You are on page 1of 29

Eur. Phys. J.

Plus (2020) 135:129


https://doi.org/10.1140/epjp/s13360-020-00167-4

Regular Article

Effective acid corrosion inhibitors for X60 steel


under turbulent flow condition based
on benzimidazoles: electrochemical, theoretical, SEM,
ATR-IR and XPS investigations

Ikenna B. Onyeachu1, Ime Bassey Obot1,a , Aeshah H. Alamri2 ,


Chinenye A. Eziukwu3
1 Center of Research Excellence in Corrosion, Research Institute, King Fahd University of Petroleum and
Minerals, Dhahran 31261, Saudi Arabia
2 Chemistry Department, College of Science, Imam Abdulrahman Bin Faisal University, P.O.
Box 76971, Dammam, Saudi Arabia
3 Department of Chemistry, Edo University, Iyamo, Edo State, Nigeria

Received: 6 June 2019 / Accepted: 4 November 2019


© Società Italiana di Fisica (SIF) and Springer-Verlag GmbH Germany, part of Springer Nature 2020

Abstract The corrosion inhibition properties of two benzimidazole derivatives 2-


(2-Bromophenyl)-1H-benzimidazole (BHB) and 2-(2-Bromophenyl)-1-methyl-1H-
benzimidazole (BMB) for X60 steel in 1 M HCl solution were investigated under
laminar and turbulent flow conditions. Experimental methods (electrochemical impedance
spectroscopy, linear polarization and potentiodynamic polarization) were complemented
with scanning electron microscopy, attenuated total reflectance infrared spectroscopy and
X-ray photoelectron spectroscopy. The results obtained all agree that the protonation of
BHB and BMB facilitated their protective adsorption on the steel. Although turbulent
flow condition increased the corrosion rate, BHB and BMB provided significant corrosion
protection for the carbon steel in the acid medium. The presence of a methyl group on the
benzimidazole ring conferred greater corrosion inhibition property on BMB when compared
with BHB. Theoretical calculations provided more insight into the inhibitor–steel interaction
at the atomic and molecular levels.

1 Introduction

Hydrochloric acid (HCl) is used in many industries for the removal of corrosion scales from
the surface of steel-based structural materials. Usually, this process also leads to serious
corrosion attack on the underlying steel substrate. A cost-effective and efficient approach to
reduce the corrosion attack during acid cleaning is to introduce corrosion inhibitors into the
acid solution.
Literature survey shows that nitrogen-based organic compounds are the most largely
employed inhibitors for steel corrosion in hydrochloric acid solution [1–11]. The nitrogen
atoms in the inhibitor molecules which are usually present in heterocyclic or aromatic rings
act as reactive centers which facilitate the adsorption of the corrosion inhibitors on the steel

a e-mail: obot@kfupm.edu.sa

0123456789().: V,-vol 123


129 Page 2 of 29 Eur. Phys. J. Plus (2020) 135:129

surface. The adsorption leads to the formation of an inhibitor layer which shields the steel
surface from contact with the acid solution. As nitrogen-based compounds, benzimidazole
derivatives have gained tremendous attention as acid corrosion inhibitors for steel because
they offer high inhibition efficiencies and also exhibit good environmental properties [2,
12–20].
The performance of a corrosion inhibitor can be affected by the inhibitor properties (like
its functional groups, steric effects, electron density at donor atoms and hydrophobicity), as
well as the solution properties like concentration and hydrodynamic (or flow) conditions [21,
22]. Most of the benzimidazole derivatives studied as corrosion inhibitors for acid corrosion
of steel have been largely investigated under static conditions. Not much research was carried
out to understand how this class of corrosion inhibitors performs under hydrodynamic condi-
tions. Studies in this direction are important because hydrodynamic conditions represent the
actual conditions under which most industrial acid-cleaning practices are conducted. While
an inhibitor may exhibit high efficiency under static corrosion condition, hydrodynamic fac-
tors like mass transport and wall shear stress could cause dramatic changes in inhibition
performance [23–25]. The influence of hydrodynamics on the efficiency of an inhibitor may
either be positive whereby corrosion inhibitor mass transport and surface coverage of the
metal surface are enhanced [26, 27], or negative whereby wall shear stress is increased and
inhibitor removal from metal surface is promoted [28–30].
The aim of the present work, therefore, is to investigate the effect of hydrodynam-
ics on the inhibitive properties of two benzimidazole derivatives: 2-(2-Bromophenyl)-1H-
benzimidazole (BHB) and 2-(2-Bromophenyl)-1-methyl-1H-benzimidazole (BMB) for steel
in 1 M HCl solution. Our earlier publication revealed that both molecules investigated in
this work showed good inhibition efficiency for mild steel in 0.5 M HCl under static con-
dition [31]. On this basis, we employ the rotating cylinder electrode (RCE) technique to
investigate the inhibitive performance of two benzimidazole derivatives under laminar and
turbulent flow conditions using rotation speeds of 200 and 1000 rpm similar to the work
reported earlier [32]. Electrochemical measurements like electrochemical impedance spec-
troscopy (EIS), linear polarization resistance (LPR) and potentiodynamic polarization (PDP)
were used. Surface characterizations techniques like scanning electron microscopy (SEM),
attenuated total reflectance infrared (ATR-IR) spectroscopy and X-ray photoelectron spec-
troscopy (XPS) were carried on the corroded and inhibitor film form on the steel surface.
Quantum chemical calculations and Monte Carlo simulations were utilized to support the
experimental results.

2 Experimental

2.1 Materials and solution preparation

The benzimidazole derivatives BHB (Fig. 1a) and BMB (Fig. 1b) were purchased from
Sigma-Aldrich. Stock solutions were prepared in isopropanol from which working concen-
trations of 300, 600 and 1200 ppm were obtained. The corrosive medium was 1 M HCl
solution, prepared by diluting analytical grade stock HCl (37%) with double-distilled water.
The rotating cylinder electrodes (RCE) employed for the corrosion testing were machined
from an API X60 cylindrical steel pipe with elemental composition as reported previously
[33]. The RCE steel coupons were prepared by abrading up to #800 grit size using waterproof
silicon carbide polishing papers, followed by thorough washing with double-distilled water,
ultrasonic cleaning in acetone [34] and drying with warm air.

123
Eur. Phys. J. Plus (2020) 135:129 Page 3 of 29 129

(a)

(b)

Fig. 1 Molecular structure of a 2-(2-Bromophenyl)-1H-benzimidazole (BHB) and b 2-(2-Bromophenyl)-1-


methyl-1H-benzimidazole (BMB)

2.2 Electrochemical measurements

The electrochemical measurements were conducted in a conventional 250-ml three-electrode


cell at room temperature and under atmospheric condition. The Ag/AgCl (3 M KCl) served
as the reference electrode, a graphite rod was used as counter electrode, and the RCE X60
steel coupon (surface area of 3.14 cm2 and thickness 0.2 cm) functioned as the working
electrode with an exposed volume of 0.628 cm3 . The electrochemical measurements were
performed using rotation speed of 200 and 1000 using the Gamry speed rotator (model RDE
710 Rotating Electrode). Given the surface area of the working electrode, the rotation speeds
of 200 and 1000 rpm are equivalent to Reynolds number (Re) of 2107 and 10,508 and wall
shear stress (τ ) of 0.176 and 2.705 Pa, respectively, according to Eqs. (1) and (2) [32, 35].
The Re value at 200 rpm indicates a fluid system in transition from laminar to turbulent
flow, while the Re value at 1000 rpm indicates fluid under real turbulence [32]. Throughout
this work, therefore, the rotation speed at 200 rpm is regarded as laminar hydrodynamic
condition, while the 1000 rpm speed is regarded as turbulent hydrodynamic condition.
ρudcylinder
Re  (1)
μ
τ  0.0791ρRe−0.3 u 2 (2)
The electrochemical measurements were conducted after allowing the working electrode
to attain an open-circuit potential (E corr ), as shown in Fig. 2. A quick glance reveals the
tendency for BMB to shift OCP more positively than BHB at the lowest concentration con-
sidered (300 ppm). The effect is attributed to the methyl group in BMB which enhances the
hydrophobic property of the benzimidazole. The EIS was acquired potentiostatically at the
E corr within the frequency range of 100 kHz to 0.1 Hz using a voltage amplitude of 10 mV.
The LPR measurement was performed within ± 10 mV/Ecorr using a scan rate of 0.167 mV/s.
For the PDP measurement, a potential range of ± 300 mV/E corr was applied with a scan rate

123
129 Page 4 of 29 Eur. Phys. J. Plus (2020) 135:129

(a)
-0.408

-0.416

Potential/Ag/AgCl (V) -0.424

-0.432

-0.440
Blank
300 ppm
-0.448
600 ppm
1200 ppm
-0.456
0 300 600 900 1200 1500 1800
Time (s)
(b) -0.404

-0.408

-0.412
Potential/Ag/AgCl (V)

-0.416

-0.420

-0.424
Blank
-0.428 300 ppm
600 ppm
1200 ppm
-0.432
0 300 600 900 1200 1500 1800
Time (s)

Fig. 2 Variation of potential with time for the attainment of open-circuit potential for X60 steel during corrosion
in 1 M HCl solution without and with different concentrations of a BHB and b BMB

of 0.167 mV/s. The electrochemical measurements were all performed using a Gamry 600
Potentiostat/Galvanostat/ZRA, and the electrochemical data were analyzed using Echem
Analyst 6.0 software package.

2.3 Surface characterization

Surface characterization of the steel surface was conducted after the steel corrosion in the 1 M
HCl solution without and with the maximum concentration of each inhibitor. The ATR-IR
spectra were recorded within the range of 400–4000 cm−1 using an IR reflectance spec-
trophotometer (PerkinElmer, Spectrum One, universal ATR attachment with a diamond and
ZnSe crystal, The Netherlands). A scanning electron microscope (SEM JEOL JSM-6610
LV) was operated under an acceleration voltage of 20 kV and irradiation current of 10 μA in
order to acquire the SEM surface images of the steel after corrosion. The XPS spectra were

123
Eur. Phys. J. Plus (2020) 135:129 Page 5 of 29 129

recorded in vacuum (below 1 × 10−8 Torr) using Al Kα radiation (1486.6 eV) as excitation
source. The takeoff angle (θ ) of the emitted photoelectrons was adjusted to 45° with respect
to the surface. A Thermo Scientific ESCALAB-250Xi spectrometer was used for the XPS
characterization.

