You are on page 1of 26

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/369141275

Alloy design for microstructural-tailored boron-modified ferritic stainless steel


to ensure corrosion and wear resistance

Article in Journal of Materials Research and Technology · March 2023


DOI: 10.1016/j.jmrt.2023.03.023

CITATIONS READS

0 35

7 authors, including:

David D. S. Silva Alexandre Nascimento


Universidade Federal de São Carlos McGill University
22 PUBLICATIONS 102 CITATIONS 13 PUBLICATIONS 158 CITATIONS

SEE PROFILE SEE PROFILE

Guilherme Yuuki Koga Guilherme Zepon


Universidade Federal de São Carlos Universidade Federal de São Carlos
57 PUBLICATIONS 652 CITATIONS 69 PUBLICATIONS 1,000 CITATIONS

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Spray forming of Advanced High Strength Steels View project

Production and performance of protective metallic coatings with high corrosion and wear resistances View project

All content following this page was uploaded by David D. S. Silva on 10 March 2023.

The user has requested enhancement of the downloaded file.


Journal Pre-proof

Alloy design for microstructural-tailored boron-modified ferritic stainless steel to


ensure corrosion and wear resistance

David D.S Silva, Alexandre R.C Nascimento, Guilherme Y Koga, Guilherme Zepon,
Claudio S Kiminami, Walter J Botta, Claudemiro Bolfarini

PII: S2238-7854(23)00484-2
DOI: https://doi.org/10.1016/j.jmrt.2023.03.023
Reference: JMRTEC 6767

To appear in: Journal of Materials Research and Technology

Received Date: 4 January 2023


Revised Date: 2 March 2023
Accepted Date: 4 March 2023

Please cite this article as: Silva DDS, Nascimento ARC, Koga GY, Zepon G, Kiminami CS, Botta WJ,
Bolfarini C, Alloy design for microstructural-tailored boron-modified ferritic stainless steel to ensure
corrosion and wear resistance, Journal of Materials Research and Technology, https://doi.org/10.1016/
j.jmrt.2023.03.023.

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

© 2023 The Author(s). Published by Elsevier B.V.


Alloy design for microstructural-tailored boron-modified ferritic stainless
steel to ensure corrosion and wear resistance

David D.S. Silvaa,*,ϯ, Alexandre R.C. Nascimentoa,ϯ, Guilherme Y. Kogab,c, Guilherme Zeponb, Claudio S.
Kiminamib, Walter J. Bottab, Claudemiro Bolfarinib,c

a
Federal University of São Carlos, Graduate Program in Materials Science and Engineering, Rodovia
Washington Luís, km 235, 13565-905, São Carlos, SP, Brazil
b
Federal University of São Carlos, Department of Materials Engineering, Rodovia Washington Luís, km
235, 13565-905, São Carlos, SP, Brazil
c
Federal University of São Carlos, Center for Characterization and Development of Materials (CCDM), Rod.
Washington Luís, São Carlos, SP, CEP 13565-905, Brazil

f
oo
Abstract

r
Ferritic stainless steels armored with borides had their endurance tested in demanding both abrasive
-p
and corrosive environment. The degradation was investigated through a very aggressive plate-on-
cylinder wear test in water-based drilling mud containing chlorides, which was designed to
re
simulate, in laboratorial scale, the corrosive tribosystem found in the wear of risers and casings used
in deep water oil exploitation. A commercially-grade 430 ferritic stainless steel containing boron
lP

content far beyond its solubility limity, namely 1.2 and 3.5 wt.%,, was processed by spray-forming
to induce a ferritic stainless steel matrix protected by a boride skeleton. Additional corrosion
na

assessment was performed by potentiodynamic polarization and electrochemical impedance


spectroscopy in 0.27 M KCl solution, since KCl is a common clay swelling inhibitor. For
ur

comparison, API 5L X80 steel was also included since this material is used in risers and casings
manufacture. The addition of boron (1.2 and 3.5 wt.%) led to the formation of microstructures
Jo

composed of an α-ferrite matrix with homogeneously distributed M2B-type borides of different


morphologies. Electrochemical tests revealed high corrosion resistance of the alloy with 1.2 wt.%
B. However, the alloy with 3.5 wt.% B showed a considerable decrease in corrosion resistance due
to Cr-depletion of the ferritic matrix given the M2B formation richer in Cr. Therefore, the alloy with
1.2 wt.% B offered the best tribocorrosion resistance.

Keywords: Stainless steel; Spray forming; CALPHAD method; Solidification; Tribocorrosion.

*
Corresponding author: david.domingos20@gmail.com (David D.S. Silva).
ϯ
These authors contributed equally to this work.

1
1. Introduction

Corrosion and wear are inherent threats in a wide array of industrial environments, resulting
in the premature degradation and eventual failure of engineering components [1]. Particularly in the
offshore sector, materials degradation could lead to costly and catastrophic failures with severe
consequences in terms of security and environmental risks during oil and gas activities [2]. The
equipment in this sector are exposed to demanding environments such as drilling components that
operates at high pressure in chloride-rich media, extreme condition responsible of lifespan reduction
caused by wear and corrosion [3]. Therefore, a strict balance of electrochemical and tribological
properties is demanded for a safe and longstanding petrochemical operation.

Stainless steel are tough, strong and corrosion resistant alloys that finds use in a variety of

f
oo
applications [4,5]. In the offshore sector, stainless steels are employed in several parts, such as
tubes, pipes, pumps, and others [6]. However, although largely used, conventional stainless steels

r
-p
may not meet the tribological requirements given the limited strengthening possibilities available to
harden the surface [7].
re
The ferritic stainless steel grade relies mostly on the Fe-Cr system containing between 11 and
lP

30 wt.% Cr, and carbon content below 0.20 wt.% [8]. They present good ductility and formability,
recognized resistance to stress corrosion cracking, and low Ni content that makes them
na

economically competitive [8,9]. Thus, the ferritic stainless steels meet most of the corrosion
ur

resistance requirements, and its hardenability is interesting from a tribological point of view.
However, the extent of hardening due to martensite transformation is limited, since the carbon
Jo

content allowed is low to avoid the risk of sensitization, and even the martensitic structure of
stainless steel is not able to guarantee enough wear resistance in harsh environments [4].

Different strategies for improving the wear resistance of ferrous alloys have been considered
[10,11]. An efficient method is based on the incorporation of hard particles in the microstructure,
such as carbides [12] or borides [13]. Among carbon and boron, the first has been largely reported
to cause high risk of sensitization of stainless steels. Boron, on the other hand, induces the
formation of borides only over the solidification due to its very limited solubility and mobility in
ferrite; thus, during solidification of the ferritic phase, the boron will segregate to the liquid, being
totally consumed to form borides from the liquid [13,14]. Indeed, spray formed boron-modified
stainless steels have been reported and characterized to display homogeneous distribution of the
reinforcing particles, which are of capital importance to the mechanical and tribological properties.
This advanced processing route allows homogeneous multiphase microstructures with equiaxed

2
grains, not achievable by most conventional casting methods [15,16]. Also, in optimal solidification
conditions, the remaining alloying elements, which render the alloy resistance to corrosion, are
homogeneously distributed within the matrix, ensuring the maintenance of the corrosion resistance.

