Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

Materials Research Express

PAPER

Corrosion response of ultra-high strength steels used for automotive


applications
To cite this article: Husnu Gerengi et al 2019 Mater. Res. Express 6 0865a6

View the article online for updates and enhancements.

This content was downloaded from IP address 144.122.7.102 on 27/05/2019 at 09:55


Mater. Res. Express 6 (2019) 0865a6 https://doi.org/10.1088/2053-1591/ab2178

PAPER

Corrosion response of ultra-high strength steels used for automotive


RECEIVED
13 February 2019
applications
REVISED
9 May 2019
ACCEPTED FOR PUBLICATION
Husnu Gerengi1 , Nuri Sen1, Ilyas Uygur1 and Moses M Solomon2
13 May 2019 1
Corrosion Research Laboratory, Department of Mechanical Engineering, Faculty of Engineering, Duzce University, 81620, Duzce,
PUBLISHED Turkey
2
24 May 2019 Center of Research Excellence in Corrosion, Research Institute, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi
Arabia
E-mail: nurisen@duzce.edu.tr and moses.solomon@kfupm.edu.sa.

Keywords: mechanical property, corrosion resistance, chloride-rich environment, pitting corrosion, Docol alloys

Abstract
The corrosion resistance properties of two ultra-high strength steels, Docol 1200 and 1400 were
studied in 3.5 wt% NaCl solution using chemical and electrochemical techniques supported by surface
characterization techniques namely scanning electron microscope (SEM), energy dispersive
spectroscopy (EDX), atomic force microscope (AFM), and optical profilometer. The mechanical
properties of both Docol 1200 and Docol 1400 uncorroded and corroded were determined by tensile
test. Results obtained reveal that the two alloys exhibit similar mechanical properties. Corrosion has
effect on the mechanical properties of the alloys. Corrosion studies reveal that both Docol 1200 and
1400 are prone to corrosion in NaCl environment. A corrosion rate of 30.6 mpy and 49.6 mpy was
recorded for a Docol 1200 at 25 °C and 60 °C, respectively. For Docol 1400, the corrosion rate
obtained at 25 °C and 60 °C was 32.7 mpy and 52.4 mpy, respectively. Docol 1200 is adjudged to
exhibit superior corrosion resistance property than Docol 1400. The optical profilometric results
disclose that Docol 1200 and Docol 1400 suffered pitting corrosion in NaCl solution. Pitting corrosion
was more severe with Docol 1400 than 1200.

1. Introduction

Improved engine efficiency, weight reduction, minimized frictional losses and using renewable energies in
automotive and transport industries can minimize significantly CO2 and NOx emissions. One way in which this
can be achieved is by using Ultra High Strength Steels (UHSS) in vehicle body structures with low thickness or by
using lightweight materials such as aluminium, magnesium or polymer matrix composites. UHSS steels like
Docol 1400, Docol 1600 are generally used for side impact beams, bumpers, battery protection for hybrid/
electric cars, tunnel reinforcements, and floor beams. Microstructure of most of these steels mainly consists of
hard and brittle docolensite in the soft ferrite matrix. Docolensite volume fraction can be as high as 30%.
However, dislocations are free to the mobile in the ferrite phase blocked on the docolensite islands due to lack of
a well-defined interface [1]. The strengthening effect is a function of the docolensite phase while the soft matrix
ensures high ductilty; other phases (bainite, perlite and residual austenite) may also be present in minute
quantities [1]. Increasing the strength of steel comes with a decrease in the ductility. This makes the material
brittle and unsuitable for some applications [2–4].
Recent advances in this class of steels allow designers to make thin-sectioned car body and structural
applications in various industries. However, reduction in thickness requires better corrosion resistance for
UHSS steels. The degradation of steel by corrosion can be seen in the form of rust, but actually, this is more than
just an orange-brownish looking layer on the surface. It always lead to the deterioration of the physical and
chemical properties as well as induced the generation of holes and voids on the surface [2]. One aspect that has a
direct influence on the corrosion property of alloys is the complex phase character. In a typical scenario, ferrite,

© 2019 IOP Publishing Ltd


Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Table 1. Chemical composition of Docol 1200 and Docol 1400


utilized as the working electrode.