2.4 Theoretical calculations

Quantum chemical calculations and the molecular geometry optimization were performed
by using the Gaussian 16 package program [36]. DFT calculations for the two benzimidazole
derivatives were conducted using Becke’s three-parameter hybrid function combined with the
Lee–Yang–Parr correlation function (B3LYP) using the cc-pVTZ basis set [37]. Gauss View
6 program was employed to visualize the computed outputs [38]. Frequency calculation was
employed to confirm the stationary points as the minima in energy and achieve the relevant
zero-point energy. The solvent effect on the molecular structures was calculated using the self-
consistent reaction field theory based on integral equation formalism polarizable continuum
model (IEF-PCM). In this model, the water molecules are treated as an expanse of dielectric
media and the inhibitors as a trapped molecule in a cavity surrounded by water molecules
[39]. The energy of the lowest unoccupied molecular orbital (E LUMO ), the energy of the
highest occupied molecular orbital (E HOMO ) and energy gap (E) for the inhibitors were
calculated based on the values of the total electronic energy upon neutral (E) and addition
(E N+1 ) or removal (E N−1 ) of an electron using the following equations:
 
E HOMO  − E N −1 − E N (3)

E LUMO  −[E N − E N +1 ] (4)

E  [E LUMO − E HOMO ] (5)

Monte Carlo simulations were used to investigate the adsorption behavior of the two
benzimidazole derivatives on Fe (110) and Fe2 O3 (001) surfaces, since the XPS analyses
revealed the presence of Fe and Fe2 O3 crystal phases. Adsorption locator module using the
Metropolis Monte Carlo simulations methodology as implemented in the Material Studio 8.0
software from BIOVIA–Accelrys Inc. USA was used for the simulation. Fe (11 0) is a densely
packed surface and has the most stabilization, so we choose Fe (11 0) surface to simulate the
adsorption process [40, 41]. The modeled surface composed of five layers of Fe (1 1 0) crystal
plane and three layers for the Fe2 O3 (001) crystal plane, respectively. The bottom four layer
atoms of the Fe (110) and Fe2 O3 (001) surface were fixed, while the top layer was allowed
to interact with the molecules investigated. A vacuum slab of 30 Å was constructed with
periodic boundary conditions on top of both Fe (110) and Fe2 O3 (001) crystal planes. The
simulations were carried out in a simulation box (18.11, 18.11, 38.11) Å and (52.57, 52.57,
70.57) Å for Fe (110) and Fe2 O3 (001) surfaces, respectively. Layers were constructed such
that each contains one inhibitor molecule and 200 water molecules. The Monte Carlo search
was adopted to compute the low configuration adsorption energy of the interaction between
the inhibitor molecules and the two surfaces in the aqueous phase. For the whole simulation
procedure, the COMPASS (Condensed-phase Optimized Molecular Potentials for Atomistic
Simulation Studies) forcefield was used to optimize the structures of all component of the
system of interest [42]. Ewald summation technique [43] was used to calculate electrostatic
interactions. The accuracy of the simulation was set to 1 × 10−5 kcal/mol. Van der Waals
interaction was calculated by atom-based summation approach. The cutoff distance, spline
width and buffer width were set at 18.5, 1 and 0.5 Å, respectively.

123
129 Page 6 of 29 Eur. Phys. J. Plus (2020) 135:129

(a) 240 (b) 2.5


Blank Blank
300 ppm 300 ppm
200 600 ppm 2.0 600 ppm
1200 ppm 1200 ppm
160 1.5

cm2)
cm-2)

120 1.0
9.931 Hz
-Zim (

49.87 Hz 2.504 Hz
80 0.5

Log
40 0.0

0 -0.5
-1 0 1 2 3 4 5
0 40 80 120 160 200 240 10 10 10 10 10 10 10

Zr ( cm-2) Log. Frequency (Hz)

(c) 160 (d) 2.5


Blank Blank
300 ppm 300 ppm
600 ppm 2.0 600 ppm
120 1200 ppm 1200 ppm

1.5
cm2)
cm-2)

80 1.0
9.931 Hz
-Zim (

0.5
Log

158.4 Hz 5.008 Hz
40
0.0

0 -0.5
-1 0 1 2 3 4 5
0 40 80 120 160 10 10 10 10 10 10 10
Log. Frequency (Hz)
Zr ( cm-2)

Fig. 3 Impedance plots for X60 steel during corrosion in 1 M HCl solution without and with different BHB
concentrations expressed in a Nyquist and b absolute impedance formats under laminar flow, and c Nyquist
and d absolute impedance plots under turbulent flow

3 Results and discussion

3.1 Electrochemical impedance spectroscopy (EIS)

The EIS measurement was conducted in order to understand the effect of BHB and BMB on
the kinetics and mechanism involved in X60 steel corrosion in the 1 M HCl solution under
the laminar and turbulent hydrodynamic conditions. The results are shown in Figs. 3 and 4
for BHB and BMB, respectively. The Nyquist plots shown in Fig. 3a, c (in the presence of
BHB) and Fig. 4a, c (in the presence of BMB) are capacitive loops with shapes of depressed
semicircles. They are characteristic shapes indicating charge transfer processes at the steel
surface due to corrosion attack and inhibitor adsorption [44]. Since the size of the Nyquist
semicircle is a measure of the corrosion resistance [45], it implies, therefore, that BHB and
BMB show significant corrosion resistance for X60 steel in the acid solution. Increasing the
concentration of the corrosion inhibitors up to 1200 ppm further provides more protection
for the steel from the acid corrosion.
The nitrogen atoms held within the imidazole ring donate lone pair of electrons which
interact with the empty d-orbitals of the iron (Fe) atoms on the steel surface [31]. In addition,
the pi-electrons from the aromatic rings as well as unpaired electrons from the electronegative
bromine substituent can interact with the Fe atoms via a donor–acceptor mechanism. These

123
Eur. Phys. J. Plus (2020) 135:129 Page 7 of 29 129

(a) 250 (b) 2.5


Blank Blank
300 ppm 300 ppm
600 ppm 2.0 600 ppm
200
1200 ppm 1200 ppm

1.5

cm2)
150
cm-2)

6.317 Hz 1.0
-Zim (

100
39.72 Hz 1.585 Hz 0.5

Log
50
0.0

0 -0.5
-1 0 1 2 3 4 5
0 50 100 150 200 250 10 10 10 10 10 10 10

Zr ( cm-2) Log. Frequency (Hz)

(c) 180 (d) 2.5


Blank Blank
300 ppm 300 ppm
150 600 ppm 2.0 600 ppm
1200 ppm 1200 ppm
120 1.5
cm2)
-Zim ( cm-2)

90 1.585 Hz 1.0

792.4 mHz
60 7.945 Hz 0.5
Log

30 0.0

0 -0.5
0 30 60 90 120 150 180 10
-1 0
10 10
1
10
2
10
3
10
4 5
10

Zr ( cm-2) Log. Frequency (Hz)

Fig. 4 Impedance plots for X60 steel during corrosion in 1 M HCl solution without and with different BMB
concentrations expressed in a Nyquist and b absolute impedance formats under laminar flow, and c Nyquist
and d absolute impedance plots under turbulent flow

processes trigger the inhibitor adsorption which leads to the formation a protective film that
shields the steel surface away from the corrosion agents in the acid solution. Nevertheless, the
steel inhibited by BMB exhibits greater corrosion resistance than BHB. This must, therefore,
be attributed to the methyl (CH3 ) group attached to the bromobenzimidazole moiety, because
alkyl substituents are well-known electron-donating groups which can activate the aromatic
ring by increasing the electron density through an inductive effect [46].
Meanwhile, the corrosion resistance provided by either BHB or BMB clearly decreases
under turbulent condition. The Nyquist semicircles are seen to decrease in their sizes. The
values of absolute impedance are also reduced due to this transition, according to Fig. 3b,
d (in the presence of BHB) and Fig. 4b, d (in the presence of BMB). It can be reasoned
that the mass transfer of corrosion agents like H+ and Cl− to the steel surface is increased,
and the increase in wall shear stress from 0.176 to 2.705 Pa also facilitates detachment of
the adsorbed corrosion inhibitor molecules from the steel surface so that active sites are
further exposed for the corrosion attack. Yet, the steel inhibited by BMB still exhibits higher
corrosion resistance than in BHB under the turbulent flow condition. This, therefore, indicates
that BMB possesses stronger attachment and surface coverage on the steel surface.
The electrical parameters for the EIS result were obtained by adopting the equivalent
circuit models in Fig. 5a (for the uninhibited steel) and Fig. 5b (with the investigated corrosion
inhibitors). The equivalent circuit model analysis gave the least values of Chi-square (χ 2 )

123
129 Page 8 of 29 Eur. Phys. J. Plus (2020) 135:129

(c) 100 (d) 250


Experimental Plot Experimental Plot
Fitting Plot Fitting Plot
80 200

150

-2)
-Zim (Ω cm-2)

60

-Zim (Ω cm
40 100

20 50

0
0
0 50 100 150 200 250
0 20 40 60 80 100
Zr (Ω cm-2)
Zr (Ω cm-2)