Recent advances on boride-reinforced stainless steel aiming at improving the wear resistance
were reported for different grades: supermartensitic [3], duplex [17], and superduplex stainless steel
[18]. Research conducted by Lopes et al. [19] investigated the addition of small quantities of boron
in the phase formation and microstructure of spray-formed AISI 430 ferritic stainless steel. The
authors concluded that it led to the formation of M2B type borides (where M = Fe, Cr). However, no
data regarding wear resistance has been reported for the boron-modified ferritic stainless steels. On
the other hand, the corrosion behavior of such modified steels is of concern as it may be affected by

f
the alloy modification. In particular, borides, which consume Cr of the matrix (as in the case of

oo
M2B) are the main reason for lowering the corrosion properties of boron-modified stainless steels

r
[20,21]. Thus, developing strategies to improve the wear resistance of stainless steels without
-p
impairing their corrosion resistance is an important scientific and technological challenge, as a
re
trade-off may exist between electrochemical and tribological properties of boron-modified steels
[22]. Despite the large progress on the boriding of stainless steels, few studies have challenged in
lP

extreme conditions where both corrosion and wear operate in synergy.


na

In this work, ferritic stainless steels armored with borides endured harsh tribocorrosion testing
as proved for the first time. Two compositions were selected and fabricated by spray forming (AISI
ur

430 ferritic stainless steel with 1.2 and 3.5 wt.% B). The microstructure and phase transformation
Jo

were investigated using thermodynamic calculations of the equilibrium phase fractions by


CALPHAD method, x-ray diffraction (XRD), and scanning electron microscopy (SEM). A plate-
on-cylinder (POC) wear test was performed in water-based drilling mud containing chlorides to
replicate the operational conditions found in-service. This test simulates, in laboratorial scale, the
corrosive tribosystem found when drilling risers and casings are worn by contact of the rotating drill
pipe with corrosive drilling mud. Since API 5L X80 steel is used in risers and casings manufacture,
this alloy was also tested as comparison. Electrochemical tests, such as electrochemical impedance
spectroscopy (EIS) and potentiodynamic polarization, were performed on the alloys in 0.27 M KCl
solution. This solution was selected based on the drilling mud, which contains KCl as a clay
swelling inhibitor. Among the alloys tested herein, the alloy with 1.2 wt.% B offered the best
combination of corrosion and wear resistance.

3
2. Materials and Methods

2.1 Thermodynamic calculations

Thermodynamic calculations of the equilibrium phase fractions between 500 °C and 1750 °C
were evaluated by the CALPHAD method, Thermo-Calc® software [23], using the TCFE7
database. All calculations were performed at pressure P = 1 atm and with the whole system
containing one mole of atoms, and the phase volume fractions plotted as a function of temperature.

2.2 Spray forming

A commercial AISI 430 FSS supplied by Villares Metals S.A., Brazil, was taken as the master
alloy, and industrially-grade Fe-B as boron source. Based on the thermodynamic calculations, two

f
oo
different alloys containing 1.2 and 3.5 wt.% B were produced, designated as F-1.2B and F-3.5B,
respectively. The F-1.2B alloy is a hypoeutectic composition while the F-3.5B alloy is a

r
-p
hypereutectic one, which dictates the sequence of solidification and, therefore, how the borides are
formed [19]. Even with boron content far above the solubility limit of stainless steel,
re
macrosegregation free homogeneous microstructure with uniform distribution of second phases is
lP

possible by spray forming, given its characteristic solidification sequency, recently demonstrated
for boron-modified stainless steels [24]. To ensure Cr contents within standard specifications of the
na

AISI 430 Steel, electrolytic Fe and commercially pure Cr were introduced in the molten in order to
compensate its reduction due to the Fe-B addition. The content of Ni was not corrected as it is not
ur

important for this alloy class. The final composition was obtained in an induction furnace using an
Jo

alumina crucible. The molten metal was spray formed using a close-coupled atomizer with nitrogen
gas. First, the molten at 1650 °C was atomized with a 0.5 MPa nitrogen pressure, being the droplets
deposited onto a 20-rpm rotating circular steel substrate located 370 mm below the atomization
nozzle, resulting in disk-shaped deposit of about 3.5 kg, with 250 mm in diameter and an average
thickness of 13 mm. Additional information regarding the spray forming apparatus may be found in
Ref. [24,25].

2.3 Structural, chemical and morphological characterization of the alloys

X-ray diffraction (XRD), using a Rigaku Geigerflex ME210GF2 model diffractometer


operating with Cu radiation, in the range of 30° – 90°, with 0.02° step size, was performed for phase
identification. Chemical composition of the alloys was checked by ICP-AES (Inductive-Coupled
Plasma Atomic Emission Spectroscopy) and AAS (Atomic Absorption Spectroscopy) only for the
boron content. Microstructural characterization was carried out in a Philips XL30 FEG scanning

4
electron microscope (SEM) equipped with an Oxford Link EDS detector and an optical microscope
Olympus BX41 M-LED. ImageJ® software was used for phase fraction quantification. For
microstructural analysis, the specimens were prepared with SiC papers up to a grit of 1500 mesh,
and mechanically polished using diamond pastes (6 - 0.25 μm), using Behara for etching according
to ASTM E407 [26]. To better reveal the borides morphologies, the α-ferrite matrix was removed
using a deep etching procedure with a solution composed of 82 ml ethyl alcohol, 10 ml
hydrochloric acid (HCl), 3 ml nitric acid (HNO3), and 5 ml iron chloride (FeCl3).

2.4 Electrochemical measurements

Three-electrode cell set-up was considered in the electrochemical testing, using a saturated
calomel electrode (SCE) and a Pt grid counter electrode, respectively, in a Gamry 600+ potentiostat.

f
oo
The working electrodes were alloys sanded on abrasive papers with grain sizes ranging from 220 to
1200 mesh, with an exposed area of 0.4 cm2. API 5L X80 specimens were also tested for

r
comparison purposes. -p
re
The electrolyte was a 0.27 M KCl solution prepared using demineralized water and high
purity KCl (>99.5%), which represent the corrosiveness of bentonite-based fluids used for drilling
lP

in oil and gas industries, as will be detailed in Section 2.5. All electrochemical experiments were
conducted in the open air at room temperature (≈ 25 °C). Before launching the electrochemical
na

measurements, the working electrodes were immersed in the test solution for 60 min, and open
ur

circuit-potential (OCP) was recorded. This time was sufficient to stabilize the potential.
Jo

Electrochemical impedance spectroscopy (EIS) around the OCP was considered to


electrochemically characterize the specimen | electrolyte interface, with ∆𝐸 = 10 mVrms, recording
10 points per decade from 105 Hz up to 10−2 Hz. The experimental impedance data (Z(ω)exp) was
assessed using an electrical equivalent circuit (EEC), which is presented with the accompanying
data, with EIS analysis aided using EC-Lab software (BioLogic®). Potentiodynamic polarization
stated at -200 mV regarding OCP in the anodic direction at a scan rate of 1 mV s−1, interrupted at 10
mA cm−2. Repeatability of the results reported herein was checked from measurements in replicate.
The corrosion potential (𝐸𝑐𝑜𝑟𝑟 ), corrosion current density (𝑖𝑐𝑜𝑟𝑟 ), corrosion current density value at
the passivation plateau (𝑖𝑝𝑎𝑠𝑠 ) and breakdown potential (𝐸𝑏 ) were extracted from the polarization
curves.

5
2.5 Mechanical and tribological measurements

Vickers hardness measurements were performed according to ASTM E384 [27], with 1000 g
load and 10 s dwell time. The average of 5 measurements was reported as the mean value within the
± 2σ error.

A dedicated wear testing machine was especially developed to perform the plate-on-cylinder
(POC) tribological tests in corrosive environment. It consisted of three separate chambers,
containing in each one a 25 x 90 x 10 mm3 machined specimen pressed against a rotating AISI 1040
steel axis (quenched and tempered, 55 HRC). The surface of specimens and axis were prepared to
average roughness of Ra = 1.7 µm. A normal force (𝐹𝑁 ) of 575.7 N was applied to the specimens by
a lever arm system. However, during the wear tests, this initial normal force is slightly increased

f
oo
due to changes in the relative position of the mass center of the arms due to the thickness loss of the
specimen. This increase may be described as a function of the loss thickness (ℎ) by the following

r
equation (𝐹𝑁 = 0.093ℎ2 + 7.495ℎ + 575.7), where the units of 𝐹𝑁 and ℎ are N and mm,
-p
respectively. In all tests, the rotation speed used was 200 rpm and 25500 m of sliding distance.
re
Mass losses were measured using a precision balance (sensitivity of 1 x 10-4 g), and the specimens
lP

were weighed before and after the tests. Further details on the wear test machine may be found in
Ref. [3].
na

The specific wear rate (𝑘), in mm3 N-1 m-1, after test was calculated according to equation (1).
ur

𝑘 = 𝑉/(𝐹𝑁 𝐿) (1)
Jo

where 𝑉 is the worn volume (mm3), 𝐹𝑁 is the normal force (N) and 𝐿 is the total sliding
distance (m). Note that the increments in normal force, as previously mentioned, were considered to
calculate the specific wear rate.

The wear tests were performed in the boron-modified stainless steels (F-1.2B and F-3.5B)
produced by spray forming. API 5L X80 steel was also tested as a comparison basis.

To simulate the wear conditions found in risers and casing, although not at the same extreme
conditions of pressure and temperature, the chambers were filled with real drilling mud, whose
composition is detailed in Table 1. Replicate measurements for each specimen were performed to
ensure repeatability. A new 1040 steel axis was used in each test. The worn surfaces of the
specimens were characterized by SEM.

6
Table 1. Composition of the drilling mud used in the POC wear tests.
Component Function Concentration (kg/m3)
Bentonite Solids suspension 25
Sodium carboxymethyl cellulose Viscosifiers 3
Potassium chloride Inhibitor of expansive clay 20

3. Results and Discussion

3.1 Thermodynamic calculations and microstructural characterization

Fig. 1 shows the equilibrium volume fraction of each phase versus temperature from
thermodynamic calculations for F-1.2B and F-3.5B alloys by the CALPHAD method. The primary

f
phase to be formed is sensitive to the B content. Addition of 1.2 wt.% B results in the formation of

oo
δ-ferrite as primary phase, while the formation of M2B takes place in the final stage of the
solidification process (Fig. 1 (a)). On the other hand, 3.5 wt.% B forms M2B as primary phase (Fig.

r
-p
1 (b)), highlighting that the F-1.2B and F-3.5B alloys correspond to a hypoeutectic and
hypereutectic composition, respectively.
re
A stable region of γ-austenite exists up to 880 °C for the F-1.2B alloy, and 850 °C for the F-
lP

3.5B alloy, which decomposes into α-ferrite and M23C6/M7C3 (formed at lower temperatures
na

resulted from solid-state reactions), that remained stable at room temperature. Thermodynamic
calculations indicate that most of the B added to the alloys is consumed by forming M2B upon
ur

solidification [19,21], being the volume fractions of the M2B calculated at 500 °C for F-1.2B and F-
Jo

3.5B alloys 18% and 47%, respectively. This quite distinct volume fraction may be attributed to the
small atomic mass of boron. Once formed from the liquid, the volume fraction of the M2B did not
change significantly during cooling through the solid state region since boron has very low
solubility in the α-ferrite matrix [28]. The concentration of B in solid solution in α-ferrite matrix for
the F-1.2B and F-3.5B alloys was 5.7 x 10-8 and 8.1 x 10-7 wt.%, respectively.

7
(a) 100 (b) 100
L Liquid (L) L
Austenite (g)
g L Liquid (L) Ferrite (d, a)
a a a Boride (M2B) L
80 Austenite (g) 80
g d Ferrite (d, a) Carbide (M7C3)
Volume fraction (%)

Volume fraction (%)


Boride (M2B) L
Carbide (M23C6)
g d
60 60 g g g
a a a
M2B M2B M2B M2B M2B

40 40
M2B
a d g

M2B M2B M2B M2B M2B g a M2B


20 20
d d M2B
g a d g a
M23C6 M7C3 d
0 0
500 750 1000 1250 1500 1750 500 750 1000 1250 1500 1750
Temperature (°C) Temperature (°C)
Fig. 1. Equilibrium volume fraction of phases calculated by the CALPHAD method at different temperatures for the (a)
F-1.2B and (b) F-3.5B alloys, whose chemical compositions are presented in Table 2.

f
oo
Table 2 shows the chemical composition of the boron-modified stainless steels (F-1.2B and F-
3.5B). The chemical composition of the precursor materials (AISI 430 stainless steel and Fe-B), as

r
well as API 5L X80 steel, is also presented in Table 2. The results indicate that the F-1.2B and F-
3.5B alloys presented nominal composition close to precursor material (AISI 430). This was made
-p
re
possible by the addition of Cr, which compensated for the compositional changes induced by Fe-B
lP

addition. It can also be noticed an increase in the silicon content in the boron-modified stainless
steels, which is a typical impurity found in the commercially graded precursors such as Fe-B and
na

AISI 430 ferritic stainless steel. The Ni-content was reduced in comparison with the AISI 430 as
already mentioned no addition of Ni was provided during melting.
ur
Jo

Table 2. Chemical composition, wt.%, of the commercial precursors (AISI 430 stainless steel and Fe-B) and of the
resulting boron-modified stainless steels. The composition of the API 5L X80 steel considered as a benchmark material
is also given.
Alloy C Si Mn Cr Ni S P Nb Co N B Fe
AISI 430 0.06 0.20 0.74 17.62 0.37 0.17 0.02 0.03 0.03 0.03 --- Bal.
Fe-B 0.30 0.57 --- --- --- --- --- --- --- --- 16.50 Bal.
F-1.2B 0.06 0.29 --- 17.20 0.19 --- --- --- --- --- 1.20 Bal.
F-3.5B 0.05 0.53 --- 18.50 0.23 --- --- --- --- --- 3.50 Bal.
*
API 5L X80 0.07 0.21 1.64 0.10 0.25 --- 0.01 0.04 --- --- --- Bal.
*
Additional elements: Al = 0.02, Ti = 0.01, Mo = 0.13, Cu = 0.02.
Note: Bal. is the Fe composition to get 100%.

X-ray diffraction patterns for the boron-modified stainless steels are shown in Fig. 2. Only
characteristic peaks of α-ferrite and M2B phases were identified. The peaks ascribed to M2B are
more pronounced in the case of F-3.5B given the larger volume fraction. However, no other phases

8
were noted, such as M23C6/M7C3 (M = Fe, Cr), even if predicted in the thermodynamic calculations,
Fig. 1. These results are in accordance with those previously reported [19], and interesting from the
corrosion resistance standpoint to avoid a high degree of sensitization nor risk of intergranular
corrosion. Despite not considered as a rapid solidification process, spray forming is effective to
suppress the formation of some stable phases, and may even induce metastable phase formation.

· ¨ M 2B

· a-ferrite
Intensity (a. u.)

¨ ·
¨ ¨ ¨ ¨ ·

f
F-3.5B

r oo
F-1.2B -p
re
30 40 50 60 70 80 90
2q (degree)
lP

Fig. 2. X-ray diffraction patterns of boron-modified stainless steels (F-1.2B and F-3.5B) indicating strong peaks of α-
ferrite and borides (M2B).
na

Fig. 3 shows SEM (BSE and SE) images of the microstructures aimed to unveil the M2B
ur

morphology of boron-modified stainless steels. Fig. 3 (a, b) correspond to F-1.2B alloy, while Fig. 3
(c, d) to F-3.5B alloy. Both microstructures are composed of a α-ferrite matrix (lighter gray) with
Jo

borides (darker gray) homogeneously distributed, see Fig. 3 (a, c). However, the different
morphologies of the M2B are worth emphasizing. For the hypoeutectic F-1.2B alloy (Fig. 3 (a)), the
borides present an eutectic microstructure typical of the last solidified phase to be formed from the
liquid. These borides form an interconnected and rigid armor within the microstructure, being
resistant also against the severe corrosive attack given their survival even after deep etching for the
stainless steel matrix dissolution, Fig. 3b. In contrast, for the hypereutectic F-3.5B alloy (Fig. 3 (c)),
the primary boride is evident due to its faceted and elongated morphology. These borides nucleate
from the liquid upon cooling as a primary phase, and share common physicochemical properties
with those formed in hypoeutetic alloys. Similar borides morphologies were also reported by our
research group in previously published studies [3,18,19], and displayed interesting electrochemical
and tribological properties as assessed separately from conventional testing procedures for a
comparative ranking of performance, which motivated this study to explore them further in
condition more representative of their potential use in tribocorrosion environment. It is also

9
important to point out the low porosity levels (< 1%) observed in the spray-formed boron-modified
steels. In addition, one could reasonably and qualitatively assess that for the case of F-3.5B alloy,
given their amount as well as faceted and elongated boride morphology, the pores are more
concentrated around the borides, see Fig. 3 (c).

To clarify the morphological differences of borides, a more detailed characterization, using a


deep etching procedure was carried out. In this type of etching, the α-ferrite matrix is removed,
leaving the boride structure in evidence, as shown in Fig. 3 (b, d). The M2B in F-1.2B alloy (Fig. 3
(b)) forms an interconnected eutectic continuous three-dimensional network surrounding the α-
ferrite matrix (hollow regions), confirming that they were formed eutectically during solidification.
Fig. 3 (d) shows the deep etching of the F-3.5B alloy highlighting the faceted primary M2B

f
morphology. According to the thermodynamic calculations, these borides should nucleate from the

oo
liquid upon cooling as a primary phase, as previously mentioned. This explains the difference in

r
morphology when compared to eutectic formation.
-p
The M2B fractions measured through metallographic examination for F-1.2B and F-3.5B
re
alloys were 19 and 45%, respectively. With the measurement methodology applied, the relative
lP

error of the area fraction evaluation of phases did not exceed 3%. These results agree with the
values of 18 and 47% predicted by the thermodynamic calculations.
na

(a) Eutectic M2B (b)


ur
Jo

α-Fe

Pores

BSE 20 μm SE 30 μm

(c) (d)

Pores

α-Fe

Primary M2B

BSE 30 μm SE 30 μm

Fig. 3. Microstructures and borides morphology of boron-modified stainless steels. (a, b) F-1.2B and (c, d) F-3.5B
alloys. BSE and SE stand for backscattered and secondary electron micrographs, respectively.

10
The chemical composition of each phase measured by EDS analysis is presented in Table 3.
The M2B is composed mainly of Fe and Cr, with smaller amounts of Si. However, it is worth
pointing out that EDS does not measure the minor elements nor the light elements such as N, B, and
C. The Cr content of the α-ferrite matrix is about 14.6 wt.% for the F-1.2B alloy, sufficient to be
labelled as stainless, and sufficient to grant a good corrosion resistance for this alloy. Only the F-
3.5B alloy presented the Cr content of the α-ferrite matrix (9.3 wt.%) lower than the FSS grades
(11-30 wt.%) [8], indicating a higher risk to corrode in chloride containing environement. EDS
results also showed that, for both alloys, the M2B presents higher Cr content than the matrix, which
may be deleterious from the corrosion resistance standpoint [29]. This same deleterious effect is
seen on the duplex stainless steel grades when the sigma phase enriched in Cr precipitates in the α-
ferrite matrix causing a decrease in the corrosion resistance [30,31] and a consequent drastic

f
oo
reduction in toughness and ductility of the material [32,33].

r
Table 3. EDS microanalyses of the α-ferrite matrix and the M2B of the spray formed boron-modified steels.
Alloy Phase
-p
Fe (wt.%) Cr (wt.%) Si (wt.%) Ni (wt.%)
re
α-ferrite matrix 84.7 14.6 0.4 0.3
F-1.2B
M2B 64.3 35.6 0.1 ---
lP

α-ferrite matrix 87.1 9.3 3.2 0.4


F-3.5B
M2B 61.6 37.9 0.5 ---
na

Note: Light elements such as N, B, and C are not considered.


ur
Jo

3.2 Electrochemical characterization

Fig. 4 presents the electrochemical responses of the boron-modified stainless steels and API
5L X80 steel in 0.27 M KCl solution from EIS and potentiodynamic polarization measurements.
Representative results from the repetitive electrochemical impedance spectroscopy tests are
presented in Fig. 4 (a, b). The Nyquist plot (Fig. 4 (a)) of F-1.2B alloy was characterized by a
truncated and enlarged semi-circle, typically observed for passive stainless steels. Note the much
superior impedance for any frequency of F-1.2B in Fig. 4 (a) compared to the F-3.5B and the API
5L X80 alloy, these latter are only observable in the zoomed inset. In general, enlarged and
unrelaxed impedance arc indicates high polarization resistance of the electrolyte-electrode system,
indication of superior resistance against faradaic processes at rest [34]. Note that the F-3.5B and the
API 5L X80 was almost completely relaxed along the frequencies probed given the relaxation
frequency () dictated by resistive/capacitive behavior be within the 105 to 10-2 Hz range.

11
Bode plots (Fig. 4 (b)) help to distinguish the different relaxation phenomena. Bode plot
indicates a higher impedance modulus at the lowest frequency (|𝑍|𝑓𝑟𝑒𝑞 = 10-2 Hz) for F-1.2B,
indicating greater corrosion resistance of this alloy. Furthermore, the F-1.2B alloy presents a single
but enlarged relaxation process in an extended frequency range, a behavior seen in passive alloys
[35]. F-3.5B and API 5L X80 alloys, in contrast, point out more clearly the existence of a single
time constant in low-frequencies. The impedance value (|𝑍|) (Fig. 4 (b)), at the frequency of 10-2
Hz, decreases from 107.9 to 7.9 kΩ cm2 when comparing the F-1.2B alloy with the F-3.5B, which
corresponds to a 92.7% reduction. Indeed, while the |𝑍| value at 10-2 Hz of the F-3.5B alloy
approaches the polarization resistance, the |𝑍| value of the F-1.2B alloy for the same frequency is
still far below to its polarization resistance and would increase in a faster pace compared to the F-
3.5B alloy as the probed frequency decreases below 10-2 Hz. The API 5L X80 alloy presented the

f
oo
lowest value, about 7.3 kΩ cm2.

r
A modified Randle’s circuit, i.e., a resistor in series with a non-ideal capacitor (a constant
-p
phase element, 𝐶𝑃𝐸) in parallel to a resistor, (𝑅𝑒 + 𝐶𝑃𝐸/𝑅𝑝 ) (Fig. 4 (c)), was selected to represent
re
the overall electrochemical processes occurring at the corroding interface. This circuit has been
reported as suitable to evaluate the experimental impedance data of stainless steel alloys in chloride-
lP

containing solutions [36–39]. The physical meaning of each analog was: 𝑅𝑒 the electrolyte
resistance and the 𝑅𝑝 the polarization resistance. Since the experimental capacitive-like response
na

often deviates from a pure capacitor, a 𝐶𝑃𝐸 was used instead, being characterized by its parameter
ur

𝑄 and exponent 𝛼. Note that the 𝐶𝑃𝐸 is considered to reflect the capacitance-like response from the
whole interface, including passive film, double-layer relaxation, faradaic reactions, etc.
Jo

Table 4 summarizes the results from fitting the impedance of the chosen equivalent circuit to
experimental EIS data, using the EC-Lab software (BioLogic®). In all cases, given the higher
concentration of small and mobile ions, K+ and Cl-, the 𝑅𝑒 values (53-57 Ω cm2) are considerably
low for testing the alloys in an aqueous medium. The F-1.2B exhibited the highest polarization
resistance, 𝑅𝑝 , amongst the studied alloys, with values ≈ 319 kΩ cm2. The greater the 𝑅𝑝 , the
greater the resistance of the material to corrosion [40], given the low related corrosion current
density as stablished by the Stern-Geary equation (icorr = B/Rp) [41]. The exponent 𝛼 (−1 ≤ 𝛼 ≤ 1)
characterizes the deviation of 𝑄 from an ideal capacitance, where -1 is characteristic for an
inductance, 1 corresponds to a capacitor and 0 corresponds to a resistor [34]. The 𝐶𝑃𝐸 of the F-
1.2B displayed capacitance-like behavior, since 𝛼 ~ 0.90. Moreover, the effective capacitance, 𝐶𝑒𝑓𝑓 ,
was relatively low, 24.5 μF cm-2, typical of passive alloys. Increasing the B content from 1.2 to 3.5
wt.% resulted in a decrease of the 𝑅𝑝 to around 9 kΩ cm2. The 𝐶𝑒𝑓𝑓 and the 𝛼 values of the F-3.5B
12
were 33 μF cm-2 and 0.70, respectively. The API 5L X80 exhibited the lowest 𝑅𝑝 value (7 kΩ cm2)
as expected, and the highest 𝐶𝑒𝑓𝑓 value of 53.3 μF cm-2. The increase of the 𝐶𝑒𝑓𝑓 value may be
related to the worse stability of the passive film on the electrode surface [42]. 𝐶𝑒𝑓𝑓 value accounts
the contributions of the passive layer and double-layer capacitances. If the capacitance value of the
passive film is small compared to the double-layer one, which is often the case of highly passivated
alloys, 𝐶𝑒𝑓𝑓 value is dominated by the capacitance of the passive film [43].

The thickness of the passive film on the alloys is inversely proportional to 𝐶𝑒𝑓𝑓 . Thus,
superior 𝑅𝑝 value and inferior 𝐶𝑒𝑓𝑓 value indicate more protective and thick products on the F-1.2B
alloy surface than on F-3.5B and API 5L X80. In particular, borides, which consume Cr in the
matrix, are the main reason for the poor corrosion properties of most B-containing stainless steels

f
oo
[20]. In addition, as previously mentioned, the F-3.5B alloy presented the Cr content in the α-ferrite
matrix lower than the FSS grades, which further contributes to the low corrosion resistance of this

r
alloy. The reproducibility of empirically determined values of 𝐶𝑒𝑓𝑓 validate that the EEC analog
-p
was effective in representing the electrochemical response of the studied alloys and, mainly,
re
revealing their differences.
lP

The potentiodynamic polarization measurements provided further comparison basis on the


corrosion differences between the alloys under investigation. Therefore, potentiodynamic
na

polarization measurements were performed to complement the results from the EIS measurements.
ur

Fig. 4 (d) shows the representative polarization curves of the boron-modified stainless steels and
API 5L X80 steel. The resulting electrochemical parameters are presented in Table 5.
Jo

Polarization testing reveals that the F-1.2B presented better corrosion properties than the other
alloys, presenting 𝐸𝑐𝑜𝑟𝑟 and 𝑖𝑐𝑜𝑟𝑟 values of -132 mVSCE and 0.01 µA cm-2, respectively. The
formation of a clear passive plateau was seen since the beginning of anodic polarization, with
associated current density, 𝑖𝑝𝑎𝑠𝑠 , at low values along an extended potential range, with disruption
caused by the formation of pits beyond the breakdown potential, 𝐸𝑏 . The 𝑖𝑝𝑎𝑠𝑠 and 𝐸𝑏 values for
this alloy are 0.06 µA cm-2 and 254 mVSCE, respectively. This behavior was expected since defects
like pores, even in reduced content, represent preferential region for pitting corrosion initiation due
to the penetration of the electrolyte through the formed channels, without neglecting the deleterious
effect due to the M2B in this alloy presenting higher Cr content than the α-ferrite matrix, as shown
in Table 3. Increasing the B content from 1.2 to 3.5 wt.% resulted in a decrease of the 𝐸𝑐𝑜𝑟𝑟 to
around -616 mVSCE, and an increase of the 𝑖𝑐𝑜𝑟𝑟 to around 0.43 µA cm-2, not showing the presence
of a passive plateau; active dissolution occurrence readily upon anodic polarization. This result may

13
be attributed to the low Cr content present in the a-ferrite matrix of F-3.5B alloy, only 9.3 wt.%,
lower than the FSS grades, as well as the deleterious effect caused by high Cr content in the M2B as
previously mentioned. A clear difference in corrosion resistance was observed for the API 5L X80
alloy. For benchmarking, it is noted that the F-1.2B alloy herein displayed greater corrosion
resistance to that of the F-3.5B and API 5L X80 alloys, highlighting the potentially promising
performance of this alloy in KCl media, found in drilling mud, see Table 1.

The results presented here suggest that the F-1.2B alloy can form a more protective
passivation film (albeit in a comparative sense) among the tested alloys (F-3.5B and API 5L X80).
As reported in previous studies for boron-modified stainless steels [3,20,29,44], the addition of
boron does not impair its corrosion resistance if the composition of the alloy is adjusted in such a

f
way that the final composition of the matrix is similar to that of the non-modified alloy.

oo
(a) 8x104 F-1.2B (b)

r
F-3.5B
105
0

6x104
API 5L X80
Fitting 0.01 Hz
-p -20
Single time constant
re
104
1x104
F-1.2B

|Z| (W cm2)
-Z" (W cm2)

Phase (°)

0.25 Hz F-3.5B
API 5L X80
lP

-40
-Z" (W cm2)

4
4x10 Fitting
5x10 3
0.01 Hz 103

-60
na

2x104 0.01 Hz
0
0 5x103 1x104 102
Z' (W cm2) -80
Single enlarged time constant
ur

0
0 2x104 4x104 6x104 8x104 10-2 10-1 100 101 102 103 104 105
2 Frequency (Hz)
Z' (W cm )
Jo

(c) (d) 10-2


F-1.2B
F-3.5B
CPE API 5L X80
10-4
Current density (A cm )

(Q, α)
-2

icorr
10-6
Re icorr Passivation
plateau

10-8 icorr

Eb
10-10

Rp Ecorr Ecorr Ecorr


10-12
-1.5 -1.0 -0.5 0.0 0.5
Potential (VSCE)
Fig. 4. (a, b) Nyquist and Bode plots, respectively of the F-1.2B and F-3.5B alloys in 0.27 M KCl solution. Symbols
represent the experimental data, and the solid cyan color lines fitting considering the electrical circuit analog approach.
The inserted graph in (a) is the Nyquist plot’s zoomed region to visualize the results better. (c) Modified Randle’s
equivalent circuit selected to represent the corrosion processes occurring at the alloys | electrolyte interface, and (d)
characteristic potentiodynamic polarization curves of the alloys. Results from commercially available API 5L X80 steel
are included in (a, b, d) for comparison. Passivation plateau is the region upon the anodic polarization where the current
14
density remains almost unchanged. 𝐸𝑏 refers to the potential from which there is a steep increase in current density
upon anodic polarization. 𝑖𝑐𝑜𝑟𝑟 was determined around 𝐸𝑐𝑜𝑟𝑟 from the Tafel extrapolation of the cathodic branch, which
exhibit ‘Tafel-like’ behavior in at least one decade of current from potentials relatively far from 𝐸𝑐𝑜𝑟𝑟 . Dashed lines are
intended as an eye guide only. (For interpretation of the references to color in this figure legend, the reader is referred to
the Web version of this article.).

Table 4. Ensemble of fitting results from electrochemical impedance spectroscopy associated with the equivalent
electrical circuits' elements for representing the electrochemical processes occurring on the alloy | electrolyte interface
determined in a solution of 0.27 M KCl. Results from the fitting approach of the equivalent electrical circuit to the
experimental data. Values presented as mean values and standard deviations from testing in replicate. Satisfactory
convergence level between the fitted results and the experimental data, given (𝜒/|𝑍|) < 10-2.
Alloy 𝑅𝑒 (Ω cm2) 𝐶𝑃𝐸 parameters 𝑅𝑝 (kΩ cm2) 𝐶𝑒𝑓𝑓 (µF cm-2)* 𝜒/|𝑍|
αf-1
𝑄 (µF S cm-2) 𝛼
F-1.2B 57 ± 0.5 52 ± 33 0.90 ± 0.02 319 ± 194 24.5 0.06 ± 0.02
F-3.5B 53 ± 4.6 292 ± 13 0.70 ± 0.01 9 ± 0.4 33 0.04 ± 0.005
API 5L X80 55 ± 0.5 220 ± 6 0.79 ± 0.01 7±1 53.3 0.07 ± 0.01

f
*
Effective capacitance, 𝐶𝑒𝑓𝑓 , calculated from the 𝑅𝑒 , 𝑅𝑝 , 𝑄 and 𝛼 values considering the Brug’s equation, 𝐶𝑒𝑓𝑓 = 𝑄1/𝛼 (𝑅𝑒−1 +

oo
𝑅𝑝−1 )(𝛼−1)/𝛼 [45].

r
-p
Table 5. Electrochemical parameters determined from potentiodynamic polarization in a solution of 0.27 M KCl.
Values presented as mean values and standard deviations from testing in replicate.
re
Sample 𝐸𝑐𝑜𝑟𝑟 (mVSCE) 𝑖𝑐𝑜𝑟𝑟 (µA cm-2) 𝑖𝑝𝑎𝑠𝑠 (µA cm-2) 𝐸𝑏 (mVSCE)
F-1.2B* -132 ± 4 0.01 ± 0.01 0.06 ± 0.05 254 ± 73
lP

F-3.5B -616 ± 2 0.43 ± 0.03 --- ---


API 5L X80 -810 ± 3 0.85 ± 0.04 --- ---
*
𝑖𝑝𝑎𝑠𝑠 refers to the current density value at the passivation plateau, i.e., region upon the anodic polarization where the current density
na

remains almost unchanged, and 𝐸𝑏 denotes breakdown potential. In this case, the potential beyond which there is a steep and
continuous increase of current density.
ur

3.3 Mechanical and tribological characterization


Jo

The wear resistance of the materials depends on several factors such as the hardness of main
microstructural constituents, strength and toughness of the matrix, size/distribution of borides, and
also, the conditions of the tribological system and the wear mechanism involved. The calculated
values of the specific wear rate (𝑘) of the boron-modified stainless steels (F-1.2B and F-3.5B) and
API 5L X80 steel after POC wear test and their respectively measured hardness are shown in Fig. 5.

The hardness increased significantly with the addition of boron from about 470 HV1000g in the
case of F-1.2B to 680 HV1000g for F-3.5B, which corresponds to a 45% increase. Among the tested
alloys, the API 5L X80 steel displayed the lowest hardness values, around 215 HV1000g. The wear
resistance increased with the boron content, as well. The specific wear rate after 25500 m of sliding
distance decreases from 1 x 10-6 to 4.6 x 10-7 mm3 N-1 m-1 when comparing the F-1.2B alloy with
the F-3.5B, which corresponds to a 54% reduction. The worn volumes measured after POC wear
tests for F-1.2B and F-3.5B alloys were 15 and 7 mm3, respectively, lower than the API 5L X80

15
alloy (201 mm3), which presented lower hardness. Therefore, as expected, a clear correspondence
among hardness, M2B content (19%, F-1.2B versus 45%, F-3.5B), worn volume, and specific wear
rate was found. This result may be attributed to a microstructure composed of a ductile α-ferrite
matrix that anchors the refined and rigid M2B improving the wear resistance of the boron-modified
stainless steels. Thus, the refinement of the boride within the matrix is a key aspect to be considered
to design wear-resistant boron-modified stainless steels [46]. Similar behavior was previously
reported [17,46].

The performance of the boron-modified steels was even similar to the CoCrWC alloys, which
are widely known wear-resistant materials used as a reference for abrasive resistance [47]. Among
the Fe-based alloys suitable for applications that require high wear resistance, white cast iron with

f
high chromium is widely employed [48]. However, it presents low toughness and, in some cases,

oo
low corrosion resistance. In contrast, although stainless steels have low wear resistance, they are

r
often required in applications in environments subject to wear due to their corrosion resistance [3].
-p
Therefore, microstructural modifications, such as the B addition adjusted-tailored in stainless steels
re
could lead to improved wear properties.
lP

2.0x10-5
Specific wear rate B  HV and K
Specific wear rate, k (mm3 N-1 m-1)

na

800
1.6x10-5 700
HV1000g 680
1.4x10-5
Hardness (HV1000g)

600
ur

500 470
1.2x10-5
400
Jo

300
215
-6 200
8.0x10
100
0
API 5L X80 F-1.2B F-3.5B
4.0x10-6 Specimen

1.0x10-6 4.6x10-7
0.0
API 5L X80 F-1.2B F-3.5B
Specimen
Fig. 5. Specific wear rate, 𝑘, of boron-modified stainless steels (F-1.2B and F-3.5B), calculated considering equation
(1) after POC wear test using a normal force of 575.7 N, a total sliding distance of 25500 m and a rotation speed of 200
rpm. Values presented as mean values and standard deviations from testing in replicate. The inserted graph is the
Vickers hardness, presented as mean values and their deviations from five measurements using 1000 g of load (HV 1000g)
in F-1.2B and F-3.5B alloys. Results for API 5L X80 steel are also included as terms of comparison. (For interpretation
of the references to color in this figure legend, the reader is referred to the Web version of this article.).

Fig. 6 shows the schematic illustration of the plate-on-cylinder (POC) wear test system used
in this study, photography of each specimen after POC wear test and their respective SEM SE
16
images of the worn surfaces. The surfaces showed different aspects, which indicate the wear
mechanisms acting in each case. When the cylinder slides in contact with the specimen, a crescent
worn groove is formed, as shown in the schematic illustration (Fig. 6 (a)).

The morphological conditions of the worn groove surfaces (Fig. 6 (b-d)), shows that when the
hard cylinder slides against the API 5L X80 steel (Fig. 6 (b)), the asperities and the abrasive
particles (bentonite) present in the fluid (tribosystem as a third body) push the surface material
forward in a process that involves considerable plastic deformation and scratching. Furthermore,
during the sliding, the material is detached from the surface. Regarding the F-1.2B (Fig. 6 (c)) and
F-3.5B (Fig. 6 (d)) alloys, the presence of hard M2B in the microstructure acts as a barrier to plastic
deformation and scratching from bentonite abrasives, resulting in higher wear resistance. For the F-

f
1.2B, adhesive wear is more preponderant as revealed by the tracks of plastic deformation of the

oo
soft α-ferrite matrix due to the contact with the AISI 1040 axis. In contrast, in the F-3.5B alloy, only

r
the presence of small holes in the worn surfaces was observed, which suggests a three-body wear
-p
mechanism is happening ((specimen)-(bentonite/M2B)-(AISI 1040 axis)).
re
Tests conducted by Tabrett et al. [49], showed that the ratio of abrasive grit size to the mean
lP

free distance of the matrix can give an idea of the protection of the matrix offered by the carbides
present in those materials. Fulcher et al. [50] showed that if the distance between the hard particles
na

is small, the carbides protect the matrix against wear and the matrix provides mechanical support to
the carbides. On the other hand, when the distance between the hard particles is large, the wear of
ur

the matrix is more pronounced. In this case, as the grain size and consequently the distance between
Jo

the hard M2B in the F-1.2B are larger than those presented by the F-3.5B (see Fig. 3), the protection
against the indentations promoted by the bentonite/M2B abrasives in the material with lower boron
content is less effective, once the area of the exposed α-ferrite matrix is higher in this material. On
the other hand, when the boron content is increased, the volume fraction of the M2B increases, and
the grain size decreases so that the distance between borides is reduced, which results in effective
protection of the soft α-ferrite matrix against the abrasive indentations, increasing hardness and
consequently decreasing worn volume and specific wear rate. In addition, Fig. 6 also indicates that
the groove depth and width decreased considerably.

17
Specimen holder
(a) (b)
Specimen

Worn groove

Wear direction
FN 50 μm

AISI 1040 axis

(c) (d)

f
r oo
-p
re
100 μm 100 μm
lP

Fig. 6. (a) Schematic illustration of the plate-on-cylinder (POC) wear test system used in this study. This system is
composed of: specimen holder; machined specimen forced with a normal force (𝐹𝑁 ) of 575.7 N against an AISI 1040
steel axis with a rotation speed of 200 rpm, which will generate a worn groove in the specimens. To simulate the
tribosystem found in the wear of risers and casings, the POC wear test was performed with real drilling mud, whose
na

composition is detailed in Table 1. (b-d) Photography of each specimen after POC wear test (total sliding distance of
25500 m) and their respective SEM SE images of the worn surfaces of API 5L X80, F-1.2B, and F-3.5B alloys,
respectively. SE image of API 5L X80 was included for comparison. Dash line included as eye guide to indicate the
ur

region of the worn surfaces where the SEM analysis was performed.
Jo

4. Conclusions

The electrochemical and tribological responses of boron-modified ferritic stainless steels were
assessed. Corrosion and wear-related parameters were appraised and highlight the combined
property portfolio of B-containing stainless steels to conventional API 5L X80 steel. From the
present study, the following conclusions may be drawn:

· It was possible to produce boron-modified ferritic stainless steels with additions of 1.2 and 3.5
wt.% B with high chemical and microstructural homogeneity by spray forming.

· The microstructures were composed of an α-ferrite matrix with M2B types borides
homogeneously distributed. With the addition of 1.2 wt.% B, the M2B were formed
eutectically (irregular morphology). In contrast, the addition of 3.5 wt.% B led to the
formation of M2B as a primary phase (faceted and elongated morphology).
18
· The addition of boron does not impair the stainless steels corrosion resistance if the
composition of the alloy is adjusted-tailored (which was the case of the alloy with 1.2 wt.%
B) in such a way that the final composition of the matrix was similar to that of the non-
modified alloy.

· Outstanding wear resistance was found for the boron-modified ferritic stainless steels,
especially the alloy with 3.5 wt.% B, which was the result of a combination of all
microstructural features: small grains, refined hard M2B well dispersed in a soft α-ferrite
matrix. Thus, the M2B were able to effectively protect the α-ferrite matrix from scratching and
material removal promoted by the tribosystem found in risers and casing.

· Despite the differences between the final microstructures, the corrosion and wear resistance of

f
the boron-modified ferritic stainless steels were considerably higher than the low alloy high

oo
strength steel API 5L X80 (often used in risers and casings manufacture).

r
· Of the alloys tested herein, the alloy with 1.2 wt.% B proved superior and was the alloy that
-p
offered the best combination of corrosion and wear resistance.
re
Funding: General financial support received from CNPq/Brazil, CAPES/Brazil and Petrobras S.A.
lP

No interference with study design and data analysis.


na

Acknowledgments: The authors would like to acknowledge the support from the Conselho
ur

Nacional de Desenvolvimento Científico e Tecnológico (CNPq/Brazil, grant number 131476/2013-


0), Coordenação de Aperfeiçoamento de Pessoal de Nível Superior (CAPES/Brazil, finance code
Jo

001), and Petrobras S.A. The Villares Metals S.A. is acknowledged for supplying the stainless steel.
The authors also acknowledge the Laboratory of Structural Characterization (LCE/DEMa/UFSCar)
and the Center for Characterization and Development of Materials (CCDM/DEMa/UFSCar) for the
general facilities.

Conflicts of Interest: The authors declare that they have no known competing financial interests or
personal relationships that could have appeared to influence the work reported in this paper.

Data availability: The data supporting the findings in this study are available within the paper. Any
further information or clarification is available from the corresponding author upon reasonable
request.

19
References
[1] Berger JE, Jorge AM, Koga GY, Roche V, Kiminami CS, Bolfarini C, et al. Influence of
chromium concentration and partial crystallization on the corrosion resistance of FeCrNiB
amorphous alloys. Mater Charact 2021;179:111369. doi:10.1016/j.matchar.2021.111369.
[2] Iannuzzi M, Barnoush A, Johnsen R. Materials and corrosion trends in offshore and subsea
oil and gas production. Npj Mater Degrad 2017;1:2. doi:10.1038/s41529-017-0003-4.
[3] Zepon G, Nascimento ARC, Kasama AH, Nogueira RP, Kiminami CS, Botta WJ, et al.
Design of wear resistant boron-modified supermartensitic stainless steel by spray forming
process. Mater Des 2015;83:214–23. doi:10.1016/j.matdes.2015.06.020.
[4] Bhadeshia H, Honeycombe R. Steels: Microstructure and Properties. 4th ed. Amsterdam:
Elsevier; 2017.

f
oo
[5] Loto RT. Study of the corrosion behaviour of S32101 duplex and 410 martensitic stainless
steel for application in oil refinery distillation systems. J Mater Res Technol 2017;6:203–12.

r
doi:10.1016/j.jmrt.2016.11.001.
[6]
-p
American Petroleum Institute. API Specification 5L. 46th ed., 2018.
re
[7] McCafferty E. Introduction to Corrosion Science. New York, NY: Springer New York; 2010.
lP

doi:10.1007/978-1-4419-0455-3.
[8] Smith WF. Structure and Properties of Engineering Alloys. 2nd ed. New York: McGraw-
na

Hill; 1993.
[9] Jr. WDC, Rethwisch DG. Materials Science and Engineering: An Introduction. 9th ed. 2013.
ur

[10] Monson PJE, Steen WM. Comparison of Laser Hardfacing with Conventional Processes.
Jo

Surf Eng 1990;6:185–93. doi:10.1179/sur.1990.6.3.185.


[11] Ahn D-G. Hardfacing technologies for improvement of wear characteristics of hot working
tools: A Review. Int J Precis Eng Manuf 2013;14:1271–83. doi:10.1007/s12541-013-0174-z.
[12] Wang Z, Lin T, He X, Shao H, Zheng J, Qu X. Microstructure and properties of TiC-high
manganese steel cermet prepared by different sintering processes. J Alloys Compd
2015;650:918–24. doi:10.1016/j.jallcom.2015.08.047.
[13] Sigolo E, Soyama J, Zepon G, Kiminami CS, Botta WJ, Bolfarini C. Wear resistant coatings
of boron-modified stainless steels deposited by Plasma Transferred Arc. Surf Coatings
Technol 2016;302:255–64. doi:10.1016/j.surfcoat.2016.06.023.
[14] Azarkevich AA, Kovalenko L V., Krasnopolskii VM. The optimum content of boron in steel.
Met Sci Heat Treat 1995;37:22–4. doi:10.1007/BF01151795.
[15] GRANT PS. Solidification in Spray Forming. Metall Mater Trans A 2007;38:1520–9.
doi:10.1007/s11661-006-9015-3.
[16] Schulz A, Uhlenwinkel V, Escher C, Kohlmann R, Kulmburg A, Montero MC, et al.
20
Opportunities and challenges of spray forming high-alloyed steels. Mater Sci Eng A
2008;477:69–79. doi:10.1016/j.msea.2007.08.082.
[17] Soyama J, Lopes TP, Zepon G, Kiminami CS, Botta WJ, Bolfarini C. Wear Resistant Duplex
Stainless Steels Produced by Spray Forming. Met Mater Int 2019;25:456–64.
doi:10.1007/s12540-018-0202-8.
[18] Soyama J, Zepon G, Lopes TP, Beraldo L, Kiminami CS, Botta WJ, et al. Microstructure
formation and abrasive wear resistance of a boron-modified superduplex stainless steel
produced by spray forming. J Mater Res 2016;31:2987–93. doi:10.1557/jmr.2016.323.
[19] Lopes TP, Soyama J, Zepon G, Nascimento ARC, Costa e Silva A, Kiminami CS, et al.
Thermodynamic Calculations for the Investigation of Phase Formation in Boron-Modified
Ferritic Stainless Steel. J Phase Equilibria Diffus 2017;38:343–9. doi:10.1007/s11669-017-

f
oo
0550-y.
[20] Ha H-Y, Jang JH, Lee T-H, Kim S-D, Lee C-H, Moon J. Enhancement of the resistance to

r
localized corrosion of type 304 borated stainless steels through hot rolling. Corros Sci
-p
2021;192:109798. doi:10.1016/j.corsci.2021.109798.
re
[21] Ha H-Y, Kim S-D, Jang JH, Lee T-H, Lee C-H, Moon J. Pitting Corrosion and Passive
lP

Behavior of Type AISI 304-based Borated Stainless Steels in a Boric Acid Solution. J
Electrochem Soc 2020;167:101506. doi:10.1149/1945-7111/ab9b95.
na

[22] Koga GY, Ferreira T, Guo Y, Coimbrão DD, Jorge Jr AM, Kiminami CS, et al. Challenges in
optimizing the resistance to corrosion and wear of amorphous Fe-Cr-Nb-B alloy containing
ur

crystalline phases. J Non Cryst Solids 2021;555:120537.


Jo

doi:10.1016/j.jnoncrysol.2020.120537.
[23] Andersson J-O, Helander T, Höglund L, Shi P, Sundman B. Thermo-Calc &amp; DICTRA,
computational tools for materials science. Calphad 2002;26:273–312. doi:10.1016/S0364-
5916(02)00037-8.
[24] Zepon G, Ellendt N, Uhlenwinkel V, Bolfarini C. Solidification Sequence of Spray-Formed
Steels. Metall Mater Trans A 2016;47:842–51. doi:10.1007/s11661-015-3253-1.
[25] Otani LB, Matsuo MM, Freitas BJM, Zepon G, Kiminami CS, Botta WJ, et al. Tailoring the
microstructure of recycled 319 aluminum alloy aiming at high ductility. J Mater Res Technol
2019;8:3539–49. doi:10.1016/j.jmrt.2019.06.030.
[26] ASTM E407-07e1: Standard Practice for Microetching Metals and Alloys 2015.
[27] ASTM E384-17: Standard Test Method for Microindentation Hardness of Materials 2017.
[28] Guo C, Kelly P. Boron solubility in Fe–Cr–B cast irons. Mater Sci Eng A 2003;352:40–5.
doi:10.1016/S0921-5093(02)00449-5.
[29] dos Santos ER, da Silva WA, Koga GY, Bolfarini C, Zepon G. Corrosion Resistant Boron-
21
Modified Ferritic and Austenitic Stainless Steels Designed by CALPHAD. Metall Mater
Trans A 2021. doi:10.1007/s11661-021-06226-4.
[30] Silva D, Lima L, Araújo A, Silva V, Raimundo R, Damasceno I, et al. The Effect of
Microstructural Changes on Mechanical and Electrochemical Corrosion Properties of Duplex
Stainless Steel Aged for Short Periods. Materials (Basel) 2020;13:5511.
doi:10.3390/ma13235511.
[31] Silva DDS, Simões TA, Macedo DA, Bueno AHS, Torres SM, Gomes RM. Microstructural
influence of sigma phase on pitting corrosion behavior of duplex stainless steel/NaCl
electrolyte couple. Mater Chem Phys 2021;259:124056.
doi:10.1016/j.matchemphys.2020.124056.
[32] Silva DDS, Sobrinho JMB, Souto CR, Gomes RM. Application of electromechanical

f
oo
impedance technique in the monitoring of sigma phase embrittlement in duplex stainless
steel. Mater Sci Eng A 2020;788:139457. doi:10.1016/j.msea.2020.139457.

r
[33] Silva DDS, Raimundo RA, Torquato RA, Faria GL, Morales MA, Simões TA, et al. Low-
-p
field magnetic analysis for sigma phase embrittlement monitoring in thermally aged 22Cr
re
duplex stainless steel. J Magn Magn Mater 2020;513:167072.
lP

doi:10.1016/j.jmmm.2020.167072.
[34] Macdonald JR. Impedance spectroscopy : emphasizing solid materials and systems. New
na

York: Wiley; 1987.


[35] Engelhardt GR, Case RP, Macdonald DD. Electrochemical Impedance Spectroscopy
ur

Optimization on Passive Metals. J Electrochem Soc 2016;163:C470–6.


Jo

doi:10.1149/2.0811608jes.
[36] Örnek C, Leygraf C, Pan J. Passive film characterisation of duplex stainless steel using
scanning Kelvin probe force microscopy in combination with electrochemical measurements.
Npj Mater Degrad 2019;3:8. doi:10.1038/s41529-019-0071-8.
[37] Morshed-Behbahani K, Najafisayar P, Pakshir M, Shahsavari M. An electrochemical study
on the effect of stabilization and sensitization heat treatments on the intergranular corrosion
behaviour of AISI 321H austenitic stainless steel. Corros Sci 2018;138:28–41.
doi:10.1016/j.corsci.2018.03.043.
[38] Fattah-alhosseini A, Vafaeian S. Comparison of electrochemical behavior between coarse-
grained and fine-grained AISI 430 ferritic stainless steel by Mott–Schottky analysis and EIS
measurements. J Alloys Compd 2015;639:301–7. doi:10.1016/j.jallcom.2015.03.142.
[39] Lv J, Luo H. Comparison of corrosion behavior between coarse grained and nano/ultrafine
grained 304 stainless steel by EWF, XPS and EIS. J Nucl Mater 2014;452:469–73.
doi:10.1016/j.jnucmat.2014.05.041.
22
[40] Frankel GS. Pitting Corrosion of Metals: A Review of the Critical Factors. J Electrochem
Soc 1998;145:2186. doi:10.1149/1.1838615.
[41] Stern M, Geaby AL. Electrochemical Polarization: I . A Theoretical Analysis of the Shape of
Polarization Curves. J Electrochem Soc 1957;104:56. doi:10.1149/1.2428496.
[42] Ebrahimi N, Momeni M, Kosari A, Zakeri M, Moayed MH. A comparative study of critical
pitting temperature (CPT) of stainless steels by electrochemical impedance spectroscopy
(EIS), potentiodynamic and potentiostatic techniques. Corros Sci 2012;59:96–102.
doi:10.1016/j.corsci.2012.02.026.
[43] Schultze JW, Lohrengel MM. Stability, reactivity and breakdown of passive films. Problems
of recent and future research. Electrochim Acta 2000;45:2499–513. doi:10.1016/S0013-
4686(00)00347-9.

f
oo
[44] Zepon G, Nogueira RP, Kiminami CS, Botta WJ, Bolfarini C. Electrochemical Corrosion
Behavior of Spray-Formed Boron-Modified Supermartensitic Stainless Steel. Metall Mater

r
Trans A 2017;48:2077–89. doi:10.1007/s11661-017-3980-6.
[45]
-p
Brug GJ, van den Eeden ALG, Sluyters-Rehbach M, Sluyters JH. The analysis of electrode
re
impedances complicated by the presence of a constant phase element. J Electroanal Chem
lP

Interfacial Electrochem 1984;176:275–95. doi:10.1016/S0022-0728(84)80324-1.


[46] Koga GY, Zepon G, Santos LS, Bolfarini C, Kiminami CS, Botta WJ. Wear Resistance of
na

Boron-Modified Supermartensitic Stainless Steel Coatings Produced by High-Velocity


Oxygen Fuel Process. J Therm Spray Technol 2019;28:2003–14. doi:10.1007/s11666-019-
ur

00961-2.
Jo

[47] Antony KC. Wear-Resistant Cobalt-Base Alloys. JOM 1983;35:52–60.


doi:10.1007/BF03338205.
[48] Schmid-Fetzer R. Phase Diagrams: The Beginning of Wisdom. J Phase Equilibria Diffus
2014;35:735–60. doi:10.1007/s11669-014-0343-5.
[49] Tabrett CP, Sare IR, Ghomashchi MR. Microstructure-property relationships in high
chromium white iron alloys. Int Mater Rev 1996;41:59–82. doi:10.1179/imr.1996.41.2.59.
[50] Fulcher JK, Kosel TH, Fiore NF. The effect of carbide volume fraction on the low stress
abrasion resistance of high Cr-Mo white cast irons. Wear 1983;84:313–25.
doi:10.1016/0043-1648(83)90272-7.

23
Declaration of interests

☒ The authors declare that they have no known competing financial interests or personal relationships
that could have appeared to influence the work reported in this paper.

☐ The authors declare the following financial interests/personal relationships which may be considered
as potential competing interests:

of
ro
-p
re
lP
na
ur
Jo

View publication stats

You might also like