Home Investigation of corrosion inhibition and adsorption properties of quinoxaline derivatives on metal surfaces through DFT and Monte Carlo simulations
Article Publicly Available

Investigation of corrosion inhibition and adsorption properties of quinoxaline derivatives on metal surfaces through DFT and Monte Carlo simulations

  • Yousif Hussein Azeez , Dyari Mustafa Mamand , Rebaz A. Omer ORCID logo , Awat Hamad Awla and Karzan A. Omar EMAIL logo
Published/Copyright: July 19, 2024

Abstract

This work presents a multiscale theoretical investigation into the potential of quinoxaline derivatives (Q1–Q6) as corrosion inhibitors for various metals (Fe(110), Cu(111), and Al(110)). Employing a combined approach combining density functional theory (DFT) and Monte Carlo simulations, we explore the relationship between molecular structure, electronic properties, and adsorption behavior. Density functional theory (DFT) and molecular dynamics simulations (MDS) were used to investigate the electronic characteristics of diverse compounds. The study included key parameters including highest occupied molecular orbital energy (EHOMO), lowest unoccupied molecular orbital energy (ELUMO), energy gap (Eg) between ELUMO and EHOMO, dipole moment, global hardness, softness (σ), ionization energy (I), electron affinity (A), electronegativity (χ), back-donation energy Eb−d, global electrophilicity (ω), electron transfer, global nucleophilicity (ε), and total energy (sum of electronic and zero-point energies). These properties, alongside adsorption energies (following the trend Q6 > Q2 > Q3 > Q4 > Q5 > Q1), are used to identify promising inhibitor candidates and establish structure–property relationships governing their effectiveness. The results suggest that inhibitor efficiency increases with a decreasing energy gap between frontier orbitals. Notably, the protonated state of Q6 exhibits high reactivity, low stability, and strong adsorption, making it a potential candidate for further exploration. This comprehensive theoretical approach offers crucial insights for the conceptual development of new and powerful corrosion inhibitors.

1 Introduction

Corrosion inhibitors are frequently used in the chemical industry as a dependable defense against acid solutions and a practical measure to stop corrosion on metal surfaces. Corrosion is an electrochemical mechanism significant in economics, industry, human life, and metal safety. As a result, metal corrosion prevention is an essential industrial and scientific problem. Various organic chemicals have long been employed as corrosion inhibitors for metals and alloys in acidic environments (Ahmed et al. 2023; Rasul et al. 2023). Compounds containing heteroatoms such as nitrogen, oxygen, and sulfur atoms, as well as aromatic rings, are the most efficient corrosion inhibitors. In different acid cycles, metal corrosion inhibition depends on the nature of the organic molecule’s interaction with the metals. Centers enabling coordination bonds between the organic molecules and the metal surface should exist (Maldonado and Amo-Ochoa 2021; Omar et al. 2023; Omer et al. 2023b). The coordination interaction has more substantial inhibitory effects. The effectiveness of an organic compound as a motivational inhibitor depends on several factors, including its chemical composition, molecule electronic structure, surface charge density, iron distribution, metal surface, and molecular electronic structure (Omer et al. 2022c; Umoren and Solomon 2017). Some chemical substances have inhibitory effects on the corrosive environment that have important economic implications. In the industrial sector, mild steel corrosion is one of the main problems. A number of stages can lead to the development of this condition, especially those that include pickling, acid washing, and etching steel. In harsh settings, a few chemical substances can be employed as corrosion inhibitors to prevent metal from corroding (Obot et al. 2019b; Omer et al. 2022a, 2022d). These inhibitors are utilized in petrochemical engineering, industrial water, and chemical cleaning solutions manufacturing because of their efficaciousness in lowering the cost of metal and alloy corrosion. One way to think of inhibitors is as changes to the kinetics of electrochemical reactions, such as the corrosion process. The usual mechanism for the organic inhibitor molecules to adsorb on metal is to stick to the metal surface and change its chemical or physical properties (Izadi et al. 2018; Liu and Shi 2009; Parlak et al. 2022). Calculations based on quantum chemistry have been demonstrated to be a promising tool for researching corrosion inhibition processes, owing to the significant correlations observed between the efficiencies of most compounds and various semi-empirical corrosion inhibitory features (El Sayed et al. 2006; Koparir et al. 2022; Omer et al. 2022b). Because quinoxaline derivatives are less toxic and biodegradable than typical inhibitors that include heavy metals that are harmful to the environment, they are regarded as environmentally friendly corrosion inhibitors (Chauhan et al. 2020). They are appealing for use in industries where environmental issues are important because of their ability to serve as green inhibitors (Olasunkanmi et al. 2016). The heterocyclic ring structure of quinoxaline derivatives is usually made up of two nitrogen atoms and one carbon atom. By using nitrogen atoms to adsorb onto metal surfaces, these compounds may create protective layers that prevent corrosive species from penetrating the metal surface (Lgaz et al. 2016). The process of inhibition might be due to the creation of barrier films on the metal surface, adsorption via π-electron interaction, or coordination complex formation with metal ions. In conclusion, because of their advantageous chemical structure, adsorption behavior, and environmental reliability, quinoxaline derivatives show promise in the suppression of corrosion. This work presents a novel multiscale theoretical approach to explore the potential of quinoxaline derivatives (Q1–Q6) see Figure 1 as corrosion inhibitors for various metals (Fe(110), Cu(111), and Al(110)). We combine DFT calculations to understand electronic properties with Monte Carlo simulations to analyze adsorption behavior. This synergy allows us to identify promising inhibitor candidates based on adsorption energies (Q6 > Q2 > Q3 > Q4 > Q5 > Q1) and establish structure–property relationships. We further investigate the impact of frontier orbital energy gap on inhibitor efficiency and highlight the promising characteristics of the protonated state of Q6. This combined approach offers valuable insights for the conceptual design of novel and potent corrosion inhibitors, setting the stage for future development efforts.