Elemental composition (%) Docol 1200 Docol 1400

C 0.067 0.143
Mn 1.586 1.145
P 0.0117 0.0114
Nb 0.002 <0.002
Al 0.047 0.041
Cu 0.054 0.168
Ti 0.034 0.033
Fe 97.736 98.049

bainite, austenite, cementite, docolensite, and some primary carbide particles may be controlled and combined
together to provide an optimum combination of high strength and good ductility.
There are only very few researches devoted to the corrosion of UHSS steels. Sarkar et al [5] performed
corrosion tests and reported that, an increase in the volume fraction of the Docolensite phase in the range of 10%
and the refinement of their phase constituents have a negative impact on the corrosion performance of the steel.
Although ferrite grains were more prone to corrosion than the pearlite phase, Osorio et al [6] measured a
deleterious effect on the corrosion resistance of a ferrite/docolensite microstructure when compared to the
ferrite/pearlite microstructure after the same heat treatment on the base steel. It was ascribed to the residual
stress from the docolensite formation and the generation of enormous microgalvanic corrosion cells occasioned
by the interconnection between the docolensite and ferrite phases. Zhang et al [7], however, reported an
opposite behaviour. In their studies on the dual-phase treatment of weathering steel 09CuPCrNi, Zhang et al
found that the ferrite/docolensite microstructure had a better corrosion resistance than the ferrite/pearlite
microstructure. Moreover, it was reported that, pearlite has a negative effect on the corrosion behaviour of steel
and that the preferred sites for nucleation of corrosion pits could be associated with the presence of different
phases [8].
Alloying is a simple and effective approach for enhancing the corrosion resistance of a metal. The content of
elements and the homogeneity are the crucial factors that affect the corrosion resistance. Townsend [9]
subjected a heat treated dual phase microstructures to a long test of up to 8 years in industrial, rural, and marine
environments. It was found that, the corrosion performance of the samples was independent of the quench-and-
temper heat treatments. According to Townsend, the corrosion performance could be estimated solely on initial
chemical composition of the alloy. Similarly, it had been shown that the inclusion of metals like Mn [10], Nb
[11, 12], etc improves the corrosion resistance of resultant alloy relative to single metal. Moreover, enhanced
corrosion resistance property had been achieved with Ti-doped Al2O3 [13].
The hot and cold deformation process includes a variety of microstructures within the steel components as
well as physical and chemical differences that can influence the corrosion response of these materials. Research
into the corrosion of these materials (UHSS) is important, especially determination and characterisation of
corrosion properties. Up to date, there is no direct comparison of fully martensitic steels and detailed research
about them. Thus, the aim of the present study is to characterize the corrosion properties of two different types
of UHSS fully martensitic steels with different chemical content, by electrochemical and surface examination
techniques, in order to investigate the influence of the chemical composition and the homogeneity on the
corrosion behaviour.

2. Materials preparation and methods

Sheets of about 1.5 mm thick Docol 1200 and Docol 1400 steel in the form of cold rolled and tempered were
purchased from Swedish Steel Company (SSAB). The nominal chemical compositions of both materials are
given in Table 1. The specimens were prepared with the aid of a water jet. The dogbone specimens used for the
mechanical properties testing experiments were 20 mm long and 30 mm in width with curved section of
140 mm×120 mm×1.5 mm as described in Figure 1. The mechanical properties of both Docol 1200 and
Docol 1400 were determined by uniaxial tensile test. Tensile tests were performed with 100 kN UTEST testing
machine. Tensile testing procedures were implemented at 0.01 s−1 and 0.1 s−1 strain rates to analyze the
metallurgical and mechanical changes at different deformation rates.
Gamry potentiostat/galvanostat/ZRA (Reference 600) instrument was used for the electrochemical
experiments (electrochemical impedance spectroscopy (EIS) and potentiodynamic polarization (PDP)). The
samples for electrochemical experiments were soldered into a copper wire to provide the electrical contact. They

