rate on the temper-
ing temperature for CHT specimens and for the specimens that were subjected to different
SZTs. There are two main tendencies apparently shown. The first one is that the tempering
treatment accelerates the corrosion rate (CR), and the higher the tempering temperature,
the more accelerated the CR. The second general tendency is that SZTs retard the CR, and
that the SZTs at −140 and −269 ◦ C act most effectively in this way.
The surfaces after the potentiodynamic measurements of differently heat-treated spec-
imens are presented in Figure 12. It is shown that the carbide/matrix boundaries are
preferentially attacked by corrosion. This mainly concerns the boundaries between coarser
eutectic and secondary carbides, as they differ from the matrix in terms of their chemistry.
The M7 C3 carbides (secondary carbides, SCs), for instance, contain 37.2 ± 0.8 mass % of Cr,
46.4 ± 0.5 mass % of Fe, and 12.7 ± 0.3 mass % of V (there are only metallic elements consid-
ered). The eutectic MC carbides are formed mainly by vanadium (73.6 ± 0.9 mass %), but
they also contain limited amounts of chromium (9.4 ± 0.5 mass %) and iron (0.8 ± 0.1 mass %).
On the other hand, the chromium content in the matrix is around 5.5 mass % only, and that of
vanadium is correspondingly much lower (up to 1%). The steel also contains small globular
carbides, after SZT in particular, see Figure 3. SEM micrographs in Figure 12c,d clearly
delineate that the boundaries between SGCs and matrix are less intensively attacked by
corrosion; alternatively, they remain well-embedded in the matrix and assist to lower the
corrosion rate at the sites where they are present in sufficiently high amounts.
281
Materials 2021, 14, 3759
Figure 11. Corrosion rate in dependence on tempering temperature for CHT specimens and for
specimens after application of different SZTs.
Figure 12. SEM micrographs showing the surfaces after the potentiodynamic polarisation measurements,
CHT (a), CHT + tempering at 530 ◦ C (b), SZT at −140 ◦ C (c), SZT at −140 ◦ C + tempering at 530 ◦ C (d).
282
Materials 2021, 14, 3759
The SEM image in Figure 13 presents the corrosion-attacked specimen that was
subjected to the SZT at −140 ◦ C. The same features as in Figure 12 are visible. This mainly
concerns the behaviour of different carbides. These carbides are highlighted in EDS maps
of chromium (SCs) and vanadium (ECs). The other EDS maps (chlorine, oxygen, sodium)
clearly demonstrate that the corrosion-exposed surface is covered by products of this
process. However, it is also seen that the corrosion products’ layer is not uniform. While
the matrix is fully covered, the carbides are attacked by the corrosion environment to a
much lower extent.
Figure 13. SEM micrograph showing the surface of the specimen that was subjected to the SZT at −140 ◦ C, after the
potentiodynamic polarisation measurements, and corresponding EDS maps of Cr, V, Na, Cl, and O.
Figure 14 shows the topography and work function mapping of the phases on the
clean surface of the examined steel specimen that was sub-zero-treated at −140 ◦ C. The
white spots on the topography image, in Figure 14a, correspond to the carbide particles.
The corresponding work function map in Figure 14b undoubtedly confirms that these sites
have higher potential than the matrix. The difference is almost fully consistent through
283
Materials 2021, 14, 3759
the whole measured area, and is about 50–60 mV. In other words, these measurements
indicate a more noble behaviour of carbides than the mixture of the martensite and retained
austenite (matrix microstructure).
Figure 14. KPFM images of the examined steel sample that was subjected to the SZT at −140 ◦ C. (a) Topography and
(b) work function mapping. Image size 50 × 50 μm2 .
4. Discussion
The obtained results infer that the corrosion resistance of the Vanadis 6 steel is gener-
ally improved by the application of sub-zero treatments. Additionally, it was demonstrated
that ameliorations in corrosion behaviour are maintained after tempering treatment, even
though the tempering generally deteriorates the corrosion behaviour. The mentioned
variations in corrosion behaviour are the topic of the following discussion.
It has been summarised recently [26] that the application of SZTs evokes a significant
reduction of retained austenite amount, refines the martensite, induces an acceleration
of the precipitation rate of transient carbides, and produces an enhanced number and
population density of small globular carbides.
findings. Moreover, they claimed that the principal explanation of improved corrosion
behaviour may be based on the fact that the reduction of interatomic spacing due to the
compressive stress on the surface facilitates the growth and maintenance of the passivation
film. It is also interesting to note that the introduction of compressive stresses enhances
the corrosion behaviour not only for ferrous alloys, but, for instance, also for aluminium
alloys [34]. Therefore, one can conclude that the high state of compression in the retained
austenite contributes to the overall improvement of corrosion resistance of SZT Vanadis 6
steel, even though the γR amount was significantly reduced by this kind of treatment.
285
Materials 2021, 14, 3759
However, the Cr content in the matrix of Vanadis 6 steel is at around 5.5 mass % after
austenitizing at 1050 ◦ C and quenching, and the results obtained by Gulbrandsen et al. [37]
and by de Waard et al. [38] are hardly comparable with the current ones from this point of
view. Nevertheless, one can assume a much stronger effect of Cr on the corrosion perfor-
mance of steels than that caused by carbon (considering the results of de Waard et al. [38],
for instance); hence, one can expect almost “no effect” of reduced amounts of carbon atoms
solutionised in martensite on the corrosion resistance at 5.5 mass % Cr.
286
Materials 2021, 14, 3759
enhanced chromium content (particle with number 1 as well as two carbides on the right
side of the image). On the other hand, the particles numbered 3 and 5 do not manifest
any significant partitioning of alloying elements, suggesting that they were formed under
diffusion-less conditions.
Figure 15. TEM micrograph showing the carbides in martensitic matrix of the specimen after
quenching followed by SZT at −140 ◦ C (a), EDS map of Cr (b), EDS map of V (c). The sites of
semi-quantitative EDS measurements are labelled and numbered in the TEM image (a).
Table 7. Recorded values of EDS measurements from sites in Figure 16a.
Therefore, enhanced amount and population density of carbides may not inevitably
lead to increasing the overall area ratio of anode (carbides) to cathode (matrix). In addition,
an opposite effect can occur, where increased carbides/matrix surface area ratio may
contribute to the retardation of corrosion since a more stable protective film on the surface
of these carbides can be formed. Experimental investigations of the effect of cementite
on the corrosion resistance of carbon steel provided a good example of a much nobler
response of cementite on corrosion attacks and confirmed improved corrosion behaviour
of the material when coated with Fe3 C [49].
287
Materials 2021, 14, 3759
Figure 16. A schematic of the corrosion attack of the Vanadis 6 steel: before testing (a), after testing, CHT steel (b), SZT
steel—overview (c), detail from (d).
288
Materials 2021, 14, 3759
film on the steels’ surfaces. On the other hand, the precipitation of M3 C carbides with
similar Cr content as the matrix has a less detrimental effect on the growth of protective
films [28].
For the Vanadis 6 steel, it has been demonstrated recently that the SZT accelerates
the precipitation rate of transient cementitic carbides at low tempering temperatures, but
these treatments suppress the precipitation of stable M7 C3 phase during tempering in the
secondary hardening temperature range [19,25].
Figures 8 and 9 show that the corrosion potential, Ecorr , decreases with increasing
tempering temperature for CHT steel, as well as for the steel which was subjected to the
SZT at −140 ◦ C. It should be mentioned here that the potentiodynamic measurements of
specimens subjected to other regimes of SZT (−75, −196, and −269 ◦ C) provided simi-
lar qualitative results. Additionally, it is shown (Table 4) that the corrosion current, Icorr ,
increases rapidly with the tempering, which is clearly reflected in the corrosion rate of
different SZT specimens (Figure 11). The mentioned changes in corrosion behaviour char-
acteristics can be ascribed to the precipitation of different carbides and the corresponding
changes in the matrix. Only cementitic particles were found in the experimental steel
after tempering within the low-tempering temperature range [19,25]. The precipitation
of M3 C does not evoke the Cr depletion of the matrix as the M3 C contain only very low
chromium amount. The only factor that increases the corrosion may be the higher num-
ber of activated sites by forming large amounts of M3 C/matrix boundaries. Increased
tempering temperature leads to precipitation of M7 C3 particles in the case of CHT spec-
imens, which reduces the number of solutionised Cr atoms in the microstructure and
thereby considerably deteriorates the corrosion characteristics. Instead, the precipitation
of M7 C3 carbides was not evidenced after SZTs, and the only consequence of the temper-
ing treatment is the increase in the number of M3 C particles and their coarsening [19].
Hence, the corrosion characteristics of SZT Vanadis 6 steel are less negatively influenced by
high-temperature tempering.
Based on the obtained results, the possible corrosion mechanism of the Vanadis 6
steel in 3.5% NaCl water solution could be delineated. As mentioned above, the steel
contains ECs (vanadium-rich, MC), SCs (chromium rich, M7 C3 ), and certain but very
limited amounts of SGCs (Figure 16a). During the corrosion tests, both the ECs/matrix
and SCs/matrix interface types are extensively attacked by the corrosion environment, and
the carbides are extracted from the surface, which enhances further corrosion (Figure 16b).
Additionally, the matrix is considerably attacked by corrosion in this case, as Figure 12a
illustrates, and the specimen surface manifests significantly enhanced roughness.
Conversely, the examined steel contains considerably enhanced population density of
SGCs after an application of SZTs. The SGCs/matrix interfaces are attacked less extensively
by the corrosion environment (Figures 12c and 16c). Moreover, the area percentage of
carbides increases at the same time by the application of SZTs. The carbides manifest more
noble behaviour than the matrix (Figure 14), and these particles are less covered by the
corrosion products (Figure 13). The resulting effect is that the corrosion rate of the SZT
specimens is lowered (Figure 11), implying that the application of SZT generally improves
the corrosion resistance of the Vanadis 6 steel in 3.5% NaCl water solution.
5. Conclusions
The microstructural changes in Cr-V ledeburitic steel Vanadis 6 with different sub-
zero treatments and tempering were investigated using SEM, TEM, and X-ray diffraction.
The corresponding changes in corrosion resistance in a 3.5% water solution of NaCl were
studied by potentiodynamic polarisation tests, calculation of the corrosion rate, and by
analysis of attacked surfaces by SEM.
The following conclusions can be expressed from the present study:
1. The austenitizing at 1050 ◦ C for 30 min followed by nitrogen gas quenching pro-
duced the microstructure composed of needle-like martensite, retained austenite, and
undissolved carbides.
289
Materials 2021, 14, 3759
Author Contributions: Conceptualization, P.J.; methodology, P.J., M.D., A.B., M.Č., software, A.B.,
D.B.; validation, P.J., A.B.; formal analysis, P.J.; investigation, P.J., M.D., A.B., M.Č.; resources, M.H.;
data curation, P.J., D.B.; writing—original draft preparation, P.J.; writing—review and editing, P.J.,
A.B.; visualization, P.J., M.H., M.D., D.B.; supervision, P.J.; project administration, M.H.; funding
acquisition, P.J. All authors have read and agreed to the published version of the manuscript.
Funding: The authors would like to acknowledge that the article is an outcome implementation of
the following two projects: scientific project VEGA 1/0112/20 and APRODIMET, ITMS: 26220120048,
supported by the Research & Development Operational Programme funded by the European Regional
Development Fund.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: The data presented in this study are available on request from the cor-
responding author. The data are not publicly available due to the fact the this is an ongoing research.
Acknowledgments: The authors would like to acknowledge that the article is an outcome imple-
mentation of the following two projects: scientific project VEGA 1/0112/20 and APRODIMET, ITMS:
26220120048, supported by the Research & Development Operational Programme funded by the
European Regional Development Fund. Special thanks are expressed to Suzanne Morsch from The
University of Manchester, for realizing the Kelvin probe force microscopy.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Jurči, P. Sub-zero treatment of cold work tool steels—Metallurgical background and the effect on microstructure and properties.
HTM J. Heat Treat. Mater. 2017, 72, 62–68. [CrossRef]
2. Akhbarizadeh, A.; Amini, K.; Javadpour, S. Effects of applying an external magnetic field during the deep cryogenic heat
treatment on the corrosion resistance and wear behaviour of 1.2080 tool steel. Mater. Des. 2012, 41, 114–123. [CrossRef]
3. Akhbarizadeh, A.; Javadpour, S.; Amini, K. Investigating the effect of electric current flow on the wear behaviour of 1.2080 tool
steel during the deep cryogenic heat treatment. Mater. Des. 2013, 45, 103–109. [CrossRef]
4. Jurči, P.; Kusý, M.; Ptačinová, J.; Kuracina, V.; Priknerová, P. Long-term sub-zero treatment of P/M Vanadis 6 ledeburitic tool
steel–a preliminary study. Manuf. Technol. 2015, 15, 41–47. [CrossRef]
5. Stratton, P.F. Optimising nano-carbide precipitation in tool steels. Mater. Sci. Eng. A 2007, 449–451, 809–812. [CrossRef]
6. Sobotová, J.; Jurči, P.; Dlouhý, I. The effect of sub-zero treatment on microstructure, fracture toughness, and wear resistance of
Vanadis 6 tool steel. Mater. Sci. Eng. A 2016, 652, 192–204. [CrossRef]
290
Materials 2021, 14, 3759
7. Das, D.; Dutta, A.K.; Ray, K.K. On the enhancement of wear resistance of tool steels by cryogenic treatment. Philos. Mag. Lett.
2008, 88, 801–811. [CrossRef]
8. Amini, K.; Akhbarizadeh, A.; Javadpour, S. Investigating the effect of the quench environment on the final microstructure and
wear behaviour of 1.2080 tool steel after deep cryogenic heat treatment. Mater. Des. 2013, 45, 316–322. [CrossRef]
9. Akhbarizadeh, A.; Shafyei, A.; Golozar, M.A. Effects of cryogenic treatment on wear behaviour of D6 tool steel. Mater. Des. 2009,
30, 3259–3264. [CrossRef]
10. Das, D.; Dutta, A.K.; Ray, K.K. Sub-zero treatments of AISI D2 steel: Part II. Wear behaviour. Mater. Sci. Eng. A 2010, 527,
2194–2206. [CrossRef]
11. Das, D.; Ray, K.K.; Dutta, A.K. Influence of temperature of sub-zero treatments on the wear behaviour of die steel. Wear 2009, 267,
1361–1370. [CrossRef]
12. Surberg, C.H.; Stratton, P.; Lingenhoele, K. The effect of some heat treatment parameters on the dimensional stability of AISI D2.
Cryogenics 2008, 48, 42–47. [CrossRef]
13. Ptačinová, J.; Sedlická, V.; Hudáková, M.; Dlouhý, I.; Jurči, P. Microstructure Toughness relationships in sub-zero treated and
tempered Vanadis 6 steel compared to conventional treatment. Mater. Sci. Eng. A 2017, 702, 241–258. [CrossRef]
14. Jurči, P.; Dománková, M.; Čaplovič, L.; Ptačinová, J.; Sobotová, J.; Salabová, P.; Prikner, O.; Šuštaršič, B.; Jenko, D. Microstructure
and hardness of sub-zero treated and no tempered P/M Vanadis 6 ledeburitic tool steel. Vacuum 2015, 111, 92–101. [CrossRef]
15. Tyshchenko, A.I.; Theisen, W.; Oppenkowski, A.; Siebert, S.; Razumov, O.N.; Skoblik, A.P.; Sirosh, V.A.; Petrov, J.N.; Gavriljuk,
V.G. Low-temperature martensitic transformation and deep cryogenic treatment of a tool steel. Mater. Sci. Eng. A 2010, 527,
7027–7039. [CrossRef]
16. Das, D.; Dutta, A.K.; Ray, K.K. Sub-zero treatments of AISI D2 steel: Part I. Microstructure and hardness. Mater. Sci. Eng. A 2010,
527, 2182–2193. [CrossRef]
17. Das, D.; Ray, K.K. Structure-property correlation of sub-zero treated AISI D2 steel. Mater. Sci. Eng. A 2012, 541, 45–60. [CrossRef]
18. Meng, F.; Tagashira, K.; Azuma, R.; Sohma, H. Role of eta-carbide precipitation’s in the wear resistance improvements of
Fe-12Cr-Mo-V-1.4C tool steel by cryogenic treatment. ISIJ Int. 1994, 34, 205–210. [CrossRef]
19. Jurči, P.; Dománková, M.; Hudáková, M.; Ptačinová, J.; Pašák, M.; Palček, P. Characterization of microstructure and tempering
response of conventionally quenched, short- and long-time sub-zero treated PM Vanadis 6 ledeburitic tool steel. Mater. Charact.
2017, 134, 398–415. [CrossRef]
20. Amini, K.; Akhbarizadeh, A.; Javadpour, S. Effect of Carbide Distribution on Corrosion Behavior of the Deep Cryogenically
Treated 1.2080 Steel. J. Mater. Eng. Perform. 2016, 25, 365–373. [CrossRef]
21. Hill, H.; Huth, S.; Weber, S.; Theisen, W. Corrosion properties of a plastic mould steel with special focus on the processing route.
Mater. Corros. 2011, 62, 436–443. [CrossRef]
22. Bílek, P.; Sobotová, J.; Jurči, P. Evaluation of the microstructural changes in Cr-V ledeburitic tool steel depending on the
austenitization temperature. Mater. Technol. 2011, 45, 489–493.
23. ASTM E975-13. Standard Practice for X-Ray Determination of Retained Austenite in Steel with Near Random Crystallographic
Orientation. In ASTM Book of Standards; ASTM: West Conshohocken, PA, USA, 2004; Volume 3.01.
24. ASTM G 102. Standard Practice for Calculation of Corrosion Rates and Related Information from Electrochemical Measurements.
In ASTM Book of Standards; ASTM: West Conshohocken, PA, USA, 1994.
25. Jurči, P.; Dománková, M.; Ptačinová, J.; Pašák, M.; Kusý, M.; Priknerová, P. Investigation of the Microstructural Changes and
Hardness Variations of Sub-Zero Treated Cr-V Ledeburitic Tool Steel Due to the Tempering Treatment. J. Mater. Eng. Perform.
2018, 27, 1514–1529. [CrossRef]
26. Jurči, P.; Ptačinová, J.; Sahul, M.; Dománková, M.; Dlouhý, I. Metallurgical principles of microstructure formation in sub-zero
treated cold-work tool steels—A review. Mater. Tech. 2018, 106, 104–112. [CrossRef]
27. Lee, J.S.; Fushimi, K.; Nakanishi, T.; Hasegawa, Y.; Park, Y.S. Corrosion behaviour of ferrite and austenite phases on super duplex
stainless steel in a modified green-death solution. Corros. Sci. 2014, 89, 111–117. [CrossRef]
28. Cheng, X.; Wang, Y.; Li, X.; Dong, C. Interaction between austenite-ferrite phases on passive performance of 2205 duplex stainless
steel. J. Mater. Sci. Technol. 2018, 34, 2140–2148. [CrossRef]
29. Lu, S.Y.; Yao, K.F.; Chen, Y.B.; Wang, M.H.; Chen, N.; Gea, X.Y. Effect of quenching and partitioning on the microstructure
evolution and electrochemical properties of a martensitic stainless steel. Corros. Sci. 2016, 103, 95–104. [CrossRef]
30. Dobelaar, J.A.L.; Herman, E.C.M.; De Wit, J.F.W. The influence of the microstructure on the corrosion behaviour of Fe-25Cr. Corros.
Sci. 1992, 33, 779–790. [CrossRef]
31. Villa, M.; Pantleon, K.; Somers, M.A.J. Evolution of compressive strains in retained austenite during sub-zero Celsius martensite
formation and tempering. Acta Mater. 2014, 65, 383–392. [CrossRef]
32. Peyre, P.; Scherpereel, X.; Berthe, L.; Carboni, C.; Fabbro, R.; Béranger, G.; Lemaitre, C. Surface modifications induced in 316L steel
by laser peening and shot-peening. Influence on pitting corrosion resistance. Mater. Sci. Eng. A 2000, 280, 294–302. [CrossRef]
33. Takakuwa, O.; Soyama, H. Effect of Residual Stress on the Corrosion Behavior of Austenitic Stainless Steel. Adv. Chem. Eng. Sci.
2015, 5, 62–71. [CrossRef]
34. Liu, X.; Frankel, G.S. Effects of compressive stress on localized corrosion in AA2024-T3. Corros. Sci. 2006, 48, 3309–3329. [CrossRef]
35. Gavriljuk, V.G.; Theisen, W.; Sirosh, V.V.; Polshin, E.V.; Kortmann, A.; Mogilny, G.S.; Petrov, Y.N.; Tarusin, Y.V. Low-temperature
martensitic transformation in tool steels in relation to their deep cryogenic treatment. Acta Mater. 2013, 61, 1705–1715. [CrossRef]
291
Materials 2021, 14, 3759
36. Fredriksson, H.; Hillert, M.; Nica, M. The decomposition of the M2 C carbide in high speed steel. Scand. J. Metall. 1979, 8, 115–122.
37. Gulbrandsen, E.; Nyborg, R.; Loland, T.; Nisancioglu, K. Effect of Steel Microstructure and Composition on Inhibition of CO2 Corrosion;
Corrosion 2000, Paper Nr. 23; NACE International: Houston, TX, USA, 2000.
38. De Waard, C.; Lotz, U.; Dugstad, A. Influence of Liquid Flow Velocity on CO2 Corrosion. A Semi-Empirical Model; Corrosion 1995,
Paper Nr. 128; NACE International: Houston, TX, USA, 1995.
39. Ansari, T.Q.; Luo, J.L.; Shi, S.Q. Multi-Phase-Field Model of Intergranular Corrosion Kinetics in Sensitized Metallic Materials. J.
Electrochem. Soc. 2020, 167, 061508. [CrossRef]
40. Hong, Y.Y.; Wang, X.Z.; Cadien, K.; Luo, J. Transient Potential Induced Anodic Dissolution of 316L Stainless Steel in Sulfuric Acid
Solution. J. Electrochem. Soc. 2019, 166, C3355–C3363. [CrossRef]
41. Rogal, L.; Dutkiewicz, J.; Szklarz, Z.; Krawiec, H.; Kot, M.; Zimowski, S. Mechanical properties and corrosion resistance of steel
X210CrW12 after semi-solid processing and heat treatment. Mater. Charact. 2014, 88, 100–110. [CrossRef]
42. Kawalec, M.; Krawiec, H. Corrosion Resistance of High-Alloyed White Cast Iron. Arch. Metall. Mater. 2015, 60, 301–303.
[CrossRef]
43. Abd El-Aziz, K.; Zohdy, K.; Saber, D.; Sallam, H.E.M. Wear and Corrosion Behavior of High-Cr White Cast Iron Alloys in Different
Corrosive Media. J. Bio Tribo Corros. 2015, 1, 25. [CrossRef]
44. Tang, X.H.; Chung, R.; Li, D.Y.; Hinckley, B.; Dolman, K. Variations in microstructure of high chromium cast irons and resultant
changes in resistance to wear, corrosion and corrosive wear. Wear 2009, 267, 116–121. [CrossRef]
45. Tang, X.H.; Chung, R.; Pang, C.J.; Li, D.Y.; Hinckley, B.; Dolman, K. Microstructure of high (45 wt.%) chromium cast irons and
their resistances to wear and corrosion. Wear 2011, 271, 1426–1431. [CrossRef]
46. Wiengmoon, A.; Pearce, J.T.H.; Chairuangsri, T. Relationship between microstructure, hardness and corrosion resistance in 20
wt.%Cr, 27 wt.%Cr and 36 wt.%Cr high chromium cast irons. Mater. Chem. Phys. 2011, 125, 739–748. [CrossRef]
47. Ferhat, M.; Benchettara, A.; Amara, S.E.; Naijar, D. Corrosion behaviour of Fe-C alloys in a Sulfuric Medium. J. Mater. Environ.
Sci. 2014, 5, 1059–1068.
48. Ďurica, J.; Ptačinová, J.; Dománková, M.; Čaplovič, L.; Čaplovičová, M.; Hrušovská, L.; Malovcová, V.; Jurči, P. Changes
in microstructure of ledeburitic tool steel due to vacuum austenitizing and quenching, sub-zero treatments at −140 ◦ C and
tempering. Vacuum 2019, 170, 108977. [CrossRef]
49. Wu, J.; Wang, B.; Zhang, Y.; Liu, R.; Xia, Y.; Li, G.; Xue, W. Enhanced wear and corrosion resistance of plasma electrolytic
carburized layer on T8 carbon steel. Mater. Chem. Phys. 2016, 171, 50–56. [CrossRef]
50. Bonagani, S.K.; Bathula, V.; Kaina, V. Influence of tempering treatment on microstructure and pitting corrosion of 13wt.% Cr
martensitic stainless steel. Corros. Sci. 2018, 131, 340–354. [CrossRef]
292
materials
Article
Effect of CO2 Partial Pressure on the Corrosion
Inhibition of N80 Carbon Steel by Gum Arabic
in a CO2-Water Saline Environment for Shale Oil
and Gas Industry
Gaetano Palumbo 1, *, Kamila Kollbek 2 , Roma Wirecka 2,3 , Andrzej Bernasik 3 and
Marcin Górny 4
1 Department of Chemistry and Corrosion of Metals, Faculty of Foundry Engineering, AGH University of
Science and Technology, 30-059 Krakow, Poland
2 Academic Centre for Materials and Nanotechnology, AGH University of Science and Technology,
Mickiewicza St. 30, 30-059 Kraków, Poland; [email protected] (K.K.);
roma.wirecka@fis.agh.edu.pl (R.W.)
3 Department of Condensed Matter Physics, Faculty of Physics and Applied Computer Science, AGH
University of Science and Technology, Mickiewicza St. 30, 30-059 Krakow, Poland; [email protected]
4 Department of Cast Alloys and Composites Engineering, Faculty of Foundry Engineering,
AGH University of Science and Technology, 30-059 Krakow, Poland; [email protected]
* Correspondence: [email protected]; Tel.: +48-12-888-27-63
Abstract: The effect of CO2 partial pressure on the corrosion inhibition efficiency of gum arabic (GA)
on the N80 carbon steel pipeline in a CO2 -water saline environment was studied by using gravimetric
and electrochemical measurements at different CO2 partial pressures (e.g., PCO2 = 1, 20 and 40 bar)
and temperatures (e.g., 25 and 60 ◦ C). The results showed that the inhibitor efficiency increased
with an increase in inhibitor concentration and CO2 partial pressure. The corrosion inhibition
efficiency was found to be 84.53% and 75.41% after 24 and 168 h of immersion at PCO2 = 40 bar,
respectively. The surface was further evaluated by scanning electron microscopy (SEM), energy
dispersive spectroscopy (EDS), grazing incidence X-ray diffraction (GIXRD), and X-ray photoelectron
spectroscopy (XPS) measurements. The SEM-EDS and GIXRD measurements reveal that the surface
of the metal was found to be strongly affected by the presence of the inhibitor and CO2 partial
pressure. In the presence of GA, the protective layer on the metal surface becomes more compact
with increasing the CO2 partial pressure. The XPS measurements provided direct evidence of the
adsorption of GA molecules on the carbon steel surface and corroborated the gravimetric results.
Keywords: high-pressure CO2 corrosion; corrosion inhibition; gum arabic; carbon steel N80
1. Introduction
Shale oil and gas are “unconventional” resources of natural oil and gas trapped in fine-grained
sedimentary rocks called shale. The rapid expansion of shale oil and gas exploration and the
development of a new technology (i.e., hydraulic fracturing (HF) techniques), has seen the popularity
of these natural resources to grow over the years. However, after years of exploitation, the oil and
gas production in the reservoir declines to result in a major economic challenge for the oil companies.
The injection of CO2 at high pressure into the wellbore is an effective method to increase the oil fields
lifetime [1–4]. This process is usually referred to as carbon dioxide flooding enhanced oil recovery
(CO2 -EOR). However, CO2 gas dissolves in the fluid to form the weak carbonic acid, which in turn
dissociates into bicarbonate and carbonate anions [4,5]. The presence of this weak acid can lead to
severe corrosion attacks on the steel structures [4–6].
Another common problem encountered in the extraction of these natural resources is the use of
aggressive fluids with high concentrations of chloride ions (e.g., fracturing fluid) [7]. In the HF process,
the fluid usually injected into the wellbore is a neutral water-based chloride solution (up to 4% of
potassium chloride) with different additives (i.e., inhibitors of scaling, thickening agents, corrosion
inhibitors, etc.) [8]. The literature reports that the presence of a high concentration of chloride ions in a
CO2 -containing fluid can exponentially accelerate the dissolution of the steel [6,9].
Carbon and low-alloys steel are often used in the construction of the pipeline in the shale oil and
gas industry infrastructures, mainly due to its durability, ductility, high strength, and low cost [7,8,10].
However, due to these harsh operating conditions encountered during the exploitation of these natural
resources, the steel is prone to corrode. One practical and relatively cheap method for controlling sweet
corrosion in the shale oil and gas industry is the use of corrosion inhibitors. Corrosion inhibitors are
substances that added to the solution greatly reduce the dissolution of the metal by forming a protective
layer on its surface. The literature reports that over the last decades the use of corrosion inhibitors
as a means to mitigate CO2 corrosion that occurs inside the carbon steel pipelines has received
a wide interest. Nitrogen-based compounds such as pyridine derivatives [11] imidazolines [12],
benzimidazole derivatives [13], and amines [14] were found to be effective corrosion inhibitors against
CO2 corrosion. However, most of these compounds are reported to be toxic and their synthesis can be
very expensive [15,16]. These drawbacks and the increase in environmental awareness have led many
researchers to focus on the use of more naturally occurring substances as corrosion inhibitors. Plant
extracts substances, such as berberine extract [17], Momordica charantia [18], Gingko biloba [19] were
successfully tested as green corrosion inhibitors in CO2 -saturated saline solutions.
The last trend of research has also seen the use of many naturally occurring polymers as green
corrosion inhibitors in various corrosive environments [10,20–22]. They are abundant in nature,
environmentally sustainable, and have an appreciable solubility. Additionally, polymers, unlike small
molecules, with their multiple adsorption sites for bonding on the metal surface, are expected to show
a higher corrosion inhibition efficiency, compared to their monomer counterpart.
