Experimental and Theoretical Investigations on Copper Corrosion Inhibition by Cefixime Drug in 1M HNO3 Solution

Abstract

Cefixime, a third-generation semi-synthetic cephalosporin antibiotic was used as a copper corrosion inhibitor in 1M HNO3 solution. The study was conducted through the weight loss technique at 298 - 318 K and theoretical studies based on quantum chemistry. The studied drug inhibited the corrosion of copper in 1M HNO3 over the cefixime concentration range (0.02 - 2 mM). The inhibition efficiency increased with an increase in the inhibitor concentration to reach 91.07% at 2 mM, but decreased with an increase in temperature. The thermodynamic functions related to the adsorption of cefixime on the copper surface and that of the metal dissolution were computed and analyzed. The results point out spontaneous adsorption, mainly through a physisorption mechanism following Langmuir adsorption isotherm model and an endothermic dissolution process. Quantum chemical calculations were also performed at B3LYP level with 6-31G (d, p) basis set and lead to molecular descriptors such as EHOMO (energy of the highest occupied molecular orbital), ELUMO (energy of the lowest unoccupied molecular orbital), ΔE (energy gap) and μ (dipole moment). The global reactivity descriptors such as χ (electronegativity), χ (global hardness), S (global softness), and ω (electrophilicity index) were derived using Koopman’s theorem and analyzed. The local reactivity parameters, including Fukui functions and dual descriptors were determined and discussed. Experimental and theoretical results were found to be in good agreement.

Share and Cite:

Diki, N. , Coulibaly, N. , Kambiré, O. and Trokourey, A. (2021) Experimental and Theoretical Investigations on Copper Corrosion Inhibition by Cefixime Drug in 1M HNO3 Solution. Journal of Materials Science and Chemical Engineering, 9, 11-28. doi: 10.4236/msce.2021.95002.

1. Introduction

Copper [1] [2] is one of the most used metals for several engineering and industrial applications including electricity and electronics, communications, pipelines for domestic and industrial water utilities, fabrication of heat exchanger tubes, and cooling water systems due to its excellent properties. Thus, copper and its alloys are deployed in various environments containing acids, alkalis, and salt solutions where they undergo corrosive attacks although the copper is relatively a noble metal [3]. Some ways of corrosion prevention are the use of more resistant alloys, protective coating, and the addition of inhibitors to the corrosive environment [4] [5]. Among numerous ways to protect the copper from acid corrosion, the most effective and practical method [6] [7] to slow down corrosion processes, is the addition of organic corrosion inhibitors to the metal’s environment. It is reported [8] [9] that the adsorption of these compounds is influenced by the electronic structure of inhibiting molecules, the steric factor, aromaticity, electron density at the donor site, the functional group, and also the polarizability of the group. Previous studies revealed that heterocyclic compounds [10] are employed as corrosion inhibitors because of the presence of various adsorption centers (O, N, S, P, and π electrons) which can help to form complexes with metal ions. Unfortunately, the use of some organic chemical inhibitors is limited by diverse reasons especially their costly synthesis, poor biodegradability, toxic and hazardous for human beings and the environment as well. For this purpose, natural products of plant origin or drugs are a better choice due to the fact that they are environmentally benign and contain incredibly rich sources of naturally synthesized organic compounds among which most of them are known to have inhibitive action. Plant extracts are often insoluble in aqueous media and extraction efficiencies are mainly low. The choice of some drugs used as corrosion inhibitors is based on the following justifications: 1) drugs are mainly soluble in aqueous media; 2) drug molecules contain oxygen, nitrogen and sulphur as active centers; 3) drugs are reportedly environmentally friendly and important in biological reactions; and 4) drugs including β-lactam antibiotics, quinolones, antifungal, and cephalosporins can be easily produced and purified [11] [12] [13] [14]. The aim of the present paper is to highlight the relationship between the calculated quantum chemical parameters and the experimentally determined inhibition efficiency of cefixime drug (Scheme 1) against copper corrosion in 1M HNO3 medium. This can be achieved by computing the most relevant electronic properties of the studied compound including EHOMO, ELUMO, energy Gap (ΔE), dipole moment (μ), electronegativity (χ), global hardness (η), fraction of electrons transferred (ΔN) and charges (δ) on atoms.

2. Experimental

2.1. Copper Samples

The copper specimens were in form of rod measuring 10 mm in length and 2.2 mm of diameter. They were cut in commercial copper of purity 95%.

Scheme 1. Chemical structure of cefixime.

2.2. Solutions

Cefixime of analytical grade was acquired from Sigma-Aldrich. Analytical grade 65% nitric acid solution from Sigma-Aldrich Chemicals was used to prepare the corrosive aqueous solution. The solution was prepared by dilution of the commercial nitric acid solution using double distilled water. The blank was a 1M HNO3 solution. Solutions of cefixime with concentrations in the range of 0.02 mM to 2 mM were prepared. Acetone of purity 99.5% was also purchased from Sigma-Aldrich Chemicals.