Figure 1: 
					Chemical structures for title compounds.
Figure 1:

Chemical structures for title compounds.

2 Computational method

Molecular imaging software GaussView5.0 was used to map the three-dimensional shapes of the molecular compounds in protonated and nonprotonated species present in the aqueous and gas phases. GaussView 5.0 package program was used to generate the primary forms of the molecules, and the Gaussian09W program was used to perform the computations (Mamand and Qadr 2021; Omer et al. 2023a). Because the DFT method accounts for electron density, it can produce the data required for a more accurate evaluation of the electronic characteristics of complicated structures. Furthermore, the Gaussian09W software’s hybrid feature B3LYP is suitable for the number of terminals and 6-31G ++ (d, p) as the basis set. The correlation between EHOMO (higher occupied molecular orbital energy), ELUMO (lower unoccupied molecular orbital energy), (Eg) bandgap energy, (I) ionization energy, (χ) electronegativity, (∆N) electron transfer, (∆E_(bd)) back-donation energy, and corrosion inhibitory activity was calculated. Electronic structural identifiers were found from geometry-optimized structures. Using the Monte Carlo approach, the adsorption for every component found in this investigation was calculated. Since it is the most stable surface reported in the literature, the Fe, Cu, and Al (110) crystal surface was chosen for this simulation (El Ghayati et al. 2021; Guo et al. 2017; Rasul et al. 2023). The Fe, Cu, and Al (110) surfaces were modeled using the four-slab model (Erdoğan et al. 2017). One hundred ten units, or 80 iron atoms per layer, were present in this model. A Monte Carlo search of the structural space of the substrate-adsorbed system was conducted with a decreasing temperature step to identify a low-energy adsorption site.

