10 1016@j Corsci 2016 02 036

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 39

Accepted Manuscript

Title: Development of a predictive model for corrosion


inhibition of carbon steel by imidazole and benzimidazole
derivatives

Author: Evelin Gutiérrez José A. Rodrı́guez Julián


Cruz-Borbolla José G. Alvarado-Rodrı́guez Pandiyan
Thangarasu

PII: S0010-938X(16)30079-8
DOI: http://dx.doi.org/doi:10.1016/j.corsci.2016.02.036
Reference: CS 6673

To appear in:

Received date: 3-2-2016


Revised date: 17-2-2016
Accepted date: 22-2-2016

Please cite this article as: Evelin Gutiérrez, José A.Rodríguez, Julián Cruz-Borbolla,
José G.Alvarado-Rodríguez, Pandiyan Thangarasu, Development of a predictive model
for corrosion inhibition of carbon steel by imidazole and benzimidazole derivatives,
Corrosion Science http://dx.doi.org/10.1016/j.corsci.2016.02.036

This is a PDF file of an unedited manuscript that has been accepted for publication.
As a service to our customers we are providing this early version of the manuscript.
The manuscript will undergo copyediting, typesetting, and review of the resulting proof
before it is published in its final form. Please note that during the production process
errors may be discovered which could affect the content, and all legal disclaimers that
apply to the journal pertain.
Development of a predictive model for corrosion
inhibition of carbon steel by imidazole and benzimidazole
derivatives

Evelin Gutiérreza, José A. Rodrígueza, Julián Cruz-Borbollaa*, José G. Alvarado-


Rodrígueza, Pandiyan Thangarasub

a
Área Académica de Química, Universidad Autónoma del Estado de Hidalgo, Unidad
Universitaria, km 4.5 Carretera Pachuca-Tulancingo, C.P. 42184 Pachuca-Hidalgo,
México.

b
Facultad de Química, Universidad Nacional Autónoma de México, Ciudad Universitaria,
México D.F., C.P 04510, México.

*
Corresponding author; jcruz@uaeh.edu.mx
Graphical Abstract

Highlights

 Different functional groups in imidazole inhibitors modify the corrosion behavior.


 Molecular properties can be associated to corrosion inhibition efficiency.
 Theoretical model predicts experimental results adequately.

Abstract

The corrosion inhibition efficiency of carbon steel was evaluated using fifteen molecules
derivated from imidazole and benzimidazole in 1.0 M HCl. The inhibition efficiency varies
from 15.6 to 89.4%, depending on the chemical and the electronic structure of the
compounds. The structural properties of the molecules under study were determined by
density functional theory. Experimental and theoretical parameters were used to construct a
model to predict the inhibition observed. The proposed model shows that inhibition
efficiency is related to the molecule volume, charge, electronegativity, and aromaticity. The
model is successfully used to predict the corrosion behavior of two additional molecules.

Keywords: Carbon steel; EIS, DFT, QSPR; Acid corrosion


1. Introduction

Corrosion of materials is one of the main problems in industry that is associated to


significant economic losses. Corrosion causes plant shutdowns, waste of valuable
resources, loss or contamination of the product, reduction in efficiency, increase of
maintenance needs, and expensive overdesign [1,2].

Hydrochloric acid used in pickling operations for carbon, alloy, and stainless steels is used
both to remove rust and scale and undesirable carbonate deposits in oil wells to encourage
the flow of crude oil or gas to the well [3]. Corrosion can be controlled by using protective
systems, devices, and treatments based on the corrosion mechanism. An effective
alternative to reduce the aggressiveness of acidic media in metals is the use of organic
inhibitors dissolved in the corrosive solution [4]. Protecting a metal involves the formation
of a film on the metal surface by an organic inhibitor [5] in order to reduce or prevent
contact with the acidic solution. Different organic compounds have been described as
corrosion inhibitors. The main interactions between these compounds and the metal surface
are [6]: i) electrostatic attraction between charged molecules and charged metal, ii)
interaction of unshared electron pairs in the molecule with metal, and iii) interaction of π
electrons with metal.

The effectiveness of corrosion inhibitors is associated to the nature of the functional groups
contained in the structure, spatial molecular structure, molecular electronic structure, and
affinity for metal surface. Additional to specific interactions between functional groups and
metal surface, the presence of π bonds and heteroatoms such as oxygen, nitrogen, and
sulfur enhance said effectiveness due to the presence of free electron pairs [5]. Nowadays
there is a growing interest to find environment friendly inhibitors [7].

Imidazole and its derivatives are considered nontoxic compounds [8,9] and they have been
studied and recognized as corrosion inhibitors. It is described that the inhibiting efficiency
increases by substituting alkyl or aryl groups in the imidazole moiety [8]. Imidazole has
different sites for interaction with the metal surface such as two nitrogen atoms, an
aromatic character, and the option to include different substituents in the molecule that
promote the interaction [10].
An alternative to study the corrosion process is the combination of experimental data with
the properties obtained by molecular modeling [11]. Density Functional Theory (DFT)
allows to calculate molecular properties such as the highest occupied molecular orbital
energy (EHOMO), the lowest unoccupied molecular orbital (ELUMO), dipole moment (µD),
electronegativity (χ), and hardness (η) among others and analyze their correlation with the
corrosion inhibition effect [12-14]. These properties are associated to inhibition efficiency
by multivariate methods.

Quantitative structure activity relationship and quantitative structure property relationship


are the most common models used for the study of steel corrosion in acidic media [15, 16].
The main drawback of these models is that they were constructed using experimental data
previously reported. The inhibition efficiencies reported are obtained using different
methodologies (electrochemical or gravimetric), different corrosion media, electrode areas,
and concentrations of the inhibitor and acidic solution.

The present work proposes the analysis of a group of molecules derived from imidazole and
benzimidazole in homogenous conditions in order to generate a model to determine the
corrosion inhibition efficiency. The inhibition rates are estimated by electrochemical
impedance spectroscopy employing fifteen structures among imidazole and benzimidazole,
and corroborate by potentiodynamic polarization measurements. Subsequently, a relation
with descriptors calculated from DFT is proposed to predict the phenomenon of corrosion
inhibition efficiency.

2. Experimental

2.1 Reagents and solutions

Solutions of imidazole and benzimidazole derivatives (0.1, 0.5, 1.0, 5.0 and 10.0 mM)
were prepared by dissolving the respective analytical grade reagent (Sigma-Aldrich, St.
Louis, Mo, USA) in 1.0 M HCl (Baker, Mexico) using deionized water obtained from a
Milli-Q Plus system (Millipore, Bedford, MA, USA) with a resistivity not less than 18.2
MΩ cm. A blank of 1.0 M HCl was used to evaluate the effect of the compounds as
corrosion inhibitors.

Table 1 shows the structures of imidazole and benzimidazole derivatives selected for the
study. They were selected considering the presence of heteroatoms, of unshared electron
pairs, and of π electrons in the molecule that can interact with the metal surface by
adsorption. The existence of halogens, electron donating and electron-withdrawing groups,
aromatic rings, different volumes and steric factor were also considered. Some of these
characteristics are associated to a positive effect on corrosion inhibition [17-19].

2.2 Electrochemical analysis

The assays were performed using a three electrode electrochemical cell. A carbon steel bar
AISI 1018 (0.32 cm2, 0.15-0.2 % C, 0.6-0.9 % Mn, 0.04 % P, 0.05 % S, as remainder %
Fe), platinum and silver-silver chloride (Ag/AgCl, 3.0 M KCl) were used as working,
auxiliary, and reference electrodes respectively. In order to renew the exposed area of the
working electrode, the surface was abraded sequentially using emery papers of 80, 600 and
1200 grid. Afterwards, it was cleaned with acetone to remove impurities and washed with
deionized water.

Electrochemical studies were performed in an Autolab electrochemical system (Autolab


PGSTAT 30, Eco Chemie B. V., Netherlands) equipped with a FRA module and controlled
by GPES and FRA software version 4.5 and 2.4, respectively. The experiments were
performed at room temperature.

Open circuit potential was measured during one hour in freshly prepared solutions of
imidazole and benzimidazole derivatives (1.0 mM) in 1.0 M HCl. Impedance
measurements were performed at open circuit potential (Eocp vs Ag/AgCl) over a frequency
range from 0.1 Hz to 1 kHz with a 10 mV peak to peak amplitude. The working electrode
was immersed in 1.0 M HCl for 30 min before each experiment with and without the
addition of inhibitors and without stirring. The inhibition for each concentration, analyzed
by electrochemical impedance spectroscopy, were calculated by using the following
equation [1,20]:

0
𝑅𝑐𝑡 −𝑅𝑐𝑡
%Inhibition = 100 ( ) (1)
𝑅𝑐𝑡

0
where 𝑅𝑐𝑡 and 𝑅𝑐𝑡 are the charge transfer resistances in the presence and absence of the
inhibitor compound, respectively.

Polarization curves were obtained through an anodic scan using linear sweep voltammetry
obtained from -0.67 to -0.26 V (vs Ag/AgCl) at 1.0 mV s-1. Corrosion potential (Ecorr, V, vs
AgAgCl) and corrosion current density (icorr, A cm-2) were obtained by extrapolation of
cathodic (βc) and anodic (βa) branches of Tafel curves. The inhibition was calculated by
employing the equation:

i0 -icorr
%Inhibition=100 ( corri0 ) (2)
corr

0
where 𝑖𝑐𝑜𝑟𝑟 and 𝑖𝑐𝑜𝑟𝑟 are corrosion current densities values without and with inhibitor,
respectively [21].

2.3 Computational procedure

DFT calculations for imidazole and benzimidazole derivatives were carried out by using the
Gaussian 09 suite [22]. The structures of the molecules used as corrosion inhibitors were
optimized with the Perdew, Burke, and Ernzerhof (PBE) [23,24] exchange correlation
functional, in all cases we used the 6-311++G** orbital basis set for the atoms [25]. This
DFT functional-basis set combination yielded reliable and consistent results in a wide
variety of chemical systems [26,27]. Quantum descriptors such as dipole moment (µD),
aromaticity indexes (bqISO, bqANS), ionization potential (I), electron affinity (A),
electronegativity (χ), hardness (ɳ), electrophilicity (ω*), electron donor capacity (ω-),
electron acceptor capacity (ω+), charge of nitrogen atom in positions 1 and 3, and of carbon
atom in position 2 of imidazole ring (qN1, qN3, qC2, respectively), and number of electrons
transferred (ΔN) were calculated for each molecule in the gas phase. It was reported that
there are not significant differences between calculations carried out in gas phase when
compared to aqueous phase [5]. Other descriptors, such as partition coefficient (log P) and
volume (V), were calculated with the program Spartan 08 V1.2.0 and Multiwfn V3.2 [28]
respectively. Minimum energy conformations were corroborated by calculating the second
derivative which showed positive frequencies.

Multiple linear regression models were obtained using MobyDigs software (version 1.0,
2014, TALETE srl). The model was constructed and validated by using a training set of
fifteen molecules and a test set of two compounds. The descriptors estimated were used as
independent variables while the inhibition achieved using a 10.0 mM solution was used as
dependent variable. The concentration used was selected because it showed the maximum
inhibition efficiency at that value for evaluated compounds.

3. Results and discussion

3.1 Open circuit potential

In order to determine the adequate immersion time, an open circuit potential study was
performed. Figure 1 shows the open circuit potentials of carbon steel surface measured in
1.0 M HCl (blank) free of inhibitor and with corrosion inhibitors (1.0 mM): (a) 2-
methylimidazole, (b) 4-(imidazol-1-yl) phenol, (c) 2-(aminomethyl) benzimidazole, (d) 6-
bromo-1H-benzimidazole. In inhibitor free solution, the potential decreases due to the
dissolution of the metal surface. In presence of inhibitors, the potential value changed
during the first 10 minutes of immersion. After this period, the potentials become almost
stable.

The efficiency of an inhibitor can be associated to the steady state potential value.
Therefore, a lower potential implies an increment of the metal dissolved and in
consequence the inhibitor is less active. The inhibition tendency of the compounds
evaluated in this section is: 6-bromo-1H-benzimidazole > 4-(imidazol-1-yl) phenol > 2-
(aminomethyl) benzimidazole > 2-methylimidazole. Since the potential is almost stable
after 10 minutes, an immersion potential of 30 minutes is selected to perform corrosion
experiments.
3.2 Electrochemical impedance spectroscopy

The phenomenon of inhibiting corrosion of carbon steel using organic inhibitors (imidazole
and benzimidazole derivatives) was investigated and quantified by electrochemical
impedance spectroscopy technique, which represents an effective method for analyzing the
characteristics of the surface of an electrode.

Each inhibitor molecule was analyzed at five concentration levels (0.1-10.0 mM). The
results obtained for the 15 molecules are presented in Table 2. The Nyquist plots for the
fifteen molecules show a depressed semicircle with intersections on the real axis (Z’),
which implies that the electron transfer process is a determining step during corrosion of
carbon steel in 1.0 M HCl [29]. This mechanism is associated to solid electrodes and
referred to frequency dispersion, attributed to the surface heterogeneity due to impurities,
distribution of activity centers, adsorption of inhibitor species and the presence of porous
layers [19]. The electron transfer resistance (𝑅𝑐𝑡 ) results from the intersection on the real
axis at low and high frequencies.

Figure 2 shows the Nyquist plots for 4-(imidazol-1-yl) phenol and 6-bromo-1H-
benzimidazole (Fig. 2(b) and 2(d), respectively). These molecules have the highest
inhibition efficiencies at a concentration of 10.0 mM. In contrast, Fig 2(a) and 2(c) shows
the plot obtained with 2-methylimidazole and 2-(aminomethyl) benzimidazole,
respectively. The inhibition efficiency of these compounds is the lowest at the same
experimental conditions. It is clear that the presence of the substituents in the structure of
the molecule has an effect on the inhibition properties.

Bode and phase angle diagrams for carbon steel exposed to 1.0 M HCl in presence and
absence of inhibitors selected are showed in Figure 3.The effect of inhibition is observed
from the low frequency impedance modulus. The low frequency impedance modulus (Z)
increases with the concentration of inhibitor, this phenomenon is related to the adsorption
of organic compound on the exposed surface [30]. The impedances at low frequencies
increase with the concentration of organic inhibitor. Highest impedance values are
associated with better inhibition efficiency, the sequence obtained is then equal than the
observed by the open potential circuit experiments.

According to Figure 3, the employed reagents presented an effect of inhibition associated to


their concentration in 1.0 M HCl, A depression of phase angle with decrease of inhibitor
concentration involves a decrease of capacitive response associated to higher corrosion
rates at low concentrations or absence of inhibitor [31].

The results obtained by electrochemical impedance spectroscopy are fitted to the electrical
equivalent circuit showed in Figure 4. The equivalent circuit employed consists of a parallel
combination of the charge-transfer resistance (𝑅𝑐𝑡 ) and a constant phase element (CPE).
This combination is collocated in series with the solution resistance (𝑅𝑠 ).

The impedance function of the CPE is defined as:

1
ZCPE  (3)
Q  i 
n

where Q is a magnitude of the CPE, n corresponds to a phase shift, ω is the angular


frequency (ω=2πf where f is the frequency at which the imaginary part of impedance is
highest) and i is the imaginary number. The double layer capacitance (Cdl) for the proposed
circuit is determined by using the following equation:

Cdl  Q  
n 1
(4)

Nyquist and Bode fitted diagrams for the molecules evaluated in section 3.2 are constructed
according to the circuit proposed. Figure 5 shows the Nyquist and Bode plots of the
experimental and fitted data at two concentrations (10.0 mM and 0.5 mM). Both diagrams
are similar, which corroborate the useful of the proposed circuit.

Values of Rct and Cdl are presented in Table 2. The values of Rct increase with the
concentration whilst Cdl values generally decreases. Cdl tendency is not clear in the case of
bad inhibitors such as 2-methylimidazole. The decrease of Cdl is associated to the increase
in thickness of electrical double layer as a result of adsorption of the inhibitor on carbon
steel [6,30,32]. Nyquist plots show a similar shape at each concentrations of inhibitor,
therefore the mechanism of corrosion is independent of the inhibitor used [33].

3.3 Potentiodynamic polarization measurements

Polarization curves were determined for 2-methylimidazole, 4-(imidazol-1-yl) phenol, 2-


(aminomethyl) benzimidazole and 6-bromo-1H-benzimidazole. Figure 6 shows the
polarization curves corresponding to analysis. Table 3 contains the electrochemical
parameters obtained by extrapolating the Tafel slopes. It is possible to observe a tendency
in the reduction of corrosion intensity with the concentration of compounds employed as
corrosion inhibitors.

The data obtained by potentiodynamic polarization, in absence and presence of the organic
compounds employed, demonstrate a similar tendency in inhibition efficiency. Once
corroborated the experimental data the construction of a theoretical model to predict
corrosion inhibition is performed.

Although the molecules selected are not good inhibitors, there is a difference between the
inhibition behavior of each compound. Structural differences such as long alkyl chains,
volume and aromaticity are mentioned as critical parameters in corrosion studies [4,6,19].
The structures derived from imidazole and benzimidazole, employed in present
investigation, are capable of modifying the corrosion of metallic surfaces. According to
Table 2, the inhibitors that contain halogens in their structure have inhibition efficiencies
higher than 55% at concentration 10.0 mM. In amines, the efficiency of compounds
containing chloride and bromide as substituents has been explained by negative inductive
effect and positive mesomeric effect in adsorption processes [34]. The effect of substituents
like chloride and fluoride in structures of imidazoline demonstrate that the presence of
halogens affects the adsorption, which is associated also to a weaker induction effect [35]
or steric effect [17].

Aromaticity is another factor with a positive contribution. The effect is evidenced in the
efficiencies obtained with benzimidazoles as well as with 2-(aminomethyl) benzimidazole
which has an efficiency of 69.1% at 10.0 mM, compared to structures like imidazole that do
not contain aromatic substituents. For example, 2-methylimidazole and 2-chloro-1H-
imidazole show 30.2%. and 57.1% efficiency of corrosion inhibition, respectively.
Inhibitors can also bond to metal surfaces by electron transfer to the metal to form a
coordinate type of bonding. Electron transfer from the adsorbed species is favored by
electron withdrawing groups in aromatic rings [1].

There are two important considerations at the moment of analyzing the effect of corrosion
inhibition in a molecule: the effect of the electronic structure, which determines the
electronic density of the molecule reaction center; and the effect of the chemical structure,
which considers structural properties of the molecule such as volume, surface area, and
disposition of substituents in the tridimensional space [36].

3.4 Adsorption isotherm

The key factor in the efficiency of an organic compound that is used as a corrosion inhibitor
is its ability to be absorbed on the surface of a metal [37]. An alternative to the association
of adsorption to a physical or chemical phenomenon is the use of adsorption isotherms,
which are capable of providing information about the mechanism through which the
inhibitor interacts with the metal surface [38]. To use the adsorption isotherms, it is
necessary to do an analysis of the degree of surface coverage (θ, dimensionless). This is
calculated as a function of changes in Rct as shown in the following equation:

Rct  Rct0
 (5)
Rct

The Langmuir (6), Frumkin (7) and Temkin (8) isotherms, allow to associate the inhibitor
concentration (C, M) with the value of θ calculated [39-41]. The Langmuir isotherm
implies that one adsorbate covers only one substrate site. This isotherm considers that the
surface free energy on all of the sites on the substrate is equal, and that the adsorbate
species do not interact with each other even when the surface coverage is closed to be
completed. If the concentration of the adsorbate is high and there are attraction or repulsive
interactions between species, then is necessary to employ Frumkin or Temkin isotherms.
The Frumkin isotherm contemplates the attractive or repulsive interactions of the adsorbed
molecular species, and assumes that adsorption sites are uniform. Temkin isotherm
considers that there is no molecular interaction between adsorbed species but the free
energy of surface sites vary with surface coverage. The mathematical models which
describe the adsorption behavior are:

C 1
 C Langmuir isotherm (6)
 K ads

  
log    log K ads  g Frumkin isotherm (7)
 1    C 

 
log    log K ads  g Temkin isotherm (8)
C 

where Kads is the equilibrium constant, and g is the molecular interaction parameter. In the
Frumkin isotherm g defines the attractive (negative value) or repulsive (positive value)
interactions between the adsorbed molecular species. In Temkin isotherm g relates to the
surface free energy changes in the surface coverage [42].

The mechanism of adsorption of the molecules used is investigated by the construction of


the isotherms mentioned before. Langmuir isotherms were selected because of the values
obtained for the determination coefficient (R2) (0.996-0.999). The R2 values obtained with
Frumkin and Temkin models are in the range of 0.362-0.983 and 0.631-0.986, respectively.

In Table 4 is possible to observe the determination coefficient (R2) calculated employing


the Langmuir isotherm for each inhibitor, as well as the Kads values. Kads values indicate an
interaction between the inhibitor and the carbon steel exposed to 1.0 M HCl [43]. The
interaction can be associated to lone pairs of electrons of nitrogen atoms, π-electrons in
each of the studied compounds, and the vacant d-orbitals of iron surface atoms [5].

The benzimidazole derivatives exist in the form of either neutral or cationic molecules in
acidic media. In the form of neutral molecules, benzimidazole derivatives can adsorb by
forming coordinate bonds between molecules and Fe atoms. In cationic form, molecular
adsorption is predominantly related to electrostatic interaction between the protonated
compounds and the steel surface negatively charged by the adsorption of chloride ions
[30,44].

3.5 DFT Calculations

The structure and molecular properties of corrosion inhibitors play an important role in the
mechanism of adsorption on the metal surface. The structures of the molecules were
optimized in gas phase with PBE/6-311++G**. All the calculated structural parameters for
this type of heterocyclic compounds, regardless the functional/basis used, are essentially
identical to those reported in the literature [26-27,45,46].

The dipole moment (μD), which is one of the quantum parameters further investigated in the
corrosion inhibition, allows for the discussion of the structure and reactivity of chemical
systems. High values of µD are associated to an increase in the probability of inhibitor
adsorption [39]. Some molecules have a positive effect on the inhibition of the process of
corrosion in metallic surfaces in spite of not obeying this tendency strictly in terms of
orbital energies (EHOMO, ELUMO) and µD [47].

The first ionization potential (I) and electron affinity (A) are some of the properties in the
system that measures the ability to donate or accept electrons. I and A are calculated
employing equations 9 and 10 respectively [48].

I  EcN 1  E0N (9)

A  EoN  EAN 1 (10)

where:

EcN 1 Energy after losing one electron (cation)

EoN Energy of the molecule in its basal state (neutral)

E AN 1 Energy after gain one electron (anion)


There is a relation between potential of ionization and electron affinity, which is associated
to EHOMO and ELUMO, respectively. In structures of thiophene, the inhibition efficiency
increases with the increase of EHOMO that is the increase of I [49].

Electronegativity (χ) is obtained from the calculation of I and A.


 I  A (11)
2

Hardness (η) can be understood as a resistance to charge transference in the system, and it
is associated I and A as:


 I  A (12)
2

In the application of corrosion inhibitors, the inhibitor with the lower value of η is expected
to display the highest value of corrosion inhibition efficiency. The adsorption by transfer of
electrons can occur at the part of the molecule where η value is minimum [50].

Electrophilicity (ω*) represents the stabilization energy of the system when it is saturated
by electrons coming from the surroundings. This index measures the propensity of
chemical species to accept electrons. ω* can be determined from chemical potential (μ) and
η, where µ is defined as µ=-χ. The following equation is used for determining ω* [51]:

(𝜒)2
𝜔∗ = (13)
2𝜂

In the application of tetrazole derivatives as corrosion inhibitors, the structure with the
lowest electrophilicity index value display the highest inhibition efficiency [50].

Two descriptors related to the ability of a specie to donate or accept electrons are included,
the electron donor capacity (ω-, equation 14) and electron acceptor capacity (ω+, equation
15) [52]:

 3I  A   I  3 A
2 2

 

(14)  

(15)
16  I  A 16  I  A
The aromaticity indexes (bqISO, bqANS) are calculated with the Nucleus-Independent
Chemical Shifts. Data are obtained by means of a single-point calculation in a system
formed by a test charge positioned at 1 Å from the ring critical point (RCP) of the aromatic
molecule [53,54]; the RCP is calculated according to the atoms in molecules theory [55,
56].

Other descriptors associated to the effect of inhibiting corrosion are aromaticity, partition
coefficient (log P), charge of the nitrogen atoms (qN1 and qN3), and charge of the carbon 2
in imidazole structure (qC2, indicated on structures of Table 1 with asterisk). Electrostatic
interaction is demonstrated to affect the adsorption of the molecules on the metallic surface.
This effect is observed during the study of amines, thioureas, and alcohols used as
inhibitors in corrosion experiments [57].

The number of electrons transferred (ΔN) is calculated using the following equation:

𝜒 𝜒
Δ𝑁 = 2(𝜂Fe−+𝜂inh )…….(16)
Fe inh

where χFe and χinh correspond to the electronegativity of steel and of the inhibitory molecule,
respectively. ηFe and ηinh correspond to the absolute hardness of the steel and the inhibitory
molecule, respectively. To calculate the fraction of electron transferred, a theoretical value
for electronegativity of bulk iron is used (χFe ≈7 eV mol-1) and a global hardness of ηFe of 0
eV mol-1. The value of ΔN<3.6 indicates a tendency of the molecule to donate electrons to
the metallic surface. In this case the inhibition efficiency increases along with the increase
of the electron donating ability of these inhibitors to the metal surface [58,59]. Piperidine
and benzimidazole derivatives employed as corrosion inhibitors show a correlation with
experimental inhibition efficiencies and the fraction of electron transferred [60,61].
However, it is shown that there is not a regular trend in the inhibition efficiency by
increasing values of ΔN [62].

Once the descriptors (Table 5) of the 15 derivatives of imidazole and benzimidazole are
determined, linear regressions are performed for each descriptor considering the corrosion
inhibition efficiency as dependent variable. In any case, a linear dependence between the
property and the efficiency of inhibition is not found.

The diversity of the structures and properties of the molecules analyzed involves different
variables to predict the behavior of a molecule as corrosion inhibitor. A model to predict
the efficiency of a molecule as corrosion inhibitor is proposed. Experimental and
theoretical results are associated to generate a multivariable equation which predicts the
corrosion inhibition of benzimidazole and imidazole derivatives.

3.5 Quantitative structure property relationship

To select the descriptors for setting up the mathematical model a correlation matrix is used
(Table 6). The correlated variables are marked in bold. These variables can then be omitted
from the model. An ideal model must not contain correlated variables to avoid redundant
information.

Multiple linear regression is employed to search a quantitative structure property


relationship model that contains the descriptors that contribute in an appropriate way to
predict the corrosion inhibition efficiency. MobyDigs program allows to obtain every
possible multiple linear regression model which can also be validated using various
statistical parameters that prevent the selection of models with correlated variables. The
validation criteria used the QUIK rule (δK), asymptotic Q2 rule (δQ), RP and RN [63].

The selected model (Figure 7) is described by the following equation:

%Inhibition=513.95-32.03χ+5334𝑏𝑞 𝐼𝑆𝑂 +0.37V+1433.78𝑞𝑁1

The experimental statistical criteria obtained with the equation in MobyDigs are: R2=92.16,
Q2=82.96, δK=0.11, δQ=0.01, RP=0.23, RN=-0.22 and Qboot=53.71.

The model is validated according to the methodology of Todeschini [63]. QUIK rule (δK) is
a criterion for rejecting models that present predictors with high collinearity. The proposed
value for this project is δK=0.02. A model should be rejected when the δK obtained
employing the experimental data is lower than the data proposed as limit. An experimental
value of δK=0.11 allows to confirm that the predictors selected for the model have not high
collinearity.

The second criterion is the asymptotic rule Q2 rule (δQ) to avoid the existence of over
fitting or the existence of not predictable samples. The established value is δQ=0. The value
obtained statistically δQ=0.01 indicates that the model is valid and it is possible to continue
with the validation. The RP and RN criteria are used to detect models that have both an
excess of good and bad predictors. The criterion indicates to reject a model when RP<tP.
For the dataset of this article a suggested tP=0.05 is used. A value of RP=0.23 calculated by
MobyDigs program confirms that the model can be accepted.

The RN criterion is employed to determine models that contain little useful variables. A
model should be rejected when RN<tN(ε), where tN(ε) is calculated from the following
equation:

  R
t N    (17)
R

where ρ is the number of variables involved in the model and ε can take values of 0.02. The
RN =-0.221 experimentally obtained is higher than the experimental value of tN(ε)=-0.248.

Two additional molecules are tested, benzimidazole and imidazole to compare the
inhibition efficiencies determined with the selected model, and the experimental values of
efficiency obtained by electrochemical impedance spectroscopy technique (Figure 8).
Experiments for these two molecules are carried out under the same conditions as the
samples employed to construct the model. The values of corrosion inhibition efficiency
obtained by electrochemical impedance spectroscopy are 27.6 and 62.4% for these
molecules.

The structures of the test molecules were optimized in the same conditions used to analyze
the fifteen molecules. Descriptors corresponding to the selected model are listed in Table 7.

Using the model, a good approximation is obtained in determining the efficiency of


corrosion inhibition of test molecules. In the case of benzimidazole, the efficiency of
inhibition obtained experimentally is 62.4%. The value obtained employing the prediction
model is 56.7%. For imidazole the experimental inhibition efficiency is 27.6% and
according to the prediction model it is 26.1%.

Imidazole as steel corrosion inhibitor reports an inhibition efficiency of 17.5 % in 1.0 M


HCl [64] and 72.7% in 0.5 M H2SO4 [10]. On the other hand, benzimidazole reports
efficiencies of 18.0 % (in 1.0 N H2SO4) [21] and 52.0 % (in 1.0 M HCl, inhibitor 250 mg L-
1
) [65]. These values are congruent with values predicted using the model proposed.

It is important to consider that values reached of corrosion inhibition are associated to


conditions of the corrosive medium, composition of the metallic surface, concentration of
inhibitor and exposition time in the aggressiveness media.

To illustrate the possible mechanism the adsorption behavior of both imidazole and
benzimidazole on the Fe (110) surface, it was used the methodology described previously
[27] with PBE functional. The Fe (110) surface was modeled using a supercell (6 × 3) with
three metal layers and a 20 Å vacuum regions to separate the slabs in the z-direction to
accommodate the heterocyclic compounds. The compounds were adsorbed on hollow and
top sites of the Fe(110) surface.

It is considered the approaching of the heterocycles in either a parallel or perpendicular


arrangement to the iron surface; the most stable arrangement is the parallel one. Then,
several interaction modes of a Fe atom with the molecules are considered; one of these
modes was to place the iminic nitrogen directly above of a Fe atom, thus allowing a fully
geometry optimization. The most favorable adsorption geometries for imidazole and
benzimidazole are shown in the Figure 9. The N−Fe average distances are 2.00 Å for
imidazole while for benzimidazole the N−Fe distances were 2.12 Å. These distances are
close to the sum of the covalent radii [rcov (Fe, N) = 1.95 Å]; the averaged Fe-C bond
distances for the two molecules are 2.12 Å, and they are close to the covalent radii sum
[rcov (Fe, C) = 1.97 Å]. These data indicate that there is an effective interaction of the five-
membered heterocyclic ring with the iron surface. Moreover, the calculated  → d charge
transfer (molecule→ surface) are 0.74 and 0.99 for imidazole and benzimidazole,
respectively; these data also supported the existence of the heterocycle/Fe(110) interaction.
4. Conclusions

Measurements of electrochemical impedance spectroscopy showed that the resistance to


charge transfer (Rct) increases in the presence of the molecules used as corrosion inhibitors
and even further when their concentration increases. However, it is clear that the presence
of substituents and also their position have a significant effect on the protection of steel
surface. It was observed that presence of halogens in the inhibitor structure promotes
corrosion inhibition efficiency. The molecules used were not good inhibitors but
differences in inhibition were clear as a consequence of interaction between the inhibitor
and carbon steel exposed to 1.0 M HCl.

The adsorption isotherms analysis showed that the mechanism of interaction between the
molecules used and the steel surface can be associated to Langmuir isotherm. The
mathematical model proposed was validated employing different criteria. According to
them it is possible to associate properties as electronegativity, aromaticity, volume, and the
charge of nitrogen atoms with the corrosion inhibition efficiency. The surface study
corroborates the importance of the variables found during interaction surface-inhibitor. The
proposed model attempts the complex aspect of the effect of functional groups during
inhibition.

Acknowledgements
E.G. acknowledges CONACyT for her fellowship. The authors wish to thank the UAEH-
PAI-Project for the financial support.

References

[1] P. R. Roberge, Handbook of corrosion engineering, first ed., McGraw-Hill, New York,
1999.

[2] Committee on Assessing Corrosion Education, Assessment of corrosion education, first


ed., The National Academies Press, 2009.

[3] R. W. Revie, Uhlig´s Corrosion handbook, third ed., Wiley, New Jersey, 2011.
[4] A. Yildirim, M. Çetin, Synthesis and evaluation of new long alkyl side chain acetamide,
isoxazolidine and isoxazoline derivatives as corrosion inhibitors, Corros. Sci. 50 (2008)
155-165.

[5] I. B. Obot, N. O. Obi-Egbedi, Theoretical study of benzimidazole and its derivatives


and their potential activity as corrosion inhibitors, Corros. Sci. 52 (2010) 657-660.

[6] I. Danaee, O. Ghasemi, G. R. Rashed, M. R. Avei, M. H. Maddahy, Effect of hydroxyl


group position on adsorption behavior and corrosion inhibition of hydroxybenzaldehyde
Schiff bases: Electrochemical and quantum calculations, J. Mol. Struct. 1035 (2013) 247-
259.

[7] V. S. Sastri, Green corrosion inhibitors. Theory and practice, first ed., John Wiley &
Sons, NJ, 2011.

[8] H. O. Curkovic, E. Stupnisek-Lisac, H. Takenouti, Electrochemical quartz crystal


microbalance and electrochemical impedance spectroscopy study of copper corrosion
inhibition by imidazoles, Corros. Sci. 51 (2009) 2342-2348.

[9] M. M. Antonijevic, M. B. Petrovic, Copper corrosion inhibitors. A review, Int. J.


Electrochem. Sci. 3 (2008) 1-28.

[10] Z. Zhang, S. Chen, Y. Li, S. Li, L. Wang, A study of the inhibition of iron corrosion
by imidazole and its derivatives self-assembled films, Corros. Sci. 51 (2009) 291-300.

[11] F. Kandemirli, S. Sagdinc, Theoretical study of corrosion inhibition of amides and


thiosemicarbazones, Corros. Sci. 49 (2007) 2118-2130.

[12] X. Li, S. Deng, X. Xie, Experimental and theoretical study on corrosion inhibition of
oxime compounds for aluminium in HCl solution, Corros. Sci. 81 (2014) 162-175.

[13] J. M. Roque, T. Pandiyan, J. Cruz, E. García-Ochoa, DFT and electrochemical studies


of tris(benzimidazole-2-ylmethyl)amine as an efficient corrosion inhibitor for carbon steel
surface, Corros. Sci. 50 (2008) 614-624.
[14] P. Udhayakala, T. V Rajendiran, S. Gunasekaran, Theoretical approach to the
corrosion inhibition efficiency of some pyrimidine derivatives using DFT method, J.
Comput. Methods Mol. Des. 2 (2012) 1-15.

[15] K. F. Khaled, Modeling corrosion inhibition of iron in acid medium by genetic


function approximation method: A QSAR model, Corros. Sci. 53 (2011) 3457-3465.

[16] S. G. Zhang, W. Lei, M. Z. Xia, F. Y. Wang, QSAR study on N-containing corrosion


inhibitors: Quantum chemical approach assisted by topological index, J. Mol. Struct.
THEOCHEM. 732 (2005) 173-182.

[17] W. Hui-Long, L. Rui-Bin, J. Xin, Inhibiting effects of some mercapto-triazole


derivatives on the corrosion of mild steel in 1.0 M HC1 medium, Corros. Sci. 46 (2004)
2455-2466.

[18] Y. Sasikumar, A. S. Adekunle, L. O. Olasunkanmi, I. Bahadur, R. Baskar, M. M.


Kabanda, I. B. Obot, E. E. Ebenso, Experimental, quantum chemical and Monte Carlo
simulation studies on the corrosion inhibition of some alkyl imidazolium ionic liquids
containing tetrafluoroborate anion on mild steel in acidic medium, J. Mol. Liq. 211 (2015)
105-118.

[19] H. Hamani, T. Douadi, M. Al-Noaimi, S. Issaadi, D. Daoud, S. Chafaa,


Electrochemical and quantum chemical studies of some azomethine compounds as
corrosion inhibitors for mild steel in 1 M hydrochloric acid, Corros. Sci. 88 (2014) 234-
245.

[20] L. R. Chauhan, G. Gunasekaran, Corrosion inhibition of mild steel by plant extract in


dilute HCl medium, Corros. Sci. 49 (2007) 1143-1161.

[21] J. Aljourani, M.A. Golozar, K. Raeissi, The inhibition of carbon steel corrosion in
hydrochloric and sulfuric acid media using some benzimidazole derivatives, Mater. Chem.
Phys.121 (2010) 320-325.

[22] D. J. F. M. J. Frisch. G. W. Trucks, H. B. Schlegel, G. E. Scuseria, M. A. Robb, J. R.


Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M.
Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zheng, J. L. Sonnenberg, M.
Had, Gaussian 09, (2009).

[23] J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made


simple, Phys. Rev. Lett. 77 (1996) 3865-3868.

[24] J. P. Perdew, K. Burke, M. Ernzerhof, Generalized gradient approximation made


simple. Phys Rev Lett. 78 (1997) 1396.

[25] L. A. Curtiss, P. C. Redfern, V. Rassolov, G. Kedziora, J. A. Pople, Extension of


Gaussian-3 theory to molecules containing third-row atoms K, Ca, Ga-Kr, J. Chem. Phys.
114 (2001) 9287-9295.

[26] R. L. Camacho, E. Aquino, J. Cruz, J. G. Alvarado, O. Olvera-Neria, N. Jayanthi, T.


Pandiyan, DFT analysis: Fe4 cluster and Fe(110) surface interaction studies with pyrrole,
furan, thiophene, and selenophene molecules, Struct. Chem. 25 (2013) 115-126.

[27] R. L. Camacho-Mendoza, E. Gutierrez-Moreno, E. Guzman-Percastegui, E. Aquino-


Torres, J. Cruz-Borbolla, J. A. Rodríguez-Ávila, J. G. Alvarado-Rodríguez, O. Olvera-
Neria, P. Thangarasu, J. L. Medina-Franco, Density functional theory and electrochemical
studies: structure−efficiency relationship on corrosion inhibition, J. Chem. Inf. Model. 55
(2015) 2391-2402.

[28] T. Lu, F. Chen, Multiwfn: A Multifunctional wavefunction analyzer, J. Comput.


Chem. 33 (2012) 580-592.

[29] X. Zhang, H. Ju, J. Wang, Electrochemical sensors, biosensors and their biomedical
applications, first ed., Elsevier, USA, 2008.

[30] 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 (2013) 271-282.

[31] A. K. Singh, S. K. Shuka, M. Singh, M. A. Quraishi, Inhibitive effect of ceftazidime


on corrosion of mild steel in hydrochloric acid solution, Mater. Chem. Phys. 129 (2011) 68-
76.
[32] D. Daoud, T. Douadi, H. Hamani, S. Chafaa, M. Al-Noaimi, Corrosion inhibition of
mild steel by two new S-heterocyclic compounds in 1 M HCl: Experimental and
computational study, Corros. Sci. 94 (2015) 21-37.

[33] Q. B. Zhang, Y. X. Hua, Corrosion inhibition of mild steel by alkylimidazolium ionic


liquids in hydrochloric acid, Electrochim. Acta. 54 (2009) 1881-1887.

[34] E. Stupnis̆ek-Lisac, A. Brnada, A. D. Mance, Secondary amines as copper corrosion


inhibitors in acid media, Corros. Sci. 42 (2000) 243-257.

[35] K. Zhang, B. Xu, W. Yang, X. Yin, Y. Liu, Y. Chen, Halogen-substituted imidazoline


derivatives as corrosion inhibitors for mild steel in hydrochloric acid solution, Corros. Sci.
90 (2015) 284-295.

[36] A. Popova, M. Christov, S. Raicheva, E. Sokolova, Adsorption and inhibitive


properties of benzimidazole derivatives in acid mild steel corrosion, Corros. Sci. 46 (2004)
1333-1350.

[37] S. K. Shukla, M. A. Quraishi, Cefotaxime sodium: A new and efficient corrosion


inhibitor for mild steel in hydrochloric acid solution, Corros. Sci. 51 (2009) 1007-1011.

[38] X. Li, S. Deng, H. Fu, Triazolyl blue tetrazolium bromine as a novel corrosion
inhibitor for steel in HCl and H2SO4 solutions, Corros. Sci. 53 (2011) 302-309.

[39] D. Daoud, T. Douadi, H. Hamani, S. Chafaa, M. Al-Noaimi, Corrosion inhibition of


mild steel by two new S-heterocyclic compounds in 1 M HCl: Experimental and
computational study, Corros. Sci. 94 (2015) 21-37.

[40] M. Lebrini, M. Lagrenée, H. Vezin, L. Gengembre, F. Bentiss, Electrochemical and


quantum chemical studies of new thiadiazole derivatives adsorption on mild steel in normal
hydrochloric acid medium, Corros. Sci. 47 (2005) 485-505.

[41] F. S. de Souza, R. S. Gonçalves, A. Spinelli, Assessment of caffeine adsorption onto


mild steel surface as an eco-friendly corrosion inhibitor, J. Braz. Chem. Soc. 25 (2014) 81-
90.
[42] M. Noel, K. I. Vasu, Cyclic voltammetry and the frontiers of electrochemistry, Aspect,
London, 1990.

[43] S. Şafak, B. Duran, A. Yurt, G. Türkoĝlu, Schiff bases as corrosion inhibitor for
aluminium in HCl solution, Corros. Sci. 54 (2012) 251-259.

[44] N. A. Negm, N. G. Kandile, E. A. Badr, M. A. Mohammed, Gravimetric and


electrochemical evaluation of environmentally friendly nonionic corrosion inhibitors for
carbon steel in 1M HCl, Corros. Sci. 65 (2012) 94-103.

[45] B. Hachula, M. Nowak, J. Kusz, Crystal and molecular structure analysis of 2-


methylimidazole, J. Chem. Crystallog. 40 (2010) 201-206.

[46] C. J. Fahrni, M. M. Henary, D.G. Van Derveer, Excited-state intramolecular proton


transfer in 2-(2’tosylaminophenyl)benzimidazole, J. Phys. Chem. 106 (2002) 7655-7663.

[47] S. Pournazari, M. H. Moayed, M. Rahimizadeh, In situ inhibitor synthesis from


admixture of benzaldehyde and benzene-1,2-diamine along with FeCl3 catalyst as a new
corrosion inhibitor for mild steel in 0.5 M sulphuric acid, Corros. Sci. 71 (2013) 20-31.

[48] A. Szabo, N. S. Ostlund, Modern quantum chemistry-Introduction to advanced


electronic structure Theory, first ed., Dover Publications, New York, 1996.

[49] A. S. Fouda, H. M. Abu-Elnader, M. S. Soliman, Study on corrosion inhibition from


aspect of quantum chemistry, Bull Korean Chem. 7 (1986) 97-99.

[50] P. Udhayakala, A. M. Samuel, T. V Rajendiran, S. Gunasekaran, DFT study on the


adsorption mechanism of some phenyltetrazole substituted compounds as effective
corrosion inhibitors for mild steel, Der Pharma Chemica 5 (2013) 111-124.

[51] M. M. Aboelnga, M. K. Awad, J. W. Gauld, M. R. Mustafa, An assessment to evaluate


the validity of different methods for the description of some corrosion inhibitors, J. Mol.
Model. 20 (2014) 2422.

[52] J. L. Gázquez, A. Cedillo, A. Vela, Electrodonating and electroaccepting powers, J.


Phys. Chem. A. 111 (2007) 1966-1970.
[53] A. Stranger, Nucleus independent chemical shifts (NICS): distance dependence and
revised criteria for aromaticity and antiaromaticity, J. Org. Chem. 71 (2006) 883-893.

[54] H. Fallah-Babher-Shaidaei, C. S. Wannere, C. Corminboeuf, R. Puchta, P. v. R.


Schleyer, Which NICS aromaticity index for planar π rings is best?, Org. Lett. 8 (2006)
863-866.

[55] T. M. Krygowski. Crystallographic studies of inter-and intramolecular interactions


reflected in aromatic character of. pi-electron systems, . J. Chem. Inf. Comput. Sci. 33
(1993) 70-78.

[56] F. Biegler-König, J. Schönbohm, D. Bayles, AIM2000 - A Program to Analyze and


Visualize Atoms in Molecules, submitted to J. Comp. Chem. 22 (2001) 545-559.

[57] S. P. Cardoso, J. A. C.P. Gomes, L.E.P. Borges, E. Hollauer, Predictive QSPR analysis
of corrosion inhibitors for super 13% Cr steel in hydrochloric acid, Brazilian J. Chem. Eng.
24 (2007) 547-559.

[58] P. Udhayakala, T. V. Rajendiran, S. Gunasekaran, A theoretical study of some


barbiturates as corrosion inhibitors for mild steel, Corros. Sci. 6 (2014) 1027-1039.

[59] P. Mourya, P. Singh, A. K. Tewari, R. B. Rastogi, M. M. Singh, Relationship between


structure and inhibition behaviour of quinolinium salts for mild steel corrosion:
Experimental and theoretical approach, Corros. Sci. 95 (2015) 71-87.

[60] Y. Karzazi, M. E. A. Beighiti, A. Dafali, B. Hammouti, A theoretical investigation of


mild steel by piperidine derivatives in hydrochloric acid solution, J. Chem. Pharm. Res. 6
(2014) 689-696.

[61] K. F. Khaled, Studies of iron corrosion inhibition using chemical, electrochemical and
computer simulation techniques, Electrochim. Acta. 55 (2010) 6523-6532.

[62] I. Ahamad, R. Prasad, M. A. Quraishi, Thermodynamic, electrochemical and quantum


chemical investigation of some schiff bases as corrosion inhibitors for mild steel in
hydrochloric acid solutions, Corros. Sci. 52 (2010) 933-942.
[63] R. Todeschini, V. Consonni, A. Mauri, M. Pavan, Detecting “bad” regression models:
multicriteria fitness functions in regression analysis, Anal. Chim. Acta. 515 (2004) 199-
208.

[64] G. Bereket, E. Hür, C. Öğretir, Quantum chemical studies on some imidazole


derivatives as corrosion inhibitors for iron in acidic medium, J. Mol. Struct. THEOCHEM.
578 (2002) 79-88.

[65] J. Aljourani, K. Raeissi, M. A. Golozar, Benzimidazole and its derivatives as corrosion


inhibitors for mild steel in 1M HCl solution, Corros. Sci. 51 (2009) 1836-1843.

Figure Captions

Figure 1. Eocp (vs Ag/AgCl) -time curves for carbon steel in 1.0 M HCl with concentration
of inhibitor 1.0 mM: (a) 2-methylimidazole, (b) 4-(imidazol-1-yl) phenol, (c) 2-
(aminomethyl) benzimidazole, (d) 6-bromo-1H-benzimidazole.

Figure 2. Nyquist plots of the best and worst inhibitor from imidazole and benzimidazole
derivatives: (a) 2-methylimidazole, (b) 4-(imidazol-1-yl) phenol, (c) 2-(aminomethyl)
benzimidazole, (d) 6-bromo-1H-benzimidazole. Blue points correspond to frequency where
imaginary impedance is highest at inhibitor concentration 10.0 mM.

Figure 3. The Bode and phase angle plots for of the best and worst inhibitor from
imidazole and benzimidazole derivatives: (a) 2-methylimidazole, (b) 4-(imidazol-1-yl)
phenol, (c) 2-(aminomethyl) benzimidazole and (d) 6-bromo-1H-benzimidazole.

Figure 4. Equivalent circuit corresponding to analyzed system.

Figure 5. Nyquist and Bode plots (experimental and fit) for carbon steel in presence of
imidazole and benzimidazole derivative solutions (10.0 mM and 0.5mM) in 1.0 M HCl. (a)
2-methylimidazole, (b) 4-(imidazol-1-yl) phenol, (c) 2-(aminomethyl) benzimidazole, (d)
6-bromo-1H-benzimidazole.
Figure 6. Tafel curves for carbon steel in 1.0 M HCl without and with various
concentrations of inhibitors: (a) 2-methylimidazole, (b) 4-(imidazol-1-yl) phenol, (c) 2-
(aminomethyl) benzimidazole, (d) 6-bromo-1H-benzimidazole.

Figure 7. Theoretical vs experimental % inhibition.

Figure 8. Nyquist plots obtained for test molecules: (a) blank, (b) imidazole, (c)
benzimidazole.

Figure 9. Adsorption geometries of imidazole and benzimidazole on Fe(110) surface.


Fig. 1

Fig. 2
Fig. 3

Fig. 4
Fig. 5
Fig. 6

Fig. 7
Fig. 8

Fig. 9
Table 1. Chemical structures of imidazole and benzimidazole derivatives employed for the generation of a
mathematical model for prediction of corrosion inhibition.

Inhibitor Name R2 R3 R4 R5 R6

2-Chloro-1H-imidazole
Cl - - - -

2-Methylimidazole
CH3 - - - -

4′-(Imidazol-1-yl)acetophenone
H H COCH3 - -

4-(1H-Imidazol-1-yl) aniline
H H NH2 - -

4'-(Imidazol-1-yl) benzaldehyde
H H CHO - -

1-(4-Bromophenyl)-1H-imidazole
H H Br - -

1-(4-Chlorophenyl) imidazole
H H Cl - -

4-(Imidazol-1-yl) phenol
H H OH - -

1-(3-Chlorophenyl) imidazole
H Cl H - -

1-(2-Chlorophenyl) imidazole
Cl H H - -

2-(Chloromethyl)benzimidazole CH2Cl - - H H

6-Bromo-1H-benzimidazole H - - H Br

1H-Benzimidazol-5-amine H - - H NH2

2-(Aminomethyl)benzimidazole CH2NH2 - - H H

1H-Benzimidazol-5-ylmethanol H - - CH2OH H
Table 2. Data obtained by electrochemical impedance spectroscopy (mean and relative standard deviation
n=3) for corrosion of carbon steel in 1.0 M HCl.

Inhibitor Cinhibitor (mM) Rs(Ω·cm2) Rct (Ω·cm2) Cdl (µF·cm-2) θ %Inhibition


Blank 0 1.19 (0.21) 162.39 (0.18) 16.60 (0.96) - -

2-Chloro-1H-imidazole 0.1 1.08 (1.17) 267.97 (1.07) 14.47 (6.99) 0.39 (2.07) 39.4 (2.1)
0.5 1.08 (1.64) 292.02 (1.48) 13.08 (7.43) 0.44 (2.33) 44.4 (2.3)
1.0 1.09 (1.34) 310.70 (0.97) 13.38 (4.90) 0.48 (1.38) 47.7 (1.4)
5.0 1.29 (8.81) 344.31 (6.25) 11.19 (7.39) 0.53 (5.09) 52.8 (5.1)
10.0 0.97 (2.15) 378.35 (1.77) 10.63 (3.88) 0.57 (1.26) 57.1 (1.3)

2-Methylimidazole 0.1 1.63 (0.54) 192.41 (0.32) 13.93 (3.66) 0.16 (1.59) 15.6 (1.6)
0.5 1.25 (0.90) 197.64 (0.56) 16.92 (7.38) 0.18 (3.16) 17.8 (3.2)
1.0 1.19 (1.54) 213.99 (1.14) 15.70 (4.55) 0.24 (4.30) 24.1 (4.3)
5.0 1.12 (0.70) 221.18 (0.42) 15.25 (5.79) 0.27 (1.19) 26.6 (1.2)
10.0 1.23 (3.09) 232.49 (2.17) 13.15 (7.49) 0.30 (5.99) 30.2 (6.0)

4′-(Imidazol-1-yl) acetophenone 0.1 0.98 (0.97) 289.02 (0.91) 13.36 (7.02) 0.44 (1.42) 43.8 (1.4)
0.5 1.06 (4.87) 380.42 (3.35) 10.83 (8.92) 0.57 (2.60) 57.3 (2.6)
1.0 1.14 (1.01) 550.11 (0.78) 9.16 (6.56) 0.70 (0.34) 70.5 (0.3)
5.0 0.89 (1.34) 713.63 (1.23) 8.85 (1.96) 0.77 (0.42) 77.2 (0.4)
10.0 0.92 (0.85) 852.43 (0.63) 7.03 (9.67) 0.81 (0.19) 80.9 (0.2)

4-(1H-Imidazol-1-yl) aniline 0.1 1.23 (2.68) 281.27 (1.83) 15.42 (9.11) 0.42 (2.94) 42.3 (2.9)
0.5 1.11 (0.29) 503.85 (0.14) 13.45 (2.13) 0.68 (0.12) 67.8 (0.1)
1.0 1.06 (0.29) 481.04 (0.12) 10.80 (9.03) 0.66 (0.07) 66.2 (0.1)
5.0 1.01 (0.29) 574.40 (0.26) 11.31 (1.46) 0.72 (0.18) 71.7 (0.2)
10.0 0.93 (5.51) 658.03 (3.89) 9.26 (2.04) 0.75 (1.43) 75.3 (1.4)

4'-(Imidazol-1-yl) benzaldehyde 0.1 1.04 (0.64) 212.68 (0.44) 14.41 (3.23) 0.24 (1.05) 23.6 (1.1)
0.5 1.06 (0.92) 454.06 (0.64) 11.39 (8.94) 0.64 (0.32) 64.2 (0.3)
1.0 1.11 (1.77) 584.49 (1.30) 9.90 (6.00) 0.72 (0.51) 72.2 (0.5)
5.0 0.93 (1.13) 828.71 (1.05) 9.12 (5.08) 0.80 (0.30) 80.4 (0.3)
10.0 1.05 (0.57) 778.29 (0.52) 7.05 (7.27) 0.79 (0.19) 79.1 (0.2)

1-(4-Bromophenyl)-1H-imidazole 0.1 1.06 (1.15) 225.77 (0.88) 15.71 (9.80) 0.28 (2.66) 28.1 (2.7)
0.5 1.06 (3.95) 298.04 (2.85) 14.01 (2.52) 0.46 (3.93) 45.5 (3.9)
1.0 1.14 (3.43) 372.61 (2.36) 11.03 (9.27) 0.56 (2.09) 56.4 (2.1)
5.0 1.03 (0.94) 840.17 (0.89) 6.50 (3.72) 0.81 (0.26) 80.7 (0.3)
10.0 0.90 (0.96) 1364.06 (0.82) 5.22 (5.77) 0.88 (0.14) 88.1 (0.1)

1-(4-Chlorophenyl) imidazole 0.1 1.09 (1.00) 215.35 (0.83) 17.50 (8.72) 0.25 (2.87) 24.6 (2.9)
0.5 1.08 (0.45) 276.82 (0.39) 13.75 (4.70) 0.41 (0.62) 41.3 (0.6)
1.0 1.57 (5.14) 333.14 (3.40) 11.63 (8.07) 0.51 (3.73) 51.3 (3.7)
5.0 1.16 (0.29) 437.59 (0.26) 10.43 (1.26) 0.63 (0.26) 62.9 (0.3)
10.0 1.31 (1.30) 777.21 (1.15) 6.55 (7.23) 0.79 (0.35) 79.1 (0.4)

4-(Imidazol-1-yl) phenol 0.1 1.15 (1.60) 257.58 (1.05) 16.70 (4.31) 0.37 (2.13) 37.0 (2.1)
0.5 1.05 (0.98) 527.40 (0.80) 12.03 (2.85) 0.69 (0.45) 69.2 (0.5)
1.0 1.22 (0.76) 752.13 (0.62) 11.01 (4.11) 0.78 (0.15) 78.4 (0.2)
5.0 0.94 (0.61) 1435.75 (0.48) 11.58 (3.64) 0.89 (0.07) 88.7 (0.1)
10.0 0.97 (0.32) 1453.39 (0.25) 9.44 (0.01) 0.89 (0.01) 88.8 (0.1)

1-(3-Chlorophenyl) imidazole 0.1 1.03 (1.64) 337.17 (1.09) 12.83 (2.79) 0.52 (0.88) 51.8 (0.9)
0.5 0.98 (0.76) 390.74 (0.69) 14.63 (3.43) 0.58 (0.59) 58.4 (0.6)
1.0 1.13 (0.99) 379.31 (0.64) 13.18 (6.52) 0.57 (0.41) 57.2 (0.4)
5.0 1.04 (7.64) 440.93 (5.57) 11.70 (9.90) 0.63 (3.77) 63.2 (3.8)
10.0 0.98 (0.52) 558.66 (0.40) 11.59 (3.22) 0.71 (0.20) 70.9 (0.2)

1-(2-Chlorophenyl) imidazole 0.1 1.00 (1.35) 290.62 (0.96) 13.64 (8.35) 0.44 (1.14) 44.1 (1.1)
0.5 1.05 (9.02) 295.64 (6.35) 13.53 (4.01) 0.45 (9.43) 45.1 (9.4)
1.0 0.94 (1.52) 337.42 (1.11) 11.39 (8.78) 0.52 (1.01) 51.9 (1.0)
5.0 0.96 (0.67) 391.83 (0.44) 10.35 (9.39) 0.59 (0.26) 58.6 (0.3)
10.0 1.16 (7.34) 439.25 (5.20) 11.51 (8.18) 0.63 (2.81) 63.0 (2.8)
2-(Chloromethyl) benzimidazole 0.1 1.07 (0.35) 768.14 (0.35) 7.93 (5.92) 0.79 (0.14) 78.9 (0.1)
0.5 0.92 (2.55) 805.24 (1.84) 7.88 (8.08) 0.80 (0.53) 79.8 (0.5)
1.0 0.98 (2.99) 854.82 (2.11) 6.64 (7.48) 0.81 (0.45) 81.0 (0.5)
5.0 0.95 (1.38) 1011.46 (0.81) 6.15 (5.03) 0.84 (0.13) 83.9 (0.1)
10.0 0.92 (3.03) 1070.97 (2.22) 6.98 (9.41) 0.85 (0.36) 84.8 (0.4)

6-Bromo-1H-benzimidazole 0.1 1.09 (2.17) 219.24 (1.45) 16.96 (6.20) 0.26 (5.03) 25.9 (5.0)
0.5 1.10 (11.34) 322.89 (7.97) 12.79 (2.94) 0.50 (7.19) 49.7 (7.2)
1.0 1.12 (0.84) 412.83 (0.71) 10.50 (3.25) 0.61 (0.52) 60.7 (0.5)
5.0 0.99 (1.16) 1161.75 (1.06) 5.97 (12.88) 0.86 (0.20) 86.0 (0.2)
10.0 0.78 (8.45) 1525.15 (5.91) 4.07 (9.97) 0.89 (0.65) 89.4 (0.7)

1H-Benzimidazol-5-amine 0.1 1.17 (0.67) 246.01 (0.64) 15.84 (2.11) 0.34 (1.55) 34.0 (1.6)
0.5 1.28 (1.24) 315.13 (0.89) 13.15 (10.79) 0.48 (0.87) 48.5 (0.8)
1.0 1.35 (0.73) 434.61 (0.63) 11.59 (3.62) 0.63 (0.50) 62.6 (0.5)
5.0 1.04 (4.39) 588.59 (3.01) 9.65 (6.49) 0.72 (1.29) 72.4 (1.3)
10.0 1.04 (1.60) 646.48 (1.18) 9.60 (2.61) 0.75 (0.34) 74.9 (0.3)

2-(Aminomethyl) benzimidazole 0.1 1.00 (0.31) 207.77 (0.26) 16.11 (7.74) 0.22 (1.65) 21.8 (1.7)
0.5 1.00 (0.38) 297.30 (0.17) 14.49 (1.99) 0.45 (0.24) 45.4 (0.2)
1.0 1.05 (13.27) 366.03 (9.37) 11.94 (4.58) 0.56 (9.25) 55.6 (9.3)
5.0 1.26 (0.41) 402.74 (0.24) 12.29 (2.90) 0.60 (0.06) 59.7 (0.1)
10.0 1.00 (1.00) 524.81 (0.69) 12.04 (4.73) 0.69 (0.39) 69.1 (0.4)

1H-Benzimidazol-5-ylmethanol 0.1 0.99 (0.88) 220.50 (0.72) 16.08 (5.11) 0.26 (1.87) 26.4 (1.9)
0.5 1.11 (0.39) 267.23 (0.23) 13.05 (1.96) 0.39 (0.55) 39.2 (0.6)
1.0 1.13 (3.15) 343.62 (2.08) 10.39 (3.25) 0.53 (2.15) 52.7 (2.2)
5.0 1.01 (1.28) 610.67 (1.16) 9.03 (5.62) 0.73 (0.48) 73.4 (0.5)
10.0 1.25 (8.37) 609.30 (5.88) 7.18 (2.11) 0.73 (2.44) 73.3 (2.4)
Table 3. Electrochemical parameters obtained from polarization curves of inhibitors in 1.0 M HCl (mean and
relative standard deviation n=3).

Cinhibitor Ecorr (V, vs Current density


Inhibitor (mM) Ag/AgCl) βc (V decade-1) βa (V decade-1) (10-4,A·cm-2) %Inhibition
Blank 0 -0.465 -0.160 (3.6) 0.215 (4.4) 2.93(2.0) -

2-Methylimidazole 0.1 -0.469 -0.168 (4.8) 0.218 (5.0) 2.57 (3.0) 12.3 (7.7)
0.5 -0.450 -0.141 (2.0) 0.189 (5.4) 2.44 (1.4) 16.9 (6.4)
1.0 -0.469 -0.133 (1.0) 0.267 (0.3) 2.23 (0.6) 23.9 (2.1)
5.0 -0.471 -0.144 (0.5) 0.248 (1.1) 2.16 (1.8) 26.4 (4.8)
10.0 -0.466 -0.124 (4.7) 0.194 (0.7) 1.99 (0.1) 31.9 (0.2)

4-(Imidazol-1-yl) phenol 0.1 -0.455 -0.131 (0.5) 0.137 (3.0) 1.84 (2.5) 37.2 (4.4)
0.5 -0.467 -0.130 (1.1) 0.151 (0.9) 1.09 (10.0) 62.9 (5.3)
1.0 -0.468 -0.107 (2.7) 0.118 (6.3) 0.60 (3.6) 79.7 (1.0)
5.0 -0.444 -0.122 (6.1) 0.132 (5.2) 0.42 (2.3) 85.8 (0.4)
10.0 -0.459 -0.117 (2.5) 0.137 (3.7) 0.40 (5.7) 86.3 (0.9)

2-(Aminomethyl) benzimidazole 0.1 -0.466 -0.128 (0.6) 0.261 (6.2) 2.24 (1.0) 23.5 (3.2)
0.5 -0.484 -0.113 (2.5) 0.143 (0.5) 1.62 (1.1) 44.7 (1.4)
1.0 -0.479 -0.108 (2.0) 0.124 (7.2) 1.31 (8.8) 55.4 (8.0)
5.0 -0.462 -0.115 (3.8) 0.109 (7.5) 1.24 (10.4) 57.7 (8.7)
10.0 -0.485 -0.092 (6.6) 0.102 (4.0) 0.91 (10.6) 69.0 (5.3)

6-Bromo-1H-benzimidazole 0.1 -0.470 -0.130 (0.5) 0.256 (2.7) 2.10 (0.6) 28.2 (1.6)
0.5 -0.471 -0.125 (3.3) 0.174 (4.2) 1.51 (6.1) 48.3 (6.0)
1.0 -0.479 -0.094 (0.7) 0.138 (1.5) 1.03 (8.8) 64.9 (5.2)
5.0 -0.465 -0.130 (3.3) 0.114 (3.8) 0.45 (3.0) 84.6 (0.6)
10.0 -0.456 -0.095 (1.5) 0.089 (1.6) 0.28 (3.3) 90.5 (0.4)

Table 4. Kads values for benzimidazole and imidazole derivatives employed as corrosion inhibitors.

Inhibitor Kads (103, L mol-1) R2


2-Chloro-1H-imidazole 3.23 ± 6.45x10-4 0.998
2-Methylimidazole 0.99 ± 1.86 x10-3 0.996
4′-(Imidazol-1-yl)acetophenone 4.71 ± 2.26 x10-4 0.999
4-(1H-Imidazol-1-yl) aniline 6.38 ± 2.71 x10-4 0.999
4'-(Imidazol-1-yl) benzaldehyde 5.64 ± 2.34 x10-4 0.999
1-(4-Bromophenyl)-1H-imidazole 1.92 ± 4.32 x10-4 0.998
1-(4-Chlorophenyl) imidazole 1.43 ± 1.35 x10-3 0.986
4-(Imidazol-1-yl) phenol 6.45 ± 7.02 x10-5 0.999
1-(3-Chlorophenyl) imidazole 3.56 ± 7.75 x10-4 0.996
1-(2-Chlorophenyl) imidazole 3.23 ± 5.19 x10-4 0.999
2-(Chloromethyl)benzimidazole 24.7 ± 5.83 x10-5 0.999
6-Bromo-1H-benzimidazole 2.30 ± 2.32 x10-4 0.999
1H-Benzimidazol-5-amine 3.51 ± 1.88 x10-4 0.999
2-(Aminomethyl) benzimidazole 1.84 ± 1.89 x10-4 0.999
1H-Benzimidazol-5-ylmethanol 2.23 ± 3.94 x10-4 0.999
Table 5. Descriptors calculated for structures of imidazole and benzimidazole derivatives.

µD I A χ η ω* ω- ω+ bqISO bqANS V
Structure %Inhibition log P qN1 qN2 qC2 ΔN
(Debye) (eV) (eV) (eV) (eV) (eV) (eV) (eV) (Adim) (Adim) (Å3)
2-Chloro-1H imidazole 57.1 3.82 8.79 -0.31 4.24 4.55 1.97 4.66 0.42 -10.05 -26.47 87.83 0.99 -0.21 -0.07 0.11 0.30
2-Methylimidazole 30.2 3.65 8.48 -0.32 4.08 4.40 1.89 4.48 0.40 -10.44 -28.37 92.80 0.09 -0.23 -0.08 0.08 0.33
4′-(Imidazol-1 yl)acetophenone 80.9 1.97 8.12 0.80 4.46 3.66 2.72 5.41 0.95 -9.52 -25.41 199.37 0.44 -0.22 -0.02 0.04 0.35
4-(1H-Imidazol-1-yl) aniline 75.3 6.10 7.43 -0.14 3.64 3.78 1.75 4.05 0.40 -9.73 -27.66 170.88 -1.03 -0.23 -0.03 0.04 0.44
4'-(Imidazol-1-yl) benzaldehyde 79.1 2.75 8.32 0.95 4.63 3.68 2.91 5.69 1.06 -9.40 -25.12 181.24 -0.33 -0.21 -0.02 0.05 0.32
1-(4-Bromophenyl)-1H-imidazole 88.1 2.33 8.08 0.00 4.04 4.04 2.02 4.54 0.50 -9.84 -26.54 178.45 0.82 -0.22 -0.03 0.04 0.37
1-(4-Chlorophenyl) imidazole 79.1 2.32 8.16 -0.06 4.05 4.11 1.99 4.53 0.48 -9.83 -26.65 173.96 0.55 -0.22 -0.03 0.04 0.36
4-(Imidazol-1-yl) phenol 88.8 4.71 7.84 -0.13 3.86 3.99 1.86 4.29 0.43 -10.10 -27.45 167.55 -0.40 -0.22 -0.03 0.04 0.39
1-(3-Chlorophenyl) imidazole 70.9 2.84 8.29 -0.04 4.12 4.16 2.04 4.63 0.50 -9.76 -26.33 173.93 0.55 -0.22 -0.02 0.04 0.35
1-(2-Chlorophenyl) imidazole 63.0 4.42 8.31 -0.11 4.10 4.21 2.00 4.57 0.47 -12.26 -28.75 174.00 0.55 -0.22 -0.03 0.04 0.35
2-(Chloromethyl)benzimidazole 84.8 2.48 8.34 -0.04 4.15 4.19 2.06 4.65 0.50 -10.53 -23.54 158.34 -0.02 -0.21 -0.08 0.09 0.34
6-Bromo-1H-benzimidazole 89.4 3.48 8.14 -0.08 4.03 4.11 1.98 4.51 0.48 -9.08 -26.21 144.01 -0.68 -0.21 -0.08 0.06 0.36
1H-Benzimidazol-5-amine 74.9 4.13 7.27 -0.19 3.54 3.73 1.68 3.92 0.38 -10.52 -25.31 136.36 -2.53 -0.22 -0.08 0.05 0.46
2-(Aminomethyl) benzimidazole 69.0 3.70 7.86 -0.13 3.86 4.00 1.87 4.30 0.43 -10.55 -24.21 155.08 -1.63 -0.23 -0.08 0.09 0.39
1H-Benzimidazol-5-ylmethanol 73.3 2.83 8.06 -0.30 3.88 4.18 1.80 4.26 0.38 -11.09 -27.09 151.74 -1.70 -0.22 -0.08 0.06 0.37
Table 6. Data correlation matrix of the DFT parameters

%Inhibitio bqIS bqAN log


qN1 qN3 qC2 ΔN
n µD I A χ η ω *
ω
-
ω
+ O S
V P
%Inhibitio
n 1.00
1.0
µD -0.22 0
-
0.4 1.0
I -0.36 6 0
-
0.4 0.0 1.0
A 0.35 2 9 0
-
0.6 0.7 0.7 1.0
χ -0.02 0 5 2 0
- -
0.0 0.6 0.6 0.0 1.0
η -0.52 4 9 6 4 0
- -
0.5 0.3 0.9 0.9 0.4 1.0
ω* 0.20 3 8 5 0 1 0
- -
0.5 0.5 0.8 0.9 0.2 0.9 1.0
ω- 0.10 7 5 8 6 3 8 0
- -
0.4 0.2 0.9 0.8 0.5 0.9 0.9 1.0
ω+ 0.25 7 2 9 1 6 9 3 0
- - -
0.2 0.0 0.4 0.2 0.3 0.3 0.3 0.4
bqISO 0.39 4 2 3 7 3 8 5 1 1.00
- - -
0.3 0.1 0.3 0.1 0.3 0.2 0.2 0.3
bqANS 0.35 2 0 5 6 2 8 4 1 0.26 1.00
- - -
0.1 0.3 0.5 0.1 0.6 0.4 0.3 0.4 1.0
V 0.61 5 2 8 6 7 2 1 8 0.10 0.26 0
-
0.3 0.7 0.1 0.6 0.4 0.3 0.5 0.2 - 0.0 1.0
log P -0.10 6 5 9 4 3 8 0 5 0.13 0.25 5 0
- -
qN1 0.3 0.4 0.3 0.5 0.1 0.4 0.4 0.3 0.0 0.2 1.0
0.42 9 5 0 2 2 1 7 6 0.24 0.33 2 8 0
- - -
qN3 0.1 0.0 0.5 0.3 0.3 0.5 0.4 0.5 - 0.5 0.5 0.0 1.0
0.32 2 1 4 7 9 0 5 1 0.25 0.32 9 0 3 0
- - - - - - -
qC2 0.0 0.3 0.3 0.0 0.5 0.1 0.0 0.2 - 0.5 0.0 0.1 0.7 1.0
-0.45 8 8 1 6 1 7 7 4 0.14 0.45 3 5 6 6 0
- - - - - - - - - - -
ΔN 0.5 0.9 0.2 0.8 0.5 0.5 0.6 0.3 - 0.2 0.7 0.5 0.0 0.3 1.0
0.28 2 8 5 5 5 3 8 8 0.07 0.03 3 3 0 9 3 0

Table 7. Descriptors obtained for test molecules used in model validation.

Sample χ(eV) bqISO V(Å3) qN1 %Inhibitionimpedance %Inhibitiontheoretical


Imidazole 4.32 -11.49 74.41 -0.22 27.6 26.1
Benzimidazole 4.04 -11.05 125.86 -0.22 62.4 56.7

38

You might also like