Fig. 5 Equivalent circuit model for X60 steel during hydrodynamic corrosion in 1 M HCl solution a without
and b with BHB and BMB inhibitors, and corresponding fitting plots c without and d with BHB and BMB
inhibitors

and associated errors after fitting with Echem Analyst 6.0 software. Similar equivalent circuit
models were adopted for the inhibition of acid corrosion of steel in a different report [47].
Based on the model, the corrosion of the uninhibited steel surface under laminar and
turbulent conditions is controlled by a solution resistance (Rs ) between the steel surface and
the reference electrode, a double-layer capacitance of charges assembled at the steel–acid
solution interface (expressed as the constant phase element, CPEdl ) and a resistance to charge
transfer across this interface (Rct ). The Rct value is a measure of the corrosion resistance of the
metal in the given solution. On the other hand, the adsorption of BHB and BMB molecules
in the inhibited solution modifies the corrosion mechanism by initiating an inhibitor film
capacitance (CPEf ) and film resistance (Rf ). In the equivalent model, the constant phase
element (CPE) is used in place of double-layer capacitance (C dl ) in order to correct for
surface imperfections of the electrode due to the corrosion and/or inhibitor adsorption [48].
However, C dl provides a better validation of an equivalent circuit model and was deduced
from the CPE using the following equation [49]:

Cdl  Yo (2π f max )n−1 , (6)

where Y o represents the magnitude of the CPEdl , n (− 1 ≤ n ≤ 1) is the CPEdl power, and
f max is the frequency (Hz) at which the imaginary impedance component in the Nyquist plot
is maximal.
The electrical parameters deduced from the equivalent circuit models are detailed in
Tables 1 and 2 without and with BHB and BMB, respectively. The results show higher
Rct values for the steel in the presence of the inhibitors under both hydrodynamic conditions
studied. The increased Rct values coincide with decrease in the C dl values and indicate that,

123
Eur. Phys. J. Plus

Table 1 Impedance parameters obtained for X60 steel during corrosion in 1 M HCl solution without and with BHB inhibitor under laminar and turbulent flow

Rotation speed Conc. (ppm) Rs ( cm2 ) CPEdl C dl (μF cm−2 ) Rct ( cm2 ) CPEf Rf ( cm2 ) (Rp  Rf + Rct ) (χ 2 ) (× 10−3 ) IE (%)
(2020) 135:129

(rpm) ( cm2 )
Y o (μF cm−2 α1 Y o (μF cm−2 α2
sn−1 ) sn−1 )

200 Blank 0.336 ± 0.02 407 ± 0.6 0.830 ± 0.01 201.9 ± 12.0 76.90 ± 0.04 – – – 76.90 ± 0.04 4.27 ± 0.10 –
300 0.349 ± 0.01 391 ± 0.2 0.619 ± 0.01 135.6 ± 8.8 107.10 ± 0.2 31 ± 0.6 0.919 ± 0.01 1.96 ± 0.01 109.01 ± 0.91 1.03 ± 0.01 29.5 ± 0.85
600 0.483 ± 0.01 380 ± 0.2 0.657 ± 0.01 92.2 ± 5.4 145.00 ± 0.3 30 ± 0.2 0.994 ± 0.02 2.00 ± 0.03 147.00 ± 0.81 0.216 ± 0.04 47.7 ± 0.42
1200 0.550 ± 0.01 343 ± 0.3 0.656 ± 0.01 82.8 ± 4.9 214.60 ± 0.3 23 ± 0.2 0.998 ± 0.01 2.03 ± 0.01 216.63 ± 0.79 0.493 ± 0.01 64.5 ± 0.14
1000 Blank 0.314 ± 0.01 544 ± 0.4 0.837 ± 0.01 260.9 ± 13.0 26.70 ± 0.6 – – – 26.77 ± 0.79 1.058 ± 0.04 –
300 0.411 ± 0.01 452 ± 0.6 0.744 ± 0.01 161.4 ± 10.1 58.53 ± 0.4 77 ± 0.6 0.998 ± 0.01 1.79 ± 0.01 60.32 ± 0.32 1.361 ± 0.06 55.7 ± 1.56
600 0.445 ± 0.01 393 ± 0.4 0.718 ± 0.01 122.6 ± 7.4 93.62 ± 0.4 62 ± 0.4 0.998 ± 0.01 1.85 ± 0.01 95.47 ± 0.91 0.281 ± 0.01 72.0 ± 0.60
1200 0.501 ± 0.01 360 ± 0.4 0.707 ± 0.01 106.5 ± 3.0 142.20 ± 0.6 49 ± 0.2 0.972 ± 0.01 2.27 ± 0.02 144.47 ± 1.11 0.668 ± 0.02 81.6 ± 0.89
Page 9 of 29
129

123
129

123
Page 10 of 29

Table 2 Impedance parameters obtained for X60 steel during corrosion in 1 M HCl solution without and with BMB inhibitor under laminar and turbulent flow

Rotation speed Conc. (ppm) Rs ( cm2 ) CPEdl C dl (μF cm−2 ) Rct ( cm2 ) CPEf Rf ( cm2 ) (Rp  Rf + Rct ) (χ 2 ) (× 10−3 ) IE (%)
(rpm) ( cm2 )
Y o (μF cm−2 α1 Y o (μF cm−2 α2
sn−1 ) sn−1 )

200 Blank 0.336 ± 0.02 407 ± 0.6 0.830 ± 0.01 201.9 ± 12.0 76.90 ± 0.04 – – – 76.90 ± 0.04 4.27 ± 0.10 –
300 0.450 ± 0.01 372 ± 0.2 0.618 ± 0.01 78.9 ± 1.6 163.20 ± 1.0 27 ± 0.6 0.919 ± 0.01 2.70 ± 0.01 165.90 ± 0.47 1.42 ± 0.01 53.7 ± 0.28
600 0.481 ± 0.01 358 ± 0.6 0.600 ± 0.01 68.6 ± 4.1 185.00 ± 0.6 30 ± 0.2 0.994 ± 0.02 2.95 ± 0.01 188.22 ± 0.82 0.216 ± 0.01 59.2 ± 0.28
1200 0.584 ± 0.01 265 ± 0.3 0.619 ± 0.01 65.3 ± 5.2 260.00 ± 0.3 19 ± 0.2 0.998 ± 0.02 3.04 ± 0.02 263.04 ± 0.80 0.400 ± 0.10 70.8 ± 0.07
1000 Blank 0.314 ± 0.01 544 ± 0.4 0.837 ± 0.01 260.9 ± 13.0 26.70 ± 0.6 – – – 26.70 ± 0.6 1.058 ± 0.04 –
300 0.357 ± 0.01 440 ± 1.2 0.736 ± 0.01 123.2 ± 8.9 87.56 ± 0.2 29 ± 0.2 0.968 ± 0.02 2.05 ± 0.01 89.61 ± 0.65 0.685 ± 0.04 70.2 ± 0.99
600 0.414 ± 0.01 366 ± 0.4 0.700 ± 0.01 115.1 ± 11.6 123.20 ± 0.4 45 ± 0.6 0.985 ± 0.01 2.01 ± 0.01 125.21 ± 0.91 1.011 ± 0.06 78.8 ± 0.56
1200 0.597 ± 0.01 261 ± 1.0 0.698 ± 0.01 86.2 ± 5.9 188.50 ± 0.2 35 ± 0.2 0.896 ± 0.01 2.86 ± 0.02 191.36 ± 0.66 1.029 ± 0.02 86.0 ± 0.56
Eur. Phys. J. Plus
(2020) 135:129
Eur. Phys. J. Plus (2020) 135:129 Page 11 of 29 129

by adsorbing on the steel surface, the inhibitor molecules lower the dielectric properties at the
steel–acid solution interface. Increasing the inhibitor concentration, for a given hydrodynamic
condition, further increases the R values and lowers the C values, implying greater protection
of the steel surface. Nevertheless, the steel exhibits greater R values and lower C values in
the presence of BMB than BHB. Assuming a total resistance (Rp ) defined as Rp  Rct + Rf
[48], then, the values of inhibition efficiency provided by the BHB and BMB (Tables 1, 2)
were calculated according to the following equation:
Rp(inhibited) − Rp(uninhibited)
%IEEIS  × 100, (7)
Rp(inhibited)
where Rp(uninhibited) and Rp(inhibited) indicate, respectively, the charge transfer resistance
recorded in the absence and presence of an inhibitor.

3.2 Polarization measurements

Polarization measurements employed in the present work include linear polarization resis-
tance (LPR) and potentiodynamic polarization (PDP). These are direct current methods which
measure a resultant current response to external potential applied to displace the corroding
metal away from its open-circuit potential (E corr ). The application of relatively safe potential
around ± 10 mV away from the E corr makes the LPR method fast, nondestructive and highly
sensitive to instantaneous corrosion phenomena. For such small overpotential with Tafel con-
stants βa and βc , the most important corrosion parameter derived from the LPR measurement
is the polarization resistance (Rp ) which varies inversely with the instantaneous corrosion
current (icorr ), according to the following equation [50]:
1 βa βc
Rp  (8)
2.303i corr βa + βc
By assuming anodic (βa ) and cathodic (βc ) Tafel constants of 120 mV/dec [50], the values
of Rp determined from the LPR experiments for the X60 steel in 1 M HCl solution with-
out and with different inhibitor concentrations under laminar and turbulent hydrodynamic
conditions are provided in Tables 3 and 4, for BHB and BMB, respectively. In the presence
of the corrosion inhibitors, the steel exhibits higher values of Rp , which become more sig-
nificant as inhibitor concentration increases. Maximum Rp can be observed under laminar
condition, increasing from 70.38 ± 0.93  cm2 in the blank solution to 193.80 ± 0.63  cm2
and 255.2 ± 0.63  cm2 in the presence of 1200 ppm BHB and BMB, respectively. The Rp
values, however, decrease under the turbulent condition. The decrease in Rp under turbulent
condition is consistent with the EIS result and strongly supports that greater supply of H+
and Cl− at the steel–solution interface promotes corrosion attack during turbulent condi-
tion. Nevertheless, the presence of BHB and BMB can lower the corrosion attack on the
steel surface and provide inhibition efficiency value of 80.50% and 86.28% under turbulent
hydrodynamic condition, respectively. Inhibition efficiency was calculated according to the
following equation:
Rp(uninhibited)
%IELPR  1 − × 100%, (9)
Rp(inhibited)
where Rp(uninhibited) and Rp(inhibited) indicate, respectively, the polarization resistance recorded
in the absence and presence of an inhibitor.
Potentiodynamic polarization (PDP) was employed to understand how BHB and BMB
studied under hydrodynamic conditions influence the cathodic and anodic half-reactions
during the X60 steel corrosion in 1 M HCl solution. The PDP results obtained are displayed as