Umoren et al. [15] studied the corrosion inhibition effect of two naturally occurring polymers such
as carboxymethyl cellulose and chitosan for API 5 L X60 steel in a CO2 saline solution at PCO2 = 1 bar.
The results showed that both inhibitors reduced the corrosion rate of the metal due to the formation of
a protective layer on its surface. Singh et al. [23] studied the corrosion inhibition effect of a modified
natural polysaccharide (e.g., guar gum + methylmethacrylate) in a 3.5 wt% NaCl solution saturated
with CO2 (e.g., PCO2 = 1 bar) at 50 ◦ C. The authors found that this modified polysaccharide acted like a
good corrosion inhibitor for P110 steel with maximum inhibition efficiency found to be 90%. However,
most of these studies were carried out at atmospheric pressure (e.g., PCO2 = 1 bar). The CO2 -EOR
process can significantly increase the dissolution of the tube. As reported by many studies, the severity
of the CO2 corrosion attack increases with an increase in CO2 partial pressure due to the increase in
acidity of the fluid [4–6]. Therefore, understanding how the CO2 partial pressure can influence the
inhibitory action of certain corrosion inhibitors in CO2 saline environments is important and can help
to minimize the material and economic losses.
Mustafa et al. [4] studied the effect of the CO2 partial pressure (e.g., 10, 40, and 60 bar) on the
corrosion inhibition of an imidazoline-based inhibitor for X52 steel exposed to CO2 water saline
solution at 60 ◦ C. The authors reported that the inhibitor efficiency of the tested inhibitor was observed
to be strongly affected by the concentration of inhibitor and CO2 partial pressure. Ansari et al. [16]
studied the influence of a modified chitosan corrosion inhibitor on J55 carbon steel in a 3.5 wt% NaCl
solution saturated with CO2 at 60 bar and 65 ◦ C, reporting a corrosion inhibition efficiency of 95%. Yet,
all inhibitors tested so far are labeled either as toxic or are expensive to synthesize.
Gum arabic (GA) is a natural polymer obtained from the Acacia trees of the Leguminosae
family [22] and it has been reported to successfully inhibit the corrosion of the steel in different
294
Materials 2020, 13, 4245
2. Experimental Procedure
2.1. Materials
The study was carried out on carbon steel (N80) with composition of (weight %): C 0.39%, Mn
1.80%, Si 0.26%, Cu 0.26%, V 0.19%, Cr 0.04%, Ni 0.04%, Al 0.03%, Mo 0.003%, Co 0.002%, Sn 0.004%,
S 0.001%, P 0.001% and the remainder Fe. Figure 1 shows that the microstructure of the N80 carbon
steel pipeline is composed of perlite and ferrite (α-Fe) phases, where the latter phase accounting for
circa 41% of the total. The samples used in this study were machined from pipeline carbon steel,
ground with silicon carbide abrasive paper up to 1200 grit, then were ultrasonically washed with
distilled water, dried with absolute alcohol.
ȱ
Figure 1. Optical micrographs of the N80 carbon steel microstructures.
All experiments were carried out in 3% of potassium chloride (KCl). Potassium salt was used
in this study instead of the more common NaCl, because in the fracturing fluid, the potassium (K+ )
ions formed a semi-permeable membrane on the shale rock and therefore, preventing the water from
entering the shale.
The tested solution was prepared from reagent grade material potassium chloride (Sigma-Aldrich)
and pure deionized water with an electrical resistivity of 0.055 μS/cm at T = 25 ◦ C. The tested inhibitor
was purchased from Sigma-Aldrich (Warsaw, Poland). The concentrations of inhibitor solution
prepared and used for the study ranged from 0.6 up to 2.0 g L−1 .
295
Materials 2020, 13, 4245
in the presence and absence of different concentrations of the inhibitor (i.e., from 0.6 up to 2.0 g L−1 )
at 25 and 60 ◦ C. Before each experiment, the tested solution was deaerated with CO2 for 2 h under
atmospheric pressure and then CO2 was purged for another 2 h at the tested pressure after the
introduction of the samples. After saturation, the pH and conductivity of the tested solution were 4.5
and 60.30 mS cm−1 at 1 bar and 25 ◦ C, respectively. To ensure homogeneous mixing, a Teflon-coated
blade agitator was used (e.g., 200 rpm). The weight loss was determined by retrieving the coupons
after 24 h of immersion by means of an analytical balance with an accuracy of ±0.1 mg. To assess the
effect of time, the samples were immersed for 168 h in the presence and absence of 1.0 g L−1 of GA at
different CO2 partial pressures (1, 20, and 40 bar). The corrosion products were removed according to
the ASTM G1-90 [30], then the specimens were ultrasonically washed with distilled water, dried with
absolute alcohol, and reweighed. In each case, the experiment was conducted thrice and the corrosion
rate (CR) in mm y−1 was obtained from the following equation:
87.6 Δm
CR (mm y−1 ) = (1)
dAt
where, Δm is the weight loss calculated form the difference between the initial (W i ) and the final (W f )
weight (mg). d is the density (7.87 g cm−3 ), A is the surface of the sample (cm−2 ) and t is the immersion
time (h). The inhibition efficiency (IE%) was determined using the following equation [20,22,26]:
CR − CRinh
IE% = × 100 (2)
CR
where CRinh and CR are the corrosion rates of the steel with and without the inhibitor, respectively.
p − Rp
Rinh
IE% = × 100 (3)
Rinh
p
where Rp inh and Rp are the values of the polarization resistances in the presence and absence of the
inhibitor, respectively. The PDP measurements were carried out at a potential of ±−0.3 V from the
OCP and a scan rate of 1 mV s−1 . The potentiodynamic parameters were determined by means of
Echem Analyst 5.21 software. The values of IE% were calculated from the measured icorr values using
the relationship [21,26]:
icorr − iinh
corr
IE% = × 100 (4)
icorr
where icorr and iinh
corr represent the values of the corrosion current densities without and with
inhibitor, respectively.
296
Materials 2020, 13, 4245
ȱ
Figure 2. Corrosion inhibitor efficiency at different concentrations of gum arabic (GA) and CO2 partial
pressures after 24 h of immersion.
The solubility of CO2 in water increases sharply with increasing the pressure of the system [31].
The high corrosion rate observed at higher CO2 partial pressures can be explained with the increase of
297
Materials 2020, 13, 4245
the acidity of the solution. In fact, in the presence of CO2 , the weak carbonic acid is formed, which in
turn dissociates in HCO− 2−
3 and in CO3 , according to the following reactions:
H2 CO3 ↔ H+ + HCO−
3 (6)
+
HCO−
3 ↔ H + CO3
2−
(7)
The corrosion process in a CO2 containing solution is controlled by the anodic reaction (Equation (8))
and the three cathodic reactions (Equation (9)–(11)) [4,5]:
2HCO− −
3 + 2e → H2 + 2CO3
2−
(10)
The pH of the solution plays an important role in determining the corrosion rate of carbon steel
in a CO2 environment. As the CO2 partial pressure increases, its solubility also increases, resulting
in an increase of the carbonic acid concentration in the solution (Equation (5)). Nesic’ predicted that
the concentrations of H2 CO3 in the solution would increase of about 40 times with changing the
pressure from PCO2 = 1 bar to PCO2 = 40 bar [31]. Increasing the concentration of carbonic acid leads
to an increase in the rate of reduction of carbonic acid and bicarbonate ions (Equations (9) and (10)),
and ultimately the anodic dissolution of the steel (Equation (8)) as reported by several studies [4,5,31].
After the addition of the inhibitor, it can be seen that the corrosion rate of the metal is greatly
reduced going from 1.28 to 0.37 mm y−1 , with a maximum corrosion inhibition efficiency found to be
71.09% at PCO2 = 1 bar, after 24 h of immersion. The data shows that in contrast to the uninhibited
solution, an increase in CO2 partial pressure has a favorable effect on the corrosion rate of the metal
in the presence of the inhibitor. It follows from Figure 2 that IE, which varies inversely with CR,
significantly increased after the addition of GA and with CO2 partial pressure, with a maximum
corrosion inhibition efficiency of 78.77% and 84.53% at PCO2 = 20 bar and PCO2 = 40 bar, after 24 h of
immersion, respectively [4].
The literature reports that GA [7,21], and in general polysaccharides-like inhibitors [15,20],
is mainly adsorbed on the metal surface in acidic condition by weak electrostatic interaction between
the protonated inhibitor molecules and the chloride ions adsorbed on the metal surface. In a weak acid
solution GA molecules are in equilibrium with their protonated molecules according to the following
reaction (see also Section 3.5.1) [7]:
+
GA + xH+ ↔ [GAHx ]x(sol )
(12)
+
where [GAH x ]x(sol )
is the protonated inhibitor in the solution. As mentioned before, an increase in CO2
partial pressure leads to an increase in the acidity of the solution [32]. The higher value of IE observed
at higher CO2 partial pressures can be ascribed to the higher concentration of H+ ions present in the
solution, which in turn leads to an increase in the number of protonated inhibitor molecules that can
be adsorbed on the metal surface. Moreover, Figure 2 also reveals that IE varies with the concentration
of the inhibitor until the system reached a state (e.g., 1.0 g L−1 of GA), in which it can be said that the
inhibitor molecules are in equilibrium with their protonated counterpart. For further increase in GA
concentration, IE remains almost stable. The results clearly demonstrate that GA has greatly reduced
the CR of the metal in the tested environment, and the high corrosion inhibition activity of GA was
influenced by both its concentration and CO2 partial pressure. The lower values of CR observed in the
298
Materials 2020, 13, 4245
presence of the inhibitor can be ascribed to its adsorption on the metal surface, covering the metal
surface and thereby, blocking the active corrosion sites on its surface [4,7,28]. The gravimetric results
are also supported by the SEM analysis presented from Figures 7–9, where it can be seen that the
surface coverage increases and the protective layer becomes more compact in the presence of GA and
with increasing CO2 partial pressure.
As the temperature rises, IE slightly decreased. This decrease may be due to the combination of
two different reasons. For instance, the solubility of CO2 decreases with increasing the temperature of
the solution [31], which can lead to a less acid environment. The pH of the solution increases slightly
and therefore shifting the equilibrium reaction Equation (12) to the left. At higher pH, the concentration
of H+ ions in the solution is smaller, which would result in the formation of fewer protonated inhibitor
molecules available for the absorption process. Another possible reason may be due to the fact that
these types of inhibitors get absorbed via electrostatic interactions (e.g., van der Waal forces) onto
the surface of the metal, and it is known that this types of interaction generally grow weaker with an
increase in temperature due to larger thermal motion [3,20]. Consequently, an increase in temperature
will increase the metal surface kinetic energy, which has a detrimental effect on the adsorption process
and encourages desorption processes [15,20].
Table S2 lists the inhibition efficiency of various corrosion inhibitors used to mitigate sweet
corrosion obtained at different immersion times and temperatures. It is worth mentioning that most of
these inhibitors are labeled either as toxic or are expensive to synthesize. Umoren et al. [15] reported the
corrosion inhibition efficiency of a commercial inhibitor to be 87 and 88% at 25 and 60 ◦ C, respectively
after 24 h of immersion. The table shows that GA, compared to other studied corrosion inhibitors,
and the commercial corrosion inhibitor, can be considered a good environmentally friendly corrosion
inhibitor for carbon steel in a CO2 -saturated saline solution. Moreover, since GA is already used as
a thickening agent in the make-up of the fracturing fluid, can also work as an active component in
corrosion inhibitor in the shale gas industry.
299
Materials 2020, 13, 4245
ȱ ȱ
(a)ȱ (b)ȱ
ȱ ȱ
(c)ȱ (d)ȱ
Figure 3. EIS plot recorded in the presence and absence of 1 g L−1 of GA after 24 h of immersion at
different CO2 -partial pressures. (a) Nyquist; (b) Bode; (c) phase angle; (d) equivalent circuit
Table 1. Electrochemical impedance parameters with and without the presence of 1.0 g L−1
concentrations of GA after 24 h of immersion.
CPEf CPEdl Rp = Rf +
Rf Rct χ2
C (g L−1 ) Rs (Ω cm2 ) Yf (mΩ−1 sn Ydl (mΩ−1 Rct IE (%)
(Ω cm2 ) (Ω cm2 ) (× 10−3 )
cm−2 )
nf
sn cm−2 )
ndl (Ω cm2 )
1 bar Blank 7.13 1.24 0.66 11.59 1.68 0.79 217.90 229.49 1.12 -
1 bar Ga 7.39 0.34 0.61 30.29 0.68 0.81 730.60 760.89 1.50 69.83
20 bar Blank 6.92 1.28 0.68 12.18 1.85 0.76 95.07 107.25 1.34 -
20 bar GA 7.32 0.10 0.78 38.89 0.55 0.82 469.90 507.79 1.28 78.68
40 bar Blank 7.49 0.15 0.70 9.44 3.99 0.78 43.87 53.31 1.11 -
40 bar GA 7.56 0.01 0.71 50.11 0.28 0.81 374.50 424.61 1.99 87.44
The presence of a time constant at HF is reported in several studies [34,35] and it is often observed
in a Fe/water system. This time constant may be due to the capacity of a porous thin layer formed
onto the metal surface. In this study, and without the inhibitor, the presence of this time constant at
HF is due to the formation of a thin layer of Fe3 C onto the metal surface. As mentioned before, the
microstructure of the tested carbon steel is composed of circa 41% of a ferritic phase and the remaining
of a perlitic phase (Figure 1). The ferritic phase is more active than the Fe3 C contained in the perlitic
phase [7], in this case, the former phase will act as an anode and the latter one as a cathode. This will
generate a micro-galvanic effect, which will eventually lead to the formation of a thin layer of Fe3 C
onto the metal surface. However, it follows from the data that both the values of Rf and Rct greatly
increased in the presence of the inhibitor, which indicated that the GA molecules were adsorbed onto
the metal surface leading to the formation of a protective layer that covers the surface, as confirmed
also from the morphological analysis (e.g., SEM-EDS and XPS). Moreover, the difference between
these two values obtained in the absence and the presence of GA increased even more with increasing
CO2 partial pressure, suggesting that this protective layer becomes more stable and compact, with
300
Materials 2020, 13, 4245
the corrosion inhibition efficiency going from 69.83% up to 87.44% at PCO2 = 1 bar and PCO2 = 40 bar,
respectively. The increase in IE observed with an increase in CO2 partial pressure agrees with the
results obtained with the gravimetric measurements and is in agreement with the ones reported in the
literature [4]. It is evident that the addition of GA had a remarkable effect on the corrosion process of
the metal and that its inhibition not only depends on the concentration of GA but also from CO2 partial
pressure. The results show that the coverage and thickness of the formed protective layer increased
with CO2 partial pressure, acting both as a barrier against the charge and the mass transfer processes
that occur onto the metal surface owing to the corrosive attack of the aggressive electrolyte.
Figure 4 and Table 2 show the potentiodynamic polarization measurements and the corrosion
kinetic parameters obtained from the polarization plots in the presence and absence of GA at different
CO2 partial pressures, respectively. As can be seen from Figure 4, the anodic polarization curve of the
blank solution does not show the typical Tafel behavior consequently, the corrosion current densities
were calculated from the extrapolation of the cathodic Tafel region.
ȱ ȱ
(a)ȱ (b)ȱ
ȱ
(c)ȱ
Figure 4. Potentiodynamic polarization parameters obtained in the absence and presence of 1.0 g L−1
of GA at different CO2 -partial pressures, after 24 h of immersion. (a) PCO2 = 1 bar, (b) PCO2 = 20 bar
and (c) PCO2 = 40 bar.
Table 2. Potentiodynamic polarization parameters obtained after 24 h of immersion without and with
1.0 g L−1 of GA.
301
Materials 2020, 13, 4245
The data shows that in absence of GA, the corrosion current density of the steel increased with an
increase in CO2 partial pressure, which is linked with the increased acidity of the solution, in agreement
with the gravimetric experiments. However, it is evident from the data that the corrosion current
density of the steel was prominently reduced after the addition of GA to the solution. Furthermore,
both the cathodic and anodic curves of the polarization curves were shifted towards lower current
densities after the addition of GA. The result suggests that the inhibitor impeded both the rate of the
anodic dissolution (Equation (8)) and the cathodic reactions (Equations (9)–(11)), by either covering
part of the metal surface and/or blocking the active corrosion sites on the steel surface. The dominant
cathodic reaction depends on the pH value of the solution. At lower pH (e.g., less than 4) the reduction
of H+ ions would be the dominant cathodic reaction (Equation (11)). At pH > 4 the dominant cathodic
reaction will be the reduction of HCO− 3 ions and H2 CO3 (Equations (9) and (10)). At higher values
of CO2 partial pressure, GA suppresses the Equation (11) (e.g., the pH of the solution is circa 3.5 at
PCO2 = 40 bar), through the formation of H-bonding between the hydroxyl groups of the inhibitor
units and the H+ ions, adsorbed onto the steel surface, as discussed in more detail in Section 3.5.2.
Moreover, after the addition of GA, the Ecorr can be seen to shift with no definite trend toward
both the anodic and cathodic regions. This result suggested that GA behaves as a mixed type inhibitor
as also reported by other studies for this inhibitor [7,21,27,28].
ȱ
Figure 5. Corrosion inhibitor efficiency obtained at different CO2 partial pressures after 24 and 168 h of
immersion at 25 ◦ C.
This behavior has also been reported by several studies [36,37]. The decrease in IE may be due
to the desorption of the inhibitor from the metal surface, which makes the protective layer unstable.
In this study, the desorption of GA is likely ascribed to its deprotonation due to the consumption of
CO2 from the tested solution because of the electrochemical reactions occurring into the system [5,38].
This leads to a decrease in the acidity of the solution and shifting the Equation (12) towards the
deprotonation of the inhibitor.
302
Materials 2020, 13, 4245
These results confirm that GA is effectively able to protect the steel surface from sweet corrosion
at high CO2 partial pressures even after a prolonged immersion time, reflecting a strong molecular
adsorption of GA on the metal surface and the formation of a stable protective layer.
where R is the gas constant (8.314 J K−1 mol−1 ), T is the absolute temperature (K). The plot of surface
coverage (θ) as a function of the logarithm of the inhibitor concentration at different CO2 partial
pressures is shown in Figure 6.
ȱ
Figure 6. Temkin’s adsorption isotherm for carbon steel (N80) pipeline steel in CO2 -saturated chloride
at different pressures.
The plot of θ vs. log C yields a straight line and the regression coefficient ranges from 0.985
to 0.996. The calculated values of adsorption parameters ΔG◦ ads , a and K at different CO2 partial
pressures are presented in Table 3 and the following notes can be written: (i) The values of ΔG◦ ads
are negative for all three pressures, indicating that the adsorption of GA on the steel surface in the
tested solution is a spontaneous process [7,21,22,28]. Furthermore, the value of ΔG◦ ads ranges between
−10.64 to −8.37 kJ mol−1 indicating that the adsorption of GA on the steel occurs through a physical
adsorption process [7,21,22,28]; (ii) The values of “a” are negative for all three pressures, indicating
that repulsion forces exist between the adsorbed inhibitor molecules in the adsorption layer, as also
303
Materials 2020, 13, 4245
reported by other studies for the same tested inhibitor [7,22]; (iii) The values of Kads increases with an
increase in CO2 partial pressure. It should be noted that Kads denotes the strength between adsorbate
and adsorbent. It can be inferred that a large value of Kads implies a more efficient adsorption process
and thus, a better corrosion inhibition efficiency [21,22]. The results suggest that the adsorption of
GA increases with an increase of the environment pressure, leading to a greater surface coverage and
consequently, a better protection performance.
Table 3. Parameters of the Temkin’s adsorption isotherm calculated from weight loss measurements
after 24 h of immersion time.
Pressure ΔGads
R2 Slope Intercept a Kads
(bar) (kJ mol−1 )
1 0.985 0.483 0.708 −2.38 29.10 −8.37
20 0.995 0.478 0.786 −2.41 44.09 −9.39
40 0.996 0.453 0.844 −2.50 72.97 −10.64
ȱ ȱ
(a)ȱ (b)ȱ
Figure 7. SEM images of the N80 carbon steel surface morphology after 24 h of immersion in
the uninhibited ((a) a lower and b higher magnification) and inhibited ((b) c lower and d higher
magnification) solution at 25 ◦ C and PCO2 = 1 bar CO2 .
304
Materials 2020, 13, 4245
ȱ ȱ
(a)ȱ (b)ȱ
Figure 8. SEM images of the N80 carbon steel surface morphology after 24 h of immersion in
the uninhibited ((a) a lower and b higher magnification) and inhibited ((b) c lower and d higher
magnification) solution at 25 ◦ C and PCO2 = 20 bar CO2 .
ȱ ȱ
(a)ȱ (b)ȱ
Figure 9. SEM images of the N80 carbon steel surface morphology after 24 h of immersion in
the uninhibited ((a) a lower and b higher magnification) and inhibited ((b) c lower and d higher
magnification) solution at 25 ◦ C and PCO2 = 40 bar CO2 .
The severity of the corrosion attack increases with an increase in CO2 partial pressure in the
blank solution, as shown in Figure 8a,b and Figure 9a,b respectively carried out at PCO2 = 20 bar and
PCO2 = 40 bar. However, it can be seen that in the presence of GA, an increase in CO2 partial pressure
led to a gradual increase in the surface coverage on the metal surface, as a result of an increase of
GA molecules adsorbed onto the metal surface (Figure 8c,d and Figure 9c,d). At higher CO2 partial
pressure (e.g., PCO2 = 40 bar, Figure 9) the protective action of the inhibitor is even more evident.
The images show that for the uninhibited solution, the surface of the metal appears severely corroded,
while the one obtained in the presence of the inhibitor shows the formation of a uniform protective
layer over its entire metal surface. The results indicate that in the presence of GA and with a gradual
increase in CO2 partial pressure, the protective layer gradually becomes more compact and thicker [4].
As discussed in Section 3.1, the solubility of CO2 increases with its partial pressure, and as a result of
this, the concentration of H+ ions into the solution also increases, hence the number of the inhibitor
molecules that can be protonated and adsorbed onto the metal surface also increases according to the
Equation (12), leading to a substantial reduction of the corrosion rate of the metal.
The morphology of the metal surface was also analyzed with the help of an energy-dispersive
spectroscopy with the result listed in Table 4. In the absence of GA, the metal surface was characterized
by a corrosion product layer mainly consisting of carbon, iron, and a small amount of oxygen elements,
305
Materials 2020, 13, 4245
indicating that this corrosion layer is mainly composed of Fe3 C. These results are in agreement with
that previously observed in the literature [5,7,39,40]. Other researchers reported that at a temperature
below 40 ◦ C, the corrosion product layer is generally composed of Fe3 C, and only little traces of FeCO3
were observed on the metal surface [4,7,39,40], as also confirmed by the GIXRD measurements shown
in Figure 10. The presence of Fe3 C on the metal surface is due to the anodic dissolution of the ferrite
phase over the cementite in the perlitic phase, which leads to an accumulation of the cementite on the
metal surface.
Weight%
Element
C O Fe Total
Polished 0.70 - 99.30 100
Blank (1 bar) 1.18 - 98.82 100
1.0 g L−1 (1 bar) 4.06 3.51 92.43 100
Blank (20 bar) 7.28 0.83 91.89 100
1.0 g L−1 (20 bar) 8.00 21.98 70.02 100
Blank (40 bar) 4.99 2.05 92.96 100
1.0 g L−1 (40 bar) 9.90 16.21 73.89 100
Figure 10. XRD spectra of corrosion product film formed on the metal surface after been exposed for
24 h without and with the presence of 1.0 g L−1 of GA at different CO2 partial pressures at 25 ◦ C.
It is worth mentioning that in the presence of GA the content of carbon and oxygen was found
to be higher than those observed for the blank solution. It should be noted that carbon and oxygen
are also the main constituents of the tested inhibitor and therefore, their higher concentration on the
protective layer formed in the presence of the inhibitor can be attributed to its adsorption onto the
metal surface, as also reported by other studies [4,7,20]. Moreover, it can be seen from the table that the
percentage of Fe decreased in the presence of GA, likely due to the overlying effect of the inhibitor layer.
The GIXRD analysis for the samples corroded in an inhibited and uninhibited solution at
PCO2 = 40 bar and at 25 ◦ C (Figure 10) shows the presence of cementite on the metal surface, although
in the presence of GA the intensity of these peaks is much weaker. This result can be explained as
follows: Fe3 C accumulates on the metal surface after the dissolution of the ferritic phase. However,
in the presence of the inhibitor, it only accumulates in small amounts on the bare metal surface at the
early stage of the experiment, since the dissolution of the ferritic phase is quickly suppressed by the
absorption of the inhibitor on the surface of the metal.
306
Materials 2020, 13, 4245
Figure 11a,b show the surface morphology for specimens corroded in the blank and inhibited
solution carried out at 60 ◦ C and PCO2 = 40 bar, after immersion the samples for 24 h in the tested
solution, without and with the presence of GA, respectively. The corrosion product layer appears to
be different for the inhibited solution compared to one observed in the presence of GA. Figure 11a
shows the presence of a porous corrosion product layer formed onto the metal surface corroded in a
free-inhibitor solution, pores which create paths for the solution to penetrate it and thereby leading
to the dissolution of the underlying metal. On the other hand, the surface of the metal corroded in
the presence of GA (Figure 11b) shows the formation of a more compact layer, which forms a better
protective barrier and thereby greatly reducing the corrosion rate of the metal.
(a)ȱ (b)ȱ
Figure 11. SEM images of the N80 carbon steel morphology after 24 h of immersion in the tested
solution at PCO2 = 40 bar and at 60 ◦ C, without (a) and with (b) the presence of GA.
EDS analysis reports high content of carbon, oxygen, and iron elements in both layers (C:11.11%,
O:6.02% and C:13.69%, O:11.0%, in the blank and inhibited solution, respectively). The GIXRD
measurements presented in Figure 12 show the characteristic XRD diffraction patterns associated
with FeCO3 . By contrast, the intensity of the iron carbonate peaks observed in the presence of GA is
almost negligible. These results suggest that the layer observed for the uninhibited solution is mainly
composed of Fe3 C and FeCO3 , while in the presence of GA is mainly composed of Fe3 C with little
traces of FeCO3 [4]. Similar behavior was also reported by Ding et al. [41] related to the study of the
effect of an imidazoline-type inhibitor against CO2 corrosion of mild steel. The authors suggested
that the formation of the corrosion inhibitor layer was able to suppress the formation of the iron
carbonate. The precipitation of FeCO3 depends on the concentration of the Fe2+ and CO2− 3 ions, pH,
and temperature. When the concentrations of Fe2+ and CO2− 3 ions exceed the solubility limit, FeCO3
will precipitate on the surface [9,40,42]. At higher temperatures, its solubility decreases, and therefore
the likelihood of its precipitation will be also higher. In a free-inhibitor solution, the dissolution of the
ferrite phase may lead to an increase in the concentration of Fe2+ ions in the bulk solution and thereby
favoring the precipitation of FeCO3 onto the surface of the metal. Conversely, in the presence of the
inhibitor, the protective layer formed onto the surface of the metal slows down the corrosion processes,
and thereby reducing the concentrations of Fe2+ ions available for the formation of FeCO3 .
307
Materials 2020, 13, 4245
Figure 12. XRD spectra of corrosion product film formed on the metal surface after been exposed for
24 h without and with the presence of 1.0 g L−1 of GA at PCO2 = 40 bar and at 60 ◦ C.
Figure 13a–d show the SEM analysis of the metal surface after 168 h of immersion in the absence
and presence of 1.0 g L−1 GA at PCO2 = 40 bar, respectively. It is apparent from the figures that a thick
porous layer covers both surface samples; although it seems that in the presence of the GA, this layer
appears denser, thus providing a higher level of protection. To analyze the condition of the metal
surface, these porous layers were removed with the help of Clark’s solution. It can be seen that both
surfaces show clear signs of corrosion attacks (Figure 13b; however, it is also clear from the figures that
in the presence of the inhibitor (Figure 13d) the surface of the metal appears to be less damaged and
smoother, with the ground scratches still visible on the surface. This result was also confirmed by the
atomic force microscopy experiments performed by Azzaoui et al. [28] concerning the use of GA as a
corrosion inhibitor in a 1 M HCl solution. The authors reported that in the uninhibited solution the
surface of the metal was found to be more corroded with an average roughness of 1.3 μm, while in the
presence of GA the average roughness was reduced to 500 nm. The authors justified this behavior
due to the formation of a more compact protective layer on the metal surface that strongly reduced
the diffusion of the aggressive substances to the metal, and thereby reducing the corrosion rate of
the metal.
ȱ ȱ
(a)ȱ (b)ȱ
Figure 13. SEM images of the N80 carbon steel morphology after 168 h of immersion in the tested
solution in the presence of 1.0 g L−1 of GA at PCO2 = 40 bar. Without (a,b) and with the inhibitor (c,d)
at 25 ◦ C.
308
Materials 2020, 13, 4245
The SEM-EDS and GIXRD result confirm that GA provides adequate protection to the metal
surface from sweet corrosion even at high CO2 partial pressures and after long immersion times.
The results are in agreement with the findings obtained with the weight loss measurements, confirming
the high inhibition efficiency value observed after a long immersion time.
X-ray photoelectron spectroscopy analysis was employed as a means to confirm the adsorption of
the tested inhibitor on the carbon steel surface. The analysis was carried out on the native inhibitor
and the steel surface after 24 h of immersion in the tested solution in the presence of 1.0 g L−1 of GA at
PCO2 = 40 bar and at 25 ◦ C. The XPS results presented in Figure 14a showed evidence of the presence
of O, C, N, and Fe on the carbon steel surface, where the O and C contents displayed the highest
amount, while the signal of N was detected with small intensity. The high-resolution peaks core levels
were analyzed through a deconvolution fitting of the complex spectra. The binding energies and the
corresponding quantification (%) of each peak component are presented in Table S4.
Figure 14. Cont.