2
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 1. Dimensions of dogbone specimen for tensile test.

were insulated to obtain 0.5 cm2 as the exposed surface area. The samples were abraded mechanically using
400–2000 grade abrasive paper, cleaned with distilled water and acetone, and dried using sample dryer. In the
electrochemical experiments, a set of three electrodes, namely a reference (Ag/Ag/Cl), a counter (platinum
mesh), and a working (Docol 1200 or Docol 1400) electrode were utilized. The corrodent was a 3.5 wt% NaCl
solution. Before the EIS experiments, the working electrode was trapped in the corrodent for 3600 s to ensure
open circuit potential (OCP) stability. The frequency range adopted for EIS experiments was 100 kHz to 0.1 Hz
and amplitude signal was 10 mV peak-to-peak. The obtained results were analyzed using Echem Analyst 6.32
program. The PDP experiments were performed by scanning the working electrode at a constant sweep rate of
1 mV s−1 between ±250 mV interval with reference to corrosion potential. The extrapolation technique was
adopted for the analysis of the polarization curves.
The weight loss measurements of the studied alloys in 3.5 wt% NaCl solution at 25 °C and 60 °C, was
performed in accordance with the ISO 8407:2009, IDT standard. Briefly, 100 ml of the corrodent was introduced
into sample bottles labeled Docol 1200 and Docol 1400, respectively. The sample bottles were placed in a
thermostated water bath maintained at 25 °C or 60 °C. External thermometer was used to ascertain that the
temperature of the corrodent was either 25 °C or 60 °C as the case may be. Upon establishing this, two pre-
weighed coupons, each of Docol 1200 and Docol 1400 (surface area=9 cm2) were freely suspended in their
respective sample bottle. The experiment was allowed to stand for 24 h. At the end of the experiment, the
coupons were retrieved, washed thoroughly under running water, rinsed in distilled water and acetone, dried
using sample dryer, and reweigh. The weight loss was calculated as the difference between the initial and final
weights. The weight loss value was used to compute the corrosion rate in mpy following equation (1):

3.45 ´ 106 ´ W
Corrosion Rate (mpy) = (1)
A´t´D

where t=time of exposure in hours, A=area in cm2, W=mass loss in grams, and D=density in g cm−3.
The surface morphology of the corroded Docol samples was studied using SEM (FEI, Model: Quanta FEG
250) coupled to an energy dispersive spectroscopy (EDX) probe (accelerator voltage equals 20 keV). AFM studies
were done using Park System X-100E AFM model. Veeco Dektak XT Bruker mechanical stylus profilometer was
used for the surface roughness studies.

3. Results and discussion

3.1. Tensile test and microstructure


The mechanical properties of the high strength Docol 1200 and Docol 1400 material before and after corrosion
were determined. The stress versus strain graphs obtained from the tensile tests experiments are presented in
Figure 2. Clearly, the studied alloys behaved alike; that is, the stress increases linearly until it reaches the yield
strength. Thereafter, the yielding plateau is observed. However, the mechanical properties appears to be sensitive

3
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 2. Engineering stress-strain curves of studied steels (a) before corrosion; (b, c) comparative graphs of before and after corrosion
for three days.

to the chemical composition of the alloy. As could be seen in the figure, the Docol 1200 reaches the yield strength
faster than the Docol 1400. Accordingly, the ultimate tensile strength (UTS), yield strength, strength coefficient,
and strain-hardening parameters for the tested UHSS specimens were obtained from the graphs and the
parameters are listed in the inserted tables. The yield strength was determined using the 0.2% offset method [14]
while the coefficient was calculated using equation (2):
s = Ke n (2)