123
129

123
Page 12 of 29

Table 3 Polarization parameters obtained for X60 steel during corrosion in 1 M HCl solution without and with BHB inhibitor under laminar and turbulent flow

Rotation speed (rpm) Conc. (ppm) PP LPR

E corr (mV/Ag/AgCl) icorr (μA/cm2 ) β a (mV/Dec.) IE (%) E corr (mV/Ag/AgCl) Rp ( cm2 ) IE (%)

200 Blank − 422 104.70 31 – − 424 ± 1.14 70.38 ± 0.93 –


300 − 434 67.42 35 35.60 − 432 ± 0.93 99.40 ± 0.91 29.25 ± 0.71
600 − 426 43.65 36 58.31 − 426 ± 1.14 143.55 ± 1.14 51.00 ± 0.20
1200 − 414 39.20 30 62.56 − 416 ± 1.20 193.80 ± 0.63 63.70 ± 0.77
1000 Blank − 407 303.60 31 – − 407 ± 0.91 24.23 ± 0.19 –
300 − 421 130.30 36 57.08 − 422 ± 0.20 57.25 ± 0.80 57.70 ± 0.63
600 − 422 83.93 36 72.36 − 422 ± 1.71 88.92 ± 0.91 72.80 ± 0.32
1200 − 450 64.94 35 78.61 − 450 ± 1.41 124.60 ± 0.90 80.50 ± 0.32
Eur. Phys. J. Plus
(2020) 135:129
Eur. Phys. J. Plus

Table 4 Polarization parameters obtained for X60 steel during corrosion in 1 M HCl solution without and with BMB inhibitor under laminar and turbulent flow

Rotation speed (rpm) Conc. (ppm) PP LPR


(2020) 135:129

E corr (mV/Ag/AgCl) icorr (μA/cm2 ) β a (mV/Dec.) IE (%) E corr (mV/Ag/AgCl) Rp ( cm2 ) IE (%)

200 Blank − 421 104.70 31 – − 424 ± 1.14 70.38 ± 0.93 –


300 − 423 45.37 33 56.67 − 422 ± 1.14 162.6 ± 0.91 56.71 ± 0.71
600 − 408 33.85 33 67.67 − 410 ± 1.73 212.2 ± 1.14 66.83 ± 0.20
1200 − 410 28.10 32 73.16 − 410 ± 1.41 255.2 ± 0.63 72.42 ± 0.77
1000 Blank − 407 303.60 30 – − 407 ± 0.91 24.23 ± 0.19 –
300 − 401 70.97 32 76.62 − 401 ± 1.73 92.1 ± 1.41 73.69 ± 0.63
600 − 403 56.65 30 81.34 − 405 ± 0.91 125.5 ± 1.41 83.08 ± 0.32
1200 − 394 38.23 28 87.40 − 396 ± 0.91 176.60 ± 0.91 86.28 ± 0.32
Page 13 of 29
129

123
129 Page 14 of 29 Eur. Phys. J. Plus (2020) 135:129

(a) -0.1
Blank
300 ppm
-0.2 600 ppm
1200 ppm

Potential/Ag/AgCl (V)
-0.3

-0.4

-0.5

-0.6

-0.7

-0.8
-7 -6 -5 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10

Current Density (A cm-2)


(b) -0.1
Blank
300 ppm
-0.2
600 ppm
1200 ppm
Potential/Ag/AgCl (V)

-0.3

-0.4

-0.5

-0.6

-0.7

-0.8
-7 -6 -5 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10

Current Density (A cm-2)

Fig. 6 Potentiodynamic polarization plots for X60 steel during hydrodynamic corrosion in 1 M HCl solution
without and with different BHB concentrations under a laminar and b turbulent flow

polarization curves in Figs. 6 and 7, respectively, for the different concentrations of BHB and
BMB. The derived electrochemical parameters are given in Tables 3 and 4. It can be observed
that, at concentrations below 1200 ppm, neither BHB nor BMB has significant influence on the
E corr of the steel in the acid solution both under laminar condition (Figs. 6a, 7a) and under
turbulent condition (Figs. 6b, 7b). However, at 1200 ppm, and especially under turbulent
condition, Fig. 6b shows that the BHB significantly shifts the E corr of the steel to less noble
direction from − 407 to − 450 mV, while the shift was significant to more noble potential
(− 394 mV) in the presence of BMB, based on Fig. 7b.
Considering the shapes of the polarization curves, the BHB and BMB obviously lower
both cathodic and anodic current densities, indicating that both inhibitors are mixed type,
and this mechanism is not disturbed by hydrodynamics. However, the relative shift of E corr
value, especially under turbulence and 1200 ppm inhibitor concentration, shows BHB as a

123
Eur. Phys. J. Plus (2020) 135:129 Page 15 of 29 129

(a) -0.1
Blank
300 ppm
-0.2
600 ppm
1200 ppm
Potential/Ag/AgCl (V) -0.3

-0.4

-0.5

-0.6

-0.7

-0.8
-7 -6 -5 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10

Current Density (A cm-2)


(b) -0.1
Blank
300 ppm
-0.2 600 ppm
1200 ppm
Potential/Ag/AgCl (V)

-0.3

-0.4

-0.5

-0.6

-0.7

-0.8
-7 -6 -5 -4 -3 -2 -1 0
10 10 10 10 10 10 10 10

Current Density (A cm-2)

Fig. 7 Potentiodynamic polarization plots for X60 steel during hydrodynamic corrosion in 1 M HCl solution
without and with different BMB concentrations under a laminar and b turbulent flow

more cathodic inhibitor and BMB as more anodic. On a closer observation, the cathodic arm
of the polarization curve for the uninhibited steel becomes more diffusion-controlled under
turbulent condition, relative to the laminar condition. The most predominant cathodic reaction
for steel during acid corrosion is the reduction of H+ ion into hydrogen gas: 2H+(aq) + 2e− →
H2(g) . Thus, turbulence definitely increases the diffusion of H+ from the solution to the steel
surface, which increases the corrosion rate. However, it is the presence of BMB (rather than
BHB) that suppresses this diffusion phenomenon under turbulent condition more significantly
(compare Figs. 6b, 7b). This must be due to a stronger adsorption and larger surface coverage
by the BMB which effectively deactivates cathodic sites for H+ reduction. The inhibition
efficiency was calculated from the more significant results as follows:

123
129 Page 16 of 29 Eur. Phys. J. Plus (2020) 135:129

i corr(inhibited)
%IEPDP  1 − × 100% (10)
i corr(uninhibited)

where icorr(uninhibited) and icorr(inhibited) indicate, respectively, the corrosion current density
recorded in the absence and presence of an inhibitor. The values of the polarization parameters
clearly confirm the inhibitive properties of BHB and BMB for the steel. Under turbulence,
for instance, the maximum inhibitor concentrations decrease the icorr value of the X60 steel
from 303.60 μA/cm2 in the blank to 64.94 μA/cm2 and 38.23 μA/cm2 in the presence of
BHB and BMB, respectively. These correspond to % IE values of 78.61% and 87.40% in the
presence of BHB and BMB, respectively. The PDP results are also consistent with the EIS
and LPR results based on the trends in inhibition efficiencies.

3.3 Surface characterization

3.3.1 ATR-IR characterization

The electrochemical measurements show that the addition of up to 1200 ppm of BHB and
BMB significantly inhibited the corrosion of X60 steel in 1 M HCl solution under laminar and
turbulent conditions. Such corrosion inhibition is usually attributed to the inhibitor adsorp-
tion on the steel surface. In order to confirm this, ATR-IR characterization was employed
to characterize the surface of the X60 steel after corrosion in the acid solution without
and with BHB and BMB. For this purpose, the X60 steel was immersed in 1 M HCl
acid solution without and with 1200 ppm of BHB and BMB under laminar condition for
24 h.
Figure 8a shows the ATR-IR spectra for pure BHB and for X60 steel surface after the
corrosion in 1 M HCl solution without and with 1200 ppm BHB. Figure 8b shows the ATR-IR
spectra for pure BMB and for X60 steel surface after the corrosion in 1 M HCl solution without
and with 1200 ppm BMB. The pure BHB and BMB show typical bromophenyl benzimidazole
peaks due to N–H stretching modes (between 3000 and 3500 cm−1 ) with corresponding N–H
in-plane bending modes around 1600 cm−1 [51]. The C=C stretching peak corresponding to
the benzene and imidazole rings occurs between 1500 and 1470 cm−1 , while the imidazole
ring bending can be seen as the somewhat triplet peaks observed around 750–780 cm−1
[51]. In addition, the C–Br stretching which usually absorbs in the interval of 650–395 cm−1
in most aromatic bromo compounds [52] can also be seen. After the steel corrosion in the
acid solution containing BHB and BMB, the bromophenyl benzimidazole peaks can still
be detected on the steel surface, as observed in Fig. 8a, b. This shows the presence of the
inhibitors on the steel surface and, hence, confirms that the inhibitors protect the X60 steel
from acid corrosion via an adsorption mechanism. On the contrary, such peaks cannot be
observed on the surface of the steel which corroded in the uninhibited acid solution. The
recurrent peak observed between 2250 and 2500 cm−1 is attributed to the residual peak from
isopropanol solvent with which the ATR-IR lens was cleaned prior to the characterization.
Nevertheless, the peak between 550 and 600 cm−1 corresponds to the presence of Fe–O bonds
[53] and is observed for the uninhibited and inhibited steel. Meanwhile, the absorption peaks
are stronger and more pronounced on the steel surface in contact with BMB and provide more
evidence that the BMB adsorbs more strongly with greater surface coverage, than BHB. The
stronger interaction between BMB and the steel surface must, therefore, be the reason that
the molecule provides a more significant corrosion inhibition for the acid corrosion of the
X60 steel, than BHB.

