You are on page 1of 23

Anal. Bioanal. Electrochem., Vol. 11, No.

3, 2019, 373-395

Analytical &
Bioanalytical
Electrochemistry
2019 by CEE
www.abechem.com

Full Paper

Computational Study and QSPR Approach on the


Relationship between Corrosion Inhibition Efficiency and
Molecular Electronic Properties of Some Benzodiazepine
Derivatives on C-steel Surface
El Hassan El Assiri,1,* Majid Driouch,1 Zakariae Bensouda,1 Mustapha Beniken,1
Ali Elhaloui,1,2 Mouhcine Sfaira1 and Taoufiq Saffaj3

1
Laboratory of Materials Engineering, Modeling and Environment, LIMME, Faculty of
Sciences Dhar El Mahraz, Sidi Mohamed Ben Abdellah University, USMBA, Po. Box 1796
Atlas Fez, Morocco
2
Laboratory of Materials, Electrochemistry and Environment, Faculty of Sciences, Ibn Tofaîl
University, Po. Box 133-14000 Kénitra, Morocco
3
Laboratory of Application Organic Chemistry, Faculty of Sciences and Techniques, Sidi
Mohamed Ben Abdellah University, USMBA, BP 2626 route d’Immouzer-Fez, Morocco

*Corresponding Author, Tel.: +212 6 74 87 08 78


E-Mail: elassirielhassan@gmail.com or elhassan.elassiri@usmba.ac.ma

Received: 7 September 2018 / Received in revised form: 18 February 2019 /


Accepted: 21 February 2019 / Published online: 31 March 2019

Abstract- The Density Functional Theory (DFT) study was used to investigate the corrosion
inhibition performance of four inhibitors namely: 2,3-dihydro-1H-pyrrolo [2,1c][1,4]
benzodiazepine-5,11(10H,11aH)-dione (BZD1); 2,3-dihydro-1H-pyrrolo [2,1c][1,4]
benzodiazepine-5,11(10H,11aH)-dithione (BZD2); 10-benzyl-pyrrolo [2,1-c][1,4]
benzodiazepine-5,11-dione (BZD=2O) and 10-benzyl-pyrrolo [2,1-c][1,4] benzodiazepine-
5,11-dithione (BZD=2S) on carbon steel using the B3LYP/6-311G(d, p) level of theory. The
geometry optimization was conducted only on the neutral form of all benzodiazepine
derivatives, without any protonated form as testified by the Marvin software. The most
relevant quantum chemical parameters according to their potential action as corrosion
inhibitors were calculated. The descriptors considered were: the EHOMO (highest occupied
molecular orbital energy), the ELUMO (lowest unoccupied molecular orbital energy), the
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 374

energy gap (∆E), the dipole moment (µ), the hardness (η), the softness (σ), the absolute
electronegativity (χ), the total energy (Etot), the ionization potential (IP), the electron affinity
(EA) and the fraction of electrons transferred (∆N). Besides, the local reactivity was analyzed
through the Fukui function in order to compare the possible sites for nucleophilic and
electrophilic attacks. The electronic properties of these inhibitors obtained by DFT were
correlated with their experimental efficiencies using two mathematical models, based-QSPR
approaches; the multiple linear regressions (MLR) and the multiple polynomial regressions
(MPR). The quantum chemical study showed that the theoretical and experimental results
were in good agreement, and the statistical results revealed that the MPR was the most
relevant and predictive model in comparison with the MLR model, with a very high
determination coefficient (R2 = 0.99), adjusted determination coefficient (R2adj = 0.99) and
predicted determination coefficient (R2pred = 0.97)

Keywords - Corrosion inhibition, Benzodiazepine, C-steel, DFT at B3LYP/6-31G**, QSPR

1. INTRODUCTION
Corrosion phenomenon is a fundamental process playing a negative role in global
economy and human health, particularly for the metal compounds and alloys. Despite its
exposure to corrosion, steel has found wide applications in a broad spectrum of industries and
machinery because of its low cost. For this reason, the corrosion inhibition of steel is
therefore a matter of theoretical as well as practical importance [1-8]. The use of inhibitors is
one of the most practical methods for protecting C-steel against this phenomenon; especially
in acidic media. The selecting of potent effective inhibitive molecules depends on their
electron donating properties and mechanism of action [9]. In fact, a number of heterocyclic
compounds containing nitrogen, oxygen, and sulfur either in the aromatic or long chain
carbon system such as benzodiazepine derivatives has been reported to be effective inhibitors
[10,11]. These organic derivatives can provide electrons to unoccupied d orbital of the metal
surface to form coordinate covalent bonds, and can also accept free electrons from the metal
surface by using their lower unoccupied orbital, which make then excellent corrosion
inhibitors [12]. On the other hand, quantum chemical approaches have proven to be very
useful in determining the molecular structure as well as elucidating the electronic structure
and the characteristics of the reactive sites [13] in order to explain the mechanism of
reactivity for corrosion inhibition process, and to understand the relationship between the
corrosion inhibition efficiency and a number of molecular indices of these organic corrosion
compounds [14-16]. The theoretical calculations results can also be used to fetch compounds
with desired properties using Quantitative Structure Property Relationships (QSPR) approach
[17,18]. This statistical method can supply useful qualitative and quantitative information, for
a better understanding of the corrosion inhibition process. In this context, the estimation of
the corrosion inhibition efficiency in terms of experimental conditions is an essential part of
the approach [19-22] which requires that the data must be obtained according to a single
experimental protocol and then suitable mathematical models with predictive ability can be
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 375

developed. Accordingly, in the present work, we were constrained to limit ourselves to four
molecules, published experimentally, taken at four different concentrations so that the
number of individuals rises to sixteen while carefully respecting the experimental conditions
for obtaining the corresponding efficiencies.
The aim of this study is to investigate computationally, in gas and aqueous phases, the
intrinsic electronic and structural properties of four Benzodiazepine derivatives denoted
hereafter, BZD1, BZD2, BZD=2O and BZD=2S. Table 1 shows the molecular structures of
the compounds already studied as C-steel corrosion inhibitors in 1 M of hydrochloric acid
solution in the range of concentration from 10-6 to 10-3 M with distinguishable efficiencies
[23,24].

Table 1. The 2D un-optimized molecular structures, IUPAC names and abbreviations of the
exploited derivatives

Structure IUPAC name Abbreviation


O 2,3-dihydro-1H-pyrrolo[2,1c][1,4] BZD1
N benzodiazepine-5,11(10H,11aH)-dione

N
H O

S 2,3-dihydro-1H-pyrrolo[2,1c][1,4] BZD2

N benzodiazepine-5,11(10H,11aH)-dithione

N
H S

O 10-benzyl-pyrrolo[2,1-c][1,4] BZD=2O
N benzodiazepine-5,11-dione

N
O

S
10-benzyl-pyrrolo[2,1-c][1,4]benzodiazepine-5,11-dithione BZD=2S
N

N
S
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 376

The investigation is conducted by the us DFT at B3LYP methods with 6-31G (d, p) basis
set, in an attempt to determine the pertinent theoretical parameters, to identify the inhibition
properties of the undertaken derivatives and to establish the correlation between their
inhibitive effect and their relevant molecular properties, by exploiting the QSPR approach,
referring to a multiple linear regression (MLR) and a multiple polynomial regression (MPR)
models, in order to estimate the corrosion inhibition efficiencies of the whole inhibitors.

2. COMPUTATIONAL METHODOLOGY

2.1. Quantum chemical calculations


The effect of molecular structure on the inhibitive action of the used benzodiazepines,
was performed by theoretical chemistry calculations using the density functional theory
(DFT) with the Beck’s three parameter exchange functional along with the Lee-Yang-Parr
non local correlation functional (B3LYP) with 6-31G (d, p) basis set, using the GAUSSIAN
03 software [25].
Computationally, in first time this study was carried out in phase gaz. Then, since the
corrosion inhibition phenomenon was studied experimentally, in the electrolytic solution; it is
necessary to include the effect of solvent in the computations. For this reason, the Self-
Consistent Reaction Field (SCRF) theory [16], with Tomasi’s Polarized Continuum Model
(PCM) was used for better approach of the experimental results obtained in aqueous phase
[26]. However, the quantum calculation, at the Gaussian 03 level, cannot represent the effect
of hydrogen chloride solution, hence water was instead used to include the solvent effect.
Molecular properties such as: the energies of the highest occupied and lowest unoccupied
molecular orbital (EHOMO and ELUMO), the energy gap (ΔE), the dipole moment (µ), the
absolute electronegativity (χ), the absolute hardness (η) and softness (σ), the ionization
potential (IP), the electron affinity (EA) and the fraction of electrons transferred from the
used molecule to the metal surface (ΔN), were determined.
Ionization potential IP is associated to EHOMO by the equation [27]:

IP   E HOMO (1)

Electron affinity (EA) is related to ELUMO by the equation [27]:

EA   E LUMO
(2)

The electronegativity and the global hardness and softness were calculated, by the use of
the finite difference approximation, as linear combinations of the calculated IP and EA [27]:

IP  EA
 (3)
2
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 377

IP  EA

2 (4)
1
 (5)

Moreover, for two chemical reagents with different electronegativites, the electronic flow
will occur from the organic molecule with the lower of χ towards the metallic surface with
the higher value of this parameter, until the chemical potentials are equal [28]. Therefore, the
fraction of electrons transferred (∆N) from the inhibitor to the metallic atom was determined
according to Pearson electronegativity scale [28]:

 Fe   inh
N  (6)
2( Fe   inh )

For the electronegativity of bulk iron the theoretical value was used  Fe  7eV and a
global hardness of Fe  0 , by assuming that for a metallic bulk IP=EA because they are
softer than the neutral metallic atoms [29].
Fukui functions were calculated since they provide an avenue for analyzing the local
selectivity of a corrosion inhibitor [30]. Their values were used to identify which atoms in the
inhibitors were more prone to undergo an electrophilic or a nucleophilic attack. The
condensed Fukui functions are then computed using the finite-difference approximation as
indicated below:

f k  Pk ( N  1)  Pk ( N ) for nucleophilic attack (7a)



f  Pk ( N )  Pk ( N  1)
k for electrophilic attack (7b)
P ( N  1)  Pk ( N  1)
f k0  k for radical attack (7c)
2

where, Pk (N), Pk (N+1) and Pk (N-1) are the natural populations in neutral, anionic and
cationic systems of the site k, respectively.

2.2. QSPR calculations


As a second part of this study, a statistical correlation between the calculated molecular
parameters and the experimental corrosion inhibition efficiency (E%) was conducted; using
the quantitative structure-property relationship (QSPR) analysis. This statistical approach was
used to establish mathematical models that relate the inhibitive activity of compounds to their
chemical structures such as electronic and structural molecular descriptors.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 378

2.2.1. Multiple linear regression model (MLR)


After their calculation, the selected descriptors were used to perform a multiple linear
regression (MLR) model, based on the minimization of the difference between experimental
and predicted values using the least-squares method, which consists to minimize the residual
sum of square (SSres), using the XLSTAT 15 and MINITAB 16 software.
The MLR is expressed in the following equation (8):

Ecal %  A X j Ci  B (8)

Where B is constant, A is a quantum chemical index coefficient; xj a chemical index


characteristic for the derivative j; Ci denotes the inhibitor concentration.

2.2.2. Multiple polynomial regression model (MPR)


To improve the quality of the correlation between the corrosion inhibition efficiency and
the molecular structure, the chosen variables in the best linear model were exploited to
evaluate a multiple polynomial regression (MPR for which a quadratic correlation between
E% and the selected descriptors was ensured. This quadratic polynomial model is expressed
by equation (9):

Ecal %  cst  (a0  a1 EHOMO  a2 ELUMO  a3  )Ci  (b0 EHOMO  b1 E LUMO  b2  )Ci
2
(9)

Where a0, a1, a2, a3, b0, b1 and b2 are quantum chemical index coefficients; EHOMO, ELUMO
and µ are quantum chemical index characteristics for the molecule j. Several parameters were
used for the validation of the two models, such as: the coefficient of correlation R, the
coefficient of determination R2, the adjusted coefficient of determination R2adj, the predicted
coefficient of determination R2pred, the predicted residual error sum of square PRESS, the
Fisher test Fstat and the standard deviation SD, with a critical probability p-value<0.05 for all
the used descriptors.

3. RESULTS AND DISCUSSION

3.1. Quantum calculations


3.1.1. Protonation study
In order to determine the predominant form present in the used electrolytic solution, it is
required to study the possibility of protonation of the four investigated benzodiazepine
derivatives. Using Marvin software, Figs. 1 and 2 present the percentage distribution of the
species as a function of pH of the medium for both BZD1 and BZD2 derivatives at (298 K).
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 379

Fig. 1. (a) distribution diagram of the BZD1 micro-species at T=298 K and at different pH of
the medium and (b) process of deportation of BZD1

Fig. 2. (a) distribution diagram of the BZD2 micro-species at T=298 K and at different pH of
the medium and (b) process of deportation of BZD2
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 380

From Figs. 1 and 2, the pKa values for both BZD1 and BZD2 molecules are 12.02 and
6.77, respectively, meaning that the neutral form of these two molecules is the predominant
form in molar hydrochloric acid solution. It is also noted that if pH>pKa, the two derivatives
BZD1 and BZD2 are deprotonated by releasing a proton H+ by a nitrogen atom.

Molecule Geometry
BZD1

BZD 2

BZD=2O

BZD=2S

Fig. 3. The selected bond lengths (Å) of the optimized studied benzodiazepine derivatives
calculated at B3LYP/6-31G** in gas (Bold) and aqueous phases (underline)
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 381

The same calculation conducted for the compounds BZD=2O and BZD=2S reveals that
neither protonation nor deprotonation, whatever the pH value of the medium.
Accordingly, all the computational calculations were carried out with neutral forms of the
four benzodiazepine derivatives BZD1, BZD2, BZD=2O and BZD=2S, using more
energetically stable geometries in both gaseous and aqueous phases.

3.1.2. Molecular geometries


The structural geometries of the four studied benzodiazepine derivatives are fully
optimized in gaseous phases at DFT method, by using of B3LYP functional together with 6-
31G (d,p) basis set, and for a higher precision of the experimental parameters, the compound
geometries were re-optimized in aqueous phase at the same level of theory using the PCM
model.
The absence of imaginary frequency in the vibrational spectra demonstrates that the
equilibrium structures correspond to the minima energy for each of these inhibitor molecules
[9,16]. The final optimized geometries are illustrated in Fig. 3 along with the main bond
lengths (Å) and only the more significant bond angles (º) and the dihedrals angles (º) are
collected in Table 2.
It can be seen from Fig. 3 that, the values of the bond lengths indicate any net divergence
between gaseous and aqueous phases for the studied compounds. A detailed analysis shows
that the substitution of an oxygen atom in BZD1 and BZD=2O with a sulphur atom in BZD2
and BZD=2S, respectively, affect both the molecular and electronic structures of the
investigated derivatives. At this level, this substitution leads to a slight increase in the bond
lengths, which shows that the average bond lengths of C12-O1 and C7-O2 are 1.226 Å and
1.234 Å between BZD1 and BZD=2O, while these averages are 1.670 Å and 1.688 Å for
C12-S1 and C7-S2 between BZD2 and BZD=2S [9]. These results suggest that the adsorption
of BZD2 and BZD=2S on the surface of metal through the sulphur atom will be easier and
louder than that through the oxygen atom of BZD1 and BZD=2O. Indeed, as the rate of
absorption is significantly correlated to the corrosion inhibition efficiency, the enhancement
of this rate increases the inhibiting efficiency of BZD2 and BZD=2S with respect to BZD1
and BZD=2O respectively.
The average of the bond lengths for C12–N1 and C7–N2 are (~1.367 Å) and (1.354 Å),
respectively, are shorter than typical carbon nitrogen single bonds (1.42–1.47 Å) in gas phase
[31] reflecting the presence of n-σ-π electron delocalization in the studied inhibitors.
The C11–C12 and C6–C7 are the lowest bonds of all the benzodiazepine bond lengths
with an average (~1.534 Å) and (~1.498 Å) respectively, probably because of the proximity
of the strongly electronegative oxygen (O1,O2) and sulphur (S1,S2) atoms linked to C12 and
C7 atoms as well as the presence of nitrogen atoms (N1,N2) which are linked to C12 and C7,
respectively, in the benzodiazepine ring.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 382

The computed structural bond lengths in aqueous phase reveal that the effect of the
solvent on the structural parameters is almost insignificant. However, a noticeable
lengthening of C=S bond with an average (~0.015 Å), in solution is probably a consequence
of the highly polarity of this bond [32].
The analysis of Table 2 indicates that the substitution of the oxygen atom by the sulfur
atom in C7 and C12 for BZD1 and BZD2 doesn’t have appreciable effect on the bond angles
and dihedral angles for BZD=2O and BZD=2S either in gas or in aqueous phases; except for
a much smaller reduction of these selected angles for BZD1 compared to BZD2 and
BZD=2O compared to BZD=2S.

Table 2. The selected bonds and dihedral angles, in degree, of the used derivatives calculated
at DFT/B3LYP in gaseous (G) and aqueous (A) phases

Angle Phase BZD1 BZD2 BZD=2O BZD=2S


C1 – N1 – C12 G 129.65 129.15 124.52 123.51
A 128.87 128.52 124.41 123.71
C7 – N2 – C11 G 125.69 124.85 125.09 124.23
A 125.43 124.31 124.80 123.63
N1– C12 – O1 G 121.19 – 122.12 –
A 121.35 – 121.94 –
N2– C7 – O2 G 121.35 – 121.88 –
A 121.26 – 121.80 –
G – 121.31 – 124.24
N1– C12 – S1
A – 121.35 – 123.49
N2– C7 – S2 G – 121.92 – 122.64
A – 121.62 – 122.21
C12– N1 – C1– C2 G 142.43 139.48 137.09 133.86
A 141.64 139.84 137.27 131.95
N2– C7 – C6– C5 G 152.89 145.51 144.49 137.47
A 150.44 143.99 143.22 137.44
N1– C12 – C11– G 179.54 176.43 178.56 174.54
C10 A 178.93 176.45 178.44 173.84

C12– C11 – N2– G 114.26 109.19 111.27 108.90


C8 A 112.90 109.15 110.69 107.20

A comparison of dihedrals angles derived from the optimized geometry for both gaseous
and aqueous phases shows no significant difference between the two phases for the studied
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 383

compounds. However, a negligible diminution is noted in liquid phase, especially, for the
dihedral angles [C12–C11–N2–C8] and [C12–N1–C1–C2] of BZD1 and BZD=2S
respectively, as well for the dihedral angle [N2–C7–C6–C5] of both derivatives BZD1 and
BZD2. This deviation is probably the result of the extreme polarity of the bonds constituting
these dihedrals.

3.1.3. Global molecular reactivity


All quantum chemical parameters used are those derived from the ground state for the
optimized geometries of the investigated derivatives.

Table 3. quantum chemical parameters of the four studied compounds using DFT-B3LYP /6-
31G** calculations, in gas (G) and in aqueous (A) phases and their experimental corrosion
inhibition efficiencies

Inhibitor Phase BZD1 BZD2 BZD=2O BZD=2S


ET (a.u.)
G -724.5105 -1370.4210 -994.8750 -1640.7806
A -724.5335 -1370.4458 -994.8946 -1640.8019
EHOMO (eV) G -6.2970 -5.5904 -6.2489 -5.5398
A -6.2736 -5.9785 -6.3522 -5.9937
ELUMO (eV) G -1.0447 -1.7718 -0.9726 -1.5664
A -0.9938 -1.8003 -1.0096 -1.7193
∆E (eV) G 5.2523 3.8186 5.2762 3.9733
A 5.2798 4.1782 5.3426 4.2744
µ (D) G 2.7012 3.7688 3.0206 4.3847
A 4.3128 6.4781 4.5117 7.0316
IP (eV) G 6.2970 5.5904 6.2489 5.5398
A 6.2736 5.9785 6.3522 5.9937
EA (eV) G 1.0447 1.7718 0.9726 1.5664
A 0.9938 1.8003 1.0096 1.7193
χ (eV) G 3.6708 3.6811 3.6107 3.5531
A 3.6337 3.8894 3.6809 3.8565
ƞ (eV) G 2.6261 1.9093 2.6381 1.9867
A 2.6399 2.0792 2.6713 2.1372
σ (eV-1) G 0.3807 0.5237 0.3790 0.5033
A 0.3788 0.4809 0.3743 0.4679
ΔN G 0.6338 0.8691 0.6423 0.8674
A 0.6375 0.7480 0.6212 0.7354
Eexp% a _ 81.50 87.90 84.20 90.00
a
Inhibition efficiencies data from [23,24]

The frontier molecular orbital theory (FMO) provides the information about the reactive sites
and the nature of orbitals available in the inhibitor molecule. According to this theory, the
chemical reactivity is a function of the interaction between the highest occupied molecular
orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) levels of reacting
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 384

species [33]. The HOMO and LUMO are often used to describe prospective donation and
retro-donation mechanism between the inhibitor molecules and metal atom. The HOMO
allows to provide information about the most likely sites able to donate electrons of the
inhibitor molecules to the appropriate vacant orbital of the metal atom to afford interactions
that will lead to adsorption and corrosion inhibition. Therefore, higher values of EHOMO
indicate better tendency towards the donation of electron, enhancing the adsorption of the
inhibitor on C-Steel and therefore better inhibition efficiency. The LUMO provides
information about the regions in a molecule that possess a high tendency to receive electrons
from electron rich compounds. In the inhibitor-metal interactions, the LUMO is a pointer to
the sites of the molecule where electrons are most likely to be received from the appropriate
occupied orbital of the metal atom. The calculated molecular parameters of the frontier
orbitals are collected in Table 3.
In relation to the effect of substitution between BZD1 and BZD2 on the one hand, and
BZD=2O and BZD=2S on the other, Table 3 shows that the higher EHOMO (-5.5904 eV) and
the lower ELUMO (-1.7718 eV) of BZD2 indicate that this inhibitor would be willing to play a
dual role, i.e. to donate and accept electrons when compared to BZD1 which has -6.2970 eV
and -1.0447 eV for EHOMO and ELUMO, respectively.
Otherwise, the gap (ΔE = ELUMO −EHOMO) is a significant parameter as a function of
reactivity of the inhibitor towards the adsorption on the metallic surface. As ∆E decreases, the
molecular reactivity increases, leading to increase in the E%. The results as indicated in Table
3 follow the order: ΔE (BZD2)<ΔE (BZD1), which shows that inhibitor BZD2 has the lowest
ΔE; this means that this derivative could have better performance as corrosion inhibitor than
the molecule BZD1.
Likewise, it can be clearly seen that EHOMO and ELUMO for BZD=2S are respectively,
higher and lower than those for BZD=2O, implying that ΔE (BZD=2S)<ΔE (BZD=2O)
which probably confirms improved reactivity for BZD=2S.
The dipole moment (μ) is another important electronic parameter which provides the
information on the molecular reactivity and polarity. In the corrosion inhibition study, this
parameter does not always show univocal trend; several researches have reported that the
dipole moment increases with increase in the inhibition efficiencies of the inhibitor
compounds [34]; but, other researchers have shown this index decreases with increase in the
inhibition efficiencies of the inhibitors [35]; in this context, there are also research works that
show this parameter does not have proper relationship with the inhibition efficiencies of the
inhibitors [36]. A clear comparison of the trend in the dipole moment of the studied inhibitors
and the trend in their corrosion inhibition efficiencies indicates that the inhibition efficiency
increases with increasing the values of this index, which could be attributed as higher polarity
compounds will facilitates electrostatic interaction between the electric field as a result of the
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 385

charged metal surface and electric moments of the molecule of inhibitor and contributes to a
better adsorption by influencing the transport process through the adsorbed layer [37,38].

HOMO LUMO

BZD1

BZD2

BZD=2O

BZD=2S

Fig. 4. The HOMO and the LUMO electron density distributions of the studied compounds
computed at B3LYP/6-31G (d, p) level in gaseous phase

Ionization potential (IA) is an essential descriptor of the chemical reactivity of atoms and
compounds. Higher IA shows that the high stability and chemical inertness and smaller IA
indicate high reactivity of the chemical species [39].
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 386

Table 4. Relevant natural populations and Fukui functions of the studied inhibitors calculated
at B3LYP/6-31G** in gaseous (G) and aqueous (A) phases

Inhibitor Atom Phase P(N) P(N+1) P(N-1) f k+ f k- f k0


BZD1 C6 G 6.162 6.266 6.149 0.104 0.013 0.058
A 6.169 6.279 6.121 0.110 0.048 0.079
C3 G 6.211 6.359 6.192 0.148 0.019 0.083
A 6.218 6.374 6.214 0.156 0.004 0.080
N2 G 7.472 7.503 7.361 0.031 0.111 0.071
A 7.457 7.494 7.416 0.037 0.041 0.039
N1 G 7.641 7.626 7.562 -0.015 0.079 0.032
A 7.643 7.639 7.520 -0.004 0.123 0.059
O2 G 8.623 8.715 8.464 0.092 0.159 0.125
A 8.687 8.781 8.589 0.094 0.098 0.096
O1 G 8.602 8.675 8.464 0.073 0.138 0.105
A 8.655 8.699 8.522 0.044 0.133 0.088
BZD2 C4 G 6.240 6.297 6.214 0.057 0.026 0.041
A 6.244 6.288 6.231 0 .044 0.013 0.028
C3 G 6.211 6.283 6.168 0.072 0.043 0.057
A 6.213 6.276 6.178 0.063 0.035 0.049
N2 G 7.397 7.429 7.354 0.032 0.043 0.037
A 7.375 7.413 7.338 0.038 0.037 0.037
N1 G 7.575 7.568 7.554 -0.007 0.021 0.007
A 7.565 7.569 7.551 0.004 0.014 0.009
S2 G 16.171 16.348 15.726 0.177 0.445 0.311
A 16.311 16.484 15.668 0.173 0.643 0.408
S1 G 16.127 16.349 15.872 0.222 0.255 0.238
A 16.244 16.455 16.111 0.211 0.133 0.172
BZD=2O C6 G 6.147 6.235 6.142 0.088 0.005 0.046
A 6.153 6.272 6.124 0.119 0.029 0.074
C3 G 6.215 6.320 6.199 0.105 0.016 0.060
A 6.218 6.373 6.209 0.155 0.009 0.054
N2 G 7.467 7.492 7.369 0.025 0.098 0.061
A 7.453 7.489 7.381 0.036 0.072 0.054
N1 G 7.462 7.449 7.405 -0.013 0.057 0.022
A 7.456 7.449 7.362 -0.007 0.094 0.043
O2 G 8.624 8.696 8.490 0.072 0.134 0.103
A 8.685 8.777 8.566 0.092 0.119 0.105
O1 G 8.628 8.679 8.508 0.051 0.120 0.085
A 8.658 8.696 8.525 0.038 0.133 0.085
BZD=2S C5 G 6.187 6.248 6.181 0.061 0.006 0.033
A 6.192 6.232 6.185 0.040 0.007 0.023
N2 G 7.467 7.413 7.356 -0.054 0.111 0.028
A 7.371 7.448 7.336 0.077 0.035 0.056
N1 G 7.462 7.433 7.378 -0.029 0.084 0.027
A 7.390 7.377 7.371 -0.013 0.019 0.003
S2 G 16.173 16.359 15.799 0.186 0.374 0.280
A 16.307 16.584 15.762 0.277 0.545 0.411
S1 G 16.146 16.320 15.877 0.174 0.269 0.221
A 16.212 16.329 15.988 0.117 0.224 0.170
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 387

From Table 3, the low ionization potential of BZD=2S (5.5398eV), BZD2 (5.5904 eV)
signifies the high corrosion inhibition efficiency of these compounds in comparison with
BZD=2O (6.2489 eV) and BZD1 (6.2970 eV), respectively. Likewise, it can be also observed
that the electron affinity (EA) increases with the increase of E%, which is consistent with the
meaning of this parameter.
The chemical hardness (η) signifies fundamentally the resistance towards the deformation
or the polarization of the electron cloud of the atoms, molecules or ions under slight
perturbation of chemical reaction; i. e. a hard inhibitor has a large ΔE and a soft inhibitor has
a small ΔE [36]. The global hardness values presented in Table 3 show that these results
corroborate the previous predictions, derived from the other chemical parameters, and justify
the high reactivity of BZD=2S and BZD2 compared to BZD1 and BZD=2O, respectively.
The fraction of electrons transferred (∆N) was also evaluated and tabulated in Table 3.
Generally, ΔN defines the tendency of electrons donation within a set of inhibitor. According
to the Lukovits’s research [40], if ΔN < 3.6, the corrosion inhibition efficiency increased with
increasing electron-donating ability at the C-Steel/electrolyte interface. Table 3 demonstrates
that the calculated values of ΔN increase in the following order; BZD=2S (0.8674)>BZD=2O
(0.6423) and BZD2 (0.8691)>BZD1 (0.6338) which indicate that these values correlate
highly with the experimental inhibition efficiencies order.

3.1.4. Local molecular reactivity


The HOMO and LUMO electron density surfaces of the studied inhibitors are shown in Fig.
4. The almost uniform electron density distribution of the frontier orbitals illustrate in Fig. 4
indicates that the absorption will probably occur on multiple reactive sites distributed along
the structure of inhibitor, especially, for the heteroatom sites, which may increase the
adsorption stability and led to the enhancement of inhibition potential of the studied
molecules.
According to Table 4, the calculated Fukui functions show that all the inhibitors have
anionic and cationic sites for attack, in gaseous as well as in solution, especially, for the
heteroatom sites. In this context, it can be observed that the S1 and S2 are the most
significant sites susceptible for electrophilic attacks in BZD2 and BZD=2S. Indeed, they have
the most elevated values of f k which attain 0.255 and 0.269 for S1 and 0.445 and 0.374 for
S2 in BZD2 and BZD=2S, respectively. This reason may clarifies the best high inhibition
efficiencies of these derivatives in comparison with BZD1 and BZD=2O, respectively. In
fact, the values of f k corresponding to the oxygen atoms are of only 0.138 and 0.120 for O1
and 0.159 and 0.134 for O2 in BZD1 and BZD=2O, respectively. On the other hand, it is to
be noted that the studied molecules have also other sites for nucleophilic and/or electrophylic
attacks for (N1, N2, C3, C4, C5 and C6), which increases the adsorption of these derivatives
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 388

onto the surface of metal and explains their higher inhibition efficiency in the hydrochloric
acid solution.

3.2. QSPR study


3.2.1. Regression equations and correlation analysis
Using the statistical software Minitab 16, the obtained equations of the two regression
models are as follows:

a. Multiple linear regression model (MLR)


The obtained regression equation is:

E cal %  2507.3164  26146.9862Ci  335.5572E HOMO  185.7866E LUMO  95.4669 (10)

Statistical characteristics of this obtained equation are:


N = 16; R2= 0.88; R2adj = 0.87; R2pred = 0.85; PRESS = 712.04; SD = 5.29

b. Multiple polynomial regression model (MPR)


The regression equation obtained with this model is:

E cal %  1081 22343183Ci  139E HOMO  89E LUMO  26


 3013464 E HOMO  Ci  1853634E LUMO  Ci  624092  Ci (11)
 7,36 10 E HOMO  C  4.30 10 E LUMO  C  7.96 10   C
8
i
2 8
i
2 8
i
2

Statistical characteristics of this obtained equation are:

N = 16; R2= 0.99; R2adj = 0.99; R2pred= 0.97; PRESS = 332.17; SD = 0.99.

The predicted Ecal % values calculated from the equations (10, 11) using the MLR and
MPR models are given in Table 5 along with the residual errors (RE).
A first comparison of the statistical characteristics values of two models, in particular,
the coefficient of determination R2, the adjusted coefficient of determination R2adj and the
standard deviation SD demonstrate that the multiple polynomial regression model (MPR) is
very efficient that the multiple linear model (MLR). This implies that it has a very good
quadratic correlation between the inhibition efficiency and the molecular structure. Indeed,
the selected parameters have significant regression coefficients and have a very important
contribution in the explanation of the MPR model.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 389

Table 5. Eexp% vs Ecal% obtained by MLR and MPR models, proposed for DFT-B3LYP/6-
31G** along with the residual error (RE)

Inhibitor 106 Ci / M Eexp% Ecal% Residual Error (RE)


MLR MPR RE.-1 RE.-2
BZD1 1 37.5 40.58 38.34 -3.08 -0.84
10 58.6 57.81 57.77 0.79 0.83
100 70.0 77.70 70.98 -7.70 -0.98
1000 81.5 74.71 81.49 6.79 0.01
BZD2 1 45.5 50.59 45.02 -5.09 0.48
10 65.7 61.63 65.23 4.07 0.47
100 75.5 69.87 76.44 5.63 -0.94
1000 87.9 80.22 87.89 7.68 0.01
BZD=2O 1 47.8 51.63 46.96 -3.83 0.84
10 61.3 58.87 60.56 2.43 0.74
100 78.6 83.80 79.17 -5.20 -0.57
1000 84.2 78.17 84.19 6.03 0.01
BZD=2 S 1 58.6 64.47 59.47 -5.87 -0.87
10 75.0 70.06 77.36 4.94 -2.36
100 83.8 76.47 84.56 7.33 -0.76
1000 90.0 85.82 89.99 4.18 0.01
RE.-1: Residual error between Ecal% obtained by using MLR (Eq. 10) and Eexp%
RE.-2: Residual error between Ecal% obtained by using MPR (Eq. 11) and Eexp%

The comparison between the experimental and calculated efficiency values is represented
in Fig. 5.

100
90
Inhibition efficiency (E %)

80
70
60
50 Eexp%
40 Epred % MNLR
30 Epred % MLR
20
10
0
1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16

Observations

Fig. 5. Eexp% and Ecal% obtained by MLR and MPR models Eq. ((10), (11)) proposed of
compounds BZD1–BZD=2O usingB3LYP/6-31G** calculations
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 390

The diagnosis of residual error (RE) between the experimental and calculated inhibition
efficiency illustrated in Table 5, shows that the MPR model explains the inhibition efficiency
with a high level of significance when compared to MLR model. This degree of significance
is demonstrated by the best concordance between the Eexp% and Ecal%, as shown in Fig. 5.
However, to choose the appropriate statistical model, it was necessary to compare the
prediction indicators of these models. Effectively, the predicted determination coefficient
(R2pred) and the predicted residual error sum of squares (PRESS) are the most used parameters
to evaluate the quality of prediction of any model. So, the obtained values of these parameters
show that the MPR is the most significant predictive model due to its low PRESS which
equals 332.17 value and high predicted determination coefficient R2pred=0.97 when compared
to MLR model.
The correlation between experimental and predicted efficiency is illustrated in Fig.6.

Fig. 6. The experimental vs. calculated E% (Eq. 10, 11) efficiencies of benzodiazepine
derivatives by the use of MLR model (a) and MPR model (b)

According to Fig. 6, the multiple linear regression curve does not give satisfactory results,
based on the gap between the experimental and calculated efficiencies, clearly identified in
the allure of the plot (a) correlating the Ecal% obtained by eq. 10 for each concentration with
the Eexp%. Indeed, it has a low determination coefficient R2=0.88 and correlation coefficient
R=0.94. On the contrary, it seems that the values of the calculated efficiencies Ecal% obtained
from the multiple polynomial regression curve (b) are in very good agreement with the
experimental efficiencies Eexp% values, with R2=0.99 and R=0.99.
In order to highlight the presence or not of a systematic error in developing the two MLR
and MPR models, Fig. 7 exemplifies the residual errors (RE) versus the studied observations.
Despite the high values of the residual error (RE.-1) presented by MLR, the random
dispersion of the RE on both sides of zero illustrated by Fig. 7 justifies that no systemic error
exists, as suggested by Jalali-Heravi and Kyani [41]. This implies that the MPR model can be
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 391

successfully applied to predict the corrosion inhibition efficiency of benzodiazepine


derivatives.

Fig. 7. The residual error (RE) between the predicted and experimental inhibition efficiency
of benzodiazepine derivatives at different concentrations

3.2.2. Variance analysis and validation


The sums of squares SS, mean squares MS, F-obs, F-stat and p-value with the degrees of
freedom DF, are summarized in ANOVA Table 6:

Table 6. ANOVA for the two models (MLR and MPR)

Model Source SS DF MS Fobs Fstat p-value


MLR Regression 2231.16564 1 2231.1656 102.80879 4.60 <0.0001**
Residual error 303.82926 14 21.702090
Total 2534.9949 15
MPR Regression 3951.31624 1 3951.3162 5203.92634 4.60 <0.0001**
Residual error 10.6301300 14 0.7593000
Total 3961.94637 15
** Indicates highly significant at level 99%

The assessment of the ANOVA results shows that:


The variation related to the regression SSreg is largely greater than the residual error SSres with
an observed Fisher test Fobs superior than the statistical Fisher test Fstat at a critical probability
(p–value<0. 0001) clearly below the tolerated risk (5%). This indicates that the null
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 392

hypothesis is wrong, and implies, that the models bring globally a significant amount of
information with 95% of confidence, which means the good quality and best prediction of
selected models, especially, for the MPR model. Indeed, it is characterized by the higher
values of the adjusted determination coefficient and the predicted determination coefficient:
R2adj=0.99 and R2pred =0.97, respectively along with the lower value of the predicted residual
error sum of square: PRESS=332.17 and the negligible value of the standard deviation:
SD=0.99.
According to the previous considerations issued from the statistical indicators and the
ANOVA analysis, it is recommended to adopt the MPR model for the prediction of the
inhibitory potential of the studied derivatives.

4. CONCLUSION
The corrosion inhibition phenomenon by new benzodiazepine derivatives has been
theoretically investigated by using of DFT quantum chemical method. The theoretical
calculations show that the benzodiazepine derivatives namely: BZD1, BZD2, BZD=2O and
BZD=2S, present good inhibition efficiencies of steel corrosion in acid medium at very low
concentrations, and E% of these organic inhibitors can be directly linked to the structure
electronic parameters. The molecule inhibition efficiencies increase in the order of
BZD2>BZD1> and BZD=2S>BZD=2O, which implies the substitution of oxygen atom by
sulphur atom leads to increase the adsorption capacity of inhibitors on the metallic surface
which significantly enhances the inhibition efficiency of the studied compounds. These
predictions are in fairly consistent with the experimental results.
Using QSPR approach, we have determined a direct relationship between the molecular
structure and the inhibition efficiency of four benzodiazepine derivatives, by the use of a
multiple linear regression (MLR) and a multiple polynomial regression (MPR) models
joining the E% to chemical quantum descriptors. The statistical results demonstrate that MPR
is the most predictive statistic model to study the inhibitors selected in this work in
comparison with MLR model. The relevance of this model is reflected by the high adjusted
determination coefficient R2adj=0.99 which indicated that about 99% of the variables are
covered in the response. Also the results indicate high correlation coefficient R=0.99,
demonstrating the good compatibility between the obtained calculated inhibition efficiencies
and the corresponding experimental values.
The established MPR model equation can be also used to predict the corrosion efficiency
of new benzodiazepine derivatives of the same family.
Acknowledgements
The authors gratefully thank the ‘‘Association Marocaine des Chimistes Théoriciens’’
(AMCT) for the means of the calculations.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 393

Symbols

DFT Density Functional Theory


B3LYP/6-31G (d,p) Becke, 3-parameter, Lee-Yang-Parr/Basis set 6-31G with
polarization functions on heavy atoms and hydrogen
EHOMO Energy of the highest occupied molecular orbital
ELUMO Energy of the lowest unoccupied molecular orbital
µ Dipole moment
Ci Concentration of inhibitor
Epred (%) Predicted corrosion inhibition efficiency
Eexp (%) Experimentally corrosion inhibition efficiency
QSPR Quantitative Structure–Property Relationships
MLR Multiple Linear Regression
MPR Multiple Polynomial Regression
N Observations
R Correlation coefficient
R2 Determination coefficient
Radj2 Adjusted determination coefficient
R2pred Predicted determination coefficient
PRESS Predicted residual error sum of squares
SS Sum of squares
SSres Residual sum of squares
SSreg regression sum of squares
SST Total sum of squares
DF Degree of freedom
MS Mean of squares
Fstat Statistic value of Fisher.
Fobs Observed value of Fisher
p-Value Critical probability
Α Critical probability value
SD Standard deviation

REFERENCES
[1] A. S. Patel, V. A. Panchal, P. T. Trivedi, and N. K. Shah, Port. Electrochim. Acta 30
(2012) 163.
[2] I. Belfilali, A. Chatouani, B. Hammouti, A. Aounti, S. Louhibi, and S. S. Al-Deyab, Int.
J. Electrochem. Sci. 7 (2012) 3997.
[3] H. Zarrok, A. Zarrouk, R. Salghi, H. Oudda, B. Hammouti, M. E. Touhami, M.
Bouachrine, and S. Boukhris, Electrochim. Acta 30 (2012) 405.
[4] F. Bentiss, M. Lebrini, N. E. Chihib, M. Abdalah, C. Jama, M. Lagrenee, and S. S. Al-
Deyab, Int. J. Electrochem. Sci. 7 (2012) 3947.
[5] M. Benabdellah, B. Hammouti, A. Warthan, S. S. Al-Deyab, C. Jama, M. Lagrenee, and
F. Bentiss, Int. J. Electrochem. Sci. 7 (2012) 3489.
[6] B. A. Suraj, M. N. Deshpande, and D. G, Kolhatkar, J. Chem. Pharm. Res. 4 (2012)
1033.
[7] S. Rajendran, C. Thangavelu, and G. Annamalai, J. Chem. Pharm. Res. 4 (2012) 4836.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 394

[8] T. A. Aljohani, B. E. Hayden, and A. Anastasopoulos, Electrochim. Acta 76 (2012)


389.
[9] Z. El Adnani, M. Mcharfi, M. Sfaira, M. Benzakour, A. T. Benjelloun, and M. E.
Touhami, Int. J. Electrochem. Sci. 7 (2012) 6738.
[10] N. O. Eddy, and S. A. Odoemelam, Adv. Nat. Appl. Sci. 2 (2008) 35.
[11] S. A. Umoren, I. B. Obot, E. E. Ebenso, and N. O. Obi-Egbedi, Int. J. Electrochem. Sci.
3 (2008) 1029.
[12] P. Udhayakala, A. Maxwell Samuel, T. V. Rajendiranand, and S. Gunasekaran, J.
Chem. Pharm. Res. 5 (2013) 142.
[13] E. Kraka, and D. Cremer, J. Am. Chem. Soc. 122 (2000) 8245.
[14] O. Mokhtari, I. Hamdani, A. Chetouani, A. Lahrach, H. El Halouani, A. Aouniti, and
M. Berrabah, J. Mater. Environ. Sci. 5 (2014) 310.
[15] S. S. Shivakumar, and K. N. Mohana, Int. J. Corros. 2013 (2013)
Doi:10.1155/2013/543204.
[16] Z. El Adnani, M. Mcharfi, M. Sfaira, M. Benzakour, A. T. Benjelloun, and M. E.
Touhami, Corros. Sci. 68 (2013) 223.
[17] L. Vera, M. Guzman, and P. Ortega-Luoni, J. Chil. Chi. Soc. 51 (2006) 1034.
[18] El H. El Assiri, M. Driouch, Z. Bensouda, F. Jhilal, T. Saffaj, M. Sfaira, and Y.
Abboud, Desalin. Water Treat. 111 (2018) 208.
[19] I. Lukovits, E. Kalman, and G. Palinkas, Corrosion 51 (1995) 201.
[20] I. Lukovits, I. Bakó, A. Shaban, and E. Kálmán, Electrochim. Acta 43 (1998) 131.
[21] F. Bentiss, M. Traisnel, H. Vezin, and M. Lagrenée, Corrosion science 45 (2003) 371.
[22] M. Lebrini, M. Lagrenee, H. Vezin, L. Gengembre, and F. Bentiss, Corros. Sci. 47
(2005) 485.
[23] W. Niouri, B. Zerga, M. Sfaira, M. Taleb, M. EbnTouhami, B. Hammouti, M. Mcharfi,
S. S. Al-Deyab, H. Benzeid, and El M. Essassi, Int. J. Electrochem. Sci. 9 (2014) 8283.
[24] W. Niouri, B. Zerga, M. Sfaira, M. Taleb, B. Hammouti, M. EbnTouhami, S. S. Al-
Deyab, H. Benzeid, and El M. Essassi, Int. J. Electrochem. Sci. 7 (2012) 10190.
[25] Gaussian 03, Revision B.01, M. J. Frisch, G. W. Trucks, H. B. Schlegel, G. E. Scuseria,
M. A. Robb, J. R. Cheeseman, J. A. Montgomery Jr., T. Vreven, K. N. Kudin, J. C.
Burant, J. M. Millam, S. S. Iyengar, J. Tomasi, V. Barone, B. Mennucci, M. Cossi, G.
Scalmani, N. Rega, G. A. Petersson, H. Nakatsuji, M. Hada, M. Ehara, K. Toyota, R.
Fukuda, J. Hasegawa, M. Ishida, T. Nakajima, Y. Honda, O. Kitao, H. Nakai, M. Klene,
X. Li, J. E. Knox, H. P. Hratchian, J. B. Cross, C. Adamo, J. Jaramillo, R. Gomperts, R.
E. Stratmann, O. Yazyev, A. J. Austin, R. Cammi, C. Pomelli, J. W. Ochterski, P. Y.
Ayala, K. Morokuma, G. A. Voth, P. Salvador, J. J. Dannenberg, V. G. Zakrzewski, S.
Dapprich, A. D. Daniels, M. C. Strain, O. Farkas, D. K. Malick, A. D. Rabuck, K.
Raghavachari, J.B. Foresman, J. V. Ortiz, Q. Cui, A. G. Baboul, S. Clifford, J.
Anal. Bioanal. Electrochem., Vol. 11, No. 3, 2019, 373-395 395

Cioslowski, B. B. Stefanov, G. Liu, A. Liashenko, P. Piskorz, I. Komaromi, R. L.


Martin, D. J. Fox, T. Keith, M. A. Al-Laham, C. Y. Peng, A. Nanayakkara, M.
Challacombe, P. M. W. Gill, B. Johnson, W. Chen, M. W. Wong, C. Gonzalez, and J.
A. Pople, Gaussian, Inc., Pittsburgh PA (2003).
[26] J. B. Foresman, and A. Frisch, Gaussian, Inc., Pittsburgh PA (USA) 15106 (1995) 96.
[27] C. C. Zhan, J. A. Nichols, and D. A. Dixon, J. Phys. Chem. A. 107 (2003) 4184.
[28] K. F. Khaled, Electrochim. Acta. 55 (2010) 6523.
[29] V. S. Sastri, and J. R. Perumareddi, Corrosion 53 (1997) 617.
[30] A. A. Siaka, N. O. Eddy, S. Idris, and L. Magaji, Res. J. Appl. Sci. 6 (2011) 487.
[31] F. H. Allen, O. Kennard, D. G. Watson, L. Brammer, A. G. Orpen, and R. J. taylor,
Chem. Soc. Perkin. Trans. 2 (1987) S1–S19.
[32] R. Sustmann, W. Sicking, and R. Huisgan, J. Am. Chem. Soc. 117 (1995) 9679.
[33] A. Y. Musa, A. H. Kadhum, A. B. Mohamad, A. B. Rohoma, and H. Mesmari, J. Mol.
Struct. 969 (2010) 233.
[34] C. M. Goulart, A. Esteves-Souza, C. A. Martinez-Huitle, C. J. F Rodriguez, M. A.M.
Maciel, and A. Echevarria, Corros. Sci. 67 (2013) 281.
[35] I. B. Obot, E. E. Ebenso, and M. M. Kabanda, J. Environ. Chem. Eng. 1 (2013) 431.
[36] N. O. Obi-Egbedi, I. B. Obot, M. I. El-Khaiary, S. A. Umoren, and E. E. Ebenso, Int. J.
Electro. chem. Sci. 6 (2011) 5649.
[37] Y. M. Tang, W. Z. Yang, X. S. Yin, Y. Liu, R. Wan, and J. T. Wang, Mater. Chem.
Phys. 116 (2009) 479.
[38] K. Babić-Samardžija, K. F. Khaled, and N. Hackerman, Appl. Surf. Sci. 240 (2005)
327.
[39] S. K. Rajak, N. Islam, and D. C. Ghosh, J. Quantum Inf. Sci. 1 (2011) 87.
[40] I. Lukovits, E. Kálmán, and F. Zucchi, Corrosion 57 (2001) 3.
[41] M. Jalali-Heravi, J. Chem. Inf. Comput. Sci. 44 (2004) 1328.

Copyright © 2019 by CEE (Center of Excellence in Electrochemistry)


ANALYTICAL & BIOANALYTICAL ELECTROCHEMISTRY (http://www.abechem.com)

Reproduction is permitted for noncommercial purposes.

You might also like