where K is the strength coefficient and n is the strain-hardening exponent. Stress hardening coefficient refers to the
magnitude of the strength of a material. For most metals, n has value in the range of 0.1–0.5 [15, 16]. A low
coefficient value implies that the material will break without being strain hardened; its high value indicates that the
material can strain harden, and that it has a ductile structure [15, 16]. Both Docol 1200 and 1400 exhibited high
yield strength, UTS, strength coefficient and strain-hardening exponent (see inserted table in Figure 2(a)). This
indicates the ability of retained and increased work hardening response under dynamic conditions and thus
promoting energy absorption. The corrosion of these alloys in 3.5 wt% NaCl solution slightly alters the mechanical
properties of the alloys. While the values of the ultimate tensile strength, yield strength, strength coefficient, and
strain hardening increase for corroded Docol 1400 (inserted table in Figure 2(c)), the values of the parameters
decrease for corroded Docol 1200 (inserted table in Figure 2(b)) relative to those of the uncorroded alloys. This
observation may due to the differences in the chemical composition of the alloys (Table 1).
The surface morphology of steel grades was determined using optical microscope. Figure 3 shows the optical
images of Docol 1200 and Docol 1400 samples. The microstructures consist of almost all fully martensitic matrix
and a little retained austenite. A comparison of Figures 3(a) to (b) disclosed that, the retained austenites are more
in Figure 3(b) as compared to Figure 3(a). This austenite retention may increase the ductility to break of the
Docol 1400 as suggested by the higher value of the strain-hardening exponent (Figure 2). A deeper insight into
the structural impact behavior of toe cap models produced from these materials with respect to strength
properties was given by Costa et al [17]. The authors used experimental standard testing program and numerical
simulation developed using an explicit dynamics software to analyse Docol 1200 and Docol 1400 for strain-rate
sensitivity. It was found that, both Docol 1200 and Docol 1400 display moderate strain-rate dependence with
slight low response from Docol 1400 due to strength properties and typical low elongation values. Nevertheless,
the alloys demonstrated good ability under dynamic crush conditions to promote and retain work hardening
effects and, even for higher deformation values, a linear and reasonable load response and absorbed energies
were observed.

4
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 3. Microstructure of the materials (a) Docol 1200, (b) Docol 1400.

Figure 4. Potentiodynamic polarization curves for the studied alloys in 3.5 wt% NaCl solution at ordinary temperature.

3.2. Corrosion resistance


The chemical and electrochemical techniques namely, weight loss, potentiodynamic polarization, and
electrochemical impedance spectroscopy were used to study the corrosion resistance property of Docol 1200
and Docol 1400 alloys in 3.5 wt% NaCl solution. Figure 4 shows the polarization curves obtained for Docol 1200
and Docol 1400 in 3.5 wt% NaCl solution at 25 °C. The two graphs are similar and infer same corrosion
mechanism. A direct relationship between the corrosion current density and the corrosion potential is observed
in the anodic region of the graphs. Huang et al [18] had recently reported similar observation for NiCu and
S690Q steel in 3.5 wt% NaCl solution. It is indicative of the active dissolution of the alloys in the studied
environment.
The polarization parameters namely, corrosion potential (Ecorr), corrosion density (icorr), anodic and
cathodic Tafel slopes (βa and βc) derived from the analysis of the curves are presented in Table 2. As could be seen
in the table, the Ecorr and icorr values for Docol 1200 are nobler compare to those of Docol 1400. For instance, the
Ecorr and icorr values for Docol 1200 are −546 mV versus Ag/AgCl and 22 μA cm−2 while those of Docol 1400
are −615 mV versus Ag/AgCl and 25 μA cm−2, respectively. This is reflective of higher corrosion resistance
property by Docol 1200 than Docol 1400 in 3.5 wt% NaCl solution. This may be due to the differences in the
chemical composition of the alloy. In a specific term, it may be due to the lesser amount of Fe in Docol 1200 than
Docol 1400 (see Table 1) as this element is sensitive to corrosion.
In previous studies, the value of Ecorr was used to define the prevalence corrosion mechanism of UHSS in
chloride containing environments [19–22]. The Ecorr value in the range of −570 to −880 mV versus reference

5
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 5. Impedance diagrams for the studied alloys in 3.5 wt% NaCl solution at ordinary temperature in (a) Nyquist and (b) Bode-
Phase angle representations.

Table 2. Polarization parameters for the studied alloys in 3.5 wt% NaCl solution at ambient
temperature.