123
Eur. Phys. J. Plus (2020) 135:129 Page 17 of 29 129

(a) 120

Fe-O bond

100

N-H bending
Intensity (a.u)
C-H stretching

C-N stretching
80
C-Br stretching
N-H stretching
C=C stretching

Imidazole ring bending


60

BHB
X60 + 1M HCl
X60 + 1M HCl + 1200ppm BHB
40
4000 3500 3000 2500 2000 1500 1000 500

Wavenumber (cm-1)
(b) 120

Fe-O bond

100
Intensity (a.u)

C-H stretching
N-H bending
80 N-H stretching C-N stretching

C=C stretching
C-Br stretching
Imidazole ring bending
60
BMB
X60 + 1M HCl
X60 + 1M HCl + 1200ppm BMB
40
4000 3500 3000 2500 2000 1500 1000 500

Wavenumber (cm-1)

Fig. 8 ATR-IR spectra obtained for a pure BHB and X60 steel after hydrodynamic corrosion in 1 M HCl
solution without and with 1200 ppm BHB and b pure BMB and X60 steel after hydrodynamic corrosion in
1 M HCl solution without and with 1200 ppm BMB

3.3.2 SEM characterization

Figure 9a reveals that the surface of the uninhibited steel has suffered severe corrosion.
Several localized pits of different sizes are scattered all over the steel surface. These are
definitely the result of intense chloride attack from the solution. Pitting corrosion was also
reported for steel during corrosion in HCl solution [54]. It has been well established that
pitting corrosion usually occurs in the presence of aggressive anionic species, especially Cl−
ions, as a result of significant solubility of many metal cations in chloride solutions [55]. The
pitting corrosion triggers an increase in the supply of Fe atoms from the steel substrate toward
the pitting front in order to replenish the Fe atoms consumed by the pitting corrosion. This
will encourage further electromigration of Cl− ions into the pit so as to maintain the charge

123
129 Page 18 of 29 Eur. Phys. J. Plus (2020) 135:129

Fig. 9 SEM surface morphology of X60 steel after hydrodynamic corrosion in 1 M HCl solution a without
and with 1200 ppm of b BHB and c BMB

123
Eur. Phys. J. Plus (2020) 135:129 Page 19 of 29 129

neutrality and balance the charge associated with the Fe2+ concentration and eventually lead
to pit propagation. The extent of pitting degradation can be seen in the higher magnification
image inserted within Fig. 9a which reveals severely deteriorated pit interior.
Conversely, the surfaces of the inhibited steel are less attacked by the pitting corrosion.
Figure 9b, c reveals the SEM micrographs of the steel surface after corrosion in the HCl
solution containing 1200 ppm BHB and BMB, respectively. From a low magnification per-
spective, it can be noticed that the adsorption of the corrosion inhibitors plays an important
role in ameliorating the tendency for pit formation. Compared with the uninhibited steel
surface, the pits formed on the inhibited steel surface are fewer, smaller in size and shallower
in depth. This indicates that once adsorbed, the BHB and BMB form an inhibitor film which
prevents the aggressive Cl− ions from making contact with the steel surface. In acid solutions
containing chloride ions, the adsorption of benzimidazole derivatives has been attributed to
a combined effect of protonation of the imidazole nitrogen atom and electrostatic interac-
tion with primarily adsorbed chloride ions on the steel surface [56]. Nonetheless, the BMB
contains the more hydrophobic methyl group which provides greater isolation of the steel
surface from the pitting agents. This is the reason that the surface of the steel surface inhibited
by BMB presents smaller and shallower pits (Fig. 9c) compared with the larger and deeper
pits on the steel inhibited by BHB (Fig. 9b), as further buttressed by comparing the high
magnification images inserted within the respective SEM micrographs.

3.3.3 XPS characterization

Deeper insight into the interaction between the benzimidazole derivatives (BHB and BMB)
and the X60 steel during hydrodynamic corrosion in 1 M HCl solution was acquired using
XPS characterization. Figure 10 shows the peak deconvolution of Fe 2p, O 1s, C 1s and
Cl 2p spectra obtained for the X60 steel after 24 h corrosion in 1 M HCl solution without
inhibitor. The C 1s peak, Fig. 10a, is observed around 284.8 eV which is consistent with the
presence of carbon [57, 58]. Figure 10b shows the Fe 2p3/2 peaks at binding energies 706.9,
711 and 719.9 eV, as well as Fe 2p1/2 peak around 724.8 eV. The peak around 706.9 eV is
characteristic of unreacted metallic Fe0 , while the peak at 711 eV with distinctive shake-up
satellite peaks around 719.9 eV (Fe 2p3/2 ) and 724.8 eV (Fe 2p1/2 ) corresponds to Fe2 O3
species [59]. The corresponding O 1s spectrum in Fig. 10c shows peaks with binding
energies 529.8 and 532.8 eV which are typical of O2− species (as in Fe2 O3 ) and adsorbed
water molecules (H2 O(ads) ), respectively [60, 61]. Furthermore, the spectrum in Fig. 10d
shows Cl 2p3/2 peak around 198.7 eV and Cl 2p1/2 peak around 200.2 eV, both of which
are indicative of chloride (Cl− ) species. In the presence of 1200 ppm of BHB (Fig. 11) and
BMB (Fig. 12), the binding energies of Fe, O and Cl peaks are similar to the observation for
the uninhibited steel. The C 1s peak, however, occurs around 285.3 eV and can be attributed
to C–N bond [62]. The N 1s peaks observed in the presence of both inhibitors appear as
triplet peaks around 398.7, 399.6 and 400.9 eV and, respectively, represent the presence of
C=N–H, C–N–H and C=N–H+ bonds [62].
The observations from XPS characterization can be applied to explain the mechanism by
which the BHB and BMB inhibit the corrosion of the X60 steel in the acid solution under
the different hydrodynamic conditions. The dissolution (corrosion) of the steel in the acid
solution is initiated by the surface adsorption of water molecules carrying the aggressive acid
ions, Cl− and H+ . In the absence of the corrosion inhibitors, more water molecules with the
corrosion agents primarily adsorb on the steel surface. This is evidenced by the higher peak
area and intensity of the O 1s peak at 532.8 eV and the Cl 2p peaks at 198.7 and 200.2 eV

123
129 Page 20 of 29 Eur. Phys. J. Plus (2020) 135:129

(a) 1600 (b) 14000


Fe 2p
C 1s
Carbon
12000 Fe2O3
Fe2O3(sat)
1400 Fe2O3(sat)

Intensity (count s )
10000
Intensity (count s )

-1
-1

Fe

8000
1200

6000

1000 4000

2000
800
288 287 286 285 284 283 282 730 725 720 715 710 705 700
Binding Energy (eV) Binding Energy (eV)

(c) 7000
O 1s
(d) 1800
Cl 2p
6000 -
1600 2p3/2 (Cl )

Intensity (Count s-1)


Intensity (count s )

5000
-1

H2O(ads) 2- -
O 1400 2p1/2 (Cl )

4000
1200

3000
1000
2000
800
1000

536 534 532 530 528 526 204 202 200 198 196

Binding Energy (eV) Binding Energy (eV)

Fig. 10 XPS spectra of X60 steel surface after hydrodynamic corrosion in 1 M HCl solution without inhibitor
showing a C 1s, b Fe 2p, c O 1s and d Cl 2p peaks

detected on the uninhibited steel surface, compared with the observation in the presence of
BHB and BMB.
Once adsorbed, the corrosive agents initiate anodic oxidation/dissolution (Fe0 → Fe2+ ) of
the steel surface, according to Eqs. (11–13) [63, 64]. During the 24 h corrosion under laminar
condition, the transition of Fe0 → Fe2+ → Fe3+ can occur based on Eq. (14). Eventually, the
Cl− ions incorporate within the Fe3+ (as Fe2 O3 ) layer and cause local dissolution that results
in serious pitting corrosion [65]. According to the point defect model [66], cationic vacancies
will be created within the corrosion layer and encourage the outward supply of more Fe atoms
from the steel matrix toward the pitting front in order to replenish the Fe atoms consumed by
the pitting process. This explains why more Fe0 at 706.9 eV are detected on the uninhibited
steel surface than the in the presence of inhibitor.
Fe + H2 O + Cl− ↔ [FeClOH]− +
(ads) + H + e

(11)

[FeClOH]−
(ads) ↔ [FeClOH](ads) + e

(12)

[FeClOH](ads) + H+ ↔ Fe2+ + Cl− + H2 O (13)


1
2Fe2+ + O2 + 2H+ → 2Fe3+ + H2 O (14)
2
In the inhibited acid solution, the adsorption of BHB and BMB films on the steel surface
is confirmed by the detection of the N 1s peaks and the shift of C 1s peak to higher binding
energy. Therefore, the nitrogen atoms in both benzimidazole derivatives are fundamentally

123
Eur. Phys. J. Plus (2020) 135:129 Page 21 of 29 129

(a) 5000 (b) 7000


C 1s
Fe 2p
Carbon Fe2O3(sat)
4000 6000
Fe2O3
Intensity (count s )
-1

Intensity (count s )
Fe2O3(sat)

-1
5000
3000

4000
2000
Fe
3000

1000
2000

0 1000
287 286 285 284 283 730 725 720 715 710 705 700
Binding Energy (eV) Binding Energy (eV)

(c) 1050
N 1s

1000
+ C-N-H
Intensity (count s )

C=N-H
-1

950 C=N-H

900

850

800
402 401 400 399 398 397
Binding Energy (eV)

(d) 8000
O 1s
(e) 1400 Cl 2p
7000
1300 -
2p3/2 (Cl )
2-
6000 O
H2O(ads)
Intensity (count s )

Intensity (count s )
-1

-1

1200
5000

4000 1100 2p1/2 (Cl-)

3000 1000

2000
900
1000
800
0
536 534 532 530 528 526 203 202 201 200 199 198 197 196

Binding Energy (eV) Binding Energy (eV)