309
Materials 2020, 13, 4245
Figure 14. XPS spectra of the native gum Arabic: (a,c,e,g). XPS spectra of the film formed on the
N80 carbon steel after 24 h exposure in CO2 at PCO2 = 40 bar in the presence of 1.0 g L−1 of GA at
25 ◦ C: (b,d,f,h).
The deconvoluted Fe2p3/2 peaks (Figure 14b) at 710.5 and 713.0 eV could be associated with the
α-Fe2 O3 or/and γ- Fe2 O3 [13]. The presence of these species is likely due to the partial decomposition
of iron carbonate. The literature reported that FeCO3 begins decomposing at temperatures below
100 ◦ C according to the following reaction [5,42]:
In the presence of CO2 or water vapor, FeO transforms into Fe3 O4 [5,42].
However, in the presence of oxygen, FeO and Fe3 O4 transform into Fe2 O3 [5,42].
The C1s spectra of the native gum arabic and the adsorbed one (Figure 14c,d, respectively) show
three main peaks. The C1s peak with binding energy at 284.8 eV could be attributed to the C–C/C–H
bonds [26,28]. The C1s peak at 286.2 eV could be attributed to the C–OH/C=O bonds related to the
different groups of GA [26,43]. This peak may also be assigned to the carbon atom bonded to nitrogen
in C–N bond [13,44] and could be related to the glycoprotein and/or to the arabinogalactan-protein
fractions of the inhibitor (Figure 15b,c, respectively). The last C1s peak with a binding energy of
287.7 eV could be associated with the presence of carbonyl type groups O–C=O/N–C=O that result
from the protonation of the GA molecule in the acid environment [28].
It is worth mentioning that no peaks assigned to Fe3 C were found with the XPS analysis in
contrast to the results reported from the GIXRD analysis, where the characteristic peaks assigned to this
compound can be seen in the presence of GA (Figure 10). Fe3 C cannot be detected since the average
depth of analysis for an XPS measurement is approximately 5 nm however, the cementite formed on the
metal surface at the early stage of the experiment is covered by a thicker layer of inhibitor (Figure 9c,d).
The deconvoluted O1s spectra of the native and adsorbed inhibitor are displayed in Figure 14e,f,
respectively. The peaks at 531.2 and 532.7 eV could be attributed to the single bonded oxygen
in C–O and the double bonded oxygen C=O and/or to the single bonded oxygen in O–C–O
respectively [4,13,26,28]. The latter peak may correspond to the carbonyl type groups and/or to
the glycosidic C(1)-O-C(4)/C(1)-O-C(6) linkages of the GA molecules (Figure 15a), as well as, in
310
Materials 2020, 13, 4245
the case of the sample exposed to the tested solution, to FeCO3 formed on the metals surface,
respectively [4,26,28]. Moreover, some authors reported that the peak at 231.2 eV could also be
attributed to the oxygen of the hydroxyl groups (–OH) [5,43], likely due to the hydroxyl groups of
the tested polysaccharide. The O1s spectrum of the adsorbed inhibitor (Figure 14f) displays an extra
peak at 529.7 eV corresponding to O2− related to the oxygen atoms bonded with Fe3+ in the Fe2 O3
oxide [4,43,44]. The O1s results are in good agreement with the findings of the Fe2p spectrum.
ȱ
ȱ
ȱ
ȱ ȱ
(a)ȱ (b)ȱ (c)ȱ
Figure 15. Structure of gum arabic: (a) arabinogalactan; (b) glycoprotein; (c) arabinogalactan-protein.
The presence of N1s peak in the survey for the adsorbed GA on the carbon steel surface (Figure 14a)
provides evidence that gum arabic was effectively adsorbed on the tested substrate surface since the
N80 carbon steel substrate does not contain nitrogen in its chemical composition. The N1s spectra of
the native and adsorbed inhibitor are presented in Figure 14g,h. Both images show the presence of a
peak at 400 and 399.8 eV attributed to the nitrogen atom bonded with the carbon atom, for the native
and adsorbed inhibitor. However, as it can be seen that the high-resolution N1s spectrum of the tested
substrate sample after the addition of GA depicts an extra peak at 397.6 eV. This extra peak can be
ascribed to the coordinated nitrogen atom of the amino group with the metal surface (N–Fe bond) [44].
Other authors also suggested that this peak could be attributed to the bond between the nitrogen of
the amino groups and the oxide layer on the metal surface (FeOx ) [45].
311
Materials 2020, 13, 4245
surface and the protonated inhibitor molecules and therefore, assisting the adsorption of GA on the
metal surface. This type of adsorption mechanism is likely the one that accounts for the most inhibition
action of the inhibitor. In fact, the results presented in this manuscript have demonstrated clearly
that the corrosion inhibition action of GA was strongly influenced by both the concentration of the
inhibitor, CO2 partial pressure, and temperature. A change in one of these two factors has a great
effect on the equilibrium reaction (Equation (12)), shifting the equilibrium towards the protonated
or the deprotonated form of the inhibitor. A shift to the right implies an increase in the number of
protonated molecules of GA available to interact with the chloride ions adsorbed on the surface and
thus, an increase in IE of the system.
ȱ
Figure 16. Schematic representation of the corrosion inhibition mechanism of the N80 carbon steel by
GA. (a) electrostatic; (b) H-bond formation; (c) chemical adsorption.
Hads + H+
(aq)
→ H2(g) (21)
The potentiodynamic measurements presented in Figure 5 showed that the cathodic current
density of the system was greatly reduced after the addition of GA in the solution, suggesting that GA
was able to suppress the hydrogen evolution reaction (Equation (11)) to some extent. Similar results
were also confirmed by other authors [21,26–28]. This assumption was also confirmed by FT-IR and
Raman measurements performed on GA [7,26] and other gum-like [20,36,47] compounds. The results
showed that the characteristic peak assigned to the hydroxyl groups of the carbohydrate units narrowed
down and/or shifted after its adsorption on the metal surface. The authors agreed that this change in
shape was likely due to a possible interaction of the hydroxyl groups of the GA molecules with the
H adsorbed on the cathodic sites of the metal surface via H-bond formation (Figure 16b). Therefore,
the high value of IE observed in this study at different CO2 partial pressure can be also ascribed
312
Materials 2020, 13, 4245
to the ability of GA to suppress one of these reactions (Equations (20)–(22)) via H-bonds formation,
thus suppressing Equation (11) and consequently the dissolution of the steel (Equation (8)).
The adsorption of GA may also be promoted by the presence of the oxide layer on the metal surface
via hydrogen bonding (Figure 16b). Studies concerning the adsorption of GA on oxide nanoparticles
(i.e., iron oxide nanoparticles [48] and zinc or aluminum oxide nanoparticles [49]) reported that GA
showed a strong affinity toward these oxide nanoparticles. The authors suggested that the adsorption
of GA on these oxide nanoparticles surface might be due to the formation of hydrogen bonds between
the functional groups of the GA molecules (e.g., hydroxyl, carboxylate, and amino) with the oxidized
surface. The XPS analysis presented in this study showed that the metal surface after 24 h of exposure
is covered by different oxide species such as Fe2 O3 and/or Fe3 O4 , (e.g., Equations (15)–(19)). Therefore,
the adsorption of GA assisted by the presence of oxide species formed on the metal surface via H-bonds
formation is an adsorption mechanism that must be also taken into account.
4. Conclusions
The corrosion inhibition effect of gum Arabic on the corrosion of carbon steel (N80) exposed in a
high-pressure CO2 -saline environment has been studied and the following conclusion can be drawn:
• The weight loss results showed that the thickening agent gum arabic was found to be an efficient
corrosion inhibitor for carbon steel in a high-pressure CO2 -saline environment. The Inhibition
efficiency increased with an increase in inhibitor concentration and CO2 partial pressure with the
maximum value of IE found to be 84.53% at PCO2 = 40 bar after 24 h of immersion. Moreover,
the weight loss results also showed that GA was effectively able to protect the steel surface from
sweet corrosion at high CO2 partial pressures (i.e., 40 bar) even after a prolonged immersion time
(i.e., 168 h) with a corrosion inhibition efficiency found to be 74.41%.
• The adsorption of GA on the carbon steel surface follows the Temkin’s adsorption isotherm model.
The negative free energy of adsorption ΔG◦ ads indicates a strong and spontaneous adsorption of
GA on the carbon steel surface. Furthermore, the value of ΔG◦ ads indicates that the GA adsorbs
mainly via physical adsorption on the metal surface.
• The SEM analysis revealed that in the presence of GA the protective layer on the metal surface
becomes more compact and dense with an increase in CO2 partial pressure. Also, the SEM analysis
revealed that after 168 h of immersion, in the presence of GA, the metal surface appeared to be
less damaged and smother.
• The XPS results confirmed the formation of a protective layer containing GA molecules and iron
oxides on the metal surface.
313
Materials 2020, 13, 4245
Author Contributions: G.P. conceived, designed, and performed the measurements, analyzed the experimental
data, wrote and edited the manuscript; K.K. performed the XRD analysis; R.W. and A.B. performed the XPS.
analysis; M.G. performed the SEM-EDS analysis. All authors have read and agreed to the published version of
the manuscript.
Funding: This research received no external funding.
Acknowledgments: RW has been partly supported by the EU Project POWR.03.02.00-00-I004/16.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Sheng, J.J. Enhanced oil recovery in shale reservoirs by gas injection. J. Nat. Gas Sci. Eng. 2015, 22, 252–259.
[CrossRef]
2. Bai, H.; Wang, Y.; Ma, Y.; Zhang, Q.; Zhang, N. Effect of CO2 Partial Pressure on the Corrosion Behavior of
J55 Carbon Steel in 30% Crude Oil/Brine Mixture. Materials 2018, 11, 1765. [CrossRef] [PubMed]
3. Bai, H.; Wang, Y.; Ma, Y.; Ren, P.; Zhang, N. Pitting Corrosion and Microstructure of J55 Carbon Steel Exposed
to CO2 /Crude Oil/Brine Solution under 2–15 MPa at 30–80 ◦ C. Materials 2018, 11, 2374. [CrossRef] [PubMed]
4. Mustafa, A.H.; Ari-Wahjoedi, B.; Ismail, M.C. Inhibition of CO2 Corrosion of X52 Steel by Imidazoline-Based
Inhibitor in High Pressure CO2 -Water Environment. J. Mater. Eng. Perform. 2012, 22, 1748–1755. [CrossRef]
5. Islam, A.; Farhat, Z.N. Characterization of the Corrosion Layer on Pipeline Steel in Sweet Environment.
J. Mater. Eng. Perform. 2015, 24, 3142–3158. [CrossRef]
6. Aristia, G.; Hoa, L.Q.; Bäßler, R. Corrosion of Carbon Steel in Artificial Geothermal Brine: Influence of
Carbon Dioxide at 70 ◦ C and 150 ◦ C. Materials 2019, 12, 3801. [CrossRef]
7. Palumbo, G.; Górny, M.; Banaś, J. Corrosion Inhibition of Pipeline Carbon Steel (N80) in CO2 -Saturated
Chloride (0.5 M of KCl) Solution Using Gum Arabic as a Possible Environmentally Friendly Corrosion
Inhibitor for Shale Gas Industry. J. Mater. Eng. Perform. 2019, 28, 6458–6470. [CrossRef]
8. Palumbo, G.; Banas, J.; Bałkowiec, A.; Mizera, J.; Lelek-Borkowska, U. Electrochemical study of the corrosion
behaviour of carbon steel in fracturing fluid. J. Solid State Electrochem. 2014, 18, 2933–2945. [CrossRef]
9. Linter, B.; Burstein, G. Reactions of pipeline steels in carbon dioxide solutions. Corros. Sci. 1999, 41, 117–139.
[CrossRef]
10. Palumbo, G.; Banaś, J. Inhibition effect of guar gum on the corrosion behaviour of carbon steel (K-55) in
fracturing fluid. Solid State Phenom. 2015, 227, 59–62. [CrossRef]
11. Tang, J.; Hu, Y.; Han, Z.; Wang, H.; Zhu, Y.; Wang, Y.; Nie, Z.; Wang, Y. Experimental and Theoretical Study on
the Synergistic Inhibition Effect of Pyridine Derivatives and Sulfur-Containing Compounds on the Corrosion
of Carbon Steel in CO2 -Saturated 3.5 wt.% NaCl Solution. Molecules 2018, 23, 3270. [CrossRef]
12. Ortega-Toledo, D.; Gonzalez-Rodriguez, J.G.; Casales, M.; Martinez, L.; Martinez-Villafañe, A. Co2 corrosion
inhibition of X-120 pipeline steel by a modified imidazoline under flow conditions. Corros. Sci. 2011, 53,
3780–3787. [CrossRef]
13. Singh, A.; Ansari, K.R.; Quraishi, M.A.; Lgaz, H. Effect of Electron Donating Functional Groups on Corrosion
Inhibition of J55 Steel in a Sweet Corrosive Environment: Experimental, Density Functional Theory, and
Molecular Dynamic Simulation. Materials 2018, 12, 17. [CrossRef] [PubMed]
14. Ghareba, S.; Omanovic, S. The effect of electrolyte flow on the performance of 12-aminododecanoic acid
as a carbon steel corrosion inhibitor in CO2 -saturated hydrochloric acid. Corros. Sci. 2011, 53, 3805–3812.
[CrossRef]
15. Umoren, S.; Alahmary, A.A.; Gasem, Z.M.; Solomon, M. Evaluation of chitosan and carboxymethyl cellulose
as ecofriendly corrosion inhibitors for steel. Int. J. Boil. Macromol. 2018, 117, 1017–1028. [CrossRef] [PubMed]
16. Ansari, K.; Chauhan, D.S.; Quraishi, M.; Mazumder, M.A.; Singh, A. Chitosan Schiff base: An environmentally
benign biological macromolecule as a new corrosion inhibitor for oil & gas industries. Int. J. Boil. Macromol.
2020, 144, 305–315. [CrossRef]
17. Lin, Y.; Singh, A.; Ebenso, E.E.; Quraishi, M.A.; Zhou, Y.; Huang, Y. Use of HPHT Autoclave to Determine
Corrosion Inhibition by Berberine extract on Carbon Steels in 3.5% NaCl Solution Saturated with CO2 .
Int. J. Electrochem. Sci. 2015, 10, 194–208.
18. Singh, A.; Lin, Y.; Liu, W.; Ebenso, E.E.; Pan, J. Extract of Momordica charantia (Karela) Seeds as Corrosion
Inhibitor for P110SS Steel in CO2 Saturated 3.5% NaCl Solution. Int. J. Electrochem. Sci. 2013, 8, 12884–12893.
314
Materials 2020, 13, 4245
19. Singh, A.; Lin, Y.; Ebenso, E.E.; Liu, W.; Pan, J.; Huang, B. Gingko biloba fruit extract as an eco-friendly
corrosion inhibitor for J55 steel in CO2 saturated 3.5% NaCl solution. J. Ind. Eng. Chem. 2015, 24, 219–228.
[CrossRef]
20. Palumbo, G.; Berent, K.; Proniewicz, E.; Banaś, J. Guar Gum as an Eco-Friendly Corrosion Inhibitor for Pure
Aluminium in 1-M HCl Solution. Materials 2019, 12, 2620. [CrossRef]
21. Bentrah, H.; Rahali, Y.; Chala, A. Gum Arabic as an eco-friendly inhibitor for API 5L X42 pipeline steel in
HCl medium. Corros. Sci. 2014, 82, 426–431. [CrossRef]
22. Umoren, S. Inhibition of aluminium and mild steel corrosion in acidic medium using Gum Arabic. Cellulose
2008, 15, 751–761. [CrossRef]
23. Singh, A.; Ansari, K.; Quraishi, M. Inhibition effect of natural polysaccharide composite on hydrogen evolution
and P110 steel corrosion in 3.5 wt% NaCl solution saturated with CO2 : Combination of experimental and
surface analysis. Int. J. Hydrogen Energy 2020, 45, 25398–25408. [CrossRef]
24. Umoren, S.; Ogbobe, O.; Igwe, I.; Ebenso, E. Inhibition of mild steel corrosion in acidic medium using
synthetic and naturally occurring polymers and synergistic halide additives. Corros. Sci. 2008, 50, 1998–2006.
[CrossRef]
25. Mobin, M.; Alam Khan, M. Investigation on the Adsorption and Corrosion Inhibition Behavior of Gum
Acacia and Synergistic Surfactants Additives on Mild Steel in 0.1 MH2 SO4 . J. Dispers. Sci. Technol. 2013, 34,
1496–1506. [CrossRef]
26. Abu-Dalo, M.A.; Othman, A.A.; Al-Rawashdeh, N.A.F. Exudate gum from acacia trees as green corrosion
inhibitor for mild steel in acidic media. Int. J. Electrochem. Sci. 2012, 7, 9303–9324.
27. Shen, C.; Alvarez, V.; Koenig, J.D.B.; Luo, J.-L. Gum Arabic as corrosion inhibitor in the oil industry:
Experimental and theoretical studies. Corros. Eng. Sci. Technol. 2019, 54, 444–454. [CrossRef]
28. Azzaoui, K.; Mejdoubi, E.; Jodeh, S.; Lamhamdi, A.; Rodríguez-Castellón, E.; Algarra, M.; Zarrouk, A.;
Errich, A.; Salghi, R.; Lgaz, H. Eco friendly green inhibitor Gum Arabic (GA) for the corrosion control of
mild steel in hydrochloric acid medium. Corros. Sci. 2017, 129, 70–81. [CrossRef]
29. Spellman, F.R. Environmental Impacts of Hydraulic Fracturing; Informa UK Limited: London, UK, 2012.
30. ASTM-G1-90, Standard Practice for Preparing, Cleaning, and Evaluation Corrosion Test Specimens; ASTM
International: West Conshohocken, PA, USA, 1999.
31. Choi, Y.-S.; Nešić, S. Determining the corrosive potential of CO2 transport pipeline in high pCO2 –water
environments. Int. J. Greenh. Gas. Control 2011, 5, 788–797. [CrossRef]
32. Li, X.; Peng, C.; Crawshaw, J.; Maitland, G.; Trusler, J.M. The pH of CO2 -saturated aqueous NaCl and
NaHCO3 solutions at temperatures between 308 K and 373 K at pressures up to 15 MPa. Fluid Phase Equilibria
2018, 458, 253–263. [CrossRef]
33. Dong, B.; Liu, W.; Zhang, Y.; Banthukul, W.; Zhao, Y.; Zhang, T.; Fan, Y.; Li, X.; Wei, L.; Yonggang, Z.; et al.
Comparison of the characteristics of corrosion scales covering 3Cr steel and X60 steel in CO2 -H2S coexistence
environment. J. Nat. Gas Sci. Eng. 2020, 80, 103371. [CrossRef]
34. Bousselmi, L.; Fiaud, C.; Tribollet, B.; Triki, E. Impedance spectroscopic study of a steel electrode in condition
of scaling and corrosion. Electrochim. Acta 1999, 44, 4357–4363. [CrossRef]
35. Bousselmi, L.; Fiaud, C.; Tribollet, B.; Triki, E. The characterisation of the coated layer at the interface carbon
steel-natural salt water by impedance spectroscopy. Corros. Sci. 1997, 39, 1711–1724. [CrossRef]
36. Roy, P.; Karfa, P.; Adhikari, U.; Sukul, D. Corrosion inhibition of mild steel in acidic medium by polyacrylamide
grafted Guar gum with various grafting percentage: Effect of intramolecular synergism. Corros. Sci. 2014, 88,
246–253. [CrossRef]
37. Saha, S.K.; Dutta, A.; Sukul, D.; Ghosh, P.; Banerjee, P. Adsorption and corrosion inhibition effect of Schiff
base molecules on the mild steel surface in 1 M HCl medium: A combined experimental and theoretical
approach. Phys. Chem. Chem. Phys. 2015, 17, 5679–5690. [CrossRef]
38. Outirite, M.; Lagrenée, M.; Lebrini, M.; Traisnel, M.; Jama, C.; Vezin, H.; Bentiss, F. Ac impedance, X-ray
photoelectron spectroscopy and density functional theory studies of 3,5-bis(n-pyridyl)-1,2,4-oxadiazoles as
efficient corrosion inhibitors for carbon steel surface in hydrochloric acid solution. Electrochim. Acta 2010, 55,
1670–1681. [CrossRef]
39. Paolinelli, L.; Perez, T.; Simison, S. The effect of pre-corrosion and steel microstructure on inhibitor
performance in CO2 corrosion. Corros. Sci. 2008, 50, 2456–2464. [CrossRef]
315
Materials 2020, 13, 4245
40. Mora-Mendoza, J.; Turgoose, S. Fe3C influence on the corrosion rate of mild steel in aqueous CO2 systems
under turbulent flow conditions. Corros. Sci. 2002, 44, 1223–1246. [CrossRef]
41. Ding, Y.; Brown, B.; Young, D.; Singer, M. Effectiveness of an Imidazoline-Type Inhibitor Against CO2
Corrosion of Mild Steel at Elevated Temperatures (120 ◦ C–150 ◦ C). In Proceedings of the CORROSION 2018,
Phoenix, AZ, USA, 15–19 April 2018; p. 22.
42. Heuer, J.; Stubbins, J. An XPS characterization of FeCO3 films from CO2 corrosion. Corros. Sci. 1999, 41,
1231–1243. [CrossRef]
43. Boumhara, K.; Tabyaoui, M.; Jama, C.; Bentiss, F. Artemisia Mesatlantica essential oil as green inhibitor for
carbon steel corrosion in 1M HCl solution: Electrochemical and XPS investigations. J. Ind. Eng. Chem. 2015,
29, 146–155. [CrossRef]
44. Bouanis, M.; Tourabi, M.; Nyassi, A.; Zarrouk, A.; Jama, C.; Bentiss, F. Corrosion inhibition performance
of 2,5-bis(4-dimethylaminophenyl)-1,3,4-oxadiazole for carbon steel in HCl solution: Gravimetric,
electrochemical and XPS studies. Appl. Surf. Sci. 2016, 389, 952–966. [CrossRef]
45. Hashim, N.Z.N.; Anouar, E.H.; Kassim, K.; Zaki, H.M.; Alharthi, A.I.; Embong, Z. XPS and DFT investigations
of corrosion inhibition of substituted benzylidene Schiff bases on mild steel in hydrochloric acid. Appl. Surf. Sci.
2019, 476, 861–877. [CrossRef]
46. Barker, R.; Burkle, D.; Charpentier, T.; Thompson, H.; Neville, A. A review of iron carbonate (FeCO3 )
formation in the oil and gas industry. Corros. Sci. 2018, 142, 312–341. [CrossRef]
47. Messali, M.; Lgaz, H.; Dassanayake, R.; Salghi, R.; Jodeh, S.; Abidi, N.; Hamed, O. Guar gum as efficient
non-toxic inhibitor of carbon steel corrosion in phosphoric acid medium: Electrochemical, surface, DFT and
MD simulations studies. J. Mol. Struct. 2017, 1145, 43–54. [CrossRef]
48. Williams, D.N.; Gold, K.A.; Holoman, T.R.P.; Ehrman, S.H.; Wilson, O.C. Surface Modification of Magnetic
Nanoparticles Using Gum Arabic. J. Nanopart. Res. 2006, 8, 749–753. [CrossRef]
49. Leong, Y.; Seah, U.; Chu, S.; Ong, B. Effects of Gum Arabic macromolecules on surface forces in oxide
dispersions. Colloids Surf. A Physicochem. Eng. Asp. 2001, 182, 263–268. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
316
materials
Article
Effect of Phase Transformation on Stress Corrosion Behavior
of Additively Manufactured Austenitic Stainless Steel
Produced by Directed Energy Deposition
Tomer Ron *, Ohad Dolev, Avi Leon, Amnon Shirizly and Eli Aghion7
Abstract: The present study aims to evaluate the stress corrosion behavior of additively manufactured
austenitic stainless steel produced by the wire arc additive manufacturing (WAAM) process. This was
examined in comparison with its counterpart, wrought alloy, by electrochemical analysis in terms of
potentiodynamic polarization and impedance spectroscopy and by slow strain rate testing (SSRT) in
a corrosive environment. The microstructure assessment was performed using optical and scanning
electron microscopy along with X-ray diffraction analysis. The obtained results indicated that in
spite of the inherent differences in microstructure and mechanical properties between the additively
manufactured austenitic stainless steel and its counterpart wrought alloy, their electrochemical
performance and stress corrosion susceptibility were similar. The corrosion attack in the additively
manufactured alloy was mainly concentrated at the interface between the austenitic matrix and
the secondary ferritic phase. In the case of the counterpart wrought alloy with a single austenitic
phase, the corrosion attack was manifested by uniform pitting evenly scattered at the external surface.
Both alloys showed ductile failure in the form of “cap and cone” fractures in post-SSRT experiments
Citation: Ron, T.; Dolev, O.; Leon, A.; in corrosive environment.
Shirizly, A.; Aghion, E. Effect of
Phase Transformation on Stress Keywords: additive manufacturing; direct energy deposition; wire arc additive manufacturing;
Corrosion Behavior of Additively 316L stainless steel; stress corrosion
Manufactured Austenitic Stainless
Steel Produced by Directed Energy
Deposition. Materials 2021, 14, 55.
https://dx.doi.org/10.3390/ 1. Introduction
ma14010055
Additive manufacturing (AM) has been considered to be a promising technology for
producing a variety of complex components in a relatively short time [1–6]. Traditional AM
Received: 23 November 2020
processing of metals mainly focuses on powder bed fusion (PBF) methods such as selective
Accepted: 21 December 2020
laser melting (SLM) and electron beam melting (EBM) [7–10]. However, PBF technologies
Published: 24 December 2020
are relatively expensive due to the high cost of raw material, high energy consumption and
Publisher’s Note: MDPI stays neu-
relatively low deposition rate. In addition, the size of the printed component is limited and
tral with regard to jurisdictional claims depends on the printing cell dimension. The inherent disadvantages of PBF technologies
in published maps and institutional highlight the need to use more affordable AM methods such as the wire arc additive
affiliations. manufacturing (WAAM) process. Comparatively, proven PBF technologies can produce a
deposition rate of 0.1 kg/h, while the deposition rate of WAAM is about 10 kg/h [11–13].
In addition, the use of relatively inexpensive wires as raw materials and an electric arc as the
energy source can reduce the cost of the printing process by 80% compared to PBF [14,15].
Copyright: © 2020 by the authors. Li- Furthermore, the dimensions of components produced by WAAM are almost unlimited [16].
censee MDPI, Basel, Switzerland. This
It should be pointed out that WAAM process can be also implemented using computer
article is an open access article distributed
numerical control (CNC) systems [17]. The almost unlimited dimensions of WAAM is due
under the terms and conditions of the
to the fact that the printing process can be performed by an external robot that is free to
Creative Commons Attribution (CC BY)
move in all directions [18–20]. However, there are some limitations related to the WAAM
license (https://creativecommons.org/
process compared to PBF technology. This includes relatively increased surface roughness,
licenses/by/4.0/).
317
and limited capabilities to produce complex structures. Currently, most of the research
activities related to the WAAM process have focused on optimizing the printing parameters
and residual stresses status [21–25], with very limited attention paid to the corrosion
performance of the obtained components. This study mainly aims at evaluating the effect
of phase transformation on stress corrosion behavior of additively manufactured austenitic
stainless steel in the form of 316L alloy produced by the WAAM process. For reference
consideration, the obtained stress corrosion behavior was compared to its counterpart
wrought alloy AISI 316L. The general corrosion performance was evaluated in terms of
potentiodynamic polarization and electrochemical impedance spectroscopy (EIS) analyses
all in 3.5% NaCl solution.
Figure 1. General appearance of the cylindrical component along with a tension specimen obtained
by the WAAM process.
318
Materials 2021, 14, 55
3. Results
The chemical composition of the welding wire, printed alloy and counterpart AISI
316L stainless steel alloy are shown in Table 1. This reveals that the composition of the
printed alloy was in line with the composition of the welding wire and quite similar to that
of the counterpart AISI alloy in terms of main alloying elements and carbon content.
Table 1. Chemical composition (wt.%) of welding wire, printed stainless steel part and counterpart
AISI 316L.
Material C Mn Si S P Cr Ni Cu Mo Co N Fe
Welding Wire 0.013 1.97 0.51 0.0010 0.018 19.22 11.50 0.15 2.42 0.07 0.092 Bal.
Printed 316L 0.024 1.85 0.44 0.0007 0.020 19.21 11.62 0.09 2.48 0.26 0.052 Bal.
AISI 316L 0.021 1.59 0.40 0.02 0.037 16.58 10.07 0.44 2.04 0.19 0.078 Bal.
X-ray diffraction analysis of the printed alloy and its counterpart AISI 316L are shown
in Figure 2. This reveals that the printed alloy was composed from an austenitic matrix
(γ-Fe) and a secondary ferritic phase (δ-Fe), as expected from a regular 316L printed al-
loy [31] while the stock material (wire) has one austenitic phase. In parallel, the counterpart
AISI 316L was composed from only an austenitic phase. In addition, significant differ-
ences were obtained between the XY-plane (building direction 0◦ ) and XZ-plane (building
direction 90◦ ) of the printed alloy in terms of peak intensity. This can be attributed to
the epitaxial characteristics of the AM process that display a preferred orientation of the
solidification process. The calculated lattice parameter of the γ-Fe and δ-Fe related to
the printed alloy were 2.88 and 3.59 Å, respectively, which basically comes in line with
the parameters found in the literature: 2.86 Å (PDF 006–096) and 3.59 Å (PDF 33–0397).
In the case of the counterpart AISI 316L, the calculated lattice parameter was 3.59 Å, as ex-
pected. Furthermore, it should be pointed out that, in contrast to the microstructure of
316L obtained by the WAAM process as presented by X. Chen et al. [25,32], no σ-phase
was observed in this study.