3 Quantum computational detail

Quantum chemical computations are crucial to exploring and developing new corrosion inhibitors. Time and cost savings are achieved by estimating the corrosion inhibitor’s capacity to prevent corrosion based on its molecular properties. Using Material Studio software, the critical quantitative properties of coconut leaf extract are identified (Anwar Omar et al. 2023; Chen et al. 2020). Collecting structural and dynamic data at the atomic and molecular levels is difficult, even though the numbers produced by experimental procedures can indicate typical properties. It is common practice to investigate corrosion inhibition methods of inhibitor molecules using simulation techniques instead of environmentally hazardous chemicals and lab equipment (Mamand and Qadr 2023a; Mamad et al. 2023a, 2023b). The main theoretical calculation methods used to analyze inhibitors are based on DFT quantum mechanics and molecular mechanics calculations. The interaction between the HOMO or LUMO of the inhibitor and metal surface can be used to calculate the chemical reactivity characteristics of most chemical reactions, so only HOMO and LUMO need to be considered when analyzing the chelation process of chemisorption, according to the frontier molecular orbital (FMO) theory (Mattsson et al. 2004). The highest occupied molecular orbital energy (EHOMO), lowest unoccupied molecular orbital energy (ELUMO), energy band gap (Eg = ELUMOEHOMO), global hardness (η), and softness (σ), ionization potential (I), electron affinity (A), electronegativity (χ), and absolute electrophilicity index (ω) are the most frequently used molecular classifications for DFT-based quantum mechanics. The symbols EHOMO and ELUMO stand for inhibitor compounds’ electron-donating and electron-accepting capabilities. A higher EHOMO value indicates that the inhibitor molecule may effectively donate electrons to the low-energy orbitals, whereas a lower ELUMO value indicates that the metal tends to receive electrons (Fergachi et al. 2019). A critical measure of how well inhibitor chemicals can halt corrosion processes is the energy differential (Eg) between LUMO and HOMO. The energy needed to remove an electron from the last occupied orbit decreases with increasing inhibition efficiency.

Moreover, the molecule stability is measured by (Eg); a lower Eg value indicates a more complex formed on the metal surface (Obot et al. 2016). According to Koopman’s theory, the inhibitor molecule’s HOMO and LUMO orbital energies are related to electron affinity and ionization potential. Equations (1) and (2) show this relationship between EHOMO against I and ELUMO versus A. The letters I and A represent a molecule’s ability to supply and absorb electrons; the higher the value, the more likely the related process is to occur. More corrosion inhibition is indicated by higher σ, χ, and ω or lower values of η, respectively (Feng et al. 2018; Hussein and Azeez 2023; Mousavi et al. 2011; Wang et al. 2016). These conclusions are primarily supported by Pearson’s hard/soft acid/base principle and the propensity of inhibitor compounds to receive and transfer electrons.

(1)I=EHOMO
(2)A=ELUMO

If the value of μ is lower, it indicates that inhibitor molecules are adsorbed close to the electrode surface. In contrast, a greater value indicates a strong connection between an inhibitor molecule and a metal surface. However, its relationship to the effectiveness of corrosion inhibition is still unknown. When ΔN > 0, as it occurs when N is determined using Pearson’s calculation method, the inhibitor molecule gathers electrons rather than transferring them to the metal (Pearson 1988). The metal surface has a greater capacity to donate electrons as ΔN < 3.6, as revealed by Lukovit (Fu et al. 2010). As N increases, this causes a rise in corrosion inhibition. Koopman’s theorem (Fu et al. 2010) has connected electronic structural parameters like electronegativity, chemical potential, and chemical hardness to Molecular Orbital Theory (MOT). According to Koopman’s theorem, EHOMO and ELUMO are correlated with electron affinity and ionization energy.

(3)χ=I+A2
(4)η=IA2
(5)Nmax=χFeχinh2η(ηFe+ηinh)

Quinoxalines can be modified to enhance their exceptional ability to bind with metallic ions, which can help create corrosion inhibitors that work well (Dhanaraj and Johnson 2015). Moreover, by adding or removing suitable electron-donating or -withdrawing functional groups, phenyl rings, alkyl chains, and so on to the quinoxaline backbone, the reduction in corrosion efficacy may be enhanced (Haque et al. 2020). Two types of corrosion reactions occur when a strong electrolytic solution attacks a metallic surface: (i) oxygen reduction, which occurs at a pH close to neutral, or anodic metal dissolution, and (ii) cathodