Theoretical Prediction and Electrochemical Evaluation of Vinylimidazole and Allylimidazole As Corrosion For Mild Steel-2015

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 12

Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


journal homepage: www.elsevier.com/locate/jiec

Theoretical prediction and electrochemical evaluation of


vinylimidazole and allylimidazole as corrosion inhibitors for mild steel
in 1 M HCl
I.B. Obot a, *, S.A. Umoren a , Z.M. Gasem a , Rami Suleiman a , Bassam El Ali b
a
b

Centre of Research Excellence in Corrosion, Research Institute, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
Department of Chemistry, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia

A R T I C L E I N F O

A B S T R A C T

Article history:
Received 15 April 2014
Received in revised form 27 May 2014
Accepted 31 May 2014
Available online 11 June 2014

Corrosion inhibition potentials of two imidazole derivatives namely, vinylimidazole (VI) and
allylimidazole (AI) for carbon steel in 1 M HCl at 25  C were predicted theoretically using quantum
chemical calculations and molecular dynamics (MD) simulations. DFT calculations indicated that VI is
more reactive towards steel surface than AI. Equilibruim adsorption behaviour of VI and AI molecules on
Fe2O3 (0 1 0) surface was investigated using molecular dynamics (MD) simulations. The equilibrium
adsorption energy followed the order: VI > AI. Theoretical conclusions were subsequently validated
experimentally using potentiodynamic polarization, linear polarization resistance, electrochemical
impedance spectroscopy, and surface analytical techniques (SEM and AFM).
2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

Keywords:
Mild steel
DFT
Vinylimidazole
Allylimidazole
Molecular dynamics simulations

Introduction
Mild steel is the most extensively investigated metal for
corrosion studies because of its wide application in different
corrosive environments. For instance, in oil industries, aqueous
acidic solution is used for descaling, acid pickling, and acid
treatment. In this case, the exposure of these metals to aqueous
acidic solution causes corrosion. The use of inhibitors is one of the
most practical approaches to protect steel from corrosion in these
acidic environments [15]. Organic inhibitors containing polar
groups (including N, S and O), heterocyclic compounds with polar
functional groups, and conjugated double bonds can effectively
inhibit the corrosion of steel, due to their chelating action and the
formation of an insoluble physical diffusion barrier on the
substrate surface [610]. The existing data reveal that most
organic inhibitors act by adsorption on the metal surface. This
adsorption is inuenced by the nature and surface charge of metal,
the type of aggressive electrolyte and the chemical structure of
inhibitors [11]. Unfortunately, most organic inhibitors are inherently toxic and potential health hazards, such as the carcinogenic
effect of aromatic heterocyclic compounds on humans [12].

* Corresponding author. Tel.: +966 533413506; fax: +966 3 860 3996.


E-mail address: obot@kfupm.edu.sa (I.B. Obot).

Imidazoles have attracted attention recently because of their


good environmental prole and their excellent corrosion inhibition performance [1319]. Most of the inhibition performances of
imidazoles were mainly studied by experimental methods, such as
weight loss measurement, polarization curves and electrochemical
impedance spectroscopy. However, theoretical studies on imidazoles as corrosion inhibitors are relatively few and is increasingly
vital to the detailed understanding of the mechanism of its
inhibition on metal surface at the molecular level [2022]. Such
studies can be undertaken by the use of quantum chemical
calculations supported by molecular dynamics simulations. These
methodologies are widely reported as important modern tool in
corrosion inhibition studies and will continue to make great
impact in corrosion science research [2327]. In addition, the use
of theoretical methodologies in predicting inhibition efcacies of
organic molecules as corrosion inhibitors for metals and alloys and
thereafter validate the prediction experimentally is scarce in the
literature [2830].
In the present communication, quantum chemical calculations in
conjuction with molecular dynamics simulations were employed to
predict the inhibition potentials of vinylimidazole (VI) and
allylimidazole (AI) on carbon steel and to understand the interactions between the inhibitor molecules and steel surface. The
inhibition performance of the two imidazole inhibitors for mild steel
in 1 M HCl were subsequently validated experimentally by using
electrochemical measurements and surface analytical techniques.

http://dx.doi.org/10.1016/j.jiec.2014.05.049
1226-086X/ 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

Experimentals
Materials and sample preparation
The mild steel specimens with the following chemical
composition (weight percentage) were used in the experiments:
C, 0.073; Mn, 1.36; P, 0.004; Ti, 0.004; S, 0.003; balance Fe. Test
coupons were cut into 3 cm  3 cm  0.25 cm dimensions. These
coupons were abraded with silicon carbide abrasive paper (from
grade no. 320 to 800), rinsed with distilled water, placed in an
ultrasonic acetone bath for about 5 min to remove possible residue
of polishing, rinsed with acetone, dried in warm air, and then
stored in moisture-free desiccators prior to use. The solutions of
1 M HCl were prepared by dilution of 37% analytical grade HCl with
double distilled water. The inhibitors tested was vinylimidazole
and allylimidazole obtained from SigmaAldrich as high grade
reagent and was used without further purication. The employed
concentration range of the inhibitors was 0.0010.01 M. All
experiments were conducted at 25  C. The molecular structures
are dipicted in Fig. 1.
Quantum chemical calculations and molecular dynamics (MD)
simulations

1329

was selected for the simulation based on the fact that the iron
metal already oxidized before introducing the acid solution [32]. VI
and AI were optimized (i.e., energy minimizing) using the Forcite
module and the Condensed-Phase Optimized Molecular Potentials
for Atomistic Simulation Studies (COMPASS) ab initio force eld
and the String Matching Algorithms Research Tool (smart) in a
simulation box (88.76  54.19  39.44 ) with periodic boundary conditions to model a representative part of the interface
devoid of any arbitrary boundary effects. The simulated annealing
procedure uses the Monte Carlo method to determine the
adsorption density and binding energy of the adsorbate on the
substrate. The Fe2O3 (0 1 0) surface was rst built and relaxed by
minimizing its energy using molecular mechanics; then the
surface area of Fe2O3 (0 1 0) was increased, and its periodicity
was changed by constructing a super cell (9  9). A vacuum slab
with a thickness of 40 was then built on the Fe2O3 (0 1 0) surface.
The six layers in the structure were chosen so that the depth of the
surface was greater than the nonbonded cutoff used in the
calculations. Once the Fe2O3 (0 1 0) surface had been built, VI and AI
were adsorbed on the metal surface using the adsorption locator
module. Extensive details of the simulation process have been
elaborated elsewhere [3335].
Electrochemical measurements

Quantum chemical calculations and molecular dynamics


simulations were conducted with Materials Studio 6.0 [31]. All
electron calculations of VI and AI molecules were accomplished by
density functional theory (DFT) hybrid (B3LYP) method with a
double zeta plus polarization (DNP) basis set and the choice of
convergence accuracy was set to ne. Frequency analysis was
performed to ensure the calculated structure was at the minimum
point on potential energy surface (without imaginary frequency). A
continuum solvation method (COSMO) was employed to study the
effect of solvent (water). DMol3 module code was employed in the
DFT calculations. The following quantum chemical parameters
were calculated from the obtained optimized molecular structure:
the energy of the highest occupied molecular orbital (EHOMO), the
energy of the lowest unoccupied molecular orbital (ELUMO), the
energy band gap (DE = ELUMO  EHOMO), the dipole moment (m),
hardness (h), softness (S) and electrophilicity (v). These parameters were employed to rank the reactive abilities of VI and AI.
Mulliken population analysis was used for the presentation of
Fukui functions.
Adsorption locator module present in Materials Studio 6.0 was
employed in the molecular dynamics simulations using simulated
annealing procedure. The adsorption of VI and AI on Fe2O3(0 1 0)
surface was simulated to nd the low-energy adsorption sites on
the iron oxide surface and to investigate the preferential
adsorption of the studied compounds. The Fe2O3 surface layer

[(Fig._1)TD$IG]

All electrochemical experiments were performed in a onecompartment cell with three electrodes connected to Gamry
instrument potentiostat/galvanostat/ZRA (Reference 3000) with a
Gamry framework system based on ESA410. Gamry applications
include software DC105 for corrosion, EIS300 for EIS measurements,
and the Echem Analyst 6.0 software package for data tting. The mild
steel was the working electrode with an exposed surface area of
0.7855 cm2 in the corrosive environment; platinum wire was used as
the counter electrode and saturated calomel electrode (SCE) as the
reference electrode. All potentials were measured versus the SCE
reference electrode. Tafel curves were obtained by changing the
electrode potential automatically from 250 to +250 mV versus the
open-circuit potential (Ecorr) at a scan rate of 1 mV s1. Linear
polarization resistance (LPR) experiments were done from 20 to
+20 mV versus Ecorr at a scan rate of 0.125 mV s1. EIS measurements
were carried out under potentiostatic conditions in a frequency
range from 100 kHz to 100 mHz, with an amplitude of 10 mV peakto-peak, using an alternating-current (ac) signal at Ecorr. All
experiments were measured after immersion for 30 min in 1 M
HCl with and without the addition of the inhibitors.
Surface morphology using scanning electron microscopy (SEM)
Morphological studies of the mild steel electrode surface were
undertaken by SEM examinations of electrode surfaces exposed to
different test solutions using a JSM-5800 LV scanning electron
microscope. Mild steel specimens of dimensions 3  3  0.25 cm3
were abraded successively with silicon carbide paper of different
grades (no. 320800) and thereafter using cloth with 1 mm
diamond paste to a near m irror nished surface. The precleaned
coupons were immersed for 18 h in the blank solutions in 1 M HCl
without and with highest concentration of VI and AI (0.01 M) at
25  C, then washed with distilled water, dried in warm air, and
submitted for SEM surface examination.
Atomic force microscopy (AFM)

Fig. 1. 2D molecular structures of (a) vinyl imidazole (VI) and (b) allylimidazole (AI).

Surface topological studies of the mild steel surface was


investigated by atomic force microscopy. For AFM analysis the mild
steel specimens of size 3  3  0.25 cm3 were immersed in the test
solution in the absence and presence of 0.01 M inhibitors (VI and

1330

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._2)TD$IG]

AI) for 3 h at ambient temperature. Then the specimens were


cleaned with distilled water, dried, and used for AFM. The AFM
analyses were carried out using a 5420 atomic force microscope
(AFM) (N9498S) (Agilent Technologies, UK).
Results and discussion
Quantum chemical study
Quantum chemical calculation has been widely used to evaluate
the inhibition performance of corrosion inhibitors, which can
quantitatively study the relationship between inhibition efciency
and molecular reactivity [3639]. With this method, the capability
of inhibitor molecules to donate or accept electrons can be
predicted with analysis of global reactivity parameters, such as the
energy gap between HOMO and LUMO, chemical hardness, dipole
moment, and electrophilicity index, etc.
The reactive abilities of VI and AI are closely related to their
frontier molecular orbtals including the highest occupied molecular orbitals (HOMO) and the lowest unoccupied molecular orbital
(LUMO). EHOMO indicates the tendency of an organic molecule to
donate electrons. The higher the value of EHOMO, the greater the
ability of a molecule to donate electrons while ELUMO indicates the
propensity of a molecule to accept electrons. The lower ELUMO is,
the greater is the ability of that molecule to accept electrons. Thus,
the binding ability of organics to the metal surface increases with
an increase in energy of the HOMO and a decrease in the value of
energy of the LUMO. The energy gap, DE, is an important
parameter which indicates the reactivity tendency of organics
towards the metal surface [40]. As DE decreases, the reactivity of
the molecule increases, leading to an increase in adsorption onto a
metal surface. A molecule with low energy gap is more polarizable
and is generally associated with high chemical reactivity and low
kinetic stability. Thus, DE, has been used in literature to
characterize the binding ability of organics to the metal surface
[41]. The reactivity of corrosion inhibitors may also be discussed in
terms of chemical hardness and softness parameters. These
quantities are often associated with the Lewis theory of acid
and bases and Pearson's hard and soft acids and bases theory [42];
a hard molecule has a large DE and therefore is less reactive; a soft
molecule has a small DE and is therefore more reactive. Adsorption
occurs most probably at the region of the molecule where softness
(S) has the highest value [43]. In the study of corrosion inhibitors
and their ability to bind to the metal surface, the inhibitor is
considered as a soft base and the metal surface as a soft acid. The
electrophilicity index (v), on the other hand denotes the electronaccepting capability of a molecule [44]. High values of v ensures
greater binding ability of corrosion inhibitors to metal surfaces.
The optimization progress curves for VI and AI are shown in
Fig. 2., while the frontier molecular orbital density distributions
are presented in Fig. 3. As can be clearly seen in Fig. 3, the HOMO
orbitals of VI and AI are entirely on the imidazole ring. In the case of
LUMO orbitals, the distribution is mainly on the imidazole ring of
VI whereas for AI it is on the allyl group. This is an indication of the
availability of more adsorption centers for VI than AI. The
electronic parameters related to the reactivity of the VI and AI
as corrosion inhibitors are reported in Table 1. It is evident from
Table 1 that VI and AI has similar EHOMO. VI on the other hand has
the lowest ELUMO and DE values, the highest S and v values, making
it to have more reactive and binding potentials towards steel
surface than AI as calculated theoretically. The dipole moments of
VI (4.76 D) and AI (5.69) are more than that of water (1.85 D) which
shows the ability of the inhibitors to displace water molecules from
the steel surface thereby inhibiting the corrosion of steel against
aqueous acidic medium. Although the dipole moment of AI is

Fig. 2. Optimization progress curves of (a) vinyl imidazole and (b) allylimidazole
calculated using DFT.

greater than that of VI from this study, some degree of confusion


exists when dealing with dipole moment data in the interpretation
of inhibition efciency data, since the scientic literature provides
both positive [2] as well as negative [45] relationships between
this parameter and inhibition efciency.
It has been reported that molecules with atoms with the highest
negative charges are often considered to have the highest tendency
to donate electrons to the metal surface [13]. Therefore, the inhibitor
is likely to interact with the metal surface through such atoms.
Table 2 shows the Mulliken population analysis performed for for VI
and AI. The results show that (VI: N1, N2, C7) and (AI: N1, N2, C3, C6,
C8) are atoms with negative charge centers that could offer electrons
to the iron atoms on the metal surface to form a coordinate bond. On
the other hand, (VI: C3, C4, C5, C6) and (AI: C4, C5, C7) are atoms of
positive centers that can accept electrons easily from 3d orbital of Fe
atoms to form feedback bonds. It can be seen from Table 2 that
although VI has less negative centers than AI, its total negative
charges are about the same (VI: 1.13) and (AI: 1.16). VI on the other
hand, has the highest total positive charges (0.378) as opposed to
(0.357) for AI. These high positive centers can be use by VI to accept
electrons from 3d orbital of Fe, thus strengthening the bond between
VI and Fe than between AI and Fe.
The Fukui function which measures reactivity in a local sense, is
by far the most important local reactivity index [46]. Using a
scheme of nite difference approximations, this procedure
condenses the values around each atomic site into a single value
that characterizes the atom in the molecule. With this approximation, the condensed Fukui function becomes:

f k qk N 1  qk Nfor nucleophilic attack




f k qk N  qk N  1for electrophilic attack

(1)

(2)

where qk is the gross charge of atom k in the molecule, N


corresponds to the number of electrons in the molecule, with N + 1

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._3)TD$IG]

1331

Fig. 3. HOMO and LUMO orbitals of VI and AI.

Table 1
Calculated quantum chemical properties for the most stable conformations of VI
and AI calculated using B3LYP/DNP level of theory in water.
Properties

VI

AI

Total energy (Ha)


EHOMO (eV)
ELUMO (eV)
DE (eV)
m (D)

321.60
6.61
0.87
5.74
4.76
2.86
0.34
9.74

363.33
6.58
0.16
6.42
5.69
3.21
0.31
7.08

corresponding to a singly-charged anion, with an electron added to


the LUMO of the neutral molecule; and N  1 corresponding to
singly-charged cation with an electron removed from the HOMO of
the neutral molecule.
An analysis of the Fukui indices for nucleophilic and electrophilic
sites are presented in Tables 3 and 4, respectively. In VI, atoms C4, C6,
C7 and in AI, atoms C6, C8 are the most susceptible sites for

N1
N2
C3
C4
C5
C6
C7
C8
TNC (Total negative charge)
TPC (Total positive charge)

Atoms

N1
N2
C3
C4
C5
C6
C7
C8

f+
VI

AI

0.025
0.075
0.093
0.160
0.037
0.107
0.211

0.023
0.033
0.017
0.022
0.055
0.177
0.018
0.277

nucleophilic attacks. On the other hand, atoms C3, C4, C5, C7 in VI and
atoms C4, C5, C7 in AI are the most probable centers for electrophilic
attacks. Nevertheless, in VI, the atom C7 has the highest value of f
whereas in AI, the atom C8 has the highest value of f. These sites are
the most reactive sites for neucleophilic attacks. As for f+,
electrophilic attacks, C5 is the highest value for both VI and AI.
This site is the most reactive for electrophilic attacks.

Table 2
Mulliken charges of VI and AI.
Atoms

Table 3
Nucleophilic Fukui function (f+) calculated for VI and AI at B3LYP/
DNP.

Table 4
Electrophilic Fukui function (f+) calculated for VI and AI at B3LYP/DNP.

Charges
VI

AI

0.383
0.492
0.003
0.226
0.038
0.111
0.258

1.133
0.378

0.383
0.498
0.041
0.035
0.269
0.058
0.053
0.183
1.163
0.357

Atoms

N1
N2
C3
C4
C5
C6
C7
C8

f+
VI

AI

0.021
0.060
0.130
0.132
0.158
0.047
0.100

0.001
0.080
0.011
0.212
0.177
0.009
0.167
0.020

1332

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

Molecular dynamics simulation


Many corrosion inhibition studies nowadays contain the use of
molecular dynamics simulation as an important tool in understanding the interaction between adsorbate-metal surface [47
50]. Thus Monte Carlo and molecular dynamics simulations were
performed on a system containing the inhibitors (VI and AI)
adsorbed on a Fe2O3 (0 1 0) surface as shown in Fig. 4. Also Fig. 5
shows the energy prole diagram for the adsorption progress of VI
and AI on Fe2O3 surface. The energy prole consist of the total
energy, average total energy, Van der Waals energy, electrostatic
energy and intermolecular energy for the systems. The Monte
Carlo docking was done on each of the 100 conformations, and
each of the docked structures were energetically relaxed. The
outputs and descriptors calculated by the Monte Carlo simulation,
including the total adsorption, rigid adsorption and deformation
energies, are presented in Table 5. The total energy is dened as the
sum of the energies of the adsorbate components. The adsorption
energy is dened as the sum of the rigid adsorption energy and the
deformation energy for the adsorbate components. The rigid
adsorption energy reports the energy, in kJ mol1, released (or
required) when the unrelaxed adsorbate components (VI and AI)
(i.e., before the geometry optimization step) are adsorbed on the
Fe2O3 surface. The deformation energy is the energy released when
the adsorbed adsorbate components are relaxed on the Fe2O3
surface. Table 5 also shows (dEads/dNi), which is the energy of
Fe2O3-adsorbate congurations where one of the adsorbate
components has been removed [51].
It has been reported that the more negative the adsorption
energies, the stronger the adsorbate-metal interaction [34]. It is

[(Fig._4)TD$IG]

clear from Table 5, that VI gave the maximum negative adsorption


energy. Therefore, VI is expected to exhibit greater inhibition
abilities as compared to AI.
In order to validate results obtained from the theoretical
studies, experimental evaluations using contemporary electrochemical techniques such as linear polarization resistance,
potentiodynamic polarization, electrochemical impedance spectroscopy (EIS) as well as surface analytical techniques like scanning
electronic microscopy (SEM) and atomic force microscopy (AFM)
were undertaken to study the inhibition performance of VI and AI
towards mild steel corrosion in 1 M HCl.
OCP versus time
Fig. 6(a) and (b) shows the variation of OCP of the working
electrode with time (1800 s) in 1 M HCl without and with optimum
concentration of VI and AI, respectively at 25  C. From the gures,
the OCP of 1 M HCl was found to be around 0.4675 V. This type of
behavior reects the breakdown of the pre-immersion air formed
oxide lm on the electrode surface and attack of electrolyte on bare
metal. The result is the attainment of a steady state potential which
corresponds to be corrosion of the bare metal [52]. This process
may be represented as follows:
Fe2 O3 6H 2e ! 2Fe2 3H2 O

(3)

With the adittion of 0.01 M VI and 0.01 M AI, the OCP shifted to a
noble direction indicating that VI and AI controls mainly anodic
metal dissolution reaction.
Potentiodynamic polarization measurements
The anodic and cathodic polarization curves for the corrosion of
carbon steel in 1 M HCl solution in the presence and absence of
varying concentrations of inhibitors (VI and AI) at 25  C are shown
in Fig. 7. The corrosion current densities and corrosion potentials
were calculated by extrapolation of linear parts of cathodic and
anodic curves to the point of intersection. The electrochemical
parameters such as corrosion potential (Ecorr), corrosion current
density (icorr), anodic Tafel slope (ba), and cathodic Tafel slope (bc),
and h (%) determined from polarization curves are summarized in
Table 6. All potentials were measured against SCE. h%, was
calculated using the equation [53]:

h%

iocorr  icorr
 100
iocorr
o

Fig. 4. Equilibrium adsorption conguration of (a) VI and (b) AI on Fe2O3(0 1 0)


surface.

(4)

where icorr and icorr are the values of corrosion current density in
the absence and presence of inhibitors, respectively.
Polarization measurements are suitable for monitoring the
progress and mechanisms of the anodic and cathodic partial
reactions as well as identifying the effect of an additive on either
partial reaction [54]. The data in Table 6 clearly show that the current
density decreases in the presence of the VI and AI investigated as
corrosion inhibitors. This is an indication that the inhibitors adsorbed
on the metal surface. It is also clear that there was more reduction in
corrosion current density in the presence of VI in all the
concentrations investigated than in the presence of AI. In acidic
media, the anodic reaction of corrosion is the passage of metal ions
from the metal surface into solution, and the cathodic reaction is the
discharge of hydrogen ions, which produces hydrogen gas or reduces
oxygen. The inhibitors may affect either the anodic reaction or the
cathodic reaction or both [55]. The anodic Tafel slope (ba) and
cathodic Tafel slope (ba) of VI and AI were observed to change
depending on the inhibitor concentration. Thus, the addition of
inhibitors in blank solution affected anodic (dissolution of carbon
steel) as well as cathodic (evolution of hydrogen) reactions. The

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._5)TD$IG]

1333

Fig. 5. Energy prole diagram for the adsorption progress of (a) VI and (b) AI on Fe2O3(0 1 0) surface.

variation in values of anodic and cathodic Tafel slopes in the


presence of inhibitors suggest that both the studied inhibitors are
mixed type inhibitors. No denite trend was observed in the shift of
Ecorr values in the presence of different concentrations of the VI and
AI, suggesting that the compounds behave as mixed-type inhibitors
[56]. However, the minor shift of Ecorr values toward positive
direction on increasing the concentration of inhibitors suggests the
predominant anodic control over the reaction. The order of
inhibitiors efciency obtained is VI > AI.

Linear polarization resistance (LPR) measurements


Stern and Geary on the basis of a detailed analysis of the
polarization curves of the anodic and cathodic reactions involved
in the corrosion of a metal, and on the assumption that both
reactions were charge-transfer controlled (transport overpotential
negligible) and that the IR drop involved in determining the
potential was negligible, derived the expression [57]:
Rp

Table 5
Outputs and descriptors calculated by the Monte Carlo simulation for adsorption of
VI and AI on Fe2O3 (0 1 0) surface.
Properties

VI

AI

Total energy (kJ mol1)


Adsorption energy (kJ mol1)
Rigid adsorption energy (kJ mol1)
Deformation energy (kJ mol1)
dEads/dNi

154.62
137.31
139.49
2.17
137.31

112.04
132.80
132.80
9.07
132.80

DE
ba bc

Di 2:3icorrba bc

(5)

where Rp is the polarization resistance determined at potentials


close to Ecorr, and ba, bc are the Tafel constants; note that in the case
of bc the negative sign is disregarded. This equation shows that the
corrosion rate is inversely proportional to Rp (or directly
proportional to the reciprocal slope of the DE versus Di curve)
at potentials close to Ecorr (typically within 10 mV), and that icorr
can be evaluated provided the Tafel constants are known.
Electrochemical corrosion kinetic parameters obtained from
polarization resistance (Rp) in 1 M HCl in the presence and absence
of VI and AI are presented in Table 7. The inhibition efciency can
be obtained as follows:

1334

[(Fig._6)TD$IG]

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._7)TD$IG]

Fig. 6. Variation of the open circuit potential (OCP) as a function of time, recorded
for a carbon steel in 1 M HCl at 25  C, in the absence and presence of highest
concentration of (a) vinyl imidazole and (b) allyl imidazole.

h% 1 

Rop
Rp

Fig. 7. Potentiodynamic polarization curves for carbon steel in 1 M HCl in the


absence and presence of different concentrations of (a) vinyl imidazole and (b) allyl
imidazole.

h%

!
 100

(6)

where Rop and Rp are the polarization resistances in the absence and
presence of inhibitors (VI and AI), respectively. The results
obtained show that the polarization resistance increases with an
increase in the concentration of VI at all concentrations but with AI,
polarization resistance increases up to 0.0075 M and thereafter
decreased when AI concentration was increased to 0.01 M in the
corrosive medium. VI has the highest inhibition efciency among
the two imidazole derivatives investigated using LPR technique.
Electrochemical impedance spectroscopy (EIS) measurements
Mechanistic information about the kinetics of elctrochemical
reactions at the surface can be obtained using EIS. This method is
most widely used to study the corrosion inhibition process. Figs. 8
and 9 show the impedance spectra represented in plots a, b and c
as Nyquist, Bode and phase angle plots, respectively, in the absence
and presence of different concentrations of VI (Fig. 8) and AI (Fig. 9)
for carbon steel corrosion in 1 M HCl. The charge transfer resistance
(Rct) and double-layer capacitance (Cdl) were obtained from the
impedance spectroscopy. The inhibition efciency is calculated
from the Rct values using the following formula [58]:

Roct  Rct
 100
Roct

(7)

where Roct and Rct are the charge-transfer resistances in the absence
and presence of the inhibitors, respectively. The calculated
electrochemical impedance parameters are given in Table 7. Data
from Table 7 show that Rct values increased while Cdl values
decreased with increase in concentrations for VI and up till
0.0075 M for AI. This may be due to the increase in the surface
coverage on the mild steel by the inhibitors, which led to the

Table 6
Potentiodynamic polarization parameters for carbon steel in 1 M HCl solution in the
absence and presence of vinyl imidazole and allyl imidazole.

h%

Ecorr
(mV/SCE)

icorr
(mA cm2)

bc

1 M HCl

470.9

72.8

101.1

Vinyl imidazole
0.001
0.0075
0.01

466.6
447.9
443.8

39.4
15.7
12.9

98.4
88.5
96.7

52.7
78.4
82.3

Allyl imidazole
0.001
0.0075
0.01

465.9
459.3
468.2

45.6
33.3
38.3

93.3
89.1
101.6

37.4
54.3
47.3

Inhibitor/concentration (M)

(mV dec1)

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

1335

Table 7
Electrochemical impedance and linear polarization parameters for carbon steel in 1 M HCl solution in the absence and presence of vinylimidazole and allylimidazole.

h%

Rp
(V cm2)

h%

3.45

314.5

461.3
1036.0
1306.0

1.93
0.37
0.24

27.3
67.6
74.3

464.7
1001.0
1415.0

32.3
68.6
77.8

435.8
555.3
444.4

2.05
1.31
2.94

22.9
65.5
24.5

410.1
497.8
444.9

23.3
36.8
29.3

Inhibitor/concentration (M)

Rs
(V cm2)

1 M HCl

2.97

15.1

0.85

335.6

Vinyl imidazole
0.001
0.0075
0.01

2.82
2.87
2.87

122.2
99.7
79.8

0.89
0.89
0.89

Allyl imidazole
0.001
0.0075
0.01

2.92
3.02
2.98

126.8
128.1
101.9

0.85
0.83
0.85

Yo
(V sn cm2)  106

increase in inhibition efciency. The the double-layer capacitance


(Cdl) is related to the thickness of the protective layer (d). This is in
accordance with the Helmholtz model and is given by the
following equation [59]:
C dl

eeo A
d

(8)

where e is the dielectric constant of the medium, e0 is the


permittivity of the free space (8.854  1012 F m1), and A is the
effective surface area of the electrode. The decrease in Cdl leads to
an increase in thickness of the double layer, which conrms that
the VI and AI molecules inhibit the corrosion by adsorption at the
mild steel/solution interface. The changes in Cdl values are due to
the replacement of water molecules by VI and AI.
The Nyquist plots (Figs. 8 and 9(a)) show depressed semicircles, indicating a non-ideal capacitive behavior of the electrochemical solid/liquid interface [60]. Such phenomenon is known as
the dispersing effect, usually attributed to the surface roughness,
the chemical in-homogeneities, the adsorption of inhibitor
molecules and the degree of polycrystallinity [61]. In addition,
the diameters of the capacitive loops increase with increasing
inhibitor concentration, which can be related to the increase of
surface coverage of inhibitive molecules on mild steel surface [53].
Thus, double layer behavior can be approximated by a constantphase element (CPE) rather than a pure capacitor. The CPE is
usually substituted for the capacitor to t the Nyquist depressed
semicircles more accurately. The admittance, YCPE, and impedance,
ZCPE of a CPE are expressed as follows [6264]:
Y CPE Y o jvn

(9)

and

Z CPE


1
1 
j  2pf max n
Yo

(10)

where Y0 is the amplitute comparable to a capacitance, j is the


square root of 1, fmax is the AC frequency at maximum and n, the
phase shift (1  n  1), when n = 0, the CPE represents pure
resistor, if n = +1, the CPE represents pure capacitor, and if n = 1,
the CPE represents inductor.
The bode plots for both VI and AI is shown in Figs. 8 and 9(b),
respectively. The increase of absolute impedance at low frequencies in Bode plots conrms the higher protection of steel with
increase in the concentration of the investigated inhibitors. This is
due to the adsorption of VI and AI on steel surface. In the same vein,
phase angle plots for VI and AI are depicted in Figs. 8 and 9(c),
respectively. It is evident from the plots that only one phase peak
close to 90 at the middle frequency point can be observed. This is
an indication that there is only a one time constant for VI and AI
[62]. Such observation may be related to the electrical double layer
formation at steel/solution interface [65,66].

Rct
(V cm2)

Cdl
mF cm2

The equivalent circuit used for the study is Randles circuit,


shown in Fig. 10. In this circuit, Rct is the charge transfer resistance,
Cdl is the double-layer capacitance, and Rs is the solution
resistance. The result of the three independent electrochemical
techniques employed are in good agreement with each other. VI
has a higher inhibition efciency than AI for mild steel corrosion in
1 M HCl.
Adsorption isotherm studies
Results so far obtained indicate that the primary mode of
interaction of VI and AI on steel surface is by adsorption. The
adsorption of organic inhibitor molecules from the aqueous
solution can be considered as a quasi-substitution process between
the organic compounds in the aqueous phase [Org(sol)] and water
molecules associated with the metallic surface [H2O(ads)] as
represented by the following equilibrium [67]:
Orgsol xH2 Oads $Orgads xH2 Osol

(11)

where x is the number of water molecules replaced by one organic


molecule. In this situation, the adsorption of VI and AI was
accompanied by desorption of water molecules from the mild steel
surface as conrmed by theoretical studies. The degree of surface
coverage (u) was evaluated from the potentiodyanamic polarization measurements as follows:

h%
100

(12)

It is necessary to determine empirically which adsorption


isotherm ts best to the surface coverage data in order to use the
corrosion rate measurements to calculate the thermodynamic
parameters pertaining to inhibitor adsorption. Attempts were
made to t surface coverage values into different adsorption
isotherm models. The models considered were [2]:
Temkinisothermexpf  u K ads C

(13)

Langmuirisothermu=1  u K ads C

(14)

Frumkinisothermu=1  u  exp2f  u K ads C

(15)

andFreundlichisothermu K ads C

(16)

where Kads is the equilibrium constant of the adsorption process, C


the inhibitor concentration and f the factor of energetic inhomogeneity. The correlation coefcient (R2) was used to choose the
isotherm that best t experimental data. Best results from the plots
were obtained for Langmuir adsorption isotherm. The linear
relationships of C/u versus C, depicted in Fig. 11, suggest that the

1336

[(Fig._8)TD$IG]

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._9)TD$IG]

Fig. 8. Impedance plots for carbon steel in 1 M HCl in the absence and presence of
different concentrations of vinyl imidazole exemplied as (a) Nyquist (b) Bode and
(c) phase angle plots.

adsorption of VI and AI on the steel surface obeyed the Langmuir


adsorption isotherm. This isotherm can be represented as:
C

1
C
K ads

(17)

The strong correlation (R2  0.96) of the Langmuir adsorption


for VI and AI was observed. This isotherm postulates that there is
no interaction between the adsorbed molecules and the energy of
adsorption is independent on the surface coverage (u). It assumes
that the solid surface contains a xed number of adsorption sites
and each holds one adsorbed species [68]. The slopes of the
straight lines obtained from the plots of Langmuir isotherm for VI

Fig. 9. Impedance plots for carbon steel in 1 M HCl in the absence and presence of
different concentrations of allyl imidazole exemplied as (a) Nyquist (b) Bode and
(c) phase angle plots.

and AI are more than unity (Table 8). So, it could be concluded that
each VI and AI unit occupies more than one adsorption site on the
steel surface. A modied Langmuir adsorption isotherm [69], could
be applied to this phenomenon, which is given by the corrected
equation:
C

n
nC
K ads

(18)

The divergence from pure monolayer adsorption can be


attributed to interactions between adsorbate species on the metal
surface as well as changes in the adsorption heat with increasing
surface coverage [70], factors which were not taken into

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

[(Fig._10)TD$IG]

1337

[(Fig._1)TD$IG]

Fig. 10. Equivalent circuit diagram used to t impedance data in the absence and
presence of VI and AI in 1 M HCl.

consideration in derivation of the isotherm.The free energy of




adsorption DGads of the inhibitors on aluminium surface was


determined using the following equation:


DGads RTlnK ads  55:5

(19)

Fig. 11. Langmuir adsorption isotherm for vinyl imidazole and allyl imidazole on
carbon steel in 1 M HCl 25  C.

where DGads is the standard free energy of adsorption, Kads is the


equilibrium constant of adsorption and the value of 55.5 is the
concentration of water in solution expressed in mol l1. The


calculated DGads and Kads results are also listed in Table 8.


The values of equilibrium adsorption constant obtained from
the Langmuir plots were 1.29  103 M1 and 3.18  103 M1, for VI


and AI, respectively. On the other hand, the values of DGads were
calculated to be 27.57 kJ/mol and 29.93 kJ/mol, for VI and AI,

Table 8
Langmuir adsorption parameters for carbon steel in 1 M HCl at 25  C.
Inhibitor
Vinyl imidazole
Allyl imidazole

DGads (kJ mol1)


27.57
29.93

Kads(M1)
3

1.29  10
3.18  103

[(Fig._12)TD$IG]

Fig. 12. SEM images of (a) polished steel (b) steel immersed in 1 M HCl (c) steel in 1 M HCl + VI (d) steel in 1 M HCl + AI.

Slope

R2

1.14
1.99

0.996
0.964

1338

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339


respectively. DGads values obtained in this study reveal that in the


presence of 1 M HCl, comprehensive adsorption involving both
physisorption and chemisorption adsorption of VI and AI on steel is
involved. The adsorption of VI and AI is not physisorption nor
chemisorption. This is because it is generally believed in the


literature that DGads values around 20 kJ mol1 or lower are


consistent with the electrostatic interaction between charged
organic molecules and the charged metal surface (physisorption);
those around 40 kJ mol1 or higher involve charge sharing or
transfer from the organic molecules to the metal surface to form a
co-ordinate type of bond (chemisorption).
Scanning electronic microscopy (SEM)
Fig. 12(a) depicts the morphologies of polished mild steel. SEM
micrographs obtained from mild steel surface after 18 h of
immersion in 1 M HCl for the untreated mild steel and treated mild
steel in 0.01 M VI and 0.01 M AI are shown in Fig. 12(b)(d),
respectively. It can be clearly observed that the mild steel surface was
strongly damaged with areas of uniform corrosionwhere the metal is
attacked more or less evenly over the entire surface. Examination on
the surface morphology of treated mild steel in 0.01 VI and 0.01 M AI

[(Fig._13)TD$IG]

reveals that the metal surface was in a better conditions by having


smooth surfaces compared to the untreated mild steel. However, VI
offered better protection to the steel surface in 1 M HCl than AI as
observed in Fig. 12(c) and (d). This may be due to the strong
interaction or adsorption of VI on mild steel surface forming a strong
barrier against corrosion.
Atomic force microscopy
The AFM is one of the foremost tools for imaging, measuring, and
manipulating matter at the nano- to micro-scale. It has become a
new method to investigate the nature of the protective layer formed
on the mild steel surface [71]. The AFM images of polished mild steel,
mild steel in 1 M HCl without and with 0.01 M VI and 0.01 M AI are
shown in Fig. 13(a)(d), respectively. It is evident from the gures
that the mild steel surface immersed in 1 M HCl appears severely
damaged than the steel surface immersed into 1 M HCl containing
the optimum concentration (0.01 M) of VI and AI. Moreover, the
values of the average roughness of mild steel immersed in blank HCl
solutionwas calculated to be 1600 nm. With the addition of VI and AI,
the average roughness was reduced to 156.3 and 296.8 nm,
respectively. This is an indication of protection abilities of the
inhibitors but with AI forming stronger lm on mild steel surface.

Fig. 13. 3D AFM images of (a) polished steel (b) steel immersed in 1 M HCl (c) steel in 1 M HCl + VI (d) steel in 1 M HCl + AI.

I.B. Obot et al. / Journal of Industrial and Engineering Chemistry 21 (2015) 13281339

Conclusions
We have employed DFT and MD simulations to predict and
ranked the inhibition efciencies of vinylimidazole and allylimidazole as possible green corrosion inhibitor for mild steel in 1 M
HCl. The ranking of inhibition efciency theoretically follows the
order: VI > AI. Electrochemical evaluations using LPR, PDP, EIS
validate the theoretical ranking. VI is a more potent corrosion
inhibitor than AI. Surface morphological studies using SEM and
AFM conrm that the steel surface containing VI in 1 M HCl was
more inhibited than the one in the presence of AI in the same
solution. This study has shown that theoretical calculations can be
used as a reliable approach to screen organic corrosion inhibitors
prior to experimental validation.
Acknowledgments
The authors gratefully acknowledged the Center of Research
Excellence in Corrosion (CORE-C), King Fahd University of
Petroleum and Minerals (KFUPM), Saudi Arabia for funding.
References
[1]
[2]
[3]
[4]
[5]
[6]
[7]
[8]
[9]
[10]
[11]
[12]
[13]
[14]
[15]
[16]
[17]
[18]
[19]
[20]
[21]
[22]
[23]

[24]

I.B. Obot, N.O. Obi-Egbedi, Mater. Chem. Phys. 122 (2010) 325.
I.B. Obot, N.O. Obi-Egbedi, S.A. Umoren, Corros. Sci. 51 (2009) 1868.
I.B. Obot, N.O. Obi-Egbedi, Corros. Sci. 52 (2010) 198.
M.M. Kabanda, L.C. Murulana, M. Ozcan, F. Karadag, I. Dehri, I.B. Obot, E.E.
Ebenso, Int. J. Electrochem. Sci. 7 (2012) 5035.
I.B. Obot, N.O. Obi-Egbedi, Curr. Appl. Phys. 11 (2011) 382.
X. Li, S. Deng, H. Fu, Corros. Sci. 51 (2011) 3241.
M. Bouklah, A. Attayibat, S. Kertit, A. Ramdani, B. Hammouti, Appl. Surf. Sci.
242 (2005) 399.
M. Kissi, M. Bouklah, B. Hammouti, M. Benkaddour, Appl. Surf. Sci. 252 (2005)
4190.
M.M. Kabanda, E.E. Ebenso, Int. J. Electrochem. Sci. 7 (2012) 8713.
E.E. Ebenso, M.M. Kabanda, L.C. Murulana, A.K. Singh, S.K. Shukla, Ind. Eng.
Chem. Res. 51 (2012) 12940.
M.M. Kabanda, I.B. Obot, E.E. Ebenso, Int. J. Electrochem. Sci. 8 (2013) 10839.
I.B. Obot, E.E. Ebenso, N.O. Obi-Egbedi, A.S. Afolabi, Z.M. Gasem, Res. Chem.
Intermed. 38 (2012) 1761.
Z. Zhang, S. Chen, Y. Li, S. Li, L. Wang, Corros. Sci. 51 (2009) 291.
H. Otma9
ci
c, E. Stupniek-Lisac, Electrochim. Acta 51 (2003) 985.
E. Stupniek-Lisac, A. Gazivoda, M. Madarac, Electrochim. Acta 47 (2002)
4189.
H.O. Curkovic, E. Stupnisek-Lisac a, H. Takenouti, Corros. Sci. 52 (2010) 398.
L. Larabi, O. Benali, S.M. Mekelleche, Y. Harek, Appl. Surf. Sci. 253 (2006) 1371.
O. Benali, L. Larabi, Y. Harek, J. Saudi Chem. Soc. 14 (2010) 231.
O. Benali, L. Larabi, M. Traisnel, L. Gengembre, Y. Harek, Appl. Surf. Sci. 253
(2007) 6130.
N. Kovacevic, A. Kokalj, Corros. Sci. 53 (2011) 909.
N. Kovacevic, A. Kokalj, J. Phys. Chem. C 115 (2011) 24189.
J.O. Mendes, E.C. da Silva, A.B. Rocha, Corros. Sci. 57 (2012) 254.
D.B. Hmamou, R. Salghi, A. Zarrouk, M.R. Aouad, O. Benali, H. Zarrok, M.
Messali, B. Hammouti, M.M. Kabanda, M. Bouachrine, E.E. Ebenso, Ind. Eng.
Chem. Res. 52 (2013) 14315.
J. Zhang, G. Qiao, S. Hu, Y. Yan, Z. Ren, L. Yu, Corros. Sci. 56 (2011) 176.