Typical microstructure of printed 316L stainless steel in the XY- and XZ-planes is
shown in Figure 3 while the microstructure of its counterpart AISI 316L alloy is introduced
in Figure 4. The microstructure of the printed alloy was composed from an austenitic
matrix and a secondary ferritic phase at the grain boundaries. The ferritic dendrites in the
XY-plane present an anisotropic morphology, while the XZ-plane introduces an epitaxial
solidification characteristic. In parallel, the counterpart AISI 316L was composed of a single
austenitic phase, as expected from conventional 316L stainless steel [33]. The two-phase
microstructure of the printed alloy was formed due to the high solidification rate of the AM
process. This can also be explained in terms of the Fe-Cr-Ni phase diagram [34] operating
under high cooling rate conditions. In addition, the inherent re-heating effect of the AM
319
Materials 2021, 14, 55
Figure 2. X-ray diffraction analysis of printed alloy and AISI 316L alloy.
Figure 3. Typical microstructures of printed 316L stainless steel obtained by optical microscopy (a) XY-plane, (b) XZ-plane.
Figure 4. Typical microstructures of AISI 316L stainless steel (a) longitudinal cross section, (b) transvers cross section.
The general macrostructure of the printed alloy in 3D is shown in Figure 5, along with
the corresponding close-up microstructures and spot chemical analysis. This clearly reveals
320
Materials 2021, 14, 55
that the microstructure in the XY-plane was quite uniform compared to the non-uniform
structure in the XZ-plane that relates to the preferred orientation of the solidification
course. In addition, the melt pool boundaries shown in Figure 5d,e clearly illustrate the
epitaxial nature of solidification in the XZ-plane. The spot chemical analyses at points 1 and
2 (Table 2) disclose the typical compositions of austenite and ferrite phases, respectively.
These compositions reflect the relatively increased amount of Ni and reduced content of Cr
in the austenitic phase and vice versa in the ferritic phase.
Figure 5. Macrostructure and corresponding microstructures of printed 316L alloy obtained by stereo microscope and SEM
(a) general macrostructure, (b,c) microstructure in the XY-plane, (d,e) microstructure in the XZ-plane.
Table 2. Spot chemical analysis of printed 316L at points 1 and 2 (Figure 5c) by EDS in wt.%.
Test
C Mn Si Cr Ni Mo Fe
Point
Point 1 0.39 2.3 0.41 17.34 12.08 2.02 Bal.
Point 2 0.25 2.12 0.39 24.86 5.43 4.17 Bal.
The mechanical properties of printed 316L and its counterpart AISI 316L in terms of
tensile strength, yield strength, elongation and hardness are shown in Table 3 along with
their typical stress-strain curves shown in Figure 6. This reveals that the strength of the
printed alloy was relatively reduced, while its ductility was increased compared to the
counterpart AISI 316L.
Table 3. The mechanical properties—UTS (Ultimate tensile strength), YP (yield point) elongation, hardness and density of
printed and AISI 316L stainless steel.
Material UTS (MPa) YP (MPa) Elongation (%) Hardness (HV) Density (gr/cm3 )
Printed 316L 552 ± 11 364 ± 17 42 ± 1 196 ± 5 7.6 ± 0.3
AISI 316L 752 ± 3 695 ± 3 37 ± 1 275 ± 9 7.8 ± 0.2
The corrosion resistance of the printed and counterpart AISI 316L in terms of poten-
tiodynamic polarization analysis is shown in Figure 7. Although the polarization curve of
the printed alloy was relatively shifted to higher corrosion currents, which reflects reduced
corrosion resistance [35,36], its break potential (Eb1 ) was relatively higher compared to the
AISI 316L (Eb2 ), which reflects an improved passivation process. Altogether, the corrosion
rate of the printed alloy in terms of Tafel extrapolation was excellent, and quite similar
to the counterpart alloy as shown in Table 4 (0.005 vs. 0.001 mmpy, respectively). In ad-
dition, as expected from metals having an active-passive transition, both alloys showed
localized corrosion attack. In the case of the printed alloy the localized corrosion attack was
concentrated at the boundaries between the austenite and ferrite phases (Figure 8a,b) [37].
321
Materials 2021, 14, 55
The corrosion attack in the counterpart AISI 316L was in the form of pitting corrosion
(Figure 8c,d) that was evenly scattered on the external surface. The counterpart alloy
presented typical pitting morphology for AISI 316L, as well as typical corrosion potential,
breakdown potential and corrosion current [38].
Figure 6. Typical stress-strain curves of printed 316L alloy and counterpart AISI 316L.
Figure 7. Potentiodynamic polarization analysis of printed 316L alloy and counterpart AISI 316L in
3.5% NaCl solution.
Table 4. Corrosion rate of tested alloys as obtained by Tafel extrapolation from potentiodynamic polarization curves.
The corrosion performance of printed and AISI 316L obtained by EIS analysis are
shown in Figure 9. The Nyquist diagrams of both alloys (Figure 9a) in terms of curve radius,
which represents the surface corrosion resistance, were quite similar. This similarity was
also maintained by the Bode magnitude diagram (Figure 9b) that introduces the solution re-
sistance. The related electrical equivalent circuit and corresponding fitting parameters (R1-
solution resistance, R2 and Q1-capacitor) [39,40] are introduced in Figure 10 and Table 5,
respectively. Altogether, the EIS analysis clearly indicates that the corrosion resistance of
printed and counterpart alloys was quite similar.
322
Materials 2021, 14, 55
Figure 8. (a,b) Close-up views of the corrosion attack at surface of printed alloy, (c,d) close-up views
of the corrosion attack at surface of counterpart AISI alloy.
Figure 9. Electrochemical impedance spectroscopy analysis of printed and AISI 316L in 3.5% NaCl solution: (a) Nyquist
diagram, (b) Bode diagram.
Figure 10. Electrical equivalent circuits for the EIS analysis shown in Figure 9.
Table 5. Corresponding fitting parameters for the EIS analysis shown in Figure 9.
323
Materials 2021, 14, 55
The stress corrosion behavior of the printed and counterpart AISI 316L in terms of
SSRT in 3.5% NaCl solution, are shown in Figures 8–10. Although the stress corrosion
performance of the two alloys was similar, as reflected by nearly equal time to failure vs.
strain rate (Figure 11), the two alloys maintain their inherent UTS and elongation properties
(Figures 12 and 13, respectively). This similarity could also be seen by the fitting equations
.m
(σUTS = C × ε ) of UTS vs. strain rate According to these fittings, the stain rate sensitively
factors (m) of the two alloys were very close: 0.007 and 0.009 for the printed and AISI 316L,
respectively. Fractography analysis of the two alloys (Figure 14a–d) clearly demonstrates
that both alloys showed ductile failure behavior in the form of “cap and cone” fractures,
as expected from 316L stainless steel alloy.
Figure 11. The effect of strain rate on time to failure of 316L produced by WAAM process in
comparison with conventional wrought alloy AISI 316L.
Figure 12. Ultimate tensile strength (UTS) versus strain rate of 316L produced by WAAM process
compared to its counterpart wrought alloy AISI 316L.
324
Materials 2021, 14, 55
Figure 13. Ductility in terms of elongation versus strain rate of 316L produced by WAAM process
compared to wrought alloy AISI 316L.
Figure 14. Fracture surface after SSRT at a strain rate of 2.5 × 10−7 (1/S) in 3.5% NaCl solution. (a,b) printed alloy,
(c,d) counterpart AISI 316L.
4. Discussion
In spite of the differences between the microstructure and mechanical properties of
WAAM 316L alloy and its counterpart AISI 316L, their corrosion performance in 3.5% NaCl
solution was quite similar. This was strongly supported by the results of potentiodynamic
polarization, EIS and stress corrosion analysis in terms of SSRT examination. Neverthe-
less, the corrosion mechanism of the two alloys was slightly different. This was clearly
demonstrated by the surface corrosion attack shown in Figure 8. According to this figure,
the localized corrosion attack in the printed alloy was mainly located at the interface be-
tween the austenitic matrix and the secondary ferritic phase (Figure 8a,b). This was mainly
attributed to the relatively reduced corrosion resistance of the ferritic phase compared
to the austenitic phase, which can induce micro-galvanic corrosion. In the case of the
counterpart AISI 316L, the localized corrosion attack was in the form of pitting corrosion
that was uniformly scattered on the surface (Figure 8c,d).
325
Materials 2021, 14, 55
Regarding the effect of strain rate on UTS and elongation under a corrosive envi-
ronment, both the printed and the counterpart AISI 316L displayed a similar response,
according to their inherent mechanical properties. The similarity in their stress corro-
sion resistance was demonstrated by nearly equal time to failure at a slow strain rate of
2.5 × 10−7 s−1 (Figure 11), where the environmental effect was most dominant. This simi-
larity was also manifested by the fractography analysis of the two alloys (Figure 14a–d)
that clearly showed ductile failure characteristics in the form of “cap and cone” fractures,
as can be expected from 316L stainless steel alloy [41].
As a final remark it should be pointed out that the similar corrosion performance of
printed WAAM 316L alloy and its counterpart AISI 316L in 3.5% NaCl solution cannot
be simply extrapolated to any different environment. This is mainly due to the inherent
differences between the microstructure of the two alloys that can affect their corrosion
behavior primarily in a more aggressive environment.
5. Conclusions
The stress corrosion behavior of additively manufactured austenitic stainless steel
(316L alloy) produced by WAAM process in terms of stress corrosion susceptibility and
electrochemical performance was similar to that of its counterpart wrought alloy. This sim-
ilarity was obtained in spite of the inherent differences in microstructure and mechanical
properties of the two alloys. The corrosion attack in the printed alloy was mainly located
at the interface between the austenitic matrix and the ferritic phase, while that of the
counterpart alloy composed of a single austenite phase was in the form of pitting corrosion
uniformly scattered on the surface. The fractography analysis of the two alloys in post-SSRT
experiments revealed that both alloys showed ductile failure in the form of “cap and cone”
fractures, as expected from austenitic stainless steel.
Author Contributions: E.A., A.S. and T.R. conceived, designed and performed the experiments; A.L.
and O.D. assisted in analyzing the data; E.A. and T.R. writing—original draft preparation and writing—
review and editing. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data sharing is not applicable to this article.
Acknowledgments: The authors thank the Kreitman School for Advanced Studies at Ben-Gurion
University of the Negev for their financial contribution in support of this research.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Leon, A.; Levy, G.K.; Ron, T.; Shirizly, A.; Aghion, E. The effect of strain rate on stress corrosion performance of Ti6Al4V alloy
produced by additive manufacturing process. J. Mater. Res. Technol. 2020. [CrossRef]
2. Rodrigues, T.A.; Duarte, V.; Avila, J.A.; Santos, T.G.; Miranda, R.; Oliveira, J. Wire and arc additive manufacturing of HSLA steel:
Effect of thermal cycles on microstructure and mechanical properties. Addit. Manuf. 2019, 27, 440–450. [CrossRef]
3. Ding, D.; Pan, Z.; Cuiuri, D.; Li, H. Wire-feed additive manufacturing of metal components: Technologies, developments and
future interests. Int. J. Adv. Manuf. Technol. 2015, 81, 465–481. [CrossRef]
4. Leon, A.; Shirizly, A.; Aghion, E. Corrosion Behavior of AlSi10Mg Alloy Produced by Additive Manufacturing (AM) vs. Its
Counterpart Gravity Cast Alloy. Metals 2016, 6, 148. [CrossRef]
5. Herzog, D.; Seyda, V.; Wycisk, E.; Emmelmann, C. Additive manufacturing of metals. Acta Mater. 2016, 117, 371–392. [CrossRef]
6. Haghdadi, N.; Laleh, M.; Moyle, M.; Primig, S. Additive manufacturing of steels: A review of achievements and challenges.
J. Mater. Sci. 2020, 56, 64–107. [CrossRef]
7. Leon, A.; Levy, G.K.; Ron, T.; Shirizly, A.; Aghion, E. The effect of hot isostatic pressure on the corrosion performance of Ti6Al4V
produced by an electron-beam melting additive manufacturing process. Addit. Manuf. 2020, 33, 101039. [CrossRef]
8. Zakay, A.; Aghion, E. Effect of Post-heat Treatment on the Corrosion Behavior of AlSi10Mg Alloy Produced by Additive Manufac-
turing. JOM 2019, 71, 1150–1157. [CrossRef]
326
Materials 2021, 14, 55
9. Rodrigues, T.A.; Duarte, V.; Miranda, R.M.; Santos, T.G.; Oliveira, J.P. Current Status and Perspectives on Wire and Arc Additive
Manufacturing (WAAM). Materials 2019, 12, 1121. [CrossRef]
10. Ding, D.; Pan, Z.; Cuiuri, D.; Li, H. A multi-bead overlapping model for robotic wire and arc additive manufacturing (WAAM).
Robot. Comput. Manuf. 2015, 31, 101–110. [CrossRef]
11. Gisario, A.; Kazarian, M.; Martina, F.; Mehrpouya, M. Metal additive manufacturing in the commercial aviation industry:
A review. J. Manuf. Syst. 2019, 53, 124–149. [CrossRef]
12. Donoghue, J.M.; Antonysamy, A.A.; Martina, F.; Colegrove, P.; Williams, S.W.; Prangnell, P. The effectiveness of combining rolling
deformation with Wire–Arc Additive Manufacture on β-grain refinement and texture modification in Ti-6Al-4V. Mater. Charact.
2016, 114, 103–114. [CrossRef]
13. Busachi, A.; Erkoyuncu, J.; Colegrove, P.; Martina, F.; Watts, C.; Drake, R. A review of Additive Manufacturing technology and
Cost Estimation techniques for the defence sector. CIRP J. Manuf. Sci. Technol. 2017, 19, 117–128. [CrossRef]
14. Bekker, A.C.M.; Verlinden, J.C.; Galimberti, G. Challenges in assessing the sustainability of wire+ arc additive manufacturing for
large structures. In Proceedings of the Solid Freeform Fabrication Symposium, Austin, TX, USA, 8–10 August 2016; pp. 406–418.
15. Cunningham, C.R.; Wikshåland, S.; Xu, F.; Kemakolam, N.; Shokrani, A.; Dhokia, V.; Newman, S. Cost Modelling and Sensitivity
Analysis of Wire and Arc Additive Manufacturing. Procedia Manuf. 2017, 11, 650–657. [CrossRef]
16. Ron, T.; Levy, G.K.; Dolev, O.; Leon, A.; Shirizly, A.; Aghion, E. Environmental Behavior of Low Carbon Steel Produced by a Wire
Arc Additive Manufacturing Process. Metals 2019, 9, 888. [CrossRef]
17. Bandari, Y.K.; Williams, S.W.; Ding, J.; Martina, F. Additive manufacture of large structures: Robotic or CNC systems. In Proceed-
ings of the 26th international solid freeform fabrication symposium, Austin, TX, USA, 10–12 August 2015; pp. 12–14.
18. Williams, S.W.; Martina, F.; Addison, A.C.; Ding, J.; Pardal, G.; Colegrove, P.A. Wire + Arc Additive Manufacturing.
Mater. Sci. Technol. 2016, 32, 641–647. [CrossRef]
19. Ron, T.; Levy, G.K.; Dolev, O.; Leon, A.; Shirizly, A.; Aghion, E. The Effect of Microstructural Imperfections on Corrosion Fatigue
of Additively Manufactured ER70S-6 Alloy Produced by Wire Arc Deposition. Metals 2020, 10, 98. [CrossRef]
20. Shirizly, A.; Dolev, O. From Wire to Seamless Flow-Formed Tube: Leveraging the Combination of Wire Arc Additive Manufactur-
ing and Metal Forming. JOM 2018, 71, 709–717. [CrossRef]
21. Derekar, K.S. A review of wire arc additive manufacturing and advances in wire arc additive manufacturing of aluminium.
Mater. Sci. Technol. 2018, 34, 895–916. [CrossRef]
22. Wu, B.; Pan, Z.; Ding, D.; Cuiuri, D.; Li, H.; Xu, J.; Norrish, J. A review of the wire arc additive manufacturing of metals: Properties,
defects and quality improvement. J. Manuf. Process. 2018, 35, 127–139. [CrossRef]
23. Pan, Z.; Ding, D.; Wu, B.; Cuiuri, D.; Li, H.; Norrish, J. Arc Welding Processes for Additive Manufacturing: A Review; Springer Science
and Business Media LLC: Berlin, Germany, 2017; pp. 3–24.
24. Wang, L.; Xue, J.; Wang, Q. Correlation between arc mode, microstructure, and mechanical properties during wire arc additive
manufacturing of 316L stainless steel. Mater. Sci. Eng. A 2019, 751, 183–190. [CrossRef]
25. Chen, X.; Li, J.; Cheng, X.; He, B.; Wang, H.; Huang, Z. Microstructure and mechanical properties of the austenitic stainless steel
316L fabricated by gas metal arc additive manufacturing. Mater. Sci. Eng. A 2017, 703, 567–577. [CrossRef]
26. Kaya, A.; Uzan, P.; Eliezer, D.; Aghion, E. Electron microscopical investigation of as cast AZ91D alloy. Mater. Sci. Technol. 2000,
16, 1001–1006. [CrossRef]
27. Aghion, E.; Gueta, Y.; Moscovitch, N.; Bronfin, B. Effect of yttrium additions on the properties of grain-refined Mg–3%Nd alloy.
J. Mater. Sci. 2008, 43, 4870–4875. [CrossRef]
28. Available online: https://www.astm.org/Standards/G129 (accessed on 21 December 2020).
29. Arnon, A.; Aghion, E. Stress Corrosion Cracking of Nano/Sub-micron E906 Magnesium Alloy. Adv. Eng. Mater. 2008, 10, 742–
745. [CrossRef]
30. Hakimi, O.; Aghion, E.; Goldman, J. Improved stress corrosion cracking resistance of a novel biodegradable EW62 magnesium
alloy by rapid solidification, in simulated electrolytes. Mater. Sci. Eng. C 2015, 51, 226–232. [CrossRef]
31. Guo, P.; Zou, B.; Huang, C.; Gao, H. Study on microstructure, mechanical properties and machinability of efficiently additive
manufactured AISI 316L stainless steel by high-power direct laser deposition. J. Mater. Process. Technol. 2017, 240, 12–22. [CrossRef]
32. Chen, X.; Li, J.; Cheng, X.; Wang, H.; Huang, Z. Effect of heat treatment on microstructure, mechanical and corrosion properties of
austenitic stainless steel 316L using arc additive manufacturing. Mater. Sci. Eng. A 2018, 715, 307–314. [CrossRef]
33. Xiong, J.; Tan, M.Y.; Forsyth, M. The corrosion behaviors of stainless steel weldments in sodium chloride solution observed using
a novel electrochemical measurement approach. Desalination 2013, 327, 39–45. [CrossRef]
34. Koseki, T.; Flemings, M.C. Solidification of undercooled Fe-Cr-Ni alloys: Part I. Thermal behavior. Met. Mater. Trans. A 1995,
26, 2991–2999. [CrossRef]
35. Itzhak, D.; Aghion, E. Corrosion behaviour of hot-pressed austenitic stainless steel in H2SO4 solutions at room temperature.
Corros. Sci. 1983, 23, 1085–1094. [CrossRef]
36. Itzhak, D.; Aghion, E. An anodic behaviour study of an analogical sintered system of austenitic stainless steel in H2SO4 solution.
Corros. Sci. 1984, 24, 145–149. [CrossRef]
37. Garcia-Cabezon, C.; Martín, F.; Blanco, Y.; De Tiedra, P.; Aparicio, M. Corrosion behaviour of duplex stainless steels sintered
in nitrogen. Corros. Sci. 2009, 51, 76–86. [CrossRef]
327
Materials 2021, 14, 55
38. Al Saadi, S.; Yi, Y.; Cho, P.; Jang, C.; Beeley, P. Passivity breakdown of 316L stainless steel during potentiodynamic polarization in
NaCl solution. Corros. Sci. 2016, 111, 720–727. [CrossRef]
39. Leon, A.; Aghion, E. Effect of surface roughness on corrosion fatigue performance of AlSi10Mg alloy produced by Selective Laser
Melting (SLM). Mater. Charact. 2017, 131, 188–194. [CrossRef]
40. Kafri, A.; Ovadia, S.; Yosafovich-Doitch, G.; Aghion, E. The Effects of 4%Fe on the Performance of Pure Zinc as Biodegradable
Implant Material. Ann. Biomed. Eng. 2019, 47, 1400–1408. [CrossRef]
41. Kubík, P.; Šebek, F.; Petruška, J.; Hůlka, J.; Park, N.; Huh, H. Comparative investigation of ductile fracture with 316L austenitic
stainless steel in small punch tests: Experiments and simulations. Theor. Appl. Fract. Mech. 2018, 98, 186–198. [CrossRef]
328
materials
Communication
Molar Ratio Effect of Sodium to Chloride Ions on the
Electrochemical Corrosion of Alloy 600 and SA508 in
HCl + NaOH Mixtures
Do Haeng Hur 1, *, Jeoh Han 1 and Jun Choi 1,2
1 Korea Atomic Energy Research Institute, 989-111, Deadeok-daero, Yuseong-gu, Daejeon 34057, Korea;
[email protected] (J.H.); [email protected] (J.C.)
2 Daesung Machinery Co., Ltd., 3Da-401, Sihwa Industrial Complex, 334, Gongdan 1-daero,
Gyeonggi-do, Siheung 15106, Korea
* Correspondence: [email protected]; Tel.: +82-42-868-8388; Fax: +82-42-868-8696
Abstract: This study aims to investigate the molar ratio effect of sodium to chloride ions on the
corrosion of an Alloy 600 steam generator tube and an SA508 tubesheet material. The corrosion
behavior was evaluated in solutions with three different molar ratios of sodium to chloride ions using
a potentiodynamic polarization method. The corrosion potentials and corrosion rates of both the
two materials were significantly decreased as the molar ratio increased from 0.1 to 10. Therefore,
it is recommended that the molar ratio control to a value of 1 is beneficial only when the crevice
chemistry has a low molar ratio with an acidic pH. The corrosion potentials and corrosion rates were
little affected by the total sodium and chloride ion concentrations. SA508 acted as an anode and its
corrosion rate was accelerated by galvanic coupling with Alloy 600.
Keywords: molar ratio index; steam generator; tubesheet; crevice chemistry; galvanic couple
1. Introduction
Stress corrosion cracking (SCC), intergranular attack (IGA) and pitting have been major degradation
modes of Alloy 600 steam generator (SG) tubes in the secondary side water environments of pressurized
water reactors (PWRs) [1–4]. Corrosion damage of the Alloy 600 tube materials is accelerated in both
acidic and alkaline environments resulting from the impurity concentration processes. The formation
of these corrosive environments is closely associated with local boiling within tube to tube support
crevices and tube to tubesheet crevices in which sludge is accumulated [5,6]. The laboratory and
field experience data have indicated that both IGA and SCC of Alloy 600 steam generator tubes are
minimized at a near neutral pH [7,8]. Therefore, it is expected that maintaining a crevice pH near
neutral reduces the corrosion damage of Alloy 600.
Based on the above backgrounds, the molar ratio control program was developed by the Electric
Power Research Institute (EPRI), of which the goal is to maintain the crevice chemistry at near neutral
pH values [7]. A molar ratio index (MRI) for managing the secondary water chemistry in PWRs was
also proposed as the following equation [7]:
Na + K
MRI = (1)
Cl + excess SO4
To maintain a desired MRI, several methods has been implemented, such as sodium source
reduction, chloride injection, and ion exchange resin manipulation [9]. This index was derived from a
viewpoint of the corrosion behavior of SG tube material itself. SG tubes are equipped in a tubesheet
by expanding the both end parts of the tubes within drilled-holes of the tubesheet. The SG tubing
is made of nickel-based alloys, while the tubesheet is SA508 low alloy steels. Therefore, an SG tube
and a tubesheet are galvanically contacted in the tube to tubesheet crevice region of a SG. In addition,
corrosion of the tubesheet material induces denting damage of the tubes at the top of the tubesheet,
accelerating SCC of the tubes [10,11]. Therefore, the corrosion behavior of the tubesheet material
should also be considered in the application of the molar ratio control. In addition, since the MRI
is a simple molar ratio of cations to anions, the total concentration of the cations and the anions is
not considered.
Potentiodynamic polarization tests provide useful information such as corrosion rate and
susceptibility of materials to corrosion in aqueous solutions, as well as pitting corrosion. Furthermore,
it has been reported that polarization behaviors of Alloy 600 and stainless steels are associated with
IGA and SCC [12–14]. In this case, however, additional IGA and SCC tests should be performed to find
a correlation at applied potentials selected from a polarization curve. In this work, an electrochemical
polarization method was used to survey the effect of the MRI on the polarization responses of an
Alloy 600 SG tube and an SA508 tubesheet material at approximately 25 ◦ C and to provide bases for
comparison with the results at higher temperatures, which will be obtained in the next phase. The effect
of impurity concentration at a constant MRI was also evaluated.
2. Experimental Methods
C Cr Fe Si Mn Ti Al Ni
0.02 15.7 10.0 0.1 0.3 0.1 0.06 Bal.
C Si Mn P S Ni Cr Mo Fe
0.199 0.051 1.52 0.005 0.006 0.987 0.232 0.582 Bal.
Specimens were cut into a size of 10 × 5 × 1 mm3 for the electrochemical corrosion tests. They were
ground using silicon carbide paper down to 1000-grit and then ultrasonically cleaned in acetone for
5 min.
To prepare a working electrode for the electrochemical test, an Alloy 600 specimen was spot-welded
to an Alloy 600 wire, while an SA508 specimen was spot-welded to a pure iron wire. The lead wire was
then shielded with a polytetrafluoroethylene (PTFE) tube for electrical insulation. A resin was coated
around the spot-weld to prevent the test solution from penetrating into any remaining crevice there.
After curing the resin, the specimen was ultrasonically cleaned in acetone for 1 min. When subtracting
the weld-junction area, the surface area exposed to the solutions during the electrochemical tests was
1.28 cm2 .
The MRIs of sodium ions to chloride ions in the test solutions were controlled to be 0.1, 1 and 10 by
the addition of HCl and NaOH into demineralized water with the resistivity near 18 MΩ·cm, as shown
in Table 3. Any other chemical species were not included to simplify the Equation (1). The total ion
330
Materials 2020, 13, 1970
concentrations of sodium and chloride ions were fixed to be 0.011 and 0.11 M at each MRI. Regardless
of the total ion concentrations, the measured solution pH was dependent on the MRI and was about 2
at the MRI 0.1, 7 at the MRI 1, and 12 at the MRI 10.
331
Materials 2020, 13, 1970
&XUUHQW'HQVLW\ $FP
05, $OOR\
05, $OOR\
05, $OOR\
3RWHQWLDO 96&(
Figure 1. Polarization curves of Alloy 600 in the solutions of the MRI 0.1, 1 and 10 at a total sodium
and chloride ion concentration of 0.011 M.
Figure 2. Scanning electron microscopy (SEM) micrographs showing the corroded surfaces of Alloy
600 after polarization scans at (a) the MRI 0.1, (b) the MRI 1, (c) magnification of the pit denoted by the
white arrow in (b), and (d) the MRI 10.
As shown in Figure 3, the corrosion potentials of SA508 were also significantly decreased as the
MRI increased from 0.1 to 10. SA508 actively dissolved at high corrosion rates without any passivation
at the MRI 0.1 and 1, whereas the alloy showed the lowest corrosion rate with a passive behavior in a
potential range of −0.190 to 0.600 V at the MRI 10. SA508 also showed an abrupt increase of current
density due to oxygen evolution near a potential of 0.600 V at the MRI 10, as did Alloy 600.
332
Materials 2020, 13, 1970
&XUUHQW'HQVLW\ $FP
05, 6$
05, 6$
05, 6$
3RWHQWLDO 96&(
Figure 3. Polarization curves of SA508 in the solutions of the MRI 0.1, 1 and 10 at a total sodium and
chloride ion concentration of 0.011 M.
Figure 4 shows the SEM micrographs of SA508 surfaces after the anodic polarization scans.
The surface exposed to the MRI 0.1 was severely and uniformly corroded enough to dissolve out the
grinding marks, which was made by emery paper during the surface finishing process. The surface at
the MRI 1 was also corroded uniformly, but less severely than at the MRI 0.1. On the contrary, it can
still clearly be seen the grinding marks on the surface exposed to the MRI 10, indicating that the anodic
dissolution rate was very low in the MRI 10 solution. Therefore, the morphologies of these corroded
surfaces were in good agreement with the polarization behaviors shown in Figure 3.
Figure 4. SEM micrographs showing the corroded surfaces of SA508 after polarization scans at (a) the
MRI 0.1, (b) the MRI 1, and (c) the MRI 10.
The important corrosion parameters from Figures 1 and 3 are summarized in Table 4. Alloy 600
showed the highest corrosion rate at the MRI 0.1, while the corrosion rates at the MRIs 1 and 10 were
nearly similar. In case of SA508, this alloy also showed the highest corrosion rate at the MRI 0.1, but the
corrosion rate at the MRIs 10 was rather smaller than that at the MRI 1. Consequently, this result
indicates that the molar ratio control method is beneficial only when the crevice chemistry has a low
MRI and pH.
Table 4. Corrosion potentials and corrosion rates of Alloy 600 and SA508 obtained from the
polarization tests.
333
Materials 2020, 13, 1970
As shown in Tables 1 and 2, Alloy 600 is a high-alloyed steel containing 15.7 wt.% Cr and 73.7 wt.%
Ni. Thus, this alloy has an excellent resistance to corrosion in overall pH ranges from acidic to alkaline.
Therefore, the difference between the corrosion rates (icorr ) at the acidic MRI 0.1 and at the alkaline MRI
10 is not so large. However, SA508 is an iron-based steel containing only 0.23 wt.% Cr and 0.58 wt.%
Mo and thus has a basically poor corrosion resistance, especially in acidic solutions. From Table 4,
the corrosion rates (icorr ) of Alloy 600 were always significantly lower than those of SA508 in all the test
conditions. In addition, there was a significant decrease of icorr for SA508 at the MRI 10 in comparison
with the MRI 1 as well as the MRI 0.1. The reason for this can be attributed to the fact that the solubility
of magnetite is significantly dependent on the pH of a solution [21,22]. The solubility of magnetite at
pH 3 is about 5 × 104 times higher than that at pH 12 in water at 100 ◦ C [21]. Therefore, the corrosion
rate of SA508 increases significantly in low pH solutions (i.e., low MRI solutions) with a high solubility
of magnetite because the corroding surface of the alloy cannot be protected by the magnetite film.
Conversely, the alloy showed a passive behavior with a low corrosion current in a potential range of
−0.190 to 0.600 V at the MRI 10 solution of pH 12, owing to a significantly low solubility of magnetite.
Figure 5 shows the potentiodynamic polarization behaviors of Alloy 600 and SA508 in the
solutions with total sodium and chloride ion concentration of 0.011 M and 0.11 M at a constant
MRI 1. The corrosion potentials of Alloy 600 and SA508 were approximately −0.450 V and −0.730 V at
both concentrations, respectively, indicating that the corrosion potentials of the two materials were
not affected by a change in the total ion concentration. The cathodic and anodic current density
of Alloy 600 was also little affected by an increase of the ion concentration from 0.011 M to 0.11 M.
However, the pitting potentials of Alloy 600 decreased from 0.390 V in 0.011 M to 0.170 V in 0.11 M.
In case of SA508, the polarization current density was nearly same in the solutions of 0.011 M and
0.11 M, when this alloy was polarized around the corrosion potential. The above results mean that the
electrochemical corrosion behavior of these materials in a region near the corrosion potentials does not
depend on the total sodium and chloride ion concentrations if the sodium to chloride molar ratio in a
solution is the same. Similar behaviors were also observed at the MRI 0.1 and 10.
&XUUHQW'HQVLW\ $FP
$OOR\ 0
$OOR\ 0
6$ 0
6$ 0
3RWHQWLDO 96&(
Figure 5. Polarization curves of Alloy 600 and SA508 in the 0.011 and 0.11 M solutions at a constant
MRI 1.
From Figure 1, Figure 3, and Figure 5, it is clear that the corrosion potential of SA508 is always
lower than that of Alloy 600 in each test condition. The anodic curve of SA508 also intersects with the
cathodic curve of Alloy 600. This result demonstrates that SA508 is an anodic member of the galvanic
couple and its corrosion rate is accelerated, when SA508 and Alloy 600 are electrically contacted.
When the two materials are coupled in equal area, the galvanic current density (icouple ) of SA508, acting
as an anode, is determined at the intersection of the anodic curve of SA508 and the cathodic curve
of Alloy 600. Figure 6 shows the effect of the MRIs on the galvanic corrosion of SA508, based on the
polarization curves. Upon coupling to an equal area of Alloy 600, the current density (icouple ) of the
334
Materials 2020, 13, 1970
coupled SA508 was increased by about 2~6 times compared to that (iSA508 ) before coupling. However,
the area of the tubesheet around a tube is much smaller than that of the tube in actual SGs, because
the tubes are densely inserted into the tubesheet to increase the heat transfer area. Consequently,
the corrosion rate of SA508 would be more accelerated by the effect of small anode (SA508) and large
cathode (Alloy 600). In addition, the galvanic corrosion rate of SA508 was little changed by the total
ion concentration at a fixed MRI as shown in Figure 6. This is because the polarization current density
of the two materials was not affected by an increase in the total ion concentration from 0.011 M to
0.11 M, as shown in Figure 5.
iSA508 0
iSA508 0
icouple 0
&XUUHQW'HQVLW\ $FP
icouple 0
05, 05, 05,
Figure 6. Effect of the MRIs on the galvanic corrosion rate of SA508 coupled to an equal area of
Alloy 600.
An SG tube can be slowly deformed by volume expansion of corrosion products due to corrosion
of the tube support materials adjacent to and around the tube, which is called denting. The denting
was attributed to concentration of chlorides and oxidants such as copper, in the crevices, leading to
rapid corrosion of the tube support materials [2,23]. SG tubes are expanded into the tubesheet of
SA508, a dissimilar metal. Therefore, based on the results obtained in this work, it is worth mentioning
that corrosion of the tubesheet is accelerated by the galvanic coupling itself without concentration of
chemical impurities in the crevices.
4. Conclusions
The electrochemical corrosion behavior of Alloy 600 and SA508 was investigated in solutions with
three different molar ratios of sodium to chloride ions. The corrosion potentials and corrosion rates
of both materials were significantly decreased as the molar ratio increased from 0.1 to 10. Therefore,
it is expected that the molar ratio control method is beneficial only when the crevice chemistry has a
low molar ratio with an acidic pH. The corrosion potentials and corrosion rates were little affected
by the total sodium and chloride ion concentrations if the alloys were polarized not far from their
corrosion potentials. Alloy 600 and SA508 acted as a cathode and an anode, respectively, when they
were electrically coupled. Therefore, this result indicates that the corrosion rate of the SA508 tubesheet
material is accelerated owing to the galvanic coupling effect itself without concentration of chemical
impurities in the crevices.
Author Contributions: D.H.H. conceived the experiments and wrote the paper; J.H. contributed analysis tools;
J.C. performed the experiments. All authors have read and agreed to the published version of the manuscript.
Funding: This work was supported by the National Research Foundation (NRF) grant funded by the government
of the Republic of Korea (NRF-2017M2A8A4015159).
Conflicts of Interest: The authors declare no conflict of interest.
335
Materials 2020, 13, 1970
References
1. Gorman, J. Corrosion problems affecting steam generator tubes in commercial water-cooled nuclear power
plants. In Steam Generators for Nuclear Power Plants; Elsevier BV: Amsterdam, The Netherlands, 2017;
pp. 155–181.
2. Staehle, R.W.; Gorman, J.A. Quantitative assessment of submodes of stress corrosion cracking on the
secondary side of steam generator tubing in pressurized water reactors: Part 1. Corrosion 2003, 59, 931–944.
[CrossRef]
3. Hur, D.H.; Choi, M.S.; Lee, D.H.; Song, M.H.; Han, J.H. Pitting corrosion and its countermeasures for
pressurized water reactor steam generator tubes. Corrosion 2006, 62, 905–910. [CrossRef]
4. Hur, D.H.; Choi, M.S.; Lee, D.H.; Song, M.H.; Han, J.H. Root causes of intergranular attack in an operating
nuclear steam generator tube. J. Nucl. Mater. 2008, 375, 382–387. [CrossRef]
5. Bahn, C.B.; Oh, S.H.; Park, B.K.; Hwang, I.S.; Rhee, I.H.; Kim, U.C.; Na, J.W. Impurity concentration behaviors
in a boiling tubesheet crevice Part I. Open crevice. Nucl. Eng. Des. 2003, 225, 129–144. [CrossRef]
6. Engelhardt, G.R.; Macdonald, D.D.; Millett, P.J. Transport processes in steam generator crevices—I. General
corrosion model. Corros. Sci. 1999, 41, 2165–2190. [CrossRef]
7. Millett, P.J. PWR Molar Ratio Control Application Guidelines. Volume 1: Summary; TR-104811-V1; Electric Power
Research Institute: Palo Alto, Houston, CA, USA, 1995; Available online: https://www.epri.com/#/pages/
product/TR-104811-V1/?lang=en- (accessed on 20 March 2020).
8. Kawamura, H.; Hirano, H.; Koike, M.; Suda, M. Inhibitory effect of boric acid on intergranular attack and
stress corrosion cracking of Alloy 600 in high-temperature water. Corrosion 2002, 58, 941–952. [CrossRef]
9. Fruzzetti, K. Pressurized Water Reactor Secondary Water Chemistry Guidelines-Revision 8; 3002010645; Electric
Power Research Institute: Palo Alto, Houston, CA, USA, 2017; Available online: https://www.epri.com/#/
pages/product/3002010645/ (accessed on 20 March 2020).
10. Choi, S.; Marks, C.; Wolfe, R. PWR steam generator tube denting at top of tubesheet. In Proceedings of the
International Conference on Nuclear Power Chemistry, Sapporo, Japan, 26–31 October 2014; p. 10137.
11. Hur, D.H.; Choi, M.S.; Lee, D.H.; Han, J.H.; Shim, H.S. Corrosion inhibition of steam generator tubesheet by
Alloy 690 cladding in secondary side environments. J. Nucl. Mater. 2013, 442, 326–329. [CrossRef]
12. Hsu, S.S.; Tai, S.C.; Kai, J.J.; Tai, C.H. SCC behavior and anodic dissolution of Inconel 600 in low concentration
thiosulfate. J. Nucl. Mater. 1991, 184, 97–106. [CrossRef]
13. Bandy, R.; Roberge, R.; van Rooyen, D. Intergranular failures of Alloy 600 in high temperature caustic
environments. Corrosion 1985, 41, 142–150. [CrossRef]
14. Lin, L.F.; Cragnolino, G.; Szklarska-Smialowska, Z.; Macdonald, D.D. Stress corrosion cracking of sensitized
Type 304 stainless steel in high temperature chloride solutions. Corrosion 1981, 37, 616–627. [CrossRef]
15. EPRI. Guidelines for PWR Steam Generator Tubing Specifications and Repair; TR-016743-V2R1; Electric Power
Research Institute: Palo Alto, Houston, CA, USA, 1999; Available online: https://www.epri.com/#/pages/
product/TR-016743-V2R1 (accessed on 27 February 2020).
16. ASME. Specification for Quenched and Tempered Vacuum-Treated Carbon and Alloy Steel Forgings for Pressure
Vssels; SA-508/SA-508M: New York, NY, USA, 1995.
17. Gnedenkov, A.S.; Sinebryukhov, S.L.; Mashtalyar, D.V.; Imshinetskiy, I.M.; Vyaliy, I.E.; Gnedenkov, S.V.
Effect of microstructure on the corrosion resistance of TIG welded 1579 alloy. Materials 2019, 12, 2615.
[CrossRef] [PubMed]
18. Dutta, R.S.; Tewari, R.; De, P.K. Effects of heat-treatment on the extent of chromium depletion and caustic
corrosion resistance of Alloy 690. Corros. Sci. 2007, 49, 303–318. [CrossRef]
19. Briant, C.L.; O’Toole, C.S.; Hall, E.L. The effect of microstructure on the corrosion and stress corrosion
cracking of Alloy 600 in acidic and neutral environments. Corrosion 1986, 42, 15–27. [CrossRef]
20. Pouraix, M. Atlas of Electrochemical Equilibria in Aqueous Solutions; NACE: Houston, TX, USA, 1974.
21. Tremaine, P.R.; LeBlanc, J.C. The solubility of magnetite and the hydrolysis and oxidation of Fe2+ in water to
300 ◦ C. J. Solut. Chem. 1980, 9, 415–442. [CrossRef]
336
Materials 2020, 13, 1970
22. Macdonald, D.D.; Shierman, G.R.; Butler, P. The Thermodynamics of Metal-Water Systems at Elevated Temperatures,
Part 2: The Iron-Water System; AECL-4137; Atomic Energy of Canada Ltd.: Pinawa, MB, Canada, 1972.
23. Fernández-Saavedra, R.; Fernández-Díaz, M.; Gómez-Mancebo, M.B.; de Diego, G.; Quejido-Cabezas, A.J.;
Gómez-Briceño, D. Hard sludge and denting on the secondary side of PWR steam generators. In Proceedings
of the International Conference on Nuclear Power Chemistry, Sapporo, Japan, 26–31 October 2014.
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
337
materials
Article
Corrosion Prediction of Weathered Galvanised Structures Using
Machine Learning Techniques
Marta Terrados-Cristos *, Francisco Ortega-Fernández, Guillermo Alonso-Iglesias, Marina Díaz-Piloneta
and Ana Fernández-Iglesias
Project Engineering Department, University of Oviedo, 33004 Oviedo, Spain; [email protected] (F.O.-F.);
[email protected] (G.A.-I.); [email protected] (M.D.-P.); [email protected] (A.F.-I.)
* Correspondence: [email protected]
1. Introduction
Academic Editor: Marián Palcut
Multiple metallic structures and equipment operate in outdoor conditions [1]. In such
cases, one of the main problems related to their stability and durability is corrosion [2,3].
Received: 10 June 2021
Accepted: 12 July 2021
World Corrosion Organization (WCO) estimates the world direct cost of corrosion to be
Published: 13 July 2021
between 1.3 and 1.4 trillion EUR, 3.1% to 3.5% of a nation’s GDP annually [4].
Corrosion is a very complex phenomenon based on the degradation of a material or
Publisher’s Note: MDPI stays neutral
its properties due to its reaction with the environment [5]. Multiple factors [6], particles [7],
with regard to jurisdictional claims in
and variables [8,9] are involved. The character of the attack and the corrosion rate are
published maps and institutional affil- consequences of the system formed by metallic materials, atmospheric environment, tech-
iations. nical parameters, and operating conditions [10]. Corrective factors are introduced in the
design phases to guarantee the structure’s integrity during its useful life [11]. However, the
difficulty of quantifying the material loss causes unnecessary over-dimensioning, leading
to superfluous costs and resources consumption [12]. Proper management of this complex
Copyright: © 2021 by the authors.
multifactorial phenomenon is key to sustainable development [13].
Licensee MDPI, Basel, Switzerland.
To ensure the integrity of the outer layer, structures are designed with physical pro-
This article is an open access article
tection. Historically, metallic zinc has provided excellent corrosion protection of steel
distributed under the terms and structures [14]. Unfortunately, corrosion damage also occurs in such systems [15]. Since
conditions of the Creative Commons corrosion leads to a mass loss, an excess thickness is often considered to ensure service
Attribution (CC BY) license (https:// life. This not only increases manufacturing cost but also does not satisfy the principles
creativecommons.org/licenses/by/ of sustainable engineering efficiency [16]. Therefore, lacking an automated monitoring
4.0/). system or predictive model, routine thickness monitoring would be required [17]. These
339
phenomena have drawn increasing attention in recent decades due to the resulting catas-
trophic accidents [18] and the growing demand for sustainable designs [19]. For an optimal
selection of materials, atmospheric aggressiveness must be considered. Depending on this,
coating needs can be set.
The current regulation regarding galvanised metallic structures (ISO 9223:2012 [20])
groups the corrosivity level of an atmosphere into six categories. After studying the effect of
corrosion on standard samples during 1 year of weathering exposure, the level of corrosion
rates achieved can be established by measuring weight losses for different materials. This
material’s loss due to corrosion is commonly used as an initial measure for determining
coating requirements. However, material loss margins are allowed within these categories,
and coating thickness designs based on them are not fixed. These margins imply variability
in the amount of material that can be translated into increased costs.
According to [20], two methods are proposed to classify the corrosivity of atmospheric
environments, depending on the availability of experimental data. When experimental
data are available, dose–response functions can be used. However, when no experimental
data are available, corrosivity category estimation using the informative procedure is rec-
ommended, and as stated in the norm, it is based on the comparison of local environmental
conditions with the description of typical atmospheric environments, which may cause mis-
interpretations [21]. Finding the optimum point between efficiency and competitive price,
while remaining within limits, is therefore challenging given the lack of characterisation of
the specific construction site.
The objective of this work is to develop machine learning models that, by analysing
real cases, predict corrosion mass loss of zinc coatings over time. The aim is to characterise
an environment without requiring long testing periods and sampling and generalising
it to any location worldwide, with the data available from international studies. This
considerably increases the existing knowledge about coated steel structure corrosion and
extends it to the full diversity of atmospheres, thereby reducing the uncertainty of its
final state.
This paper starts with a state-of-the-art analysis. Then, it explains the creation of the
database through the characterisation of each sample. Next, the applied methodology
is explained, and modelling and evaluation techniques are defined. Finally, results are
discussed, and the conclusions obtained in this research are proposed.
2. Literature Review
There is a wide range of corrosion problems in the industry, resulting from the differ-
ent combinations of materials, environments, and service conditions [22]. Therefore, the
concern about corrosion is not new. The science of atmospheric corrosion started with
Faraday in the nineteenth century [23]. Another important contribution was made by
Vernon who began systematic experiments in atmospheric corrosion in the 1920s [24]. In
1986, Benarie and Lipfert published their work on atmospheric corrosion [25], relating this
phenomenon to the concentration of certain pollutants and pH of the rain. Subsequently,
Feliu et al. developed regression equations for mild steel, zinc, copper, and aluminium [26].
There are several kinetic corrosion models that attempt to predict atmospheric cor-
rosion over time: the general linear model [27], the power function models [28], and the
power-linear models [29]. However, the corrosion process is influenced by multiple en-
vironmental factors [30]. Therefore, these corrosion kinetic models are valid at specific
locations. When the environmental condition changes, the model may no longer be appli-
cable [31]. It would be interesting to classify the aggressiveness of different atmospheres,
which would allow preventive measures to be taken. Therefore, it is important to introduce
the interaction parameters between environmental factors and corrosion rates for their
efficient prediction.
In accordance with this approach, the ISOCORRAG program was launched in 1986 [32].
The ISO 156 technical committee developed this project with the intention of obtaining
sufficient information to standardise atmospheric corrosion on metals and alloys. Four
340
Materials 2021, 14, 3906
international standards were created as a result of this project: ISO 9223 [21], ISO 9224 [33],
ISO 9225 [20], and ISO 9226 [34]. Since then, these standards have served as practical
guidelines and aids for the design of both structures and their corrosion protection. In
September 1987, the Executive Body for the Convention on Long-Range Transboundary
Air Pollution (CLRTAP) decided to launch an International Cooperation Program with
the United Nations European Economic Commission (ICP/UNECE) [35] whose objective
was to carry out a quantitative assessment of the effect of pollutants on atmospheric cor-
rosion [6]. In addition, a third cooperative program was launched, named MICAT [36]
(Ibero-American Atmospheric Corrosivity Map). Its objective was to understand the mech-
anisms that take place when this phenomenon occurs, to generate, with the data obtained,
mathematical models to calculate corrosion as a function of climate condition or pollu-
tant levels [13]. The three projects evaluated corrosion by measuring mass loss and were
based on what was indicated in the standard for measuring SO2 or Cl− levels and other
pollutant concentrations.
In 1992, the ASTM (American Society for Testing and Materials) published a study
discussing an alternative method for measuring corrosion penetration, with models that
are tighter and more rational than the traditional potential model [37]. In 2003, several
workers compiled atmospheric exposure data from many research reports and journal
articles [38]. R.E. Melchers, an engineer at Newcastle University, focused on studying the
corrosion of metals in marine atmospheres in his studies in 2008 [39] and 2013 [40]. Later,
Morcillo et al. [27] made a comprehensive compilation in the scientific literature on weath-
ering steel atmospheric corrosion [6]. In addition, they developed Damage Functions to
know the damage that a metallic structure can suffer depending on weathering conditions.
In the subsequent years, there have been local experimental studies to characterise this
phenomenon, such as those in Greece [41] and the Czech Republic [42].
The dose–response function is the most widely used. It directly correlates the influ-
encing environmental factors with the corrosion parameters [43]. The basic form of this
function follows the simple linear [36,44] or logarithmic–linear relationships [45]. However,
many researchers also started to depart from judging the effect of each environmental
factor separately and established a new multi-factor combination model [46,47]. A response
surface model (RSM) takes into account the interactive effect and the non-linearity of the
atmospheric corrosion process and allows a better approximation compared to conven-
tional dose–response function models [48]. The models offer a closer approximation of
corrosion rate by introducing different input variables. Temperature, humidity, sulphur
dioxide concentration, and chloride concentration are typically used.
In conclusion, there are different options to predict corrosion rates of metals based
on experimental input data. However, for the cases when pollutants’ concentration is
unknown, the options are limited. Time and cost constraints make the development of
these measurements difficult as they would be unrepresentative when only completed at a
specific point in time. As the environmental conditions continuously change, it is necessary
to know their distribution over larger distances and longer periods of time. All corrosion
related research carried out so far showed that there are certain factors that clearly influence
the corrosion process. Regarding atmospheric corrosion, the factors include temperature,
relative humidity, precipitation level, and pollutant concentrations (SOx , Cl− , etc.) [49,50].
A combination of parameters, such as Time of Wetness (TOW), is also used. TOW represents
the fraction of time when relative humidity exceeds 80% and ambient temperature is above
0 ◦ C (h/year) [51].
Climate has a significant influence on corrosion since some of the factors mentioned
above depend on the climatic zone. A Köppen–Geiger classification [52] is the most popular
technique for climate characterisation. According to this method, six precipitation levels
can be distinguished [52]: desert (0), steppe (1), totally humid (2), summer dry (3), winter
dry (4), and monsoon (5). Temperature and relative humidity are easily analysable climatic
variables, and their values are generally accessible. There are also additional factors besides
climate, mainly derived from human activities, whose importance is also significant. It is
341
Materials 2021, 14, 3906
evident that the most populated and most-developed areas with accumulations of vehicles
and high industrial activity have greater corrosive potential. It is also known that materials
situated in areas closer to the sea tend to have a worse corrosion performance. Therefore, it
is necessary to include these additional factors as well as they are critical for the successful
operation of the model.
ISO 9223 and ISO 9224 standards are highlighted for this project. First, ISO 9223:2012 [20]
divides the corrosivity of atmospheres into 6 categories. Each of these categories corre-
sponds to a different corrosion level. For zinc, data are shown in Table 1.
Table 1. Corrosion rates of zinc for first-year exposure for different corrosivity categories according
to ISO 9223:2012.
342
Materials 2021, 14, 3906
Second, ISO 9224:2012 proposes a relationship for long-term corrosion exposures. This
relationship is based on the power function according to the following equation:
D = rcorr tb (1)
In Equation (1), rcorr is the first-year corrosion rate, t is the number of years to be
analysed, and b is the environment and metal-specific time exponent.
3.1.1. Variables
Willing to characterise any location worldwide, its atmospheric corrosivity and climate
need to be considered. For this work, three specific types of atmospheric environments
have been introduced as binary synthetic variables, trying to represent the behaviour of
sulphates-related pollution and chlorides deposition:
• Industrial/Non-industrial: industrial are areas with fossil fuel combustion industries
(refineries, thermal power plants, etc.).
• Marine/Non-marine: this characterisation has been made according to the distance from
the coast, considering as Marine any location within 15 km from the seashore [53,54].
• Urban/Rural: locations with more than 5000 inhabitants or 300 inhabitants per square
kilometre have been considered urban locations [55].
Regarding the climate characterisation, temperature, relative humidity, TOW, and
Köppen–Geiger level of precipitation were the main characteristics, unified in a simple,
accessible, and complete way. Therefore, a total of seven numeric predictor variables
were set for the model: mean annual temperature, mean annual relative humidity, TOW,
precipitation, industrial, marine, and urban. The variable to be predicted was the zinc
corrosion loss during first-year exposure, directly taken from experimental studies, and its
atmospheric corrosivity category, based on the standard. Each sample was characterised,
following the rules mentioned above, as explained in Figure 2.
Variables: 0
ISOCORRAG
Corrosion loss
Experimental studies
ISO 9223:2012
Atmospheric Corrosivity Categories DATABASE
STANDARD
Atmospheric
Marine / Industrial / Urban /
environment
Non-marine Non-industrial Rural
characterization
Figure 2. Flow chart for database creation and future locations characterisation.
343
Materials 2021, 14, 3906
Figure 3. Frequency graphical analysis of the categorical variables. All possible atmospheric environ-
ment combinations are represented and coloured by precipitation type.
(a) (b)
Figure 4. Analysis of continuous variables at each location. (a) Distribution of mean annual relative humidity. (b) Distribu-
tion of mean annual temperature.
344
Materials 2021, 14, 3906
3.2. Methodology
The methodology followed in this paper consisted of 6 phases (Figure 5). The prepara-
tory stage (stage zero) in the previous subsection was concluded with the creation of the
database. Then, the remaining five phases included modelling and testing. The first step for
data pre-processing was to identify input variable’s importance for better understanding
their behaviour and obtaining additional information regarding their usefulness in the
final model. This was completed using Multivariate Adaptive Regression Splines (MARS,
Step 1). Then, the next phase was to define the first-year corrosion loss of galvanised
steel. Self-Organising Maps (SOM) were used, including various layers (supersom) of both
supervised and unsupervised learning. The next two steps used the result of the various
layers of this algorithm. The first layer has been the result of using unsupervised SOM,
according to the relationships between the 7 main variables. Zinc corrosion loss during
first year of exposure (Corr_Zn, in μm) was the output variable to be predicted (Step 2).
DATABASE 0
1 2
Data pre-processing: MARS Model input variables
MODELING
First-year corrosion prediction
4
Long term corrosion loss: ISO 9224:2012 equation + Newton’s Method
TEST
Minimum, maximum and most probable Reliability of the prediction
values
Figure 5. Flow chart showing the methodology followed in this paper. The six phases proposed are exposed as shown.
345
Materials 2021, 14, 3906
Techniques
• Multivariate Adaptive Regression Splines (MARS)
One of the most widely used algorithms for solving adaptive computing problems
is MARS [56]. This method consists of approximating an unknown function by the linear
combination of a set of basic functions (products of the model variables) [57]. Among the
key points of the algorithm, it stands out that it autonomously selects the relevant variables
and interactions between them for each subregion. Thus, the dimensionality reduction of
the problem is performed directly by the model, with the advantage of being locally carried
out. Precisely, this benefit can be used to analyse the relevance of the variables likely to
subsequently participate in the model.
• Self-Organising Maps (SOM)
The clustering model, known as SOM, is an unsupervised Artificial Neural Network
(ANN) presented in 1982 by T. Kohonen [58]. This model is based on certain evidence
discovered at brain level and performs a reduction of the dimensionality of the input
space to produce topologically ordered maps. This type of network has competitive,
unsupervised learning. The network itself is in charge of self-organising and discovering
common features, regularities, correlations, or categories in the input data [59,60].
Figure 6 shows the architecture of the model and how each input neuron is connected
to one of the output neurons by weights (w, according to Kohonen’s notation). The output
neurons will therefore have an associated vector of weights which is called the reference
vector (or codebook), also constituting the average vector of the category represented by
the output neuron [61,62].
wij
v1 v.. vn
Figure 6. General example of SOM model’s topography. Dimensions are expressed by x and y;
v1–n represent each one of the input neurons, and wij is the weight of each vector according to
Kohonen’s notation.
SOM’s utility lies in the holistic visual interpretation of the output rather than in
understanding the underlying processes [63]. Roughly speaking, the output layer (i.e., the
self-organising map itself) contains neurons organised in a rectangular or hexagonal lattice
to represent the entire dataset [58].
The goal of this learning is to categorise the data fed into the network. Similar values
are classified into the same category and, therefore, should activate the same output neuron.
Since this is an unsupervised method, classes or categories must be created by the network
itself through correlations between the input data [64]. However, SOM can also be used
for pattern recognition (supervised learning). The information is given at the end of
the training: if classification is involved, as in this case, the winner-takes-all strategy is
used. This principle can be extended to more layers, generating super-organised maps
(supersom). For each layer, a similarity level is calculated, and the individual similarities
are combined into a single value which is used to determine the winner node.
346
Materials 2021, 14, 3906
• Newton’s method
This nonlinear regression uses Newton’s Surface gradients, which is an unconstrained
linear regression method based on that gradient. The gradient information is provided by
analytically computed gradients. Design variables are modified, while their impact on the
objective function is analysed [65].
• Euclidean distance model
The operation of this model is based on Euclidean distances (d E ). This is a non-
negative function used to calculate the distance between two points P = (p1 ; p2 ; . . . ; pn )
and Q = (q1 ; q2 ; . . . ; qn ) on an n-dimensional space [66]. It works on the basis of the
Pythagoras Theorem (Equation (2)) [67]. Results evaluation using this method involves
checking that the model gives a 100 % quality in all the cases studied, i.e., that it perfectly
finds its counterpart.
n
d E (P, Q) = 2 2
( p1 − q1 ) + · · · + ( p n − q n ) = ∑ ( p i − q i )2 (2)
i =1
To summarise, Table 3 shows the different algorithms used in each phase of the data
mining process.
Phase Algorithm
Pre-processing data MARS
Modelling
• Corrosivity category prediction superSOM
• First-year corrosion prediction superSOM
• Long-term corrosion prediction Newton method
Quality evaluation Distance model
100
90
80
70
60
50 Mean
40 GCV
Importance [%]
30 RSS
20
10
0
347
Materials 2021, 14, 3906
Each neuron, filled or not, is represented by a codebook. These neurons are arranged
in such a way that nearby neurons represent points closer to each other. Analysing the
result of the average corrosion values per neuron along the mesh, it can be clearly seen how
the mesh is growing towards the lower right corner. Figure 9 shows this result; the larger
the circle size, the higher the average corrosion. Keeping the neighbourhood properties, a
uniform behaviour is shown, which indicates good training results.
Figure 9. Mean corrosion values per neuron. Corrosion loss in μm per year is represented by
circle size.
between the values so that the C5 are in contact with C4, C4 with C3, etc., demonstrating
an optimal training.
(a) (b)
Figure 10. Corrosion zones according to the environment. (a) Corrosion representation (larger circle, more corrosion).
(b) Corrosivity category representation, according to ISO 9223:2012 standard.
The predicted first-year corrosion rates using SOM trained network were compared
with real values. A satisfactory correlation has been obtained (Figure 11), although not all
points perfectly matched their counterpoints. The ideal situation would be if the predicted
values all lied on the diagonal line. The points tend to be located on the upper side of the
graph, meaning that predictions are conservative, and the decisions made based on them
can provide greater safety.
Figure 11. Predicted first-year corrosion values in micron vs. real first-year corrosion values. The
dashed line is the regression line (R2 = 0.7728). The points situated on the diagonal grey line represent
an optimal training.
From the trained network, it is possible to determine the corrosion rate of any situation
to be studied. When introducing a new case to the model, it finds the node that most
closely resembles its input variables. Thus, the output of the model is the corrosion rate
of that node. The uncertainty range is also given, including the minimum and maximum
349
Materials 2021, 14, 3906
values within each neuron. This can be seen with the following example for a case with the
characteristics defined in Table 4.
The case falls into the neuron indicated in Figure 12, which consists of 10 examples.
Figure 12. Case study example: the cake portions shown at each node show the contribution of each
variable; the larger the size, the greater its final weight.
Table 5 shows all results obtained. Different conclusions can be made by selecting
the maximum (Corr_max), minimum (Corr_min), and average (Corr_avg) values of the
examples in one single neuron. As a result, when the values with the most or least corrosion
occurring within the projects in the neuron are chosen, the optimistic and pessimistic
predictions can be obtained. Alternatively, β-distribution is used to determine the ‘most
probable’ rate of Corr_Zn, using the maximum, minimum, and average values. On the
other hand, the category is awarded by the weighted average of the categories in each case.
In this case, since all cases are C3, C3 is its category.
Corr_min Corr_avg Corr_max Range Given by the Model Category Range Given by ISO Standard
1.22 1.578 1.91 1.22–1.91 C3 100% 0.7–2.1
Comparing the range given by the model with the range given by the existing standard,
it is observed that the latter represents a much higher uncertainty for each corrosivity
category. Extending this comparison to the entire study scope, possible model predictions
for each category, clustered on similar values and represented by boxplots, can be presented
(Figure 13). Although not all categories are equally distributed, they show, in general,
narrower intervals.
This study is presented as a possible alternative to the informative procedure of the
ISO standard when there is no experimental data available. The results of the informative
procedure regarding atmospheric categorisation provide a range of mass losses for each
material. The current trend among companies and engineers, when no specific experi-
mental information is available, is to use the highest value of each category to make their
decisions. Since corrosion loss values are directly related to the required coating thickness,
350
Materials 2021, 14, 3906
the higher the corrosion loss value, the more coating is required. A coating thickness can
thus be directly determined by the predicted material’s loss.
Figure 13. Comparison between each category range offered by the standard using the informative procedure and the
possible mean values and uncertainties offered by the model, represented by clustered boxplots on each category.
The material requirement for coatings can be compared with the largest measurement
proposed by the standard in each category and with the value predicted by the model.
Following the example above, when using a Zn-coating of 1.6 μm (Corr_avg) instead of 2.1
μm (maximum in the range given by ISO), a 24% reduction in material’s costs is obtained.
It is then proposed to carry out this comparison for the rest of the points studied. From a
more conservative perspective, comparing the maximum predicted value (Corr_max) with
the maximum proposed by the standard using the informative method can also be used. In
this way, uncertainties are also considered. By performing this for all data studied during
the evaluation phase, an average saving of 16% in coating material is obtained.
Figure 14 compares the distribution of relative errors of both models. The nonlinear
regression relative error is represented by a solid black line and the standard formula’s
relative error (ISO 9224) by a blue dashed line. A more uniform distribution is achieved in
the nonlinear regression model.
Figure 14. Comparison between Nonlinear regression and standard’s formula relative errors.
351
Materials 2021, 14, 3906
Results obtained above show high prediction reliability. Cases similar to the one under
study have been found in the database. The model could also give a satisfactory result for
a case that is not included in the database. Ideally, the results obtained with the proposed
methodology should be compared with the results obtained with existing methods in the
literature. However, since the innovative premise of this study is based on adapting the
input variables to avoid the need for pollutant-specific data, such a comparison cannot be
made. One of the differentiating factors of this classifier model is that to obtain a corrosion
loss rate, values for pollutant concentrations are not needed. Consequently, it may be
concluded that the different algorithms developed are a good alternative for technicians
and engineers to make informed decisions based on their level of risk acceptance. To sum
up, given a specific location and based on the available data, these models can determine
the Zn-coating thickness needed for a successful short- and long-term corrosion resistance,
providing the most probable, optimistic, and pessimistic predictions.
5. Conclusions
In the present work, various models for predicting galvanised coated steel corrosion
damage of metal structures exposed to weathering have been developed. The following
conclusions can be drawn from this research.
The application of a supersom algorithm is considered for first-year corrosion predic-
tion, which allows categorising any environment while obtaining a predicted value, with
satisfactory results. In the cases when no experimental data are available, the model can be
an alternative to the conventional informative method based on pollutant input variables.
The model presented in this work could help civil engineering companies to optimise the
ratio between the minimum coating required and maximum service life, thus contributing
to a significant lifetime extension of steel structures.
The main limitation of the model is that it lacks statistical metrics to evaluate the
performance. To solve this and explore the performance and quality of the predictions, a
quality model based on Euclidean distances was proposed. A long-term corrosion predic-
tion was also optimised based on standards ISO 9224:2012 formula and the exponential
coefficient with Newton’s method.
To cover all different atmospheric environments, more specific characterisations are
required. The future research will focus on including the development of physical variables,
such as wind speed and wind direction. It is also important to feed the model with more
352
Materials 2021, 14, 3906
examples from the lesser-represented categories, as there are notable differences between
C3/C4 categories and the remainder of the cases. Adding new metallic materials will also
be explored, following the same methodology, possibly leading to the development of new
prediction models.
Author Contributions: Conceptualisation, F.O.-F. and A.F.-I.; methodology, G.A.-I. and M.D.-P;
validation, G.A.-I. and M.T.-C; writing—original draft preparation, M.T.-C.; writing—review and
editing, M.T.-C. and M.D.-P.; supervision, F.O.-F. All authors have read and agreed to the published
version of the manuscript.
Funding: This research received no external funding.
Institutional Review Board Statement: Not applicable.
Informed Consent Statement: Not applicable.
Data Availability Statement: Data sharing is not applicable to this article.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Wang, Z.; Wang, M.; Jiang, J.; Lan, X.; Wang, F.; Geng, Z.; Tian, Q. Atmospheric Corrosion Analysis and Rust Evolution Research
of Q235 Carbon Steel at Different Exposure Stages in Chengdu Atmospheric Environment of China. Scanning 2020, 2020, e9591516.
[CrossRef] [PubMed]
2. Emetere, M.E.; Afolalu, S.A.; Amusan, L.M.; Mamudu, A. Role of Atmospheric Aerosol Content on Atmospheric Corrosion of
Metallic Materials. Int. J. Corros. 2021, 2021, e6637499. [CrossRef]
3. Michael Schutze, R.B. Corrosion Resistance of Steels, Nickel Alloys, and Zinc in Aqueous Media: Waste Water, Seawater, Drinking Water,
High-Purity Water; John Wiley and Sons Ltd.: Hoboken, NJ, USA, 2016; ISBN 3-527-34069-6.
4. Hays, G.F. Now Is the Time; World Corrosion Organization: New York, NY, USA, 2010.
5. Ahmad, Z. Chapter 2—Basic Concepts in Corrosion. In Principles of Corrosion Engineering and Corrosion Control; Ahmad, Z., Ed.;
Butterworth-Heinemann: Oxford, UK, 2006; pp. 9–56, ISBN 978-0-7506-5924-6.
6. Chico, B.; De la Fuente, D.; Díaz, I.; Simancas, J.; Morcillo, M. Annual Atmospheric Corrosion of Carbon Steel Worldwide. An
Integration of ISOCORRAG, ICP/UNECE and MICAT Databases. Materials 2017, 10, 601. [CrossRef]
7. Yin, C.; Cheng, X.; Liu, X.; Zhao, M. Identification and Classification of Atmospheric Particles Based on SEM Images Using
Convolutional Neural Network with Attention Mechanism. Complexity 2020, 2020, e9673724. [CrossRef]
8. Hembrara, O.V.; Andreikiv, O.E. Effect of Hydrogenation of the Walls of Oil-and-Gas Pipelines on Their Soil Corrosion and
Service Life. Mater. Sci. 2012, 47, 598–607. [CrossRef]
9. Doyle, G.; Seica, M.V.; Grabinsky, M.W. The Role of Soil in the External Corrosion of Cast Iron Water Mains in Toronto, Canada.
Can. Geotech. J. 2003, 40, 225–236. [CrossRef]
10. Kusmierek, E.; Chrzescijanska, E. Atmospheric Corrosion of Metals in Industrial City Environment. Data Brief 2015, 3, 149–154.
[CrossRef] [PubMed]
11. Xu, Y.; Liu, L.; Zhou, Q.; Wang, X.; Tan, M.Y.; Huang, Y. An Overview of Major Experimental Methods and Apparatus for
Measuring and Investigating Erosion-Corrosion of Ferrous-Based Steels. Metals 2020, 10, 180. [CrossRef]
12. Lazorenko, G.; Kasprzhitskii, A.; Nazdracheva, T. Anti-Corrosion Coatings for Protection of Steel Railway Structures Exposed to
Atmospheric Environments: A Review. Constr. Build. Mater. 2021, 288, 123115. [CrossRef]
13. Morcillo, M.; Chico, B.; Fuente, D.; Simancas, J. Looking Back on Contributions in the Field of Atmospheric Corrosion Offered by
the MICAT Ibero-American Testing Network. Int. J. Corros. 2012, 2012, 824365. [CrossRef]
14. National Institute of Standars and Technology; American Bureau of Shiping; Colorado School of Mines; Mineral Management
Service; Office of Pipeline Safety. Coatings for Corrosion Protection: Offshore Oil and Gas Operation Facilities, Marine Pipeline and Ship
Structures; U.S. Department of Transportation: Washington, DC, USA, 2004.
15. Peabody, A.W. Control of Pipeline Corrosion; Bianchetti, R., Ed.; National Association of Corrosion Engineers (NACE): Houston, TX,
USA, 2001; ISBN 1-57590-092-0.
16. Arriba-Rodriguez, L.; Villanueva-Balsera, J.; Ortega-Fernandez, F.; Rodriguez-Perez, F. Methods to Evaluate Corrosion in Buried
Steel Structures: A Review. Metals 2018, 8, 334. [CrossRef]
17. Naz, M.Y.; Ismail, N.I.; Sulaiman, S.A.; Shukrullah, S. Electrochemical and Dry Sand Impact Erosion Studies on Carbon Steel. Sci.
Rep. 2015, 5, 16583. [CrossRef]
18. Kubzova, M.; Krivy, V.; Kreislova, K. Influence of Chloride Deposition on Corrosion Products. Procedia Eng. 2017, 192, 504–509.
[CrossRef]
19. Moins, B.; France, C.; Van den Bergh, W.; Audenaert, A. Implementing Life Cycle Cost Analysis in Road Engineering: A Critical
Review on Methodological Framework Choices. Renew. Sustain. Energy Rev. 2020, 133, 110284. [CrossRef]
353
Materials 2021, 14, 3906
20. ISO. ISO 9225:2012: Corrosion of Metals and Alloys—Corrosivity of Atmospheres—Measurement of Environmental Parameters Affecting
Corrosivity of Atmospheres; ISO/TC 156; ISO: Geneva, Switzerland, 2012.
21. ISO. ISO 9223:2012: Corrosion of Metals and Alloys—Corrosivity of Atmospheres—Classification, Determination and Estimation; ISO/TC
156; ISO: Geneva, Switzerland, 2012.
22. National Association of Corrosion Engineers; Baboian, R. NACE Corrosion Engineer’s Reference Book; NACE International: Houston,
TX, USA, 2016; ISBN 978-1-5231-0657-8.
23. Ahmad, Z. Chapter 1—Introduction to Corrosion. In Principles of Corrosion Engineering and Corrosion Control; Ahmad, Z., Ed.;
Butterworth-Heinemann: Oxford, UK, 2006; pp. 1–8, ISBN 978-0-7506-5924-6.
24. Vernon, W.H.J. First (Experimental) Report to the Atmospheric Corrosion Research Committee (of the British Non-Ferrous Metals
Research Association). Nature 1925, 115, 417. [CrossRef]
25. Benarie, M.; Lipfert, F.L. A General Corrosion Function in Terms of Atmospheric Pollutant Concentrations and Rain PH. Atmos.
Environ. 1967 1986, 20, 1947–1958. [CrossRef]
26. Feliu, S.; Morcillo, M. The Prediction of Atmospheric Corrosion from Meteorological and Pollution Parameters. Corros. Sci. 1993,
34, 403–414. [CrossRef]
27. Morcillo, M.; Chico, B.; Díaz, I.; Cano, H.; de la Fuente, D. Atmospheric Corrosion Data of Weathering Steels. A Review. Corros.
Sci. 2013, 77, 6–24. [CrossRef]
28. De la Fuente, D.; Castaño, J.G.; Morcillo, M. Long-Term Atmospheric Corrosion of Zinc. Corros. Sci. 2007, 49, 1420–1436.
[CrossRef]
29. Panchenko, Y.M.; Marshakov, A.I. Long-Term Prediction of Metal Corrosion Losses in Atmosphere Using a Power-Linear Function.
Corros. Sci. 2016, 109, 217–229. [CrossRef]
30. Leygraf, C.; Wallinder, I.; Tidblad, J.; Graedel, T. Atmospheric Corrosion, 2nd ed.; John Wiley & Sons: Hoboken, NJ, USA, 2016;
p. 374.
31. Cai, Y.; Xu, Y.; Zhao, Y.; Ma, X. Atmospheric Corrosion Prediction: A Review. Corros. Rev. 2020. [CrossRef]
32. Dean, S.W.; Knotkova, D.; Kreislová, K. ISOCORRAG International Atmospheric Exposure Program: Summary of Results; ASTM
International: West Conshohocken, PA, USA, 2011.
33. ISO. ISO 9224:2012 Corrosion of Metals and Alloys—Corrosivity of Atmospheres—Guiding Values for the Corrosivity Categories; ISO/TC
156; ISO: Geneva, Switzerland, 2012.
34. ISO. ISO 9226:2012 Corrosion of Metals and Alloys—Corrosivity of Atmospheres—Determination of Corrosion Rate of Standard Specimens
for the Evaluation of Corrosivity; ISO/TC 156; ISO: Geneva, Switzerland, 2012.
35. Swedish Corrosion Institute. UN/ECE International Cooperative Programme on Effects on Materials Including Historic and Cultural
Monuments; Report no. 1: Technical Manual; Swedish Corrosion Institute: Stockholm, Sweden, 1988.
36. Morcillo, M. Atmospheric Corrosion in Ibero-America: The MICAT Project. Atmos. Corros. 1995. [CrossRef]
37. McCuen, R.H.; Albrecht, P.; Cheng, J. A New Approach to Power-Model Regression of Corrosion Penetration Data. In Corrosion
Forms and Control for Infrastructure; ASTM International: West Conshohocken, PA, USA, 1992. [CrossRef]
38. Albrecht, P.; Hall, T.T., Jr. Atmospheric Corrosion Resistance of Structural Steels. J. Mater. Civ. Eng. 2003, 15, 2–24. [CrossRef]
39. Melchers, R.E. A New Interpretation of the Corrosion Loss Processes for Weathering Steels in Marine Atmospheres. Corros. Sci.
2008, 50, 3446–3454. [CrossRef]
40. Melchers, R.E. Long-Term Corrosion of Cast Irons and Steel in Marine and Atmospheric Environments. Corros. Sci. 2013, 68,
186–194. [CrossRef]
41. Titakis, C.; Vassiliou, P. Evaluation of 4-Year Atmospheric Corrosion of Carbon Steel, Aluminum, Copper and Zinc in a Coastal
Military Airport in Greece. Corros. Mater. Degrad. 2020, 1, 8. [CrossRef]
42. Kreislova, K.; Knotkova, D. The Results of 45 Years of Atmospheric Corrosion Study in the Czech Republic. Materials 2017, 10, 394.
[CrossRef]
43. Tidblad, J. Atmospheric Corrosion of Metals in 2010–2039 and 2070–2099. Atmos. Environ. 2012, 55, 1–6. [CrossRef]
44. Knotkova, D.; Boschek, P.; Kreislova, K. Results of ISO CORRAG Program: Processing of One-Year Data in Respect to Corrosivity
Classification. Atmos. Corros. 1995. [CrossRef]
45. Panchenko, Y.M.; Marshakov, A.I.; Nikolaeva, L.A.; Kovtanyuk, V.V.; Igonin, T.N.; Andryushchenko, T.A. Comparative Estimation
of Long-Term Predictions of Corrosion Losses for Carbon Steel and Zinc Using Various Models for the Russian Territory. Corros.
Eng. Sci. Technol. 2017, 52, 149–157. [CrossRef]
46. Cole, I.S.; Muster, T.H.; Azmat, N.S.; Venkatraman, M.S.; Cook, A. Multiscale Modelling of the Corrosion of Metals under
Atmospheric Corrosion. Electrochim. Acta 2011, 56, 1856–1865. [CrossRef]
47. Nguyen, M.N.; Wang, X.; Leicester, R.H. An Assessment of Climate Change Effects on Atmospheric Corrosion Rates of Steel
Structures. Corros. Eng. Sci. Technol. 2013, 48, 359–369. [CrossRef]
48. Gomes, H.M.; Awruch, A.M. Comparison of Response Surface and Neural Network with Other Methods for Structural Reliability
Analysis. Struct. Saf. 2004, 26, 49–67. [CrossRef]
49. Ahmad, Z. Chapter 10—Atmospheric Corrosion. In Principles of Corrosion Engineering and Corrosion Control; Ahmad, Z., Ed.;
Butterworth-Heinemann: Oxford, UK, 2006; pp. 550–575. ISBN 978-0-7506-5924-6.
50. Vargel, C. Chapter C.2—The Parameters of Atmospheric Corrosion. In Corrosion of Aluminium; Vargel, C., Ed.; Elsevier:
Amsterdam, The Netherlands, 2004; pp. 241–257, ISBN 978-0-08-044495-6.
354
Materials 2021, 14, 3906
51. Schindelholz, E.; Kelly, R.G. Wetting Phenomena and Time of Wetness in Atmospheric Corrosion: A Review. Corros. Rev. 2012, 30.
[CrossRef]
52. Kottek, M.; Grieser, J.; Beck, C.; Rudolf, B.; Rubel, F. World Map of the Köppen-Geiger Climate Classification Updated. Meteorol.
Z. 2006, 15, 259–263. [CrossRef]
53. Committee MT-014 (Corrosion Of Metals). AS 4312-2008 Atmospheric Corrosivity Zones in Australia; Standards Australia: Sydney,
NSW, Australia, 2008.
54. Chico, B.; Otero, E.; Mariaca, L.; Morcillo, M. La Corrosión En Atmósferas Marinas. Efecto de La Distancia a La Costa. Rev. Metal.
1998, 34. [CrossRef]
55. Goerlich, G.F.J.; Cantarino, M.I. Estimaciones de la población rural y urbana a nivel municipal. Estad. Esp. 2015, 57, 5–28.
56. Friedman, J.H. Multivariate Adaptive Regression Splines. Ann. Stat. 1991, 19, 1–67. [CrossRef]
57. Vanegas, J.; Vásquez, F. Multivariate Adaptative Regression Splines (MARS), Una Alternativa Para El Análisis de Series de
Tiempo. Gac. Sanit. 2017, 31, 235–237. [CrossRef]
58. Oja, E.; Kaski, S. Kohonen Maps, 1st ed.; Elsevier Science: Amsterdam, The Netherlands, 1999.
59. Wehrens, R.; Buydens, L. Self- and Super-Organizing Maps in R: The Kohonen Package. J. Stat. Softw. 2007, 21, 1–19. [CrossRef]
60. Villmann, T.; Bauer, H.-U. Applications of the Growing Self-Organizing Map11This Work Has Been Supported by Deutsche
Forschungsgemeinschaft, SFB 185 “Nichtlineare Dynamik”, TP E6. Neurocomputing 1998. [CrossRef]
61. Diazaraque, J.M.M. Los Mapas Auto-Organizados de Kohonen (SOM). Available online: https://docplayer.es/9172924-Los-
mapas-auto-organizados-de-kohonen-som.html (accessed on 15 April 2021).
62. Pachghare, V.; Kulkarni, P.; Nikam, D. Intrusion Detection System Using Self Organizing Maps. In Proceedings of the 2009
International Conference on Intelligent Agent & Multi-Agent Systems (IAMA 2009), Chennai, India, 22–24 July 2009; pp. 1–5.
63. Heasley, E.L.; Millington, J.D.A.; Clifford, N.J.; Chadwick, M.A. A Waterbody Typology Derived from Catchment Controls Using
Self-Organising Maps. Water 2020, 12, 78. [CrossRef]
64. Kohonen, T. Self-Organizing Maps, 3rd ed.; Springer Series in Information Sciences; Springer: Berlin/Heidelberg, Germany, 2001;
ISBN 978-3-540-67921-9.
65. Shanno, D.F. Conditioning of Quasi-Newton Methods for Function Minimization. Math. Comput. 1970, 24, 647–656. [CrossRef]
66. Bronshtein, I.; Semendiaev, K. Manual de Matemáticas para Ingenieros y Estudiantes; Mir: Moscow, Russia; Rubiños-1860: Madrid,
Spain, 1993; ISBN 978-84-8041-022-9.
67. Bourbaki, N. Topological Vector Spaces: Chapters 1–5; Springer: Berlin/Heidelberg, Germany, 2002; ISBN 978-3-540-42338-6.
355
materials
Article
Oxidation of Al-Co Alloys at High Temperatures
Patrik Šulhánek 1 , Marián Drienovský 1 , Ivona Černičková 1 , Libor Ďuriška 1 ,
Ramūnas Skaudžius 2 , Žaneta Gerhátová 1 and Marián Palcut 1, *
1 Faculty of Materials Science and Technology in Trnava, Slovak University of Technology in Bratislava,
J. Bottu 24, 91724 Trnava, Slovakia; [email protected] (P.Š.); [email protected] (M.D.);
[email protected] (I.Č.); [email protected] (L.Ď.); [email protected] (Ž.G.)
2 Faculty of Chemistry and Geosciences, Vilnius University, Naugarduko g. 24, 01513 Vilnius, Lithuania;
[email protected]
* Correspondence: [email protected]
Abstract: In this work, the high temperature oxidation behavior of Al71 Co29 and Al76 Co24 alloys
(concentration in at.%) is presented. The alloys were prepared by controlled arc-melting of Co and
Al granules in high purity argon. The as-solidified alloys were found to consist of several different
phases, including structurally complex m-Al13 Co4 and Z-Al3 Co phases. The high temperature
oxidation behavior of the alloys was studied by simultaneous thermal analysis in flowing synthetic
air at 773–1173 K. A protective Al2 O3 scale was formed on the sample surface. A parabolic rate
law was observed. The rate constants of the alloys have been found between 1.63 × 10−14 and
8.83 × 10−12 g cm−4 s−1 . The experimental activation energies of oxidation are 90 and 123 kJ mol−1
for the Al71 Co29 and Al76 Co24 alloys, respectively. The oxidation mechanism of the Al-Co alloys
is discussed and implications towards practical applications of these alloys at high temperatures
are provided.
1. Introduction
Co-based superalloys are promising materials for high temperature structural applications because
of their high melting points and favorable mechanical properties [1–3]. Applications of these alloys
include gas turbines, aircraft engines, and chemical reactors [4–6]. The Co-based superalloys are
often alloyed with chromium to provide oxidation resistance [7,8]. The superalloys alloyed with
Cr form a compact chromia scale (Cr2 O3 ) on their surface. Nevertheless, at high temperatures and
high oxygen partial pressures, the Cr2 O3 scale is prone to degradation. During long-term oxidation,
volatile high-valent oxides of Cr, such as CrO2 and CrO3 , start to form at the expense of Cr2 O3 .
This effect is called “chromia evaporation” and is often pronounced in humid atmospheres [9]. The loss
of protective chromia scale leads to a reduced life span of the Co-based superalloys. Several authors
have, therefore, investigated the possibility of improving the high temperature oxidation stability of
Co superalloys by alloying with Al [10–12]. Al-based alloys form a protective oxide scale composed of
alumina (Al2 O3 ). Al2 O3 has a lower growth rate compared to Cr2 O3 and is non-volatile. Furthermore,
Al is a non-transition element. It has a smaller tendency to form complex oxides with transition metals
compared to Cr, thereby reducing the risk of scale spallation over time. The formation of alumina scales
may be achieved by pack aluminizing the alloy’s surface [13–15]. The application of Al-Co coatings
could significantly extend the alloy’s lifetime [16–22].
Aluminides are aluminum-based intermetallic compounds with transition metals.
Cobalt aluminides are interesting for high temperature applications since they possess a combination
of high melting points and good corrosion resistance. At ~18–30 at.% Co, different structurally complex
aluminides in the Al-Co binary system have been observed (Figure 1, [23,24]). These include Al9 Co2
(P21 /C), Al5 Co2 (P63 /mmc), Z-Al3 Co (P2/m) and family of Al13 Co4 phases containing m-Al13 Co4
(C2/m), O-Al13 Co4 (Pmn21 ), O’-Al13 Co4 (Pnma), Y1 -Al13 Co4 (C2/m) and Y2 -Al13 Co4 (Immm) [25–34].
Although the individual cobalt aluminides are brittle [35], the Al-Co precipitates may significantly
strengthen the Al alloys [36,37]. The presence of Al5 Co2 and Al13 Co4 intermetallic compounds (IMCs)
may also be beneficial in Co-based alloys as they may greatly improve the alloy’s wear resistance [38].
Most aluminides form protective alumina scales with large resistance against corrosion [39].
Figure 1. Phase diagram of the Al-Co binary system, redrawn from [23,24].
High temperature corrosion studies of cobalt aluminides are limited. Metal oxidation is
a heterogeneous reaction taking place in several elementary steps [40]. In the first step, a gaseous
oxygen molecule is transported to the metal surface. Upon approaching the solid phase, the adsorption
of oxygen molecules to the metal substrate occurs. Subsequently, the oxygen is dissociated to atoms
and later reduced to the O2− anions. In parallel, oxidation of metal atoms takes place at the metal-oxide
interface. Recently, M. Wardé et al. studied the adsorption of oxygen on the Al9 Co2 (001) and Al13 Co4
(100) surfaces at high temperatures and reduced oxygen pressures [41]. At the surfaces, only Al–O
bonding was observed. Al–O distances were also calculated from first principles [41,42]. The Al–O
lengths were shorter in comparison with Co–O distances. The obtained Al–O distances were in
agreement with the typical distances of oxygen adsorption on the Al (111) surface, as well as with the
Al–O distances in Al2 O3 .
Al-rich Al-Co alloys belong to a relatively new group of complex metallic alloys (CMA, [43,44]).
These materials contain structurally complex phases including quasicrystals. The structurally complex
phases have non-periodically ordered atomic arrangements [45,46]. Consequently, the properties of
CMA are different from those observed in traditional materials [43]. The quasicrystalline surfaces
have a good adhesion and low coefficient of friction [47]. Owing to their high hardness and good
oxidation resistance, the Al-TM alloys (TM = transition metal) are suitable for high temperature
coatings [48–51]. The Al-Co CMAs are also interesting for catalytic and hydrogen generation
applications [52–55]. The corrosion studies of Al-Co alloys are limited. Lekatou et al. investigated the
358
Materials 2020, 13, 3152
corrosion behavior of an Al82 Co18 (metal concentrations are given in at.%) alloy in saline solution [56].
Three methods of alloy preparation were investigated: casting, arc-melting and free sintering. The alloy
prepared by arc-melting was found to be the most corrosion-resistant. The Al82 Co18 alloy was composed
of (Al), Al9 Co2 and m-Al13 Co4 . The complex intermetallic m-Al13 Co4 in the alloy was found to have
the highest corrosion resistance. Recently, we have investigated the corrosion behavior of as-solidified
and near equilibrium Al-Co alloys in various environments [24,57,58]. The alloys were composed
of various intermetallic phases. In HCl and NaCl solutions, a pitting corrosion occurred. A higher
corrosion resistance of structurally complex Z-Al3 Co phase in Cl-containing electrolytes was observed.
The difference in the corrosion behavior could be ascribed to the strong covalent character of metallic
bonds in the structurally complex Al-Co phases which prevents aluminum diffusion. The studies also
suggest that the existence of an electrical contact between different alloy phases play an important role
in the overall alloy’s corrosion behavior.
High temperature oxidation studies of Al-Co alloys have been limited to Co-rich alloys only.
Zhang et al. investigated the oxidation behavior of Co-5 at.% Al and Co-10 at.% Al alloys at 973 and
1073 K [59,60]. The oxide scales were primarily composed of cobalt oxide and cobalt-aluminum oxide.
The oxides grown on these alloys were relatively thick (>10 μm after 24 h). Only a limited amount of
Al2 O3 was found at the inner side of the scales. Irving et al. studied the oxidation behavior of Co-xAl
alloys (0 < x < 32.4 at.%) at 1073–1273 K [61]. The authors found that 20–25 at.% Al is necessary to form
a continuous alumina scale. As such, larger Al concentrations are needed to improve the corrosion
resistance of Al-Co alloys.
To our best knowledge, oxidation studies of Al-rich Al-Co alloys have not been reported yet.
In the present work, we aim to study the oxidation behavior of Al-rich Al-Co alloys in air at
773–1173 K. Two alloys, Al71 Co29 and Al76 Co24 (composition in at.%) were prepared by arc-melting.
The composition of the Al71 Co29 alloy was chosen close to Al5 Co2 (71.4 at.%). The composition of the
Al76 Co24 alloy was close to Al13 Co4 (76.5 at.%). The high temperature oxidation of the alloys was
studied with the aim to identify the role of the alloy’s chemical composition and microstructure on the
overall corrosion behavior.
359
Materials 2020, 13, 3152
INCA software (Oxford Instruments Nanoanalysis, Bucks, UK). Regimes of secondary electrons
(SE) and backscatter electrons (BSE) were used during imaging. The accelerating voltage of the
electron beam was 20 kV and working distance was 15 mm. Furthermore, a Panalytical Empyrean
PIXCel 3D diffractometer was used for the phase analysis (Malvern Panalytical Ltd., Malvern, UK).
The diffractometer was working with Bragg–Brentano geometry and used CoKα1 radiation. The X-ray
beam was generated at 40 kV and 50 mA.
Table 1. Chemical compositions of microstructure constituents observed in the as-solidified Al71 Co29
and Al76 Co24 alloys.
The microstructure of the as-solidified Al76 Co24 alloy is presented in Figure 3 in BSE imaging mode.
In this alloy, three different microstructure constituents have been found (light grey, medium grey and
dark grey). The light grey constituent in the Al76 Co24 alloy has 74.4 at.% Al and 25.6 at.% Co (Table 1).
As such, its chemical composition is comparable to chemical compositions of the dark grey constituent
observed in the as-solidified Al71 Co29 alloy (Figure 2). The medium grey constituent contains 75.2 at.%
360
Materials 2020, 13, 3152
Al and 24.8 at.% Co. The dark grey constituent of the as-solidified Al76 Co24 alloy has 81.5 at.% Al and
18.5 at.% Co. The black areas located within the dark grey constituent are pores (Figure 3).
361
Materials 2020, 13, 3152
Figure 4. Cross section of the Al71 Co29 alloy after oxidation at 1173 K for 30 h (a) and energy-dispersive
x-ray spectrometer X-max (EDS) element maps for O (b), Al (c) and Co (d).
Figure 5. Cross section of the Al76 Co24 alloy after oxidation at 1173 K for 30 h (a) and EDS element
maps for O (b), Al (c) and Co(d).
The phase constitution of the oxide scale was studied by room temperature X-ray diffraction.
The diffraction patterns of the Al71 Co29 alloy are presented in Figure 6. In the alloy, peaks corresponding
to θ-Al2 O3 have been identified. θ-Al2 O3 is a metastable alumina phase [64,65]. θ-Al2 O3 structures are
based on a cubic close packing of oxygen anions. The cubic close packing of oxygen anions of θ-Al2 O3 is
deformed monoclinic. The thermodynamically stable α-Al2 O3 adopts a corundum structure. θ-Al2 O3
is a transition phase. It transforms into stable forms of alumina during long term annealing [65].
Metastable θ-Al2 O3 has been found in oxidized Al-Cu-Fe alloys studied previously [66]. It was formed
initially with an orientational relationship to the substrate. At 1173 K, θ-Al2 O3 was found to slowly
transform into α-Al2 O3 with an increasing oxidation time (70 h, [66]).
At 1173 K, an orientation of θ-Al2 O3 in (002) crystallographic plane has been found (Figure 6).
The same behavior was also observed for the Al76 Co24 alloy (Figure 6). This observation indicates
a preferential crystal growth. The morphology of alumina scale formed on the Al71 Co29 and Al76 Co24
alloys after oxidation at 1173 K for 30 h is given in Figure 7. The scale had a blade-like structure.
362
Materials 2020, 13, 3152
The platelet-like scale morphology is indicative of rapid outwards growth. Alumina scales grow by
counter-diffusion of aluminum and oxygen [67]. The ions, however, diffuse faster in polycrystalline
alumina at near-atmospheric oxygen partial pressures. The scale morphology is indicative of rapid
diffusion through the scale. A grain boundary diffusion was probably the preferred transport path for
the Al3+ and O2− ions in the scale.
Figure 6. Room temperature XRD patterns of the scales formed on the Al71 Co29 and Al76 Co24 alloys.
For the discussion of the peak marked with an asterisk (*), please refer to the article text.
Figure 7. Blade-like morphology of alumina scale formed on the oxidized Al71 Co29 alloy (a) and
Al76 Co24 alloy (b) during oxidation at 1173 K.
The un-indexed peak next to θ-Al2 O3 (002), marked with an asterisk in Figure 6, is an alumina
peak. The closest match was found for hexagonal form of Al2 O3 (reference code 98-017-3713, [68]).
Nevertheless, it should also be mentioned that θ-Al2 O3 has a disordered structure [69–72]. As such,
the peak could also be a result of stacking faults (twinning) or other structural defects in θ-Al2 O3 [73–75].
The precise peak assignment was not possible, owing to the difficulty to unambiguously distinguish
the various Al2 O3 polymorphs by XRD technique alone.
The alumina scale was well adherent to the substrate (Figures 4a and 5a). Nevertheless, locally,
a detachment of the scale on the Al71 Co29 alloy was observed (Figure 8). A scale delamination was
found preferentially around β-AlCo dendrites. The situation is shown in Figure 8c. An explanation of
the layer spallation could reside in a mechanical stress developed during oxide growth. The stress is
formed due to different molar volumes of the oxide and the underlying original metal substrate [76].
The stress generated in the oxide during excessive growth may lead to crack formation in the scale.
The pores, cracks and other defects facilitate the access of molecular oxygen to the metal substrate.
363
Materials 2020, 13, 3152
Figure 8. Scale microstructure around β-AlCo dendrites of the Al71 Co29 alloy: (a) an overview;
(b) a detailed view; (c) cross section image of the scale.
Another interesting feature was the porosity of β-AlCo dendrites located beneath the spalled scale
of the Al71 Co29 alloy (Figure 8b). The voids in β-AlCo were not observed before oxidation (Figure 2).
The voids were thus formed during reaction, probably because of rapid aluminum outward diffusion
from the metal surface. During oxidation, metallic species diffuse out from the alloy bulk to the
alloy/oxide interface. The Al atoms, however, leave behind their vacant sites. The vacancies may have
coalesced into larger defects, giving rise to the observed macroscopic porosity. The oxide spallation
has not been observed on the Al76 Co24 alloy. The Al76 Co24 alloy is composed of structurally complex
intermetallic phases (Z-Al3 Co, m-Al13 Co4 and Al9 Co2 , Figure 3). The surfaces of these phases are
typically Al-rich [77]. Aluminum necessary for the scale growth was readily available at the surface of
complex intermetallics. As such, the diffusion of Al atoms from these phases was less rapid compared
to β-AlCo.
The scale of the Al76 Co24 alloy did not show any spallation. It was well adherent to the substrate
and had a wave-like morphology (Figure 5). The wave-like morphology of the scale could be indicative
of epitaxial growth. The XRD pattern showed a preferential orientation of θ-Al2 O3 grains in (002)
crystallographic plane (Figure 6). The θ-Al2 O3 (400) peak was not observed. The Al76 Co24 alloy was
primarily composed of m-Al13 Co4 , with small amounts of Z-Al3 Co and Al9 Co2 (Table 1). A preferred
orientation is typical for m-Al13 Co4 phase [26,78]. It is therefore likely that the θ-Al2 O3 phase was
formed with an orientation relationship to the m-Al13 Co4 phase. This may be reflected in the preferential
orientation of the θ-Al2 O3 grains, as a result of which the intensities of the peaks of this phase change
and some may disappear [79,80].
The oxide layer on the Al71 Co29 alloy, on the other hand, was more uniform (Figure 4).
The preferential layer growth is a self-limited process driven by atomic diffusion, and surface
energy minimization [81,82]. It requires a certain concentration and certain mobility of Al atoms on the
surface. The Al71 Co29 alloy was primarily composed of Al5 Co2 . The Al76 Co24 alloy, on the other hand,
was mainly composed of m-Al13 Co4 (Table 1). The surface of Al5 Co2 is terminated at specific bulk
layers (Al-rich puckered layers) where various fractions of specific sets of Al atoms are missing [83].
The surface of m-Al13 Co4 has a higher density of Al atoms when compared to Al5 Co2 . Previous
364
Materials 2020, 13, 3152
investigations on Al13 Co4 (100) showed that it was terminated by a dense aluminum topmost layer [84].
Therefore, the conditions for the atomic diffusion and subsequent oxide growth on the Al76 Co24 and
Al71 Co29 alloys were different. The preferential oxide growth was less favored on the Al71 Co29 alloy.
The mass gain of the samples was recorded by simultaneous thermogravimetry (TGA).
Thermogravimetric curves of the Al71 Co29 and Al76 Co24 alloys are presented in Figures 9 and 10.
The mass gain of the samples was increasing with increasing time. The kinetic curves obeyed
a parabolic behavior.
Figure 9. Mass gain of the Al71 Co29 alloy during isothermal oxidation in flowing synthetic air.
Figure 10. Mass gain of the Al76 Co24 alloy during isothermal oxidation in flowing synthetic air.
365
Materials 2020, 13, 3152
Table 2. Parabolic rate constants of the Al71 Co29 and Al76 Co24 alloys in air.
Oxidation is a thermally activated process. The parabolic rate constants increase with increasing
temperature. The activation energy of oxidation thus could be obtained from the following equation
EA
log kp = log A − 0.434 (2)
RT
In this equation, A is a constant, EA is the activation energy, R is the molar gas constant
(8.3144 J K−1 mol−1 ) and T is the absolute temperature (in K). The plot of rate constants at different
temperatures is presented in Figure 11. The rate constants follow the Arrhenius-type behavior.
Activation energy has been found from the slope of lines given in Figure 11. The activation energy
of the Al71 Co29 alloy is 90 kJ mol−1 and the activation energy of the Al76 Co24 alloy is 129 kJ mol−1 .
These activation energies are comparable to activation energies for Al oxidation reported by previous
studies [85,86].
Figure 11. Temperature dependence of parabolic rate constants of the Al-Co alloys.
Results presented above show that a protective scale has been formed on the surface of complex
metallic alloys. The rate constants were relatively low, and a thin alumina scale was found on the
surface (Figures 4 and 5). Alumina is formed by the following reaction
4 2
Al + O2 → Al2 O3 (3)
3 3
366
Materials 2020, 13, 3152
The Gibbs energy (ΔG) of reaction (3) is given in Figure 12. ΔG(Al2 O3 ) is very low which indicates
a strong affinity of aluminum towards oxygen. In principle, cobalt oxidation in the Al-Co alloys is also
possible. This reaction can be given by the following equation
Figure 12. Gibbs free energies of metal oxidation reactions at elevated temperatures, redrawn from [54].
Nevertheless, ΔG reaction (4) is considerably larger compared to Gibbs free energy for aluminum
oxidation (Figure 12). A further oxidation of CoO is even more energetically demanding [87,88].
Therefore, CoO tends to decompose in reaction with Al. The reaction can be expressed by the
following equation
4 2
Al + 2CoO → Al2 O3 + Co (5)
3 3
In this disproportionation reaction, CoO is reduced, and Al oxidized. The Gibbs energy of
reaction (5) is negative. Therefore, the selective oxidation of aluminum in the Al-Co alloys is
thermodynamically possible.
The oxidation of Co-rich Al-Co alloys was previously studied by Irving et al. [61]. The alloys
were studied in the as-cast state. The authors studied several Co-xAl alloys with x = 0–32 at.%.
Alloys with small Al concentration (< 10 at.%) formed a single CoO layer. Cobalt oxide layer grew with
a considerably higher corrosion rate compared to Al2 O3 . At intermediate Al concentrations (10–20 at.%),
the authors found that an inner layer of Al2 O3 started to form below the outer CoO scale. With increasing
aluminum concentration, a continuous Al2 O3 scale has been developed. The comparison of the present
results with those from literature is given in Figure 13. Our data complement the previous results of
Irving et al. The continuous Al2 O3 scale forms a barrier to cobalt diffusion. It hinders the nucleation
and growth of cobalt oxides. Irving et al. found that a protective alumina scale can be formed when Al
concentration 24 at.% at 1173 K is reached. Comparable minimum Al concentrations required to form
the external alumina scale were also found for the Ni-Al and Fe-Al alloys [89].
The comparison of the present results with previously studied complex metallic alloys is provided
in Figure 14. Kinetics of oxidation of Al-Cu-Fe and Al-Pd-Mn quasicrystal surfaces was studied in
synthetic air [66,90]. High temperature oxidation kinetics of Al-Cr-Fe complex metallic alloys was
studied in pure oxygen [91]. Our data are comparable to Al-Cu-Fe and Al-Pd-Mn alloys. The parabolic
rate constants of the Al-Cr-Fe complex metallic alloys are lower compared to the remainder of the
367
Materials 2020, 13, 3152
alloys. The oxidation resistance of the Al-Fe-TM (TM = Cr, Cu) alloys is related to the chemical
composition of the oxide scale. The scale found in Al-Cu-Fe alloys was alumina. The scale formed in
Al-Cr-Fe complex metallic alloys, however, was composed of Al2 O3 and (Al0.9 Cr0.1 )2 O3 . The second
scale component provided an additional barrier against corrosion. Chromium as a third alloying
element may improve the overall oxidation resistance of the alloy. The corrosion resistance of alumina
forming alloys alloyed with chromium is higher compared to alloys without Cr. When a sufficient Cr
concentration is available, a complete chromia scale can be formed on top of the alumina scale [92,93].
The duplex Al2 O3 /Cr2 O3 scale has an outstanding corrosion resistance.
Figure 13. Variation of parabolic rate constants of Al-Co alloys with increasing aluminum atomic fraction.
Figure 14. Parabolic rate constants for metal oxidation of Al-TM complex metallic alloys.
Previous authors also studied the microstructure evolution of the oxide scale. In early oxidation
stages, γ-Al2 O3 on the Al63 Cu25 Fe12 alloy was formed with an orientational relationship to the
underlying Al-Cu-Fe quasicrystal [66]. γ-Al2 O3 continued to grow as θ-Al2 O3 until the oxide
layer of several hundred nanometers has been formed. θ-Al2 O3 was later transformed into
the thermodynamically stable α-Al2 O3 . α-Al2 O3 continued to grow with a nodular morphology.
The oxide-nodule formation changed the growth mechanism. A protective layer formation was no
longer observed. A massive spallation occurred after few days of oxidation [66]. A spallation of Al2 O3
368
Materials 2020, 13, 3152
scale from the oxidized Al-Cr-Fe surfaces at high temperatures was also observed [91]. The massive
oxide spallation has not been observed in the present study. However, the oxide spallation was observed
locally on β-AlCo dendrites (Figure 8). It is possible that further stresses in the scale could develop
during long term annealing. Therefore, further experiments on the Al-Co complex metallic alloys are
required to study the effects of long-term annealing and/or thermal cycling on the oxidation behavior.
These studies could shed further light into the long-term oxidation resistance for practical applications
of the Al-Co alloys at high temperatures.
4. Conclusions
In the present work, the oxidation behavior of the Al71 Co29 and Al76 Co24 alloys (concentration in at.%)
was studied at 773–1173 K in air. The alloys were prepared by rapid solidification of Al and Co lumps
in argon. The alloys were studied in as-solidified state. The following conclusions can be drawn:
1. The alloys were composed of different microstructure constituents. The Al76 Co24 alloy was
composed of Al9 Co2 , m-Al13 Co4 and Z-Al3 Co. The Al71 Co29 alloy consisted of Z-Al3 Co, Al5 Co2
and β-AlCo. The precipitation sequences of the constituents were explained based on the
equilibrium Al-Co phase diagram and rapid solidification processes taking place during casting.
2. During oxidation in air, aluminum in the alloys was selectively oxidized and a protective alumina
scale was formed on the alloy surfaces. The oxidation kinetics followed a parabolic rate law.
The rate constants of the alloys were between 1.63 × 10−14 and 8.83 × 10−12 g cm−4 s−1 , depending
on the annealing temperature. The activation energies of oxidation were 90 kJ mol−1 for the
Al71 Co29 alloy and 123 kJ mol−1 for the Al76 Co24 alloy, respectively.
3. The scale of the Al76 Co24 alloy was adherent to the substrate and had a wave-like morphology.
At 1173 K, a preferential orientation of θ-Al2 O3 in (002) crystallographic plane was found.
The scale on the Al71 Co29 alloy was more uniform and spallation was observed locally on the
dendritic β-AlCo.
4. The oxidation kinetics of the Al71 Co29 and Al76 Co24 alloys was comparable to previously studied
Al24 Co76 and Al32 Co68 alloys where a continuous alumina scale had been formed. The increased
Al concentration contributes to the alloy’s corrosion resistance. The continuous Al2 O3 scale forms
a barrier to cobalt diffusion, and it hinders the nucleation and growth of cobalt oxides.
Author Contributions: Conceptualization, M.P., and M.D.; methodology, M.P., M.D., and I.Č.; formal analysis,
P.Š.; investigation, P.Š., M.D., I.Č., L.Ď., R.S., Ž.G., and M.P.; resources, M.D., and M.P.; data curation, P.Š., M.P.,
M.D., I.Č., and L.Ď.; writing—original draft preparation, M.P.; writing—review and editing, M.D., L.Ď., Ž.G.,
and I.Č.; supervision, M.D., R.S., L.Ď., I.Č., and M.P.; project administration, M.P.; funding acquisition, M.P.
All authors have read and agreed to the published version of the manuscript.
Funding: This research was supported by the Grant Agency VEGA of the Slovakian Ministry of Education,
Research, Science and Sport project no. 1/0490/18 and by the Slovak Research and Development Agency project
no. APVV-15-0049.
Conflicts of Interest: The authors declare no conflict of interest.
References
1. Sato, J.; Omori, T.; Oikawa, K.; Ohnuma, I.; Kainuma, R.; Ishida, K. Cobalt-Base High-Temperature Alloys.
Science 2006, 312, 90–91. [CrossRef]
2. Ma, Q.; Wang, W.; Dong, C. Co-Al-W-based superalloys: A summary of current knowledge. Cailiao Daobao/
Mater. Rep. 2020, 34, 3157–3164. [CrossRef]
3. Pollock, T.M.; Dibbern, J.; Tsunekane, M.; Zhu, J.; Suzuki, A. New Co-based γ-γ high-temperature alloys.
JoM 2010, 62, 58–63. [CrossRef]
4. Betteridge, W.; Shaw, S.W.K. Development of superalloys. Mater. Sci. Technol. 1987, 3, 682–694. [CrossRef]
5. Coutsouradis, D.; Davin, A.; Lamberigts, M. Cobalt-based superalloys for applications in gas turbines.
Mater. Sci. Eng. 1987, 88, 11–19. [CrossRef]
6. Klarstrom, D.L. Wrought cobalt- base superalloys. J. Mater. Eng. Perform. 1993, 2, 523–530. [CrossRef]
369
Materials 2020, 13, 3152
7. Moffat, J.P.; Whitfield, T.; Christofidou, K.A.; Pickering, E.; Jones, N.; Stone, H.J. The Effect of Heat Treatment
on the Oxidation Resistance of Cobalt-Based Superalloys. Metals 2020, 10, 248. [CrossRef]
8. Liu, L.; Wu, S.-S.; Dong, Y.; Lü, S. Effects of alloyed Mn on oxidation behaviour of a Co-Ni-Cr-Fe alloy
between 1050 and 1250 ◦ C. Corros. Sci. 2016, 104, 236–247. [CrossRef]
9. Holcomb, G.R. Steam Oxidation and Chromia Evaporation in Ultrasupercritical Steam Boilers and Turbines.
J. Electrochem. Soc. 2009, 156, C292–C297. [CrossRef]
10. Forsik, S.A.J.; Rosas, A.O.P.; Wang, T.; Colombo, G.A.; Zhou, N.; Kernion, S.J.; Epler, M.E. High-Temperature
Oxidation Behavior of a Novel Co-Base Superalloy. Met. Mater. Trans. A 2018, 49, 4058–4069. [CrossRef]
11. Klein, L.; Bauer, A.; Neumeier, S.; Göken, M.; Virtanen, S. High temperature oxidation of γ/γ -strengthened
Co-base superalloys. Corros. Sci. 2011, 53, 2027–2034. [CrossRef]
12. Yan, H.-Y.; Vorontsov, V.; Dye, D. Effect of alloying on the oxidation behaviour of Co-Al-W superalloys.
Corros. Sci. 2014, 83, 382–395. [CrossRef]
13. Goward, G.W.; Cannon, L.W. Pack Cementation Coatings for Superalloys: A Review of History, Theory,
and Practice. J. Eng. Gas Turbines Power 1988, 110, 150–154. [CrossRef]
14. Goward, G. Protective coatings—Purpose, role, and design. Mater. Sci. Technol. 1986, 2, 194–200. [CrossRef]
15. Goward, G. Progress in coatings for gas turbine airfoils. Surf. Coat. Technol. 1998, 108, 73–79. [CrossRef]
16. Streiff, R.; Boone, D.H. Corrosion resistant modified aluminide coatings. J. Mater. Eng. 1988, 10, 15–26.
[CrossRef]
17. Meier, G.; Pettit, F. High-temperature corrosion of alumina-forming coatings for superalloys. Surf. Coat. Technol.
1989, 39, 1–17. [CrossRef]
18. Liu, P.; Liang, K.; Gu, S. High-temperature oxidation behavior of aluminide coatings on a new cobalt-base
superalloy in air. Corros. Sci. 2001, 43, 1217–1226. [CrossRef]
19. Rahman, A.; Jayaganthan, R.; Chandra, R.; Ambardar, R. Microstructural Characterization and Cyclic Hot
Corrosion Behaviour of Sputtered Co-Al Nanostructured Coatings on Superalloy. Oxid. Met. 2011, 76,
307–330. [CrossRef]
20. Rahman, A.; Jayaganthan, R.; Chandra, R.; Ambardar, R. High temperature degradation behavior of sputtered
nanostructured Co-Al coatings on superalloy. Appl. Surf. Sci. 2013, 265, 10–23. [CrossRef]
21. Rahman, A.; Chawla, V.; Jayaganthan, R.; Chandra, R.; Ambardar, R. Degradation behaviour of sputtered
Co-Al coatings on superalloy. Mater. Chem. Phys. 2013, 138, 49–62. [CrossRef]
22. Jiang, J.; Zhou, T.; Shao, W.; Zhou, C. Interdiffusion behavior and lifetime prediction of Co-Al coating on
Ni-based superalloy. J. Alloys Compd. 2019, 786, 920–929. [CrossRef]
23. Stein, F.; He, C.; Dupin, N. Melting behaviour and homogeneity range of B2 CoAl and updated thermodynamic
description of the Al-Co system. Intermetallics 2013, 39, 58–68. [CrossRef]
24. Priputen, P.; Palcut, M.; Babinec, M.; Mišík, J.; Černičková, I.; Janovec, J. Correlation between Microstructure
and Corrosion Behavior of Near-Equilibrium Al-Co Alloys in Various Environments. J. Mater. Eng. Perform.
2017, 26, 3970–3976. [CrossRef]
25. Grushko, B.; Wittenberg, R.; Bickmann, K.; Freiburg, C. The constitution of aluminum-cobalt alloys between
Al5Co2 and Al9Co2. J. Alloys Compd. 1996, 233, 279–287. [CrossRef]
26. Priputen, P.; Kusy, M.; Drienovský, M.; Janičkovič, D.; Čička, R.; Černičková, I.; Janovec, J. Experimental
reinvestigation of Al-Co phase diagram in vicinity of Al13Co4 family of phases. J. Alloys Compd. 2015, 647,
486–497. [CrossRef]
27. Cooper, M.J. The electron distribution in the phases CoAl and NiAl. Philos. Mag. 1963, 8, 811–821. [CrossRef]
28. Burkhardt, U.; Ellner, M.; Grin, Y.; Baumgartner, B. Powder diffraction refinement of the Co2Al5 structure.
Powder Diffr. 1998, 13, 159–162. [CrossRef]
29. Ma, X.L.; Li, X.Z.; Kuo, K.H. A family of τ-inflated monoclinic Al13Co4 phases. Acta Crystallogr. Sect. B
Struct. Sci. 1995, 51, 36–43. [CrossRef]
30. Freiburg, C.; Grushko, B.; Wittenberg, R.; Reichert, W. Once More about Monoclinic Al13Co4. Mater. Sci. Forum
1996, 228, 583–586. [CrossRef]
31. Grin, J.; Burkhardt, U.; Ellner, M.; Peters, K. Crystal structure of orthorhombic Co4Al13. J. Alloys Compd.
1994, 206, 243–247. [CrossRef]
32. Fleischer, F.; Weber, T.; Jung, D.; Steurer, W. o’-Al13Co4, a new quasicrystal approximant. J. Alloys Compd.
2010, 500, 153–160. [CrossRef]
370
Materials 2020, 13, 3152
33. Sugiyama, K.; Genba, M.; Hiraga, K.; Waseda, Y. The structure of Y-Al13-xCo4 (x = 0.8) analyzed by
single crystal X-ray diffraction coupled with anomalous X-ray scattering. J. Alloys Compd. 2010, 494, 98–101.
[CrossRef]
34. Boström, M.; Rosner, H.; Prots, Y.; Burkhardt, U.; Grin, Y. The Co2Al9 Structure Type Revisited. Z. Anorg.
Allg. Chem. 2005, 631, 534–541. [CrossRef]
35. Heggen, M.; Deng, D.; Feuerbacher, M. Plastic deformation properties of the orthorhombic complex metallic
alloy phase Al13Co4. Intermetallics 2007, 15, 1425–1431. [CrossRef]
36. Samuel, F.H. A study of the microstructure, mechanical properties and failure behaviour of Al-Li-Co
melt-spun ribbons: Effect of Al9Co2 phase particle precipitation. J. Mater. Sci. 1986, 21, 3097–3107. [CrossRef]
37. Lekatou, A.G.; Sfikas, A.; Karantzalis, A.E. The influence of the fabrication route on the microstructure and
surface degradation properties of Al reinforced by Al 9 Co 2. Mater. Chem. Phys. 2017, 200, 33–49. [CrossRef]
38. Wolf, W.; Schulz, R.; Savoie, S.; Bolfarini, C.; Kiminami, C.S.; Botta, W. Structural, mechanical and thermal
characterization of an Al-Co-Fe-Cr alloy for wear and thermal barrier coating applications. Surf. Coat. Technol.
2017, 319, 241–248. [CrossRef]
39. Hagel, W.C. The oxidation of iron, nickel and cobalt-base alloys containing aluminum. Corrosion 1965, 21,
316–326. [CrossRef]
40. Young, D.J. Oxidation of Pure Metals. In High Temperature Oxidation and Corrosion of Metals, 2nd ed.; Elsevier:
Amsterdam, The Netherlands, 2016; Chapter 3; pp. 85–144. [CrossRef]
41. Warde, M.; Herinx, M.; Ledieu, J.; Loli, L.S.; Fournée, V.; Gille, P.; Le Moal, S.; Barthés-Labrousse, M.-G.
Adsorption of O2 and C2H (n = 2, 4, 6) on the Al9Co2(0 0 1) and o-Al13Co4(1 0 0) complex metallic alloy
surfaces. Appl. Surf. Sci. 2015, 357, 1666–1675. [CrossRef]
42. Villaseca, S.A.; Loli, L.N.S.; Ledieu, J.; Fournée, V.; Gille, P.; Dubois, J.-M.; Gaudry, É. Oxygen adsorption
on the Al9Co2(001) surface: First-principles and STM study. J. Phys. Condens. Matter 2013, 25, 355003.
[CrossRef] [PubMed]
43. Dubois, J.-M. Properties- and applications of quasicrystals and complex metallic alloys. Chem. Soc. Rev. 2012,
41, 6760. [CrossRef] [PubMed]
44. Wolf, W.; Bolfarini, C.; Kiminami, C.S.; Botta, W. Designing new quasicrystalline compositions in
Al-based alloys. J. Alloys Compd. 2020, 823, 153765. [CrossRef]
45. Steurer, W. Twenty years of structure research on quasicrystals. Part I. Pentagonal, octagonal, decagonal and
dodecagonal quasicrystals. Z. Krist. Cryst. Mater. 2004, 219, 391–446. [CrossRef]
46. Barber, E.M. Chemical Bonding and Physical Properties in Quasicrystals and Their Related Approximant
Phases: Known Facts and Current Perspectives. Appl. Sci. 2019, 9, 2132. [CrossRef]
47. Rabson, D. Toward theories of friction and adhesion on quasicrystals. Prog. Surf. Sci. 2012, 87, 253–271.
[CrossRef]
48. Balbyshev, V.; King, D.; Khramov, A.; Kasten, L.; Donley, M. Investigation of quaternary Al-based quasicrystal
thin films for corrosion protection. Thin Solid Films 2004, 447, 558–563. [CrossRef]
49. Moskalewicz, T.; Dubiel, B.; Wendler, B. AlCuFe(Cr) and AlCoFeCr coatings for improvement of elevated
temperature oxidation resistance of a near-α titanium alloy. Mater. Charact. 2013, 83, 161–169. [CrossRef]
50. Parsamehr, H.; Chang, S.-Y.; Lai, C.-H. Mechanical and surface properties of aluminum-copper-iron
quasicrystal thin films. J. Alloy Compd. 2018, 732, 952–957. [CrossRef]
51. Parsamehr, H.; Chen, T.-S.; Wang, D.-S.; Leu, M.-S.; Han, I.; Xi, Z.; Tsai, A.-P.; Shahani, A.J.; Lai, C.-H.
Thermal spray coating of Al-Cu-Fe quasicrystals: Dynamic observations and surface properties. Materialia
2019, 8, 100432. [CrossRef]
52. Chatelier, C.; Garreau, Y.; Piccolo, L.; Vlad, A.; Resta, A.; Ledieu, J.; Fournée, V.; De Weerd, M.-C.; Picca, F.-E.;
De Boissieu, M.; et al. From the Surface Structure to Catalytic Properties of Al5Co2(210): A Study Combining
Experimental and Theoretical Approaches. J. Phys. Chem. C 2020, 124, 4552–4562. [CrossRef]
53. Piccolo, L.; Chatelier, C.; De Weerd, M.-C.; Morfin, F.; Ledieu, J.; Fournée, V.; Gille, P.; Gaudry, E. Catalytic
properties of Al13TM4 complex intermetallics: Influence of the transition metal and the surface orientation
on butadiene hydrogenation. Sci. Technol. Adv. Mater. 2019, 20, 557–567. [CrossRef] [PubMed]
54. Meier, M.; Ledieu, J.; Fournée, V.; Gaudry, E. Semihydrogenation of Acetylene on Al5Co2 Surfaces.
J. Phys. Chem. C 2017, 121, 4958–4969. [CrossRef]
55. Soler, L.; Macanás, J.; Muñoz, M.; Casado, J. Aluminum and aluminum alloys as sources of hydrogen for fuel
cell applications. J. Power Sources 2007, 169, 144–149. [CrossRef]
371
Materials 2020, 13, 3152
56. Lekatou, A.G.; Sfikas, A.; Karantzalis, A.E.; Sioulas, D. Microstructure and corrosion performance of
Al-32%Co alloys. Corros. Sci. 2012, 63, 193–209. [CrossRef]
57. Palcut, M.; Priputen, P.; Kusý, M.; Janovec, J. Corrosion behaviour of Al-29at%Co alloy in aqueous NaCl.
Corros. Sci. 2013, 75, 461–466. [CrossRef]
58. Palcut, M.; Priputen, P.; Šalgó, K.; Janovec, J. Phase constitution and corrosion resistance of Al-Co alloys.
Mater. Chem. Phys. 2015, 166, 95–104. [CrossRef]
59. Zhang, H.H.; Xiang, J.; Wang, W.; Shi, Y.-J. The Oxidation of Co-5Al Alloys in 1 Atm of Pure O2 at
700 and 800 ◦ C. Adv. Mater. Res. 2011, 295, 319–322. [CrossRef]
60. Zhang, H.H.; Xiang, J.H.; Xu, X.C.; Wang, C. Comparison of Oxidation Behavior of Binary Co-10X
(x = Al, Si, Cr) Alloys at 973 and 1073K. Appl. Mech. Mater. 2013, 395, 238–242. [CrossRef]
61. Irving, G.N.; Stringer, J.; Whittle, D.P. The high-temperature oxidation resistance of Co-Al alloys. Oxid. Met.
1975, 9, 427–440. [CrossRef]
62. Agliozzo, S.; Brunello, E.; Klein, H.; Mancini, L.; Hartwig, J.; Baruchel, J.; Gastaldi, J. Extended investigation of
porosity in quasicrystals by synchrotron X-ray phase contrast radiography—I: In icosahedral AlPdMn grains.
J. Cryst. Growth 2005, 281, 623–638. [CrossRef]
63. Brunello, E.; Agliozzo, S.; Klein, H.; Mancini, L.; Härtwig, J.; Baruchel, J.; Gastaldi, J. Extended investigation
of porosity in quasicrystals by synchrotron X-ray phase contrast radiography: Part II, in grains of other
alloys and structures than AlPdMn. J. Cryst. Growth 2005, 282, 228–235. [CrossRef]
64. Prescott, R.; Graham, M.J. The oxidation of iron-aluminum alloys. Oxid. Met. 1992, 38, 73–87. [CrossRef]
65. Chevalier, S. Formation and growth of protective alumina scales. In Developments in High Temperature Corrosion
and Protection of Materials, 1st ed.; Woodhead Publishing: Abington, Cambridge, UK, 2008; Chapter 10;
pp. 290–328. [CrossRef]
66. Wehner, B.I.; Köster, U. Microstructural Evolution of Alumina Layers on an Al-Cu-Fe Quasicrystal during
High-Temperature Oxidation. Oxid. Met. 2000, 54, 445–456. [CrossRef]
67. Tolpygo, V.; Clarke, D.R. Microstructural evidence for counter-diffusion of aluminum and oxygen during the
growth of alumina scales. Mater. High Temp. 2003, 20, 261–271. [CrossRef]
68. Dan’Ko, A.J.; Rom, M.A.; Sidelnikova, N.S.; Nizhankovskiy, S.V.; Budnikov, A.T.; Grin’, L.A.; Kaltaev, K.S.-O.
Transformation of the corundum structure upon high-temperature reduction. Crystallogr. Rep. 2008, 53,
1112–1118. [CrossRef]
69. Trunov, M.A.; Schoenitz, M.; Zhu, X.; Dreizin, E.L. Effect of polymorphic phase transformations in Al2O3
film on oxidation kinetics of aluminum powders. Combust. Flame 2005, 140, 310–318. [CrossRef]
70. Zhou, R.S.; Snyder, R.L. Structures and transformation mechanisms of the η, γ and θ transition aluminas.
Acta Crystallogr. Sect. B Struct. Sci. 1991, 47, 617–630. [CrossRef]
71. Kovarik, L.; Bowden, M.; Shi, D.; Washton, N.M.; Andersen, A.; Hu, J.Z.; Lee, J.; Szanyi, J.; Kwak, J.-H.;
Peden, C.H.F. Unraveling the Origin of Structural Disorder in High Temperature Transition Al2O3: Structure
of θ-Al2O3. Chem. Mater. 2015, 27, 7042–7049. [CrossRef]
72. Jbara, A.S.; Othaman, Z.; Saeed, M. Structural, morphological and optical investigations of θ-Al2O3 ultrafine
powder. J. Alloys Compd. 2017, 718, 1–6. [CrossRef]
73. Wang, Y.; Bronsveld, P.; De Hosson, J.T.M.; Djuričić, B.; McGarry, D.; Pickering, S. Twinning in θ Alumina
Investigated with High Resolution Transmission Electron Microscopy. J. Eur. Ceram. Soc. 1998, 18, 299–304.
[CrossRef]
74. Kovarik, L.; Bowden, M.; Andersen, A.; Jaegers, N.R.; Washton, N.; Szanyi, J. Quantification of High
Temperature Transitions and Their Phase Transformations. Available online: https://chemrxiv.org/articles/preprint/
Quantification_of_High_Temperature_Transition_Al2O3_and_Their_Phase_Transformations/12584783 (accessed
on 4 July 2020).
75. Boumaza, A.; Favaro, L.; Ledion, J.; Sattonnay, G.; Brubach, J.; Berthet, P.; Huntz, A.; Roy, P.; Tétot, R.
Transition alumina phases induced by heat treatment of boehmite: An X-ray diffraction and infrared
spectroscopy study. J. Solid State Chem. 2009, 182, 1171–1176. [CrossRef]
76. McCafferty, E. High-Temperature Gaseous Oxidation. In Introduction to Corrosion Science; Springer: New York,
NY, USA, 2009; pp. 453–476. [CrossRef]
77. Ledieu, J.; Gaudry, E.; Fournée, V. Surfaces of Al-based complex metallic alloys: Atomic structure, thin film
growth and reactivity. Sci. Technol. Adv. Mater. 2014, 15, 34802. [CrossRef] [PubMed]
372
Materials 2020, 13, 3152
78. Korte-Kerzel, S.; Schnabel, V.; Clegg, W.; Heggen, M. Room temperature plasticity in m-Al13Co4 studied by
microcompression and high resolution scanning transmission electron microscopy. Scr. Mater. 2018, 146,
327–330. [CrossRef]
79. Wang, W.; Yang, W.; Liu, Z.; Lin, Y.; Zhou, S.; Qian, H.; Wang, H.; Lin, Z.; Li, G. Epitaxial growth
and characterization of high-quality aluminum films on sapphire substrates by molecular beam epitaxy.
CrystEngComm 2014, 16, 7626–7632. [CrossRef]
80. Alessandri, M.; Piagge, R.; Caniatti, M.; Del Vitto, A.; Wiemer, C.; Pavia, G.; Alberici, S.; Bellandi, E.; Nale, A.
Structural and Chemical Investigation of Annealed Al2O3 Films for Interpoly Dielectric Application in Flash
Memories. ECS Trans. 2006, 3, 183–192. [CrossRef]
81. Queraltó, A.; De La Mata, M.; Arbiol, J.; Obradors, X.; Puig, T. Disentangling Epitaxial Growth Mechanisms
of Solution Derived Functional Oxide Thin Films. Adv. Mater. Interfaces 2016, 3, 1600392. [CrossRef]
82. Mattox, D.M. Atomistic Film Growth and Some Growth-Related Film Properties. In Handbook of Physical
Vapor Deposition (PVD) Processing, 2nd ed.; Elsevier: Oxford, UK, 2010; pp. 333–398. [CrossRef]
83. Meier, M.; Ledieu, J.; De Weerd, M.-C.; Huang, Y.-T.; Abreu, G.J.P.; Pussi, K.; Diehl, R.; Mazet, T.; Fournée, V.;
Gaudry, E. Interplay between bulk atomic clusters and surface structure in complex intermetallic compounds:
The case study of the Al5Co2(001) surface. Phys. Rev. B 2015, 91, 085414. [CrossRef]
84. Anand, K.; Fournée, V.; Prevot, G.; Ledieu, J.; Gaudry, E. Nonwetting Behavior of Al-Co Quasicrystalline
Approximants Owing to Their Unique Electronic Structures. ACS Appl. Mater. Interfaces 2020, 12, 15793–15801.
[CrossRef]
85. Gulbransen, E.A.; Wysong, W.S. Thin Oxide Films on Aluminum. J. Phys. Chem. 1947, 51, 1087–1103.
[CrossRef]
86. Smeltzer W, W. Oxidation of An Aluminum-3 Per Cent Magnesium Alloy in the Temperature Range 200–550
◦ C. J. Electrochem. Soc. 1958, 105, 67–71. [CrossRef]
87. Sabat, K.C.; Paramguru, R.K.; Pradhan, S.; Mishra, B.K. Reduction of Cobalt Oxide (Co3O4) by Low
Temperature Hydrogen Plasma. Plasma Chem. Plasma Process. 2014, 35, 387–399. [CrossRef]
88. Chattopadhyay, B.; Wood, G.C. The transient oxidation of alloys. Oxid. Met. 1970, 2, 373–399. [CrossRef]
89. Stott, F.H. Influence of alloy additions on oxidation. Mater. Sci. Technol. 1989, 5, 734–740. [CrossRef]
90. Wehner, B.; Köster, U.; Rüdiger, A.; Pieper, C.; Sordelet, D. Oxidation of Al–Cu–Fe and Al–Pd–Mn quasicrystals.
Mater. Sci. Eng. A 2000, 294, 830–833. [CrossRef]
91. Prud’Homme, N.; Ribot, P.; Herinx, M.; Gille, P.; Bauer, B.; De Weerd, M.-C.; Barthés-Labrousse, M.-G.
High temperature oxidation of AlCrFe complex metallic alloys. Corros. Sci. 2014, 89, 118–126. [CrossRef]
92. Irving, G.N.; Stringer, J.; Whittle, D.P. The Oxidation Behavior of Co-Cr-Al Alloys at 1000 ◦ C. Corrosion 1977,
33, 56–60. [CrossRef]
93. Wood, G.C.; Stott, F.H. The influence of aluminum additions on the oxidation of Co-Cr alloys at
1000 and 1200 ◦ C. Oxid. Met. 1971, 3, 365–398. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
373
materials
Article
Effect of Mo Addition on the Chemical Corrosion
Process of SiMo Cast Iron
Marcin Stawarz 1, * and Paweł M. Nuckowski 2
1 Department of Foundry Engineering, Silesian University of Technology, 7 Towarowa Street,
44-100 Gliwice, Poland
2 Department of Engineering Materials and Biomaterials, Silesian University of Technology,
18A Konarskiego Street, 44-100 Gliwice, Poland; [email protected]
* Correspondence: [email protected]; Tel.: +48-32-338-5532
Abstract: The study was carried out to evaluate five SiMo cast iron grades and their resistance to
chemical corrosion at elevated temperature. Corrosion tests were carried out under conditions of an
actual cyclic operation of a retort coal-fired boiler. The duration of the study was 3840 h. The range of
temperature changes during one cycle was in the range of 300–650 ◦ C. Samples of SiMo cast iron
with Si content at the level of 5% and variable Mo content in the range 0%–2.5% were used as the
material for the study. The examined material was subjected to preliminary metallographic analysis
using scanning microscopy and an Energy dispersive spectroscopy (EDS) system. The chemical
composition was determined on the basis of a Leco spectrometer and a Leco carbon and sulfur
analyzer. The examination of the oxide layer was carried out with the use of Scanning electron
microscope (SEM), EDS, and X-ray diffraction (XRD) methods. It was discovered that, in the analyzed
alloys, oxide layers consisting of Fe2 O3 , Fe3 O4 , SO2 , and Fe2 SiO4 were formed. The analyzed oxide
layers were characterized by high adhesion to the substrate material, and their total thickness was
about 20 μm.
Keywords: chemical corrosion; SiMo cast iron; fayalite; hematite; magnetite; maghemite; sulfur
dioxide
1. Introduction
Corrosive behavior of SiMo cast iron in the air and flue gases was presented in the works [1–5].
When we subject pure iron to the corrosion process at elevated temperatures and in the ambient
atmospheric air, a multilayer oxide structure composed of FeO, Fe3 O4 , and Fe2 O3 is formed [5].
For ductile cast iron, an oxide layer is formed, located both in the material and in the surface layer as a
result of migration of Fe atoms [5]. After introducing an alloying element in the form of Si into cast
iron, a SiO2 compound is formed at the metal–oxide layer point of contact, which constitutes a barrier
to further oxidation processes [5]. Simultaneously, SiO2 can react with O, Fe, and FeO. The result of
these reactions may be the formation of fayalite, Fe2 SiO4 [6–8]. In the paper [9], the author writes that
the oxide layer on the surface of SiMo cast iron is composed of the following sub-layers situated from
the outside to the inside of the material: Fe2 O3 , Fe3 O4 , FeO, FeO + Fe2 SiO4 .
The oxide layer adheres well to the base material and the inner layer consisting of FeO + Fe2 SiO4 [9].
The higher the silicon content in the base material, the faster the oxide layer forms. A number of studies
concerning the corrosion resistance of SiMo cast iron focus on a relatively short time of exposure to
oxidation (500 h on average) [10]. These studies are conducted mainly in terms of the use of SiMo
cast iron in automotive castings, as described by Rouczka [11] and many other authors [12–17]. SiMo
cast iron is an increasingly popular material, and research on this material is also conducted with a
focus on optimizing the manufacturing process. In their work, Guzik et al. [18] write about the method
of introducing two flexible hoses with the diameter of Ø 9 mm; one filled with a FeSi + Mg mixture,
and the other with a graphitizing modifier for the treatment drum ladle. Guzik et al. [18] describe it as
a new method of secondary treatment of ferritic cast iron production of SiMo type. This method can be
used for the production of ductile iron melted in an induction furnace [18,19].
SiMo cast iron can also be successfully used in other areas of industry: exhaust parts for combustion
engines, turbocharger housings and rotors, gas turbine components, molds for the glass industry, molds
for aluminum alloys, zinc, forging dies, heat treatment furnace components, aluminum melting furnace
components, and waste incineration furnaces. This happens wherever elevated operating temperatures
and gases resulting from the combustion, e.g., of solid fuels are involved. A good example of such a
system is a coal-fired retort furnace. Nyashina and her team write about the problems related to the
emission of pollutants during the combustion process [20]. Released into the atmosphere with exhaust
gases, nitrogen oxides (mainly NO and NO2 ) are the main reason why photochemical smog appears,
which reaches the stratosphere to act as a catalyst for ozone layer depletion. Rapid oxidation of NOx
and SOx and their interaction with water vapor in the atmosphere generates tiny droplets of sulfuric
(H2 SO4 ) and nitric (HNO3 ) acids [20]. Sulfuric acid causes significant losses in the ecosystem, which
has been mentioned by many authors [21]. It is important to optimize the combustion process of solid
fuels in boilers by improving the materials from which these boilers are built. For the above reasons, in
this work, studies of resistance to chemical corrosion of SiMo cast iron were carried out during actual
operation of a retort boiler. The duration of the study was 3840 hours. To date, the corrosion resistance
of SiMo cast iron during the operation of a retort boiler has not been described in the literature. Due to
its properties, it can be successfully used for manufacturing furnace elements fired with solid, liquid,
or gaseous fuels.
376
Materials 2020, 13, 1745
Figure 1. Scheme of a single retort with (yellow) SiMo samples placed above the stoker.
In order to determine the phase composition of the studied material, X-ray diffraction analyses
were carried out with the use of an X’Pert Pro multipurpose x-ray diffractometer by Panalytical (Almelo,
The Netherlands). The measurements were conducted utilizing filtered radiation of a cobalt anode
lamp (λKα = 0.179 nm) as well as a PIXcel 3D detector on the diffracted beam axis. The diffraction
lines were recorded in the Bragg–Brentano geometry in the angular scope of 10◦ –120◦ (2θ), with the
step of 0.026◦ and the step time of 100 s. Furthermore, to obtain more precise information from the
surface oxide layer, grazing incidence diffraction (GID) geometry with a proportional detector on
the diffracted beam axis was used. In this geometry, a primary X-ray beam was set at a constant,
low angle (1.5o ) related to the sample plane, which affected the corresponding slope of diffraction
vectors related to the normal to surface. This allowed us to obtain during the measurement a constant
penetration depth of the X-ray beam, limited mainly to the surface oxide layer. The analysis of the
obtained diffraction patterns was made in the Panalytical High Score Plus software (ver.: 3.0e), with
the dedicated Panalytical Inorganic Crystal Structure Database (PAN-ICSD).
The analysis of the structure and the chemical composition was performed on a Phenom ProX
scanning microscope (Phenom-World Eindhoven, Noord-Brabant, Netherlands) equipped with an
energy-dispersive X-ray spectrometer (EDS).
3. Results
377
Materials 2020, 13, 1745
carbide (Mo2 C), a carburized zone (Figure 2e), and the passive layer and the loose oxide layer on the
top surface of the samples. Microstructure components are highlighted in Figure 2.
D E
F G
H
Figure 2. Microstructure of SiMo cast iron after the corrosion test cycle. SiMo cast iron with addition of:
(a) 0.01% Mo, (b) 0.47% Mo, (c) 1.09% Mo, (d) 1.92% Mo, (e) 2.51% Mo. SEM.
378
Materials 2020, 13, 1745
In all cases (SiMo samples 1–5), the thickness of passive layers is around 10 μm. For elevated
molybdenum content (Figure 2e, SiMo melt 5, 2.51% Mo) in the near-surface layer, the carburized
zone in the form of black inclusions distributed in the vicinity of Mo2 C carbide precipitates is clearly
visible. All the cases considered are characterized by a cohesive passive layer tightly adhering to the
sample. No cracks nor defects in the passive layer were observed, even after the process of preparing
metallographic sections (cutting, grinding, and polishing).
D E F
&Ğ &
G H
K ^ŝ
Figure 3. Microstructure of SiMo cast iron, 0.01% Mo (a). Elements decomposition maps. Map for
(b) Fe, (c) C, (d) O, and (e) Si.
The characteristic phenomenon of these alloys is their ability to perform so-called “self-healing”.
Removing the passive layer creates a new one in its place. The average thickness of the loose oxide
layer was 10 μm for the tested samples, while the thickness of the passive layer was also around 10 μm.
Of course, the thickness of this layer depends on the local conditions on the surface of the sample
exposed to the corrosive agents.
379
Materials 2020, 13, 1745
D E F
&Ğ &
G H
K ^ŝ
Figure 4. Microstructure of SiMo 3 cast iron, 1.09% Mo (a). Map for (b) Fe, (c) C, (d) O, and (e) Si.
D E F
&Ğ K
G H I
^ŝ DŽ
Figure 5. View of the carburized layer. SiMo sample 5. 2.51% Mo (a). Map for (b) Fe, (c) O, (d) C, (e) Si,
and (f) Mo.
380
Materials 2020, 13, 1745
Figure 5 shows the analysis of the sample area (for the SiMo melt 5), where a significant degree of
carburation of the casting material layer was observed. The layer is placed under the passive layer.
This can be explained by the high concentration of carbon in the corrosive atmosphere of the furnace,
and the penetration of carbon atoms through the passive layer towards the center of the casting.
Figure 7. Diffractogram from a sample. Blue chart, Bragg–Brentano geometry; red chart, grazing
incidence diffraction (GID) geometry. SiMo 3. 1.09% Mo.
381
Materials 2020, 13, 1745
From the results of the analysis presented in Figure 7, the following components of the
surface oxide layer were identified. The compounds identified were, among others: α - Fe2 O3
(hematite) with a hexagonal lattice R-3c (98-002-2505), α - Fe3 O4 (magnetite) with a cubic lattice Fd-3m
(98-026-3007), γ - Fe2 O3 (maghemite) with a cubic lattice Fd-3m (98-017-2905), and a small share of SO2
(an orthorhombic lattice).
4. Discussion
The presented photos of SiMo cast iron samples (Figure 2), after a series of chemical corrosion
resistance tests, clearly show the microstructure of the cast iron with the passive layer and the loose
oxide layer on the surface of the samples. The obtained results are analogous to the results presented
by M. Ekström in his paper [5]. The results obtained by M. Ekström were obtained under different test
conditions in a flue gas atmosphere. For all cases analyzed in the study, the thickness of the oxide and
passive layers oscillates around the level of 20 μm. For elevated molybdenum content (SiMo melt 5,
2.51% Mo) in the near-surface layer, the carburized zone, in the form of black inclusions distributed
in the vicinity of Mo2 C carbide precipitates, is clearly visible. The layer is placed under the passive
layer, which is clearly visible in Figure 5. We found no description of a similar case in the available
literature. The obtained effect can be explained by the high concentration of carbon in the corrosive
atmosphere of the furnace; penetration of carbon atoms through the passive layer in the direction of
the casting center, combined with the high concentration of molybdenum in the casting, resulted in the
accumulation of carbon atoms around the precipitates rich in molybdenum (Mo2 C carbides).
The results obtained from the XRD analysis allowed for the following components of the surface
oxide layer to be described, among others: α - Fe2 O3 (hematite), α - Fe3 O4 (magnetite), γ - Fe2 O3
(maghemite), and SO2 [22–24].
The X-ray diffraction patterns of magnetite and maghemite are very similar. This is related to
their similar structures. Both of these two oxide phases crystallize in the cubic system and their lattice
parameters are very close. For this reason, it is difficult to differentiate these structures. However,
some works [25,26] report that the maghemite phase gives two additional small diffraction lines at
23.77o (210) and 26.10o (211). One of these, line (210), was identified even on a diffractogram obtained
in Bragg–Brentano geometry. The formed magnetite and maghemite oxide layers do not show the
much higher intensity of line (113) in relation to the highest hematite line (104). This result can be
explained by the similar crystal growth of these oxide phases. Also, some authors report that the
formation of maghemite (γ - Fe2 O3 ) is a result of oxidation of the magnetite (α - Fe3 O4 ) [27,28]. One of
the variables deciding the quantitative share of hematite, magnetite, and maghemite is the range of
oxidation temperature described by M. Marciuš et al. [29].
5. Conclusions
Based on the conducted study, it can be stated that SiMo cast iron is fully resistant to chemical
corrosion during retort furnace operations. All SiMo cast iron samples were characterized by a cohesive
oxidized layer, consisting of a passive layer on the casting material side and an oxide surface layer.
The two layers adhered quite well to one another. No cracks nor defects in any of the elements
of the oxidized layer were observed. The surface oxide layer was found to consist of the following
compounds: Fe2 O3 , Fe3 O4 , and SO2 .
For the sample with increased Mo content, a significant carburization of the near-surface layer of
the sample was observed, especially in the areas adjacent to molybdenum carbide. The carburized
edge zone of the sample did not affect the corrosion resistance of SiMo cast iron.
The areas strongly enriched with silicon are the fayalite Fe2 SiO4 resulting from the reaction of
SiO2 with O, Fe, and FeO.
Due to the relatively low operating temperature range, we suggest that the Mo content in the
alloy be reduced to the range of 0%–0.5% Mo.
382
Materials 2020, 13, 1745
Author Contributions: research concept, M.S.; conducting the experiment, M.S.; microstructure studies, M.S. and
P.M.N.; diffractive analysis, P.M.N.; writing—original manuscript preparation, M.S.; writing—review and edition,
M.S. and P.M.N. All authors have read and agreed to the published version of the manuscript.
Funding: This research received no external funding.
Acknowledgments: This publication was financed by the statutory subsidy of the Faculty of Mechanical
Engineering of the Silesian University of Technology in 2019.
Conflicts of Interest: The authors report no conflicts of interest.
References
1. Tholence, F.; Norell, M. High-Temperature Corrosion of Cast Irons and Cast Steels in Dry Air.
Mater. Sci. Forum 2001, 369, 197–204. [CrossRef]
2. Tholence, F.; Norell, M. AES characterization of oxide grains formed on ductile cast irons in exhaust
environments. Surf. Interface Anal. 2002, 34, 535–539. [CrossRef]
3. Tholence, F.; Norell, M. Nitride precipitation during high temperature corrosion of ductile cast irons in
synthetic exhaust gases. J. Phys. Chem. Solids 2005, 66, 530–534. [CrossRef]
4. Tholence, F.; Norell, M. High Temperature Corrosion of Cast Alloys in Exhaust Environments I-Ductile Cast
Irons. Oxid. Met. 2007, 69, 13–36. [CrossRef]
5. Ekström, M.; Szakálos, P.; Jonsson, S. Influence of Cr and Ni on High-Temperature Corrosion Behavior of
Ferritic Ductile Cast Iron in Air and Exhaust Gases. Oxid. Met. 2013, 80, 455–466. [CrossRef]
6. Yang, Y.L.; Cao, Z.Y.; Qi, Y.; Liu, Y.B. The Study on Oxidation Resistance Properties of Ductile Cast Irons for
Exhaust Manifold at High Temperatures. Adv. Mater. Res. 2010, 97, 530–533. [CrossRef]
7. Choe, K.H.; Lee, S.M.; Lee, K.W. High Temperature Oxidation Behavior of Si-Mo Ferritic Ductile Cast Iron.
Mater. Sci. Forum 2010, 654, 542–545. [CrossRef]
8. Stawarz, M.; Janerka, K.; Dojka, M. Selected Phenomena of the In-Mold Nodularization Process of Cast Iron
That Influence the Quality of Cast Machine Parts. Processes 2017, 5, 68. [CrossRef]
9. Henderieckx, G.D. Silicon Cast Iron; Gietech BV: Terneuzen, The Netherlands, 2009.
10. Brady, M.P.; Muralidharan, G.; Leonard, D.; Haynes, J.A.; Weldon, R.G.; England, R.D. Long-Term Oxidation
of Candidate Cast Iron and Stainless Steel Exhaust System Alloys from 650 to 800 ◦ C in Air with Water Vapor.
Oxid. Met. 2014, 82, 359–381. [CrossRef]
11. Roucka, J.; Abramova, E.; Kana, V. Properties of type SiMo ductile irons at high temperatures.
Arch. Metall. Mater. 2018, 63, 601–607.
12. Zeytin, H.K.; Kubilay, C.; Aydin, H.; Ebrinc, A.A.; Aydemir, B. Effect of microstructure on exhaust manifold
cracks produced from SiMo ductile iron. J. Iron Steel Res. Int. 2009, 16, 32–36. [CrossRef]
13. Mohammadi, F.; Alfantazi, A. Corrosion of ductile iron exhaust brake housing in heavy diesel engines.
Eng. Fail. Anal. 2013, 31, 248–254. [CrossRef]
14. Cygan, B.; Stawarz, M.; Jezierski, J. Heat treatment of the SiMo iron castings—Case study in the automotive
foundry. Arch. Foundry Eng. 2018, 18, 103–109.
15. Li, D. Mixed graphite cast iron for automotive exhaust component applications. China Foundry 2017, 14,
519–524. [CrossRef]
16. Unkić, F.; Glavaš, Z.; Terzić, K. The influence of elevated temperatures on microstructure of cast irons for
automotive engine thermo-mechanical loaded parts. Mater. Geoenvironment 2009, 56, 9–23.
17. Ibrahim, M.; Nofal, A.; Mourad, M.M. Microstructure and Hot Oxidation Resistance of SiMo Ductile Cast
Irons Containing Si-Mo-Al. Met. Mater. Trans. A 2016, 48, 1149–1157. [CrossRef]
18. Guzik, E.; Kopyciński, D.; Wierzchowski, D. Manufacturing of Ferritic Low-Silicon and Molybdenum Ductile
Cast Iron with the Innovative 2PE-9 Technique. Arch. Met. Mater. 2014, 59, 687–691. [CrossRef]
19. Vaško, A.; Belan, J.; Tillová, E. Static and Dynamic Mechanical Properties of Nodular Cast Irons.
Arch. Metall. Mater. 2019, 64, 185–190.
20. Nyashina, G.S.; Kuznetsov, G.V.; Strizhak, P. Energy efficiency and environmental aspects of the combustion
of coal-water slurries with and without petrochemicals. J. Clean. Prod. 2018, 172, 1730–1738. [CrossRef]
21. Bhadra, B.N.; Jhung, S.H. Oxidative desulfurization and denitrogenation of fuels using metal-organic
framework-based/-derived catalysts. Appl. Catal. B Environ. 2019, 259, 118021. [CrossRef]
383
Materials 2020, 13, 1745
22. Avvakumov, E.G.; Golovin, A.V.; Paukshtis, E.A.; Ivanov, V.P.; Kolomiichuk, V.N.; Kustova, G.N.; Burgina, E.B.;
Litvak, G.S.; Cherepanova, S.V.; Tsybulya, S.V.; et al. Golden Book of Phase Transitions, 1st ed.; Tomaszewski
P.E.: Wroclaw, Poland, 2002; pp. 1–123.
23. Ju, S.; Cai, T.; Lu, H.-S.; Gong, C.-D. Pressure-Induced Crystal Structure and Spin-State Transitions in
Magnetite (Fe3O4). J. Am. Chem. Soc. 2012, 134, 13780–13786. [CrossRef] [PubMed]
24. Post, B.; Schwartz, R.S.; Fankuchen, I. The crystal structure of sulfur dioxide. Acta Crystallogr. 1952, 5,
372–374. [CrossRef]
25. Kim, W.; Suh, C.-Y.; Cho, S.-W.; Roh, K.-M.; Kwon, H.; Song, K.; Shon, I.-J. A new method for the
identificationand quantification of magnetite-maghemite mixtureusing conventional X-ray diffraction
technique. Talanta 2012, 94, 348–352. [CrossRef] [PubMed]
26. Darezereshki, E.; Ranjbar, M.; Bakhtiari, F. One-step synthesis of maghemite (γ-Fe2O3) nano-particles by wet
chemical method. J. Alloy. Compd. 2010, 502, 257–260. [CrossRef]
27. Long, R.; Zhou, S.; Wiley, B.J.; Xiong, Y. Oxidative etching for controlled synthesis of metal nanocrystals:
Atomic addition and subtraction. Chem. Soc. Rev. 2014, 43, 6288–6310. [CrossRef]
28. Li, C.; Wei, Y.; Liivat, A.; Zhu, Y.; Zhu, J. Microwave-solvothermal synthesis of Fe3O4 magnetic nanoparticles.
Mater. Lett. 2013, 107, 23–26. [CrossRef]
29. Marciuš, M.; Ristić, M.; Ivanda, M.; Music, S. Formation of Iron Oxides by Surface Oxidation of Iron Plate.
Croat. Chem. Acta 2012, 85, 117–124. [CrossRef]
© 2020 by the authors. Licensee MDPI, Basel, Switzerland. This article is an open access
article distributed under the terms and conditions of the Creative Commons Attribution
(CC BY) license (http://creativecommons.org/licenses/by/4.0/).
384
MDPI
St. Alban-Anlage 66
4052 Basel
Switzerland
Tel. +41 61 683 77 34
Fax +41 61 302 89 18
www.mdpi.com