2.3. Weight Loss Technique

Prior to all measurements, the copper samples were mechanically abraded with different grade emery papers (1/0, 2/0, 3/0, 4/0, 5/0, and 6/0). The specimens were washed thoroughly with double distilled water, degreased and rinsed with acetone and dried in a moisture-free desiccator. Weight loss measurements were carried out in a beaker of 100 mL capacity containing 50 mL of the test solution. The immersion time for weight loss was 1 h at a given temperature. In order to get good reproducible data, parallel triplicate experiments were performed accurately and the mean value of the weight loss was used to assess the corrosion rate (CR), the degree of surface coverage (θ) and the inhibition efficiency (IE) using Equations (1)-(3) respectively:

C R = Δ m S t (1)

θ = C R 0 C R C R 0 (2)

I E ( % ) = ( C R 0 C R C R 0 ) × 100 (3)

where CR0 and CR (expressed in mg·h−1·cm−2) are respectively the corrosion rate without and with cefixime, ∆m is the weight loss, S is the total surface of the copper specimen and t is the immersion time.

2.4. Quantum Chemical Approach

The present quantum chemistry calculations were performed with Gaussian 09 series of program package [15]. For this purpose, we used Becke’s three parameter exchange functional along with the Lee-Yang-Parr non-local correlation functional (B3LYP) [16] using 6-31G (d, p) basis Set. Figure 1 shows the optimized chemical structure of cefixime.

In order to explore the theoretical-experimental consistency, quantum chemical calculations were performed and the electronic properties resulting from this study were analyzed through global reactivity parameters and local reactivity parameters.

2.4.1. Global Reactivity Parameters

Recently, DFT has become a very powerful technique to probe the inhibitor/surface interaction and to analyze experimental data by providing insights into the chemical reactivity and selectivity [17] [18]. So, the molecular descriptors namely the highest occupied molecular orbital energy (EHOMO), the lowest unoccupied molecular orbital energy (ELUMO), the energy gap (ΔΕ = ΕLUMOΕHOMO) and the dipole moment (μ) were calculated. The reactivity descriptors, including ionization energy (I), electron affinity (A), electronegativity (χ), hardness (η), softness (S), the fraction (ΔN) of electrons transferred and the electrophilicity index (ω) were also calculated. According to Koopman’s theorem [19], the ionization energy (I) can be approximated as the negative of the highest occupied molecular orbital (HOMO) energy (Equation (4)):

I = E HOMO (4)

The negative of the lowest unoccupied molecular orbital (LUMO) energy is related to the electron affinity A and is obtained from Equation (5):

A = E LUMO (5)

The electronegativity [20] is obtained using the ionization energy I and the electron affinity A by Equation (6):

χ = I + A 2 (6)

The hardness which is the reciprocal of the electronegativity was obtained by Equation (7):

Figure 1. Optimized structure of cefixime calculated by B3LYP/6-31G (d, p).

η = I A 2 (7)

when the organic molecule is in contact with the metal, electrons flow from the system with lower electronegativity to that of higher electronegativity until the chemical potential becomes equal. The fraction of electrons transferred, ΔN, was estimated according to Equation (8) [21]:

Δ N = Φ Cu χ i n h 2 ( η Cu + η i n h ) (8)

The values of experimental work function Φ Cu = 4.98 eV [22] and hardness η Cu = 0 [23] (since for bulk metallic atoms I = A) were considered to calculate ΔN.

The global electrophilicity index, introduced by Parr [24] is given by Equation (9):

ω = χ 2 2 η (9)

2.4.2. Local Reactivity Parameters

The local selectivity of a corrosion inhibitor [25] is generally assessed using Fukui functions which enable us to distinguish each part of the studied compound on the basis of its chemical behavior due to different substituent functional groups. The Fukui function is defined as the derivative of the electronic density ρ(r) with respect to the number N of electrons:

f ( r ) = ( ρ ( r ) N ) v ( r ) (10)

The condensed Fukui functions provide information about atoms in a molecule that have a tendency to either donate (nucleophilic character) or accept (electrophilic character) an electron or a pair of electrons [26]. The nucleophilic and electrophilic Fukui function for an atom k [27] can be computed using a finite difference approximation as seen in Equations (11)-(12) respectively:

f k + = [ q k ( N + 1 ) q k ( N ) ] for nucleophilic attack (11)

f k = [ q k ( N ) q k ( N 1 ) ] for electrophilic attack (12)

where q k ( N + 1 ) , q k ( N ) and q k ( N 1 ) are the charges of the atoms on the systems with (N+ 1), N and N − 1 electrons respectively.

Recently, it has been reported [28] that a new descriptor has been introduced [29] [30] which allows the determination of individual sites within the molecule with particular behaviors. A mathematical analysis reveals that dual descriptor is a more accurate tool than nucleophilic and electrophilic Fukui functions [31]. This descriptor is defined through Equation (13):

Δ f ( r ) = ( f ( r ) N ) v ( r ) (13)

The condensed form [29] of the dual descriptor is given by Equation (14):

Δ f k ( r ) = f k + f k (14)

when Δ f k ( r ) > 0 , the process is driven by a nucleophilic attack and atom k acts as an electrophile; conversely, when, Δ f k ( r ) < 0 the process is driven by an electrophilic attack on atom k acts as a nucleophile. The dual descriptor Δ f k ( r ) is defined within the range [−1; 1], what really facilitates local reactivity interpretation [31].

3. Results and Discussion

3.1. Weight Loss Measurements

3.1.1. Effect of Inhibitor Concentration and Temperature on Corrosion Rate

The corrosion rate curves of copper without and with the addition of cefixime in 1M HNO3 at different temperatures are presented in Figure 2. This figure indicates that corrosion rate of copper in the studied medium, increases with increasing temperature. But this evolution is moderated when the concentration of the studied inhibitor increases, revealing the effectiveness of the molecule as a corrosion inhibitor for copper in 1M HNO3. These results could be interpreted as the formation of a film barrier which isolates copper from its aggressive environment [32] [33].

Figure 3 displays the evolution of inhibition efficiency versus temperature. The inhibition efficiency reaches a value of 91.07% for the concentration of 2 mM at 298 K (Figure 3). The inhibitive action of the studied drug may be due mainly to the presence of heteroatoms such as oxygen, nitrogen, sulfur and aromatic rings with π-bonds in cefixime structure and its ability with these heteroatoms to create a protective film on the metal corroding surface [34].

As shown in Figure 3, inhibition efficiency decreases when the temperature rises. This occurs due to the desorption effect of the inhibitor molecules [35].

Figure 2. Evolution of corrosion rate with temperature for different concentrations of cefixime.

Figure 3. Inhibition efficiency versus temperature for different concentrations of cefixime.

Moreover, decrease in inhibition efficiency with temperature rise can also be attributed to increased solubility of the protective film and/or any reaction product precipitated on the surface of the metal [36]. As a result, the metal surface becomes more accessible to corrosive attack.

3.1.2. Adsorption Isotherms and Thermodynamic Functions

The basic information on the interaction between the inhibitor and the metal can be provided by the adsorption isotherm. The adsorption isotherms tested in this work are the models of Langmuir, Temkin, Freundlich, El-Awady and Flory Huggins. By fitting the degree of surface coverage (θ) and the inhibitor concentration (Figure 4), the best adsorption isotherm obtained graphically is Langmuir adsorption isotherm with a strong correlation (R2 > 0.999) and the slopes of the straight lines are close to unity.

The obtained Langmuir adsorption parameters for different temperatures are presented in Table 1.

Assumptions of Langmuir relate the concentration of the adsorbate in the bulk of the electrolyte (Cinh) to the degree of surface coverage (θ) as Equation:

C i n h θ = 1 K a d s + C i n h (15)

where Cinh is cefixime concentration and Kads is the equilibrium constant of the adsorption process.

In order to evaluate the strength of the interactions between the inhibitor molecules and the metal surface, the values of adsorption equilibrium constant

Kads were computed using the intercepts of the straight lines on C i n h θ -axis. The

calculated adsorption equilibrium constant was related to the standard free energy of adsorption by the following equation [37]:

Δ G a d s 0 = R T ln ( 55.5 K a d s ) (16)

Figure 4. Langmuir adsorption isotherm for cefixime on copper surface in 1M HNO3.

Table 1. Regression parameters of Langmuir isotherm.

In the above equation [38], 55.5 is the concentration of water in mol·L−1, T is the absolute temperature while R is the universal gas constant. The values of Δ G a d s 0 and the other adsorption thermodynamic functions are gathered in Table 2.

The negative values of Δ G a d s 0 indicate that the adsorption process is spontaneous and the adsorbed layer on the aluminum surface is stable [9]. Generally, [39] [40], values of −20 kJ·mol−1 or less negative are consistent with the electrostatic interactions between the charged metal and the inhibitor i.e. physisorption. The values around −40 kJ·mol−1 or more negative are associated with chemisorption, as a result of sharing or transfer of unshared electron pair or π-electrons of organic molecules to the metal surface to form a coordinate type of bond. In the present work, Δ G a d s 0 change ranges from –38.00 to –36.14 kJ·mol−1 indicating both physisorption and chemisorption. The standard adsorption enthalpy change Δ H a d s 0 and the standard adsorption entropy change Δ S a d s 0 are correlated with standard Gibbs free energy through the relation:

Δ G a d s 0 = Δ H a d s 0 T Δ S a d s 0 (17)

Δ H a d s 0 and Δ S a d s 0 are obtained respectively as the intercept and the negative of the slope of the straight line obtained by plotting Δ G a d s 0 versus T (Figure 5). The change in adsorption enthalpy Δ H a d s 0 is negative, showing an exothermic process. The literature [41] stated that an exothermic process means either physisorption

Table 2. Adsorption thermodynamic functions.

Figure 5. Δ G a d s 0 versus T for the adsorption of cefixime on copper in 1M HNO3.

or chemisorption. Therefore, this result confirms that the process of adsorption is both physisorption and chemisorption. The change in standard adsorption entropy Δ S a d s 0 is positive, meaning that disorder increases during the adsorption process. This situation can be attributed to desorption of water molecules replaced by the inhibitor.

3.1.3. Effect of the Temperature and Activation Parameters

The activation energy ( E a ) can be obtained by using Arrhenius equation (Equation (18)):

log C R = log A E a 2.303 R T (18)

where CR is the corrosion rate and A is the Arrhenius pre-exponential constant. A plot of log C R versus 1 T yields a straight line (Figure 6) with E a 2.303 R as slope and log A as intercept.

The other activation parameters for the corrosion process were calculated from the Arrhenius equation:

log ( C R T ) = [ log ( R h ) + Δ S a * 2.303 R ] Δ H a * 2.303 R T (19)

Figure 6. Arrhenius plots for copper corrosion in 1M HNO3 without and with different concentrations of cefixime.

where Δ S a * is the change in activation entropy, Δ H a * is the change in activation enthalpy and is the Avogadro number and h is the Planck’s constant.

Figure 7 gives the plot of log ( C R T ) versus 1 T .

The slope A H a * 2.303 R T and the intercept [ log ( R h ) + Δ S a * 2.303 R ] of each

straight-line lead to the values of activation enthalpy change and activation entropy change (Table 3).

From Table 3, it seems that E a and Δ H a * varied in the same manner increasing with the concentration, probably due to the thermodynamic relation between them ( Δ H a * = E a R T ). It can be seen that the values of E a are higher in the inhibited solutions than those in uninhibited solutions. On the other hand, the higher values of E a in the presence of inhibitor compared to that in its absence and the decrease of the inhibition efficiency (IE) with the increase in temperature can be interpreted as an indication of predominant physisorption process [4] [42] [43]. Moreover, the positive signs of Δ H a * pointed out the endothermic effect of the copper dissolution process. The value of Δ S a * is higher for the inhibited solution than that for the uninhibited solution. This phenomenon suggested that a randomness decrease occurred from reactants to the activated complex. This might be the result of the adsorption of organic inhibitor molecules from the acidic solution which could be seen as a quasi-substitution process between the inhibitor in the aqueous phase and water molecules at the metal surface [44].

3.2. Quantum Chemistry Study

3.2.1. Global Reactivity

The values of selected quantum chemical parameters computed for the studied cefixime molecule by using DFT methods are listed in Table 4.

According to the frontier molecular orbital (FMO) theory, of chemical reactivity,

Figure 7. Transition state plots for copper corrosion in 1M HNO3 with or without different concentrations of cefixime.

Table 3. Calculated activation energy, activation enthalpy change and activation entropy change.

Table 4. Molecular and global reactivity descriptors of cefixime.

transition of electron is due to interaction between highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) of reacting species [45] [46]. These HOMO and LUMO energies [47] are very important in defining the reactivity of a given molecule. EHOMO [48] is often associated with the electron donating ability of a molecule. High EHOMO values indicate that the molecule has a tendency to donate electrons to appropriate acceptor molecules with low energy empty molecular orbital. This high value indicates the tendency to donate electrons to empty molecular orbital of copper ions (Cu2+: [Ar]3d9). Increasing values of the EHOMO facilitate adsorption (and therefore inhibition) by influencing the transport process through the adsorbed layer [49]. On the other hand, the LUMO energy [48], indicates the ability of the molecule to accept electrons. The lower the value of EHOMO, the more probable it is that the molecule accepts electrons. In this work, the higher value of EHOMO (−5.8680 eV) and the lower value of ELUMO (−2.0510 eV) could involve charge sharing or charge transfer from the inhibiting molecule to the metal surface [50]. That could explain good inhibition efficiency of cefixime. In the same way, low values of the energy gap (ΔΕ = ΕLUMO ΕHOMO) lead to high inhibition efficiencies because [36] the energy to remove an electron from the last occupied orbital will be low. Literature [51] revealed that excellent corrosion inhibitors are organic compounds which not only offer electrons to unoccupied orbital of metal but also accept free electrons from the metal. A molecule with a low energy gap [36] is more polarizable and is generally associated with high chemical reactivity and is considered a soft molecule. The value of ΔE(3.8170 eV) is low compared with that of many molecules in the literature, suggesting good inhibition efficiency.

Another important electronic parameter calculated is the dipole moment (μ) that results from non-uniform distribution of charges on atoms in the molecule. Several authors point out that low values of dipole moment [52] promote accumulation of the inhibitor molecules in the surface layer and therefore higher inhibition efficiency. However, many papers indicate that inhibition efficiency increases with rising values of dipole moment. On the other hand, the survey of the literature [53] [54], reveals that several irregularities were appeared in the case of correlation of dipole moment with inhibitor efficiency. So, in general [55], there is no significant relationship between dipole moment values and inhibition efficiencies.

Absolute hardness and softness are important parameters to measure the molecular stability and reactivity of a molecule. The chemical hardness fundamentally represents the resistance towards the deformation or polarization of the electron cloud of atoms, ions or molecules under small perturbation of chemical reaction. A hard molecule has a large energy gap and a soft molecule has a small energy gap [56]. In the present paper, the studied molecule has a low hardness value (1.9085 eV) and a high value of softness (0.5240 (eV)1) when compared [57] [58] with molecules in the literature.

The ionization potential (I) and the electronic affinity (A) are respectively (5.8680 eV) and (2.0510 eV). This low value of (I) and the high value of electron affinity indicate the capacity of the molecule both to donate and accept electron. The electronegativity (χ) indicates the capacity of a system to attract electrons. In our work the low value of the electronegativity of the studied molecule ( χ = 3.9595 eV ) when compared to that of copper ( χ Cu = 4.98 eV ) shows that copper has the better attraction capacity. Then the low value of hardness (1.9085 eV) confirms the relatively higher value of the fraction of electrons transferred (ΔN = 0.0840) indicating a possible motion of electrons from the inhibitor to the metal.

The electrophilicity index measures the propensity of chemical species to accept electrons; a high value of electrophilicity index describes a good electrophile while a small value of electrophilicity index describes a good nucleophile. In this work the obtained value (ω = 4.1073 eV) shows the good capacity of cefixime to accept electrons. The HOMO and LUMO diagrams of the studied drug are presented in Figure 8.

As seen in Figure 8, the density HOMO is distributed around the five-membered ring containing heteroatoms whereas, the LUMO density is distributed almost homogeneously throughout the molecule. So, these regions are probably the most active areas where transfer of electrons could be achieved (from the inhibitor molecule to copper or vice-versa).

3.2.2. Local Reactivity

Local reactivity descriptors including atomic charges, condensed Fukui functionsand dual descriptor are collected in Table 5.

It can be clearly observed from shaded rows in Table 5 that (15C) with the maximum value of f k + and positive value of Δ f k ( r ) is the most probable nucleophilic attack site, while (35C) with the maximum value of f k and negative value of Δ f k ( r ) is the most probable electrophilic attack site.

Figure 8. HOMO and LUMO diagrams of cefixime by B3LYP/6-31G (d, p).

Table 5. Calculated Mulliken atomic charges, Fukui functions and dual descriptor by DFT B3YLP 6-31/G (d, p).

4. Conclusion

Cefixime drug shows good inhibition properties for copper corrosion in 1M HNO3. The inhibition efficiency increases with increasing concentration in the studied inhibitor but decreases with a rise in temperature. Adsorption of cefixime follows the Langmuir adsorption isotherm. The adsorption and activation thermodynamic functions reveal a spontaneous adsorption process of the studied molecule onto copper and both physisorption and chemisorption mechanisms with the predominance of physisorption. Quantum chemical calculations data from B3LYP/6-31G (d, p) fit well the experimental results.

Conflicts of Interest

The authors declare no conflicts of interest regarding the publication of this paper.

References

[1] Baharia, H.S., Ye, F., Carrillo, E.A.T., Leliopoulos, C., Savaloni, H. and Dutta, J. (2020) Chitosan Nanocomposite Coatings with Enhanced Corrosion Inhibition Effects for Copper. International Journal of Biological Macromolecules, 162, 1566-1577.
https://doi.org/10.1016/j.ijbiomac.2020.08.035
[2] Echihi, S., Benzbiria, N., Belghiti, M.E., Fal, M.E., Boudalia, M., Essassi, M.E., et al. (2021) Corrosion Inhibition of Copper by Pyrazole Pyrimidine Derivative in Synthetic Seawater: Experimental and Theoretical Studies. Materials Today: Proceedings, 37, 3958-3966.
https://doi.org/10.1016/j.matpr.2020.09.264
[3] Yao, C., Chen, S., Wang, L., Deng, H. and Tong, S. (2019) Low Cost and Rapid Fabrication of Copper Sulfides Nanoparticles for Selective and Efficient Capture of Noble Metal Ions. Chemical Engineering Journal, 373, 1168-1178.
https://doi.org/10.1016/j.cej.2019.05.070
[4] Diki, N.Y.S., Bohoussou, K.V., Kone, M.G.-R., Ouedraogo, A. and Trokourey, A. (2018) Cefadroxil Drug as Corrosion Inhibitor for Aluminum in 1 M HCl Medium: Experimental and Theoretical Studies. IOSR Journal of Applied Chemistry, 11, 24-36.
[5] Diki, N.Y.S., Gbassi, K.G., Ouedraogo, A. Berté, M. and Trokourey, A. (2018) Aluminum Corrosion Inhibition by Cefixime Drug: Experimental and DFT Studies. Journal of Electrochemical Science and Engineering, 8, 303-320.
https://doi.org/10.5599/jese.585
[6] Hazazi, O.A. and Abdallah, M. (2013) Prazole Compounds as Inhibitors for Corrosion of Aluminum in Hydrochloric Acid. International Journal of Electrochemical Science, 8, 8138-8152.
[7] El-Haddad, M.N. and Fouda, A.S. (2015) Electroanalytical, Quantum and Surface Characterization Studies on Imidazole Derivatives as Corrosion Inhibitors for Aluminum in Acidic Media. Journal of Molecular Liquids, 209, 480-486.
https://doi.org/10.1016/j.molliq.2015.06.005
[8] Naqvi, I., Saleemi, A.R. and Naveed, S. (2011) Cefixime: A Drug as Efficient Corrosion Inhibitor for Mild Steel in Acidic Media. Electrochemical and Thermodynamic Studies. International Journal of Electrochemical Science, 6, 146-161.
[9] Cang, H., Fei, Z., Shao, J., Shi, W. and Xu, Q. (2013) Corrosion Inhibition of Mild Steel by Aloes Extract in HCl Solution Medium. International Journal of Electrochemical Science, 8, 720-734.
[10] Cohen, S.L., Brusic, V.A., Kaufman, F.B., Frankel, G.S., Motakef, S. and Rush, B. (1990) X-Ray Photoelectron Spectroscopy and Ellipsometry Studies of the Electrochemically Controlled Adsorption of Benzotriazole on Copper Surfaces. Journal of Vacuum Science & Technology A, 8, 2417-2424.
https://doi.org/10.1116/1.576708
[11] Ebenso, E.E., Eddy, N.O. and Odiongenyi, A.O. (2009) Corrosion Inhibition and Adsorption Properties of Methocarbamol on Mild Steel in Acidic Medium. Portugaliae Electrochimca Acta, 27, 13-22.
https://doi.org/10.4152/pea.200901013
[12] Eddy, N.O., Ebenso, E.E. and Ibok, U.J. (2010) Adsorption, Synergistic Inhibitive Effect and Quantum Chemical Studies of Ampicillin (AMP) and Halides for the Corrosion of Mild Steel in H2SO4. Journal of Applied Electrochemistry, 40, 445-456.
https://doi.org/10.1007/s10800-009-0015-z
[13] Eddy, N.O., Odoemelam, S.A. and Mbaba, A.J. (2008) Inhibition of the Corrosion of Mild Steel in HCl by Sparfloxacin. African Journal of Pure and Applied Chemistry, 2, 132-138.
[14] Eddy, N.O. and Ebenso, E.E. (2010) Adsorption and Quantum Chemical Studies on Cloxacillin and Halides for the Corrosion of Mild Steel in Acidic Medium. International Journal of Electrochemical Science, 5, 731-750.
[15] Frisch, M.J., Trucks, G.W., Schlegel, G.E.S.H.B., Robb, M.A., Cheeseman, J.R., Scalmani, G., et al. (2009) Gaussian 09. Revision A.02, Gaussian Inc., Wallingford.
[16] Lee, C., Yang, W. and Parr, R.G. (1988) Development of the Colle Salvetti Correlation-Energy Formula into Afunctional of the Electron Density. Physical Review B, 37, 785-789.
https://doi.org/10.1103/PhysRevB.37.785
[17] Rodriguez-Valdez, L.M., Martinez-Villafane, A. and Glossman-Mitnik, D. (2005) CHIH-DFT Theoretical Study of Isomeric Thiatriazoles and Their Potential Activity as Corrosion Inhibitors. Journal of Molecular Structure: THEOCHEM, 716, 61-65.
https://doi.org/10.1016/j.theochem.2004.10.082
[18] Patel, N.S., Beranek, P., Nebyla, M., Pribyl, M. and Snita, D. (2014) Inhibitive Effects by Some Benzothiazole Derivatives on Mild Steel Corrosion in 1 N HCl. International Journal of Electrochemical Science, 9, 3951-3960.
[19] Koopmans, T. (1934) über die Zuordnung von Wellenfunktionen und Eigenwerten zu den Einzelnen Elektronen Eines Atoms. Physica, 1, 104-113.
https://doi.org/10.1016/S0031-8914(34)90011-2
[20] Parr, R.G., Donnelly, R.A., Levy, M. and Palke, W.E. (1978) Electronegativity: The Density Functional Viewpoint. Journal of Chemical Physics, 68, 3801-3807.
https://doi.org/10.1063/1.436185
[21] Kokalj, A. and Kovačević, N. (2011) On the Consistent Use of Electrophilicity Index and HSAB-Based Electron Transfer and Its Associated Change of Energy Parameters. Chemical Physics Letters, 507, 181-184.
https://doi.org/10.1016/j.cplett.2011.03.045
[22] Michaelson, H.B. (1977) The Work Function of the Elements and Its Periodicity. Journal of Applied Physics, 48, 4729-4734.
https://doi.org/10.1063/1.323539
[23] Sastri, V.S. and Perumareddi, J.R. (1997) Molecular Orbital Theoretical Studies of Some Organic Corrosion Inhibitors. Corrosion, 53, 617-622.
https://doi.org/10.5006/1.3290294
[24] Parr, R.G., Szentpaly, L.V. and Liu, S. (1999) Electrophilicity Index. Journal of the American Chemical Society, 121, 1922-1924.
https://doi.org/10.1021/ja983494x
[25] Fuentealba, P., Perez, P. and Contreras, R. (2000) On the Condensed Fukui Function. Journal of Chemical Physics, 113, 2544-2551.
https://doi.org/10.1063/1.1305879
[26] Kouakou, V., Niamien, P.M., Yapo, A.J. and Trokourey, A. (2016) Copper Corrosion Inhibition in 1 M Nitric Acid: Adsorption and Inhibitive Action of Theophylline. Chemical Science Review and Letters, 5, 131-146.
[27] Eddy, N.O., Stoyanov, S.R. and Ebenso, E.E. (2010) Fluoroquinolones as Corrosion Inhibitors for Mild Steel in Acidic Medium. International Journal of Electrochemical Science, 5, 1127-1150.
[28] Yeo, M., Niamien, P.M., Bilé, E.B.A. and Trokourey, A. (2018) Thiamine Hydrochloride as a Potential Inhibitor for Aluminium Corrosion in 1.0M HCl: Mass Loss and DFT Studies. Journal of Computational Methods in Molecular Design, 8, 13-25.
[29] Morell, C., Grand, A. and Torro-Labbé, A. (2006) Theoretical Support for Using the Δf(r) Descriptor. Chemical Physics Letters, 425, 342-346.
https://doi.org/10.1016/j.cplett.2006.05.003
[30] Morell, C., Grand, A. and Torro-Labbé, A. (2005) New Dual Descriptor for Chemical Reactivity. Journal of Physical Chemistry A, 109, 205-212.
https://doi.org/10.1021/jp046577a
[31] Martinez-Araya, J.I. (2015) Why Is the Dual Descriptor a More Accurate Local Reactivity Descriptor than Fukui Functions? Journal of Mathematical Chemistry, 53, 451-465.
https://doi.org/10.1007/s10910-014-0437-7
[32] Diki, N.Y.S., Diomandé, G.G.D., Akpa, S.J., Ouedraogo, O., Pohan, L.A.G., Niamien, P.M. and Trokourey, A. (2018) Aluminum Corrosion Inhibition by 7-(Ethylthio-benzimidazolyl) Theophylline in 1M Hydrochloric Acid: Experimental and DFT Studies. International Journal of Applied Pharmaceutical Sciences and Research, 3, 41-53.
https://doi.org/10.21477/ijapsr.3.4.1
[33] Diki, N.Y.S., Coulibaly, N.H., Yao, J.N.S.T. and Trokourey, A. (2019) Thermodynamic and DFT Studies on the Behavior of Cefadroxil Drug as Effective Corrosion Inhibitor of Copper in One Molar Nitric acid Medium. Journal of Materials and Environmental Science, 10, 926-938.
[34] Ouedraogo, A., Diki, N.Y.S., Bohoussou, K.V., Soro, D. and Trokourey, A. (2018) Copper Corrosion Inhibition by Cefuroxime Drug in 1 M Nitric Acid. Chemical Science Review and Letters, 7, 427-437.
[35] Anbarasi, C.M. and Divya, G. (2017) A Green Approach to Corrosion Inhibition of Aluminium in Acid Medium Using Azwain Seed Extract. Materials Todays: Proceedings, 4, 5190-5200.
https://doi.org/10.1016/j.matpr.2017.05.026
[36] Obot, I.B., Obi-Egbedi, N.O. and Umoren, S.A. (2009) Adsorption Characteristics and Corrosion Inhibitive Properties of Clotrimazole for Aluminium Corrosion in Hydrochloric Acid. International Journal of Electrochemical Science, 4, 863-877.
[37] Keles, H., Keles, M., Dehri, I. and Serinday, O. (2008) Adsorption and Inhibitive Properties of Aminobiphenyl and Its Schiff Base on Mild Steel Corrosion in 0.5 M HCl Medium. Colloids and Surfaces, A: Physicochemical and Engineering Aspects, 320, 138-145.
https://doi.org/10.1016/j.colsurfa.2008.01.040
[38] Villamil, R.F.V., Corio, P., Rubin, J.C. and Agostinho, S.M.L. (1999) Effect of Sodium Dodecylsulfate on Copper Corrosion in Sulfuric Acid Media in the Absence and Presence of Benzotriazole. Journal of Electroanalytical Chemistry, 472, 112-119.
https://doi.org/10.1016/S0022-0728(99)00267-3
[39] Issadi, S., Douadi, T, Zouaoui, A., Chafaa, S., Khan, M.A. and Bouet, G.(2011) Novel Thiophene Symmetrical Schiff Base Compounds as Corrosion Inhibitor for Mild Steel in Acidic Media. Corrosion Science, 53, 1484-1488.
https://doi.org/10.1016/j.corsci.2011.01.022
[40] Ehteshamzadeh, M., Shahrabi, T. and Hosseini, M. (2006) 3,4-Dimethoxy-Benzaldehydethio-Semicarbazone as Corrosion Inhibitor. Anti-Corrosion Methods and Materials, 53, 296-302.
https://doi.org/10.1108/00035590610692572
[41] Durnie, W., Marco, R.D., Jefferson, A. and Kinsella, B. (1999) Development of a Structure-Activity Relationship for Oil Field Corrosion Inhibitors. Journal of the Electrochemical Society, 146, 1751-1756.
https://doi.org/10.1149/1.1391837
[42] Umoren, S.A. (2008) Inhibition of Aluminium and Mild Steel Corrosion in Acidic Medium Using Gum Arabic. Cellulose, 15, Article No. 751.
https://doi.org/10.1007/s10570-008-9226-4
[43] Chaubey, N., Savita, Singh, V.K. and Quraishi, M.A. (2017) Corrosion Inhibition Performance of Different Bark Extracts on Aluminium in Alkaline Solution. Journal of the Association of Arab Universities for Basic and Applied Sciences, 22, 38-44.
https://doi.org/10.1016/j.jaubas.2015.12.003
[44] Afia, L., Salghi, R., Zarrouk, A., Zarrok, H., Benali, O., Hammouti, B., et al. (2012) Inhibitive Action of Argan Press Cake Extract on the Corrosion of Steel in Acidic Media. Portugaliae Electrochimica Acta, 30, 267-279.
https://doi.org/10.4152/pea.201204267
[45] Lukovits, I., Palfi, K., Bako, I. and Kalman, E. (1997) LKP Model of the Inhibition Mechanism of Thiourea Compounds. Corrosion, 53, 915-919.
https://doi.org/10.5006/1.3290275
[46] Musa, A.Y., Kadhum, A.H., Mohamad, A.B., Rohoma, A.B. and Mesmari, H. (2010) Electrochemical and Quantum Chemical Calculations on 4,4-dimethyloxazolidine-2-thione as Inhibitor for Mild Steel Corrosion in Hydrochloric Acid. Journal of Molecular Structure, 969, 233-237.
https://doi.org/10.1016/j.molstruc.2010.02.051
[47] Ebenso, E.E., Arslan, T., Kandemirli, F., Caner, I.N and Love, I. (2010) Quantum Chemical Studies of Some Rhodamine Azosulpha Drugs as Corrosion Inhibitors for Mild Steel in Acidic Medium. International Journal of Quantum Chemistry, 110, 1003-1018.
https://doi.org/10.1002/qua.22249
[48] Popova, A., Christov, M. and Deligeorgiev, T. (2003) Influence of the Molecular Structure on the Inhibitor Properties of Benzimidazole Derivatives on Mild Steel Corrosion in 1 M Hydrochloric Acid. Corrosion, 59, 756-764.
https://doi.org/10.5006/1.3277604
[49] Özcan, M. and Dehri, I. (2004) Electrochemical and Quantum Chemical Studies of Some Sulphur-Containing Organic Compounds as Inhibitors for the Acid Corrosion of Mild Steel. Progress in Organic Coatings, 51, 181-187.
https://doi.org/10.1016/j.porgcoat.2004.07.017
[50] Lebrini, M., Lagrenee, M., Vezin, H., Gengembre, L. and Bentiss, F. (2005) Electrochemical and Quantum Chemical Studies of New Thiadiazole Derivatives Adsorption on Mild Steel in Normal Hydrochloric Acid Medium. Corrosion Science, 47, 485-505.
https://doi.org/10.1016/j.corsci.2004.06.001
[51] Fleming, I. (1976) Frontier Orbitals and Organic Chemical Reactions. John Wiley and Sons, New York, 249.
[52] Khalil, N. (2003) Quantum Chemical Approach of Corrosion Inhibition. Electrochimica Acta, 48, 2635-2640.
https://doi.org/10.1016/S0013-4686(03)00307-4
[53] Khaled, K.F., Babic-Samardziza, K. and Hackerman, N. (2005) Theoretical Study of the Structural Effects of Polymethylene Amines on Corrosion Inhibition of Iron in Acid Solutions. Electrochimica Acta, 50, 2515-2520.
https://doi.org/10.1016/j.electacta.2004.10.079
[54] Bereket, G., Hur, E. and Ogretir, C. (2002) Quantum Chemical Studies on Some Imidazole Derivatives as Corrosion Inhibitors for Iron in Acidic Medium. Journal Molecular Structure: THEOCHEM, 578, 79-88.
https://doi.org/10.1016/S0166-1280(01)00684-4
[55] Gece, G. (2008) The Use of Quantum Chemical Methods in Corrosion Inhibitor Studies Corrosion Science, 50, 2981-2992.
https://doi.org/10.1016/j.corsci.2008.08.043
[56] Gece, G. and Bilgic, S. (2009) Quantum Chemical Study of Some Cyclic Nitrogen Compounds as Corrosion Inhibitors of Steel in NaCl Media. Corrosion Science, 51, 1876-1878.
https://doi.org/10.1016/j.corsci.2009.04.003
[57] Udhayakala, P., Rajendiran, T.V. and Gunasekaran, S. (2012) Theoretical Approach to the Corrosion Inhibition Efficiency of Some Pyrimidine Derivatives Using DFT Method. Journal of Computional Methods in Molecular Design, 2, 1-15.
[58] Udhayakala, P. and Rajendiran, T.V. (2015) A Theoretical Evaluation on Benzothiazole Derivatives as Corrosion Inhibitors on Mild Steel. Der Pharma Chemica, 7, 92-99.

Copyright © 2024 by authors and Scientific Research Publishing Inc.

Creative Commons License

This work and the related PDF file are licensed under a Creative Commons Attribution 4.0 International License.