1339

[25] R.S. Oguike, A.M. Kolo, A.M. Shibdawa, H.A. Gyenna, ISRN Phys. Chem. (2013),
doi:http://dx.doi.org/10.1155/2013 /175910.
[26] S. John, J. Joy, M. Prajila, A. Joseph, Mater. Corros. 62 (2011) 1031.
[27] I.B. Obot, E.E. Ebenso, M.M. Kabanda, J. Environ. Chem. Eng. 1 (2013) 431.
[28] J. Zhang, F. Niu, C. Li, M, Du, J. Surfact. Deterg. (2014), doi:http://dx.doi.org/
10.1007/s11743-013-1515-8.
[29] K.F. Khaled, M.A. Amin, Corros. Sci. 51 (2009) 2098.
[30] M.G.V. Satyanarayana, V. Himabindu, Y. Kalpana, M.R. Kumar, K. Kumar, J. Mol.
Struct. (Theochem.) 912 (2009) 113.
[31] Materials Studio, Revision 6.0, Accelrys Inc., San Diego, USA, 2011.
[32] A.Y. Musa, R.T.T. Jalgham, A.B. Mohamad, Corros. Sci. 56 (2012) 176.
[33] S. John, M. Kuruvilla, A. Joseph, RSC Adv. 3 (2013) 8929.
[34] K.F. Khaled, J. Appl. Electrochem. 41 (2011) 423.
[35] K.F. Khaled, J. Chim. Acta. 1 (2012) 66.
[36] N.O. Obi-Egbedi, I.B. Obot, M.I. El-Khaiary, J. Mol. Struct. 1002 (2011) 86.
[37] E.E. Ebenso, M.M. Kabanda, T. Arslan, M. Saracoglu, F. Kandemirli, L.C.
Murulana1, A.K. Singh, S.K. Shukla, B. Hammouti, K.F. Khaled, M.A. Quraishi, I.B.
Obot, N.O. Eddy, Int. J. Electrochem. Sci. 7 (2012) 5643.
[38] M.M. Kabanda, L.C. Murulana, M. Ozcan, F. Karadag, I. Dehri, I.B. Obot, E.E.
Ebenso, Int. J. Electrochem. Sci. 7 (2012) 5035.
[39] N.O. Obi-Egbedi, I.B. Obot, M.I. El-Khaiary, S.A. Umoren, E.E. Ebenso, Int. J.
Electrochem. Sci. 7 (2012) 56495675.
[40] I.B. Obot, N.O. Obi-Egbedi, A.O. Eseola, Ind. Eng. Chem. Res. 50 (2011) 2098.
[41] A. Aytac, S. Bilgic, G. Gece, N. Ancin, S.G. Oztas, Mater. Corros. 63 (8) (2012) 729.
[42] R.G. Pearson, J. Chem. Educ. 64 (1987) 561.
[43] M.K. Awad, M.R. Mustafa, M.M. Abo Elnga, J. Mol. Struct. (Theochem.) 959
(2010) 66.
[44] R.G. Parr, L. Sventpaly, S. Liu, J. Am. Chem. Soc. 121 (1999) 1922.
[45] Y.M. Tang, Y. Chen, W.Z. Yang, W. Liu, Z.S. Yin, J.T. Wang, J. Appl. Electrochem. 38
(2008) 1553.
[46] K. Fukui, Science 218 (1982) 747.
[47] Y. Tang, X. Yang, W. Yang, Y. Chen, R. Wan, Corros. Sci. 52 (2010) 242.
[48] S. Xia, M. Qiu, L. Yu, F. Liu, H. Zhao, Corros. Sci. 50 (2008) 2021.
[49] Y. Tang, L. Yao, C. Kong, W. Yang, Y. Chen, Corros. Sci. 53 (2011) 2046.
[50] L. Feng, H. Yang, F. Wang, Electrochim. Acta 58 (2011) 427.
[51] K.F. Khaled, J. Chim. Acta 1 (2012) 59.
[52] D.K. Yadav, D.S. Chauhan, I. Ahamad, M.A. Quraishi, RSC Adv. 3 (2013) 632.
[53] S.A. Umoren, Z.M. Gasem, I.B. Obot, Ind. Eng. Chem. Res. 52 (2013) 14855.
[54] E.E. Oguzie, C.E. Ogukwe, J.N. Ogbulie, F.C. Nwanebu, C.B. Adindu, I.O. Udeze, K.
L. Oguzie, F.C. Eze, J. Mater. Sci. (2012) 3592.
[55] X. Li, S. Deng, H. Fu, G. Mu, Corros. Sci. 51 (2009) 620.
[56] M.A. Hegazy, Corros. Sci. 51 (2009) 2610.
[57] M. Stern, A.L. Geary, J. Electrochem. Soc. 104 (1957) 56.
[58] K.F. Khaled, M.M. Al-Qahtani, Mater. Chem. Phys. 113 (2009) 150.
[59] C. Bataillon, S. Brunet, Electrochim. Acta (1994) 455.
[60] S. Kumar, D. Sharma, P. Yadav, M. Yadav, Ind. Eng. Chem. Res. (2013) 14019.
[61] O. Ghasemi, I. Danaee, G.R. Rashed, M.R. Avei, M.H. Maddahy, J. Cent. South
Univ. 20 (2013) 301.
[62] D.K. Yadav, M.A. Quraishi, Ind. Eng. Chem. Res. 51 (2012) 14966.
[63] T.J. Edison, M.G. Sethuraman, ISRN Electrochem. (2013), doi:http://dx.doi.org/
10.1155/2013/256086.
[64] A.K. Satapathy, G. Gunasekaran, S.C. Sahoo, K. Amit, P.V. Rodrigues, Corros. Sci.
51 (2009) 2848.
[65] R.A. Bustamante, G.N. Silva, M.A. Quijano, H.H. Hernandez, M.R. Romo, A. Cuan,
M.P. Pardave, Electrochim. Acta 54 (2009) 5393.
[66] M. Outirite, M. Lagrenee, M. Lebrini, M. Traisnel, C. Jama, H. Vezin, F. Bentiss,
Electrochim. Acta 55 (2010) 1670.
[67] N.A. Negm, Y.M. Lkholy, M.K. Zahran, S.M. Tawk, Corros. Sci. 52 (2010) 3523.
[68] S.A. Ali, M.T. Saeed, S.U. Rahman, Corros. Sci. 45 (2003) 253.
[69] R.F.V. Villamil, P. Corio, J.C. Rubin, S.M.L. Agostinho, J. Electroanal. Chem. 472
(1999) 112.
[70] E.E. Oguzie, Y. Li, F.H. Wang, Colloid Interface Sci. 310 (2007) 90.
[71] M.A. Quraishi, A. Singh, V. Singh, D.K. Yadav, A.K. Singh, Mater. Chem. Phys. 122
(2010) 114.

You might also like