Alloy βa (mV dec−1) βc (mV dec−1) Ecorr (mV/Ag/AgCl) icorr (μA cm−2)

Ducol 1200 70 430 −546 22


Ducol 1400 92 650 −615 25

Table 3. Electrochemical impedance parameters for the studied alloys in 3.5 wt%
NaCl solution at ambient temperature.

Æ
−1 n −2
Alloy Rs (Ω cm ) 2
Y0 (Ω s cm ) α 0„α „1 Rct (Ω cm2)

Docol 1200 28.94 942.20 0.73 772.50


Docol 1400 28.43 1034.00 0.71 697.40

was associated with anodic dissolution and hydrogen induced cracking. In our case, the Ecorr for Docol 1200 and
1400 is −546 mV versus Ag/AgCl and −615 mV versus Ag/AgCl, respectively and fall within this Ecorr range.
Figure 5 shows the electrochemical impedance graphs obtained for Docol 1200 and Docol 1400 in the
studied corrosive medium at 25 °C in (a) Nyquist and (b) Bode formats. The alloys exhibited electrochemical
characteristics typically of a charge controlled corrosion mechanism [23, 24]. That is, a single and imperfect
capacitive loop at high frequencies in the Nyquist graphs (Figure 5(a)). This loop corresponds to a one-time

6
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 6. (a) Nyquist and (b) Bode representations of the experimental and fitted plots for the studied alloys. Inset: the equivalent
circuit used for the fitting.

Table 4. Weight loss results for the studied samples exposed to


3.5 wt% NaCl environment for 24 h at 25 °C and 60 °C.

Weight loss (g) Corrosion rate (mpy)

Sample 25 °C 60 °C 25 °C 60 °C

Docol 1200 0.0151 0.0245 30.6456 49.6215


Docol 1400 0.0161 0.0258 32.6751 52.3613

constant in the Bode representation (Figure 5(b)). The semi-circular graph obtained for Docol 1200 is larger
than that of Docol 1400 meaning that the rate of charge transfer in Docol 1200 was slower compared to Docol
1400. The electrochemical impedance data were analyzed using a simple Rs(QRct) equivalent circuit (see
Figure 6(a) inset). It consists of a solution resistance (Rs), a charge transfer resistance (Rct), and a constant phase
element (CPE). The CPE can be modeled using the following equation [25]:
ÆCPE = [Y0 (jw )a ]-1 (3)
where ÆCPE is the impedance, Y0 the CPE constant, α the CPE exponent, j is an imaginary number (−1)1/2, and
w the angular frequency in rad/s. The selected equivalent circuit gave good fitting as could be seen in Figure 6.
The results obtained from the analysis are given in Table 3. The results are in excellent agreement with the results
from polarization studies (Table 2) and Docol 1200 alloy exhibits a higher corrosion resistance than Docol 1400.
The charge transfer resistance for Docol 1200 is 772.50 Ω cm2 while that of Docol 1400 is 697.40 Ω cm2. The
superior corrosion resistance of Docol 1200 over Docol 1400 is also obvious in the Bode-phase graphs
(Figure 5(b)). The Docol 1200 displayed the Bode modulus and the phase toward nobler values compare to

7
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 7. SEM images showing the surface morphology of Docol 1200 (a), (c) and 1400 (b), (d) before and after immersion in 3.5 wt%
NaCl solution at ordinary temperature for 6 h.

Table 5. SEM-EDX results after and before experiments.