Fig. 11 XPS spectra of X60 steel surface after hydrodynamic corrosion in 1 M HCl solution with 1200 ppm
BHB inhibitor showing a C 1s, b Fe 2p, c N 1s, d O 1s and e Cl 2p peaks

the active centers for interaction between the steel surface and the corrosion inhibitors.
The detection of C=N–H+ at 400.9 eV also confirms that the benzimidazole derivatives
are protonated in the acid solution. Meanwhile, the adsorption of the inhibitors reduces the
intensity of O 1s peak at 532.8 eV, relative to the blank solution. This confirms that the
incoming inhibitors displace primarily adsorbed water molecules from the steel surface, in
agreement with literature reports [67, 68]. The isolation of water molecules also implies
shielding the steel surface from partaking in the corrosion process; hence, the lower intensity
of Fe2 O3 peaks was detected on the surface of steel in the presence of BHB and BMB.
Nevertheless, the inhibitor layers will exhibit some porosity depending on the extent of their
adsorption and surface coverage. The steel surface inhibited by BMB inhibitor exhibits lower
intensity of H2 O(ads) , Cl− , Fe and Fe2 O3 peaks and, therefore, must be better covered by the

123
129 Page 22 of 29 Eur. Phys. J. Plus (2020) 135:129

5000 6000
C 1s Fe 2p
(a) Carbon
(b) Fe2O3(sat)
4000 5000
Fe2O3
Intensity (count s )

Intensity (count s )
Fe2O3(sat)
-1

-1
3000 4000

2000 3000

1000 Fe
2000

0 1000
287 286 285 284 283 730 725 720 715 710 705 700

2
4
3
7
8
9
5
0
1
+
C
=
-Nu
o
c
Its
-1 B
id
HE
n
r
g
(y
e
)V
Binding Energy (eV) Binding Energy (eV)

1050
N 1s
(c)
C-N-H
1000
+
I n te n s i ty (c o u n t s )
-1

C=N-H C=N-H

950

900

850
402 401 400 399 398 397

.u
(a
iy
s
te
8
3
4
5
6
0
1
H
O)In
B
d
E
r
g
V
-2 7
8
9
1
0
2
p
1 B
id
)-lCE
n
r
g
(y
e
V
.u
a
Its
Binding Energy (eV)
/3
2

6000 1200
O 1s Cl 2p
(d) O
2- (e) -
2p3/2 (Cl )
5000
1100
H2O(ads)
4000
I n te n si ty (a . u )

I n te n s i ty (a . u )

1000
3000 -
2p1/2 (Cl )
900
2000

1000 800

0 700
536 534 532 530 528 202 201 200 199 198 197
Binding Energy (eV) Binding Energy (eV)

Fig. 12 XPS spectra of X60 steel surface after hydrodynamic corrosion in 1 M HCl solution with 1200 ppm
BMB inhibitor showing a C 1s, b Fe 2p, c N 1s, d O 1s and e Cl 2p peaks

BMB than BHB. This phenomenon, attributed to the larger size and greater hydrophobicity of
the BMB courtesy of the CH3 group, therefore, explains why the BMB displays more effective
corrosion inhibition behavior for the X60 steel during corrosion in 1 M HCl solution under
hydrodynamic flow conditions.

3.4 Theoretical calculations

3.4.1 Quantum chemical calculation

Figure 13 shows the optimized structure as well as the HOMO and LUMO frontier molec-
ular orbital density distributions of the two benzimidazole derivatives. The E HOMO depicts

123
Eur. Phys. J. Plus (2020) 135:129 Page 23 of 29 129

Optimized
Inhibitor HOMO LUMO
Structure
(a) (b) (c)

BPBA

BPMA

Fig. 13 a Fully optimized geometry, b HOMO and c LUMO frontier molecular orbitals of BHB and BMB

electron-rich centers in the corrosion inhibitors which are responsible for electron dona-
tion to the Fe atoms in the X60 steel. The E LUMO describes sites on the inhibitor molecule
which accepts electrons back from the d-orbitals of the Fe atoms into the anti-bonding
orbitals. These energy parameters, including the energy gap, E  E LUMO − E HOMO
and the molecular dipole moment (μ), are important quantum chemical parameters which
describe the reactivity of the inhibitor molecules. They can also describe the mechanism
by which the corrosion inhibitors interact with the steel surface [69, 70]. In Fig. 13b, c,
it can be seen that the E HOMO and E LUMO orbitals in the BHB are spread over the benz-
imidazole ring and other parts of the molecule, whereas in the case of BMB, the E HOMO
is more concentrated around the benzimidazole ring. The E LUMO orbitals spread over the
entire molecule for the BMB inhibitor. Therefore, the benzimidazole ring in the BHB and
BMB provides the sites for interaction with the X60 steel during corrosion inhibition in 1 M
HCl solution. The quantum chemical descriptors derived for the BHB and BMB molecules
are presented in Table 5. It can be seen that both molecules show no significant differ-
ence in the values of their E HOMO , but the BMB shows more negative E LUMO values
which also imparts lower E values. This indicates that although both BHB and BMB
may have similar tendency to donate electrons to the steel surface, the BMB has greater
tendency to accept back electrons from the Fe atoms on the steel surface. This indicates
higher chemical reactivity. Such is important because, apart from donating electrons to the
metal atoms, an effective corrosion inhibitor should also be able to accept electrons from
the oxidized metal atom into its non-bonding orbital of the inhibitor molecule. Furthermore,
BMB exhibits lower dipole moment than the BHB, which implies increase in corrosion inhi-
bition property, as was emphasized elsewhere [71]. There is no consensus in the literature
on the importance of dipole moment to the inhibition process. Some authors have reported
that increase in dipole moment enhances inhibition efficiency, while other has reported the
reverse.

123
129 Page 24 of 29 Eur. Phys. J. Plus (2020) 135:129

Table 5 Quantum chemical Inhibitor E HOMO (eV) E LUMO (eV) E (eV) μ (Debye)
descriptors for BHB and BMB
calculated at B3LYP/cc-pvtz G BHB − 6.0249 − 2.0427 3.9822 6.4236
level in aqueous phase
BMB − 6.0855 − 3.0208 3.0647 5.4931

(a) (b)

(c) (d)

Fig. 14 Equilibrium adsorption configurations of a BHB, b BMB on Fe (110) surface and c BHB, d BMB on
Fe2 O3 (001) surfaces all in aqueous phase

3.4.2 Molecular simulations

The XPS result showed that the surface of the steel contained Fe and Fe2 O3 after hydrody-
namic corrosion in the 1 M HCl solution without and with the BHB and BMB. The mode of
adsorption and the corresponding energy involved in the adsorption were, therefore, inves-
tigated by modeling the optimized BHB and BMB molecules on Fe(110) and Fe2 O3 (001)
surfaces using the Monte Carlo simulation method. The top view and side view adsorption
configuration of the molecules on the different Fe surface can be seen in Fig. 14a–d. It can be
clearly seen that both BHB and BMB adopt parallel orientations which allow the molecules
to lie flat on the Fe surface. This is highly expected for molecules which exhibit effective cor-
rosion inhibition properties based on their surface coverage of the metal surface. The energy
values computed from the simulation are presented in Table 6. The result indicates highly
negative energy values for the corrosion inhibitor molecules on both surfaces. The adsorption
energies obtained for the molecules are more negative that the energies are calculated for
water molecules. This implies that the BHB and BMB are more energetically favorable to
displace water molecules from the steel surface during the corrosion inhibition process. This
result strongly correlates with the XPS result which revealed lower intensities of adsorbed
water molecules in the presence of the inhibitors. Similarly, the BMB exhibits more negative
adsorption energies than the BHB both on the Fe (110) and Fe2 O3 (001) planes. This is

123
Eur. Phys. J. Plus (2020) 135:129 Page 25 of 29 129

Table 6 Adsorption energy of Inhibitor Fe (110) Fe2 O3 (001)


inhibitors and water calculated
using the Monte Carlo simulation Adsorption Adsorption Adsorption Adsorption
for BHB and BMB on Fe (1 1 0) energy of energy of energy of energy of
and Fe2 O3 (001) surfaces (in inhibitor water inhibitor water
kJ mol−1 )
BHB − 646.45 − 60.25 − 187.89 − 20.46
BMB − 674.14 − 59.44 − 208.41 − 29.06

confirmation of the more effective corrosion inhibition behavior displayed by the BMB for
the X60 steel during hydrodynamic corrosion in naturally aerated 1 M HCl solution.

3.5 Mechanism of inhibition

According to the results obtained from this work, the corrosion of X60 steel in naturally aer-
ated 1 M HCl solution under hydrodynamic condition is initiated by the adsorption of water
molecules carrying H+ and Cl− ions, as confirmed by XPS characterization. Once adsorbed,
an electrochemical reaction will occur whereby H+ reduction dominates as the cathodic half-
reaction, while Fe oxidation into Fe2+ dominates as the anodic half-reaction. These electro-
chemical half-reactions occur more rapidly under a turbulent hydrodynamic condition which
supplies the H+ and Cl− ions more abundantly to the steel surface. The rapid supply of cor-
rosive ions translates into increased Fe dissolution (Fe2+ formation), i.e., increased corrosion
rate. It also implies increased attack by Cl− ions which lead to serious localized pitting corro-
sion. In the naturally aerated acid solution, the abundance of oxygen in the corrosion system
can convert the formed Fe2+ species into more stable Fe3+ species (in the form of Fe2 O3 ).
In the acid solution, the benzimidazole derivatives, BHB and BMB, are protonated on
their nitrogen heteroatoms, as confirmed by XPS characterization. This enables them to
competitively adsorb on the cathodic sites of the steel surface so as to deactivate H+ ion
reduction reaction. The protonation also enables them to adsorb on the anodic sites (facilitated
by electrostatic interaction with the already adsorbed Cl− ions). Thus, the inhibitors function
as mixed-type inhibitors mitigating both cathodic and anodic electrochemical half-reactions
on the steel surface. Their adsorption on the steel surface is also confirmed by the ATR-IR
characterization. During adsorption, they displace primarily adsorbed water molecules and
isolate the steel surface from the corrosive agents. In this way, they also reduce the tendency for
further oxidation of Fe2+ species into Fe3+ species. These are confirmed by the low intensities
of O 1s, Cl 2p and Fe 2p peaks in the XPS of the inhibited steel. Accordingly, the steel surface
becomes less exposed to localized pitting corrosion, based on SEM characterization.
Once adsorbed, the presence of a methyl substituent in the BMB confers larger surface
coverage, stronger adsorption and higher repulsion of aqueous corrosion agents. Hence,
while computational results show that both inhibitors assume flat adsorption orientations on
the Fe surface (which is highly favorable for efficient inhibition), BMB shows lower energy
difference between E HOMO and E LUMO , higher adsorption energy and inhibition efficiency
than BHB.

4 Conclusions

The corrosion inhibition properties of two benzimidazole derivatives 2-(2-Bromophenyl)-


1H-benzimidazole (BHB) and 2-(2-Bromophenyl)-1-methyl-1H-benzimidazole (BMB) for

123
129 Page 26 of 29 Eur. Phys. J. Plus (2020) 135:129

X60 steel in 1 M HCl solution have been assessed under laminar and turbulent hydrodynamic
conditions at room temperature and natural atmospheric condition. The following conclusions
can be drawn from the results:

1. The benzimidazole nitrogen and the aromatic rings in BHB and BMB function as active
centers which facilitate their adsorption on the steel surface.
2. The adsorption of BHB and BMB increases the corrosion resistance of X60 steel in 1 M
HCl solution at room temperature under laminar and turbulent hydrodynamic conditions,
and corrosion inhibition efficiency increases with inhibitors concentration.
3. Turbulent hydrodynamics increase corrosion rate by increasing the transport of H+ and
Cl− ions to the steel surface and also increasing wall shear stress that can delaminate the
adsorbed inhibitors.
4. The presence of a methyl group confers greater corrosion inhibition properties on BMB
than BHB.

Acknowledgements Ikenna B. Onyeachu acknowledges the Center of Research Excellence in Corrosion


(CoRE–C), at the Research Institute of King Fahd University of Petroleum and Minerals, Saudi Arabia, for the
award of a Postdoctoral Fellowship. Aeshah H. Alamri acknowledges the Scientific and High-Performance
Computing Center at Imam Abdulrahman Bin Faisal University, Dammam, Saudi Arabia, for providing tech-
nical services and resources for this project.

Compliance with ethical standards

Conflict of interest On behalf of all authors, the corresponding author states that there is no conflict of interest.

References

1. I.B. Obot, I.B. Onyeachu, N. Wazzan, A.H. Alamri, Theoretical and experimental investigation of two
alkyl carboxylates as corrosion inhibitor for steel in acidic medium. J. Mol. Liq. 279, 190–207 (2019)
2. I.B. Obot, N.O. Obi-Egbedi, Theoretical study of benzimidazole and its derivatives and their potential
activity as corrosion inhibitors. Corros. Sci. 52, 657–660 (2010)
3. I.B. Obot, I.B. Onyeachu, A.M. Kumar, Sodium alginate: a promising biopolymer for corrosion protection
of API X60 high strength carbon steel in saline medium. Carbohydr. Polym. 178, 200–208 (2017)
4. C. Verma, L.O. Olasunkanmi, E.E. Ebenso, M.A. Quraishi, I.B. Obot, Adsorption behavior of
glucosamine-based, pyrimidine-fused heterocycles as green corrosion inhibitors for mild steel: exper-
imental and theoretical studies. J. Phys. Chem. C 120, 11598–11611 (2016)
5. I.B. Obot, E.E. Ebenso, M.M. Kabanda, Metronidazole as environmentally safe corrosion inhibitor for
mild steel in 0.5 M HCl: experimental and theoretical investigation. J. Environ. Chem. Eng. 1, 431–439
(2013)
6. M.A. Chidiebere, L. Nnanna, C.B. Adindu, K.L. Oguzie, B. Okolue, B.I. Onyeachu, Inhibition of acid
corrosion of mild steel using delonix regia leaves extract. Int. Lett. Chem. Phys. Astron. 69, 74–86 (2016)
7. B. Mernari, H. El Attari, M. Traisnel, F. Bentiss, M. Lagrenee, Inhibiting effects of 3,5-bis(N-pyridyl)-
4-amino-1,2,4-triazoles on the corrosion for mild steel in 1 M HCl medium. Corros. Sci. 40(2), 391–399
(1998)
8. F. Bentiss, C. Jama, B. Mernari, H. El Attari, L. El Kadi, M. Lebrini, M. Traisnel, M. Lagrenee, Corrosion
control of mild steel using 3, 5-bis (4-methoxyphenyl)-4-amino-1, 2, 4-triazole in normal hydrochloric
acid medium. Corros. Sci. 51(8), 1628–1635 (2009)
9. K.R. Ansari, M.A. Quraishi, A. Singh, Corrosion inhibition of mild steel in hydrochloric acid by some
pyridine derivatives: an experimental and quantum chemical study. J. Ind. Eng. Chem. 25, 89–98 (2015)
10. F. El-Hajjaji, M. Messali, A. Aljuhani, M.R. Aouad, B. Hammouti, M.E. Belghiti, D.S. Chauhan, M.A.
Quraishi, Pyridazinium-based ionic liquids as novel and green corrosion inhibitors of carbon steel in acid
medium: electrochemical and molecular dynamics simulation studies. J. Mol. Liq. 249, 997–1008 (2018)

123
Eur. Phys. J. Plus (2020) 135:129 Page 27 of 29 129

11. M. Lagrenee, B. Mernari, M. Bouanis, M. Taisnel, F. Bentiss, Study of the mechanism and inhibiting
efficiency of 3, 5-bis (4-methylthiophenyl)-4H-1, 2, 4-triazole on mild steel corrosion in acidic media.
Corros. Sci. 44, 573–588 (2002)
12. Y. Tang, F. Zhang, S. Hu, Z. Cao, Z. Wu, W. Jing, Novel benzimidazole derivatives as corrosion inhibitors
of mild steel in the acidic media. Part I: gravimetric, electrochemical, SEM and XPS studies. Corros. Sci.
74, 271–282 (2013)
13. X. Wang, Y. Wan, Y. Zeng, Y. Gu, Investigation of benzimidazole compound as a novel corrosion inhibitor
for mild steel in hydrochloric acid solution. Int. J. Electrochem. Sci. 7, 2403–2415 (2012)
14. X. Wang, The inhibition effect of bis-benzimidazole compound for mild steel in 0.5 M HCl solution. Int.
J. Electrochem. Sci. 7, 11149–11160 (2012)
15. I. Ahamad, M.A. Quraishi, Bis (benzimidazol-2-yl) disulphide: an efficient water soluble inhibitor for
corrosion of mild steel in acid media. Corros. Sci. 51, 2006–2013 (2009)
16. M. Yadav, S. Kumar, T. Purkait, L.O. Olasunkanmi, I. Bahadur, E.E. Ebenso, Electrochemical, thermo-
dynamic and quantum chemical studies of synthesized benzimidazole derivatives as corrosion inhibitors
for N80 steel in hydrochloric acid. J. Mol. Liq. 213, 122–138 (2016)
17. A. Dutta, S.K. Saha, U. Adhikari, P. Banerjee, D. Sukul, Effect of substitution on corrosion inhibition
properties of 2-(substituted phenyl) benzimidazole derivatives on mild steel in 1 M HCl solution: a
combined experimental and theoretical approach. Corros. Sci. 123, 256–266 (2017)
18. Z. Cao, Y. Tang, H. Cang, J. Xu, G. Lu, W. Jing, Novel benzimidazole derivatives as corrosion inhibitors
of mild steel in the acidic media. Part II: theoretical studies. Corros. Sci. 83, 292–298 (2014)
19. X. Wang, H. Yang, F. Wang, An investigation of benzimidazole derivative as corrosion inhibitor for mild
steel in different concentration HCl solutions. Corros. Sci. 53, 113–121 (2011)
20. J. Aljourani, K. Raeissi, M.A. Golozar, Benzimidazole and its derivatives as corrosion inhibitors for mild
steel in 1 M HCl solution. Corros. Sci. 51, 1836–1843 (2009)
21. S. Ramachandran, B.L. Tsai, M. Blanco, H. Chen, Y. Tang, W.A. Goddard, Self-assembled monolayer
mechanism for corrosion inhibition of iron by imidazolines. Langmuir 12, 6419–6428 (1996)
22. B.P. Binks, P.D.I. Fletcher, I.E. Salama, Quantitative prediction of the reduction of corrosion inhibitor
effectiveness due to parasitic adsorption onto a competitor surface. Langmuir 27, 469–473 (2011)
23. H. Ashassi-Sorkhabi, E. Asghari, Effect of hydrodynamic conditions on the inhibition performance of
l-methionine as a green inhibitor. Electrochim. Acta 54, 162–167 (2008)
24. X. Jiang, Y.G. Zheng, W. Ke, Effect of flow velocity and entrained sand on inhibitor performances of two
inhibitors for CO2 corrosion of N80 steel in 3% NaCl solution. Corros. Sci. 47, 2636–2658 (2005)
25. B.R. Tian, Y.F. Cheng, Electrochemical corrosion behavior of X-65 steel in the simulated oil and sand
slurry. I: effects of hydrodynamic condition. Corros. Sci. 50, 773–779 (2008)
26. P. Bommersbach, C. Alemany-Dumont, J.P. Millet, B. Normand, Hydrodynamic effect on the behavior
of a corrosion inhibitor film: characterization by electrochemical impedance spectroscopy. Electrochim.
Acta 51, 4011–4018 (2006)
27. N. Ochoa, F. Moran, N. Pebere, B. Tribollet, Influence of flow on the corrosion inhibition of carbon steel
by fatty amines in association with phosphonocarboxylic acid salts. Corros. Sci. 47, 593–604 (2005)
28. T. Douadi, H. Hamani, D. Daoud, M. Al-Noaimi, S. Chafaa, Effect of temperature and hydrodynamic
conditions on corrosion inhibition of an azomethine compounds for mild steel in 1 M HCl solution. J.
Taiwan Inst. Chem. Eng. 71, 388–404 (2017)
29. A. Kosari, M.H. Moayed, A. Davoodi, R. Parvizi, M. Momeni, H. Eshghi, H. Moradi, Electrochemical
and quantum chemical assessment of two organic compounds from pyridine derivatives as corrosion
inhibitors for mild steel in HCl solution under stagnant condition and hydrodynamic flow. Corros. Sci.
78, 138–150 (2014)
30. E. Barmatov, T. Hughes, M. Nagl, Efficiency of film-forming corrosion inhibitors in strong hydrochloric
acid under laminar and turbulent flow conditions. Corros. Sci. 92, 85–94 (2015)
31. I.B. Obot, A. Madhankumar, S.A. Umoren, Z.M. Gasem, Surface protection of mild steel using ben-
zimidazole derivatives: experimental and theoretical approach. J. Adhes. Sci. Technol. 29, 2130–2152
(2015)
32. B. Poulson, Electrochemical measurements in flowing solutions. Corros. Sci. 23, 391–430 (1983)
33. I.B. Onyeachu, I.B. Obot, A.A. Sorour, M.I. Abdul-Rashid, Green corrosion inhibitor for oilfield appli-
cation I: electrochemical assessment of 2-(2-pyridyl) benzimidazole for API X60 steel under sweet envi-
ronment in NACE brine ID196. Corros. Sci. 150, 183–193 (2019)
34. X. Peng, Y. Zhang, J. Zhao, F. Wang, Electrochemical corrosion performance in 3.5% NaCl of the
electrodeposited nanocrystalline Ni films with and without dispersions of Cr nanoparticles. Electrochim.
Acta 51, 4922–4927 (2006)

123
129 Page 28 of 29 Eur. Phys. J. Plus (2020) 135:129

35. J. Kestin, H.E. Khalifa, R.J. Correla, Tables of the dynamic and kinematic viscosity of aqueous NaCl
solutions in the temperature range 20–150 °C and the pressure range 0.1–35 MPa. J. Phys. Chem. Ref.
Data 10(1), 71–87 (1981)
36. M.J. Frisch, G.W. Trucks, H.B. Schlegel, G.E. Scuseria, M.A. Robb, J.R. Cheeseman, G. Scalmani et al.,
Gaussian 16, Revision A. 03 (Gaussian Inc., Wallingford, 2016)
37. A.D. Becke, Density functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 98(7),
5648–5652 (1993)
38. R. Dennington, T.A. Keith, J.M. Millam, GaussView, Version 6 (Semichem Inc., Shawnee Mission, 2016)
39. B. Mennucci, J. Tomasi, R. Cammi, J.R. Cheeseman, M.J. Frisch, F.J. Devlin, S. Gabriel, P.J. Stephens,
Polarizable continuum model (PCM) calculations of solvent effects on optical rotations of chiral
molecules. J. Phys. Chem. A 106(25), 6102–6113 (2002)
40. S. Satoh, H. Fujimoto, H. Kobayashi, Theoretical study of NH3 adsorption on Fe (110) and Fe (111)
surfaces. J. Phys. Chem. B 110(10), 4846–4852 (2006)
41. L. Guo, S. Zhu, S. Zhang, Q. He, W. Li, Theoretical studies of three triazole derivatives as corrosion
inhibitors for mild steel in acidic medium. Corros. Sci. 87, 366–375 (2014)
42. F. Bazooyar, F.A. Momany, K. Bolton, Validating empirical force fields for molecular-level simulation of
cellulose dissolution. J. Comput. Theor. Chem. 984, 119–127 (2012)
43. P.P. Ewald, Die Berechnung optischer und elektrostatischer Gitterpotentiale. Ann. der Physic 369(3),
253–287 (1921)
44. S. Kumar, D. Sharma, P. Yadav, M. Yadav, Experimental and quantum chemical studies on corrosion
inhibition effect of synthesized organic compounds on N80 Steel in hydrochloric acid. Ind. Eng. Chem.
Res. 52, 14019–14029 (2013)
45. A.A. Mohammed, K.F. Khaled, Q. Moshen, H.A. Arida, A study of the inhibition of iron corrosion in
HCl solutions by some amino acids. Corros. Sci. 52, 1684–1695 (2010)
46. C.K. Ingold, Structure and mechanism in organic chemistry (Cornell University Press, Ithaca, 1959)
47. S.A. Haladau, S.A. Umoren, S.A. Ali, M.M. Solomon, A.I. Mohammed, Synthesis, characterization and
electrochemical evaluation of anticorrosion property of a tetrapolymer for carbon steel in strong acid
media. Chin. J. Chem. Eng. (2018). https://doi.org/10.1016/j.cjche.2018.07.015
48. Z.B. Wang, H.X. Hu, Y.G. Zheng, Synergistic effects of fluoride and chloride on general corrosion behavior
of AISI 316 stainless steel and pure titanium in H2 SO4 solutions. Corros. Sci. 130, 203–217 (2018)
49. C.H. Hsu, F. Mansfeld, Technical note: concerning the conversion of the constant phase element parameter
Yo into a capacitance. Corrosion 57, 747–748 (2001)
50. M. Stern, A.L. Geary, Electrochemical polarization, 1. A theoretical analysis of the shape of polarization
curves. J. Electrochem. Soc. 104(12), 751–752 (1957)
51. T.S. Xavier, N. Rashid, I.H. Joe, Vibrational spectra and DFT study of anticancer active molecule 2-
(4-Bromophenyl)-1H-benzimidazole by normal coordinate analysis. Spectrochim. Acta A 78, 319–326
(2011)
52. G. Varsanyi, Assignments for Vibrational Spectra of Seven Hundred Benzene Derivatives, vol. 1 (Adam
Hilger, London, 1974)
53. C. Birsan, D. Predoi, E. Andronescu, IR and thermal studies of iron oxide nanoparticles in a bioceramic
matrix. J. Optoelectron. Adv. Mater. 9(6), 1821–1824 (2007)
54. E.A. Noor, A.H. Al-Moubaraki, Corrosion behavior of mild steel in hydrochloric acid solution. Int. J.
Electrochem. Sci. 3, 806–818 (2008)
55. J. Galvele, Transport processes in passivity breakdown—II. Full hydrolysis of the metal ions. Corros. Sci.
21(8), 551–579 (1981)
56. F. Zhang, Y. Tang, Z. Cao, W. Jing, Z. Wu, Y. Chen, Performance and theoretical study on corrosion
inhibition on 2-(4-pyridyl) benzimidazole for steel in hydrochloric acid. Corros. Sci. 61, 1–9 (2012)
57. M.M. Solomon, S.A. Umoren, I.B. Obot, A.A. Sorour, H. Gerengi, Exploration of dextran for application
as corrosion inhibitor for steel in strong acid Environment: effect of molecular weight, modification, and
temperature on efficiency. ACS Appl. Mater. Interfaces 10, 28112–28129 (2018)
58. M.M. Solomon, H. Gerengi, S.A. Umoren, N.B. Essien, U.B. Essien, E. Kaya, Gum Arabic-silver nanopar-
ticles composite as a green anticorrosive formulation for steel corrosion in strong acid media. Carbohydr.
Polym. 181, 43–55 (2018)
59. F. Bentiss, M. Traisnel, L. Gengembre, M. Lagrenee, Inhibition of acidic corrosion of mild steel by
3.5-diphenyl-4H-1,2,4-triazole. Appl. Surf. Sci. 161, 194–202 (2000)
60. W. Zhang, C. Kong, G. Lu, Super-paramagnetic nano-Fe3 O4 /graphene for visible-light-driven hydrogen
evolution. Chem. Commun. 51, 10158–10161 (2015)
61. B.I. Onyeachu, X. Peng, E.E. Oguzie, I. Digbo, Characterizing the electrochemical corrosion behavior of
a Ni-28 wt% Al composite coating in 3.5% NaCl solution. Port. Electrochim. Acta. 33(2), 69–83 (2015)

123
Eur. Phys. J. Plus (2020) 135:129 Page 29 of 29 129

62. M. Chevalier, F. Robert, N. Amusant, M. Traisnel, C. Ross, M. Lebrini, Enhanced corrosion resistance
of mild steel in 1 M hydrochloric acid solution by alkaloids extracts from Aniba rosaeodora plant:
electrochemical, phytochemical and XPS studies. Electrochim. Acta 131, 96–105 (2014)
63. R.J. Chin, K. Nobe, Electrodissolution kinetics of iron in chloride solutions III. Acidic solutions. J.
Electrochem. Soc. 119(11), 1457–1461 (1972)
64. D.R. MacFarlane, S.I. Smedley, The dissolution mechanism of iron in chloride solutions. J. Electrochem.
Soc. 133(11), 2240–2244 (1986)
65. T.P. Hoar, D.C. Mears, G.P. Rothwell, The relationships between anodic passivity, brightening and pitting.
Corros. Sci. 4, 279–289 (1965)
66. C.Y. Chao, L.F. Lin, D.D. Macdonald, A point defect model for anodic passive films I. Film growth
kinetics. J. Electrochem. Soc. 128, 1187–1194 (1981)
67. J.O. Bockris, A. Swinkels, A adsorption of n-decylamine on solid metal electrodes. J. Electrochem. Soc.
111(6), 736–743 (1964)
68. A.C. Maduabuchi, C.E. Ogukwe, K.L. Oguzie, C.N. Eneh, E.E. Oguzie, Corrosion inhibition and adsorp-
tion behavior of Punica granatum extract on mild steel in acidic environments and theoretical studies.
Ind. Eng. Chem. Res. 51, 668–677 (2012)
69. I.B. Obot, Z.M. Gasem, Theoretical evaluation of corrosion inhibition performance of some pyrazine
derivatives. Corros. Sci. 83, 359–366 (2014)
70. J. Zhang, G. Qiao, S. Hu, Y. Yan, Z. Ren, L. Yu, Theoretical evaluation of corrosion inhibition performance
of imidazoline compounds with different hydrophilic groups. Corros. Sci. 56, 176–183 (2011)
71. L. Guo, W. Dong, S. Zhang, Theoretical challenges in understanding the inhibition mechanism of copper
corrosion in acid media in the presence of three triazole derivatives. RSC Adv. 4, 41956–41967 (2014)

123

You might also like