Docol 1200 Docol 1400

Elements (%) Before experiment After experiment Before experiment After experiment

Na — 11.94 — 10.54
Cl — 5.73 — 5.10
O — 10.00 — 14.99
Nb 0.44 0.45 0.53 0.52
Al 0.21 0.36 0.37 0.57
C 2.63 1.52 3.68 1.73
Mn 1.83 1.36 1.65 1.04
Fe 72.42 53.70 78.66 49.04

Docol 1400. Worthy of mentioning in Table 3 also, is the fact that, the α value is close to unity (i.e. 0.73 for Docol
1200 and 0.71 for Docol 1400). When α=1, the Y0 is identical to an ideal capacitor [26]. It could be said that,
the metal/solution interface was capacitive in nature. Again, the parameter α can be used as a measure of surface
inhomogeneity [27]. The smaller α value recorded for Docol 1400 (0.71) relative to that of Docol 1200 (0.73)
infers a more heterogeneous surface. Furthermore, the Y0 value reveals the characteristics of the metal surface
film [28]. The smaller Y0 value for Docol 1200 (942.20 Ω−1 sn cm−2) compared to that of Docol 1400 (1034.00
Ω−1 sn cm−2) is indicative of a more compact corrosion product film on Docol 1200 surface than on Docol 1400
surface [28].

8
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 8. AFM pictures showing the surface morphology of Docol (a), (b) 1200 and (c), (d) 1400 before and after immersion in
3.5 wt% NaCl solution at ordinary temperature for 6 h.

To study the effect of temperature on the corrosion resistance property of the alloy in 3.5 wt% NaCl solution,
weight loss experiments were undertaken at 25 °C and 60 °C and the results obtained are given in Table 4. From
the table, it is observed that weight loss and corrosion rate, as expected increased with increasing temperature.
The alloys corrode in an alarming rate of up to 30.6456 mpy at 25 °C and 49.6215 mpy at 60 °C for Docol 1200.
For Docol 1400, the corrosion rate is 32.6751 mpy at 25 °C and 52.3613 mpy at 60 °C. The implication is that,
Docol steels requires corrosion control measures to be in place when intended for application in a corrosive
environment. The weight loss results are in good agreement with the electrochemical results (Tables 2 and 3).

3.3. Surface observation


The corrosion behaviour of Docol 1200 and 1400 was also studied employing various surface techniques.
Figure 7 shows the SEM images of the two alloys samples before and after immersion in 3.5 wt% NaCl solution
for 24 h. The corresponding EDX results are given in Table 5. Obviously, the smooth sample images in
Figures 7(a) and (b) became rougher after exposure to the corrodent (Figures 7(c) and (d)) indicating severe
corrosion. The corrosion was more serious in the case of Docol 1400 than 1200, which is in agreement with other
experimental results (Tables 2–4). A quick observation of the SEM-EDX results in Table 5 discloses that the
primary contributor to the weight loss as well as the corrosion of the alloys in the studied environment is the Fe
component. About 38% and 26% of Fe was lost by corrosion in the case of a Docol 1400 and 1200, respectively.
The severe corrosion of these alloys is also evidenced in the AFM results shown in Figure 8. The average value of
profile deviation from the mean line (Ra) and the average peak to valley height (Rz) increased remarkably after
immersion in 3.5 wt% NaCl solution. For instance, the Ra and Rz were in the range of 3.840–1.704 nm and
17.554–11.464 nm, respectively for Docol 1200 before immersion in the aggressive solution. After immersion,

9
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 9. Optical profilometer results for Docol 1200 after immersion in 3.5 wt% NaCl solution at ordinary temperature for 6 h.

the Ra and Rz increased to 345 45–111 600 nm and 143 721–325 837 nm, respectively. Similar observations were
noted for aluminum 6060 and 6082 alloys exposed to artificial acid rain solution [29].
Localized corrosion, which often results in pit formation, is a serious challenge for steels in chloride
containing environment. To examine the extent of pitting corrosion for Docol 1200 and 1400 in 3.5 wt% NaCl
solution, optical profilometry experiments were undertaken. The results obtained are presented in Figures 9 and
10. Clearly, the alloys suffered serious pitting corrosion in studied medium. Different degrees of pits can be seen
in the figures. Docol 1400 (Figure 10) is also found to be prone to pitting corrosion than Docol 1200 (Figure 9).
The deepest pit in the case of Docol 1200 is about 2.0267 μm depth while that of Docol 1400 is about 5.1327 μm.

4. Conclusions

Based on the results obtained in this investigation, the following conclusions are drawn.

1. Docol 1200 and 1400 possess similar mechanical properties.


2. Both Docol 1200 and 1400 are prone to corrosion in NaCl environment.
3. The corrosion rate of Docol 1200 may reach 31 mpy at 25 °C and 50 mpy at 60 °C.
4. The corrosion rate of Docol 1400 may reach 33 mpy at 25 °C and 52 mpy at 60 °C.
5. Docol 1200 is more resistance to corrosion than Docol 1400 in NaCl environment.
6. Docol 1200 and Docol 1400 will suffer pitting corrosion in NaCl solution. Pitting corrosion is severe for
Docol 1400 than 1200.

10
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

Figure 10. Optical profilometer results for Docol 1400 after immersion in 3.5 wt% NaCl solution at ordinary temperature for 6 h.

Acknowledgments

The authors gratefully acknowledge the financial support for this work by the Duzce University Research Fund
(Project No: 2016.06.05.457) and express special thanks to Ertugrul Kaya, Mine Kurtay and Mesut Yıldız for
preparation of the working electrodes.

ORCID iDs

Husnu Gerengi https://orcid.org/0000-0002-9663-4264


Moses M Solomon https://orcid.org/0000-0002-3251-8846

References
[1] Caprili S, Salvatore W, Valentini R, Ascanio C and Luvarà G 2018 A new generation of high-ductile Dual-Phase steel reinforcing bars
Constr. Build. Mater. 179 66–79
[2] Moreno D E F 2014 Influence of microstructure on the corrosion performance of dual phase steels. Technische Universiteit Eindhoven
geboren te Bogota Master Thesis
[3] Furukava T, Tanino M and Morikawa H 1984 Effects of composition properties processing factors on the steels of as-hot-rolled dual-
phase Transactions ISIJ 24 113–21
[4] Sen N, Karaağaç I and Kurgan N 2016 Experimental research on warm deep drawing of HC420LA grade sheet material Int. J. Adv.
Manuf. Technol. 87 3359–71
[5] Sarkar P P, Kumar P, Manna M K and Chakraborti P C 2005 Microstructural influence on the electrochemical corrosion behavior of
dual-phase steels in 3.5 wt% NaCl solution Mater. Lett. 59 2488–91
[6] Osorio W R, Peixeto L C, Garcia L R and Garcia A 2009 Electrochemical corrosion response of a low carbon heat treated steel in a NaCl
solution Materials and Corrosion-Werkstoffe Und Korrosion 60 804–12

11
Mater. Res. Express 6 (2019) 0865a6 H Gerengi et al

[7] Zhang C L, Cai D, Liao B, Zhao T and Fan Y 2004 A study on the dual-phase treatment of weathering steel 09CuPCrNi Mater. Lett. 58
1524–9
[8] Zhao Y T, Yang S W, Shang C J, Wang X M and Liu W 2007 The mechanical properties and corrosion behaviors of ultra-low carbon
microalloying steel Materials Science and Engineering a-Structural Materials Properties Microstructure and Processing 454 695–700
[9] Townsend H E 2000 Atmospheric corrosion performance of quenched-and-tempered, high strength weathering steel Corrosion 56
883–6
[10] Kim M J and Kim J G 2015 Effect of manganese on the corrosion behavior of low carbon steel in 10wt% sulfuric acid Int. J. Electrochem.
Sci. 10 6872–85 (http://www.electrochemsci.org/papers/vol10/100906872)
[11] Wang D, Fu Y, Hu M, Jiang D, Gao X, Wang Q, Yang J, Sun J and Weng L 2018 Effect of Nb content on the microstructure and
corrosion resistance of the sputtered Cr–Nb–N coatings J. Alloys Compd. 740 510–8
[12] Wu Y, Wu Q, Zhu S, Pu Z, Zhang Y, Wang Q, Lang D and Zhang Y 2016 Effect of niobium element on the electrochemical corrosion
behaviour of depleted uranium J. Nucl. Mater. 478 7–12
[13] Comakli O, Yazıcı M, Yetim T, Yetim A F and Celik A 2018 Effect of Ti amount on wear and corrosion properties of Ti-doped Al2O3
nanocomposite ceramic coated CP titanium implant material Ceram. Int. 44 7421–8
[14] Zhou H, Wang W, Wang K and Xu L 2019 Mechanical properties deterioration of high strength steels after high temperature exposure
Constr. Build. Mater. 199 664–75
[15] Uygur I 2004 Tensile behavior of powder metallurgy processed (Al–Cu–Mg–Mn)/Si Cp composites Iranian Journal of Science and
Technology Transaction B-Engineering 28 239–48
[16] Uygur I, Evans W J, Bache M and Gulenc B 2004 The fatigue behaviour of SiC particulate reinforced 2124 aluminium matrix
composites Metallofizika i Novejsie Tehnologii 26 927–39
[17] Costa S L, Mendoca J P and Peixinho N 2016 Study on the impact behaviour of new safety toe cap model made of ulta high strength
steels Mater. Design 91 143–54
[18] Huang W H, Yen H W and Lee Y L 2019 Corrosion behavior and surface analysis of 690MPa-grade offshore steels in chloride media J
Mater Res Technol 8 1476–85
[19] Jianhua L, Qiang G, Mei Y and Songmei L 2014 SCC investigation of low alloy ultra-high strength steel 30CrMnSiNi2A in 3.5 wt% NaCl
solution by slow strain rate technique Chin. J. Aeronaut. 27 1327–33
[20] Parkins R N 1980 Predictive approaches to stress corrosion cracking failure Corros. Sci. 20 147–66
[21] Liu Z Y, Li X G and Cheng Y F 2012 Mechanistic aspect of near-neutral pH stress corrosion cracking of pipelines under cathodic
polarization Corros. Sci. 55 54–60
[22] Liu Z Y, Li X G, Du C W, Zhai G L and Cheng Y F 2008 Stress corrosion cracking behavior of X70 pipe steel in an acidic soil environment
Corros. Sci. 50 2251–7
[23] Xia G, Jiang X, Zhou L, Liao Y, Duan M, Wang H, Pu Q and Zhou J 2015 Synergic effect of methyl acrylate and N-cetylpyridinium
bromide in N-cetyl-3-(2-methoxycarbonylvinyl) pyridinium bromide molecule for X70 steel protection Corros. Sci. 94 224–36
[24] Solomon M M, Gerengi H and Umoren S A 2017 Carboxymethyl cellulose/silver nanoparticles composite: synthesis, characterization
and application as a benign corrosion inhibitor for St37 steel in 15% H2SO4 medium ACS Appl. Mater. Interfaces 9 6376–89
[25] Gerengi H, Solomon M M, Öztürk S, Yıldırım A, Gece G and Kaya E 2018 Evaluation of the corrosion inhibiting efficacy of a newly
synthesized nitrone against St37 steel corrosion in acidic medium: experimental and theoretical approaches Materials Science &
Engineering C 93 539–53
[26] Wu K, Zhou X, Wu X, Lv B, Jing G and Zhou Z 2019 Understanding the corrosion behavior of carbon steel in aminofunctionalized
ionic liquids for CO2 capture assisted by weight loss and electrochemical techniques Int. J. Greenhouse Gas Control 83 216–27
[27] Umoren S A, Solomon M M, Eduok U M, Obot I B and Israel A U 2014 Inhibition of mild steel corrosion in H2SO4 solution by coconut
coir dust extract obtained from different solvent systems and synergistic effect of iodide ions: ethanol and acetone extracts J. Environ.
Chem. Eng. 2 1048–60
[28] Sha J Y, Ge H H, Wan C, Wang L T, Xie S Y, Meng X J and Zhao Y Z 2019 Corrosion inhibition behaviour of sodium dodecyl benzene
sulphonate for brass in an Al2O3 nanofluid and simulated cooling water Corros. Sci. 148 123–33
[29] Gerengi. H, Bereket G and Kurtay M 2016 A morphological and electrochemical comparison of the corrosion process of aluminum
alloys under simulated acid rain conditions J. Taiwan Inst. Chem. Eng. 58 509

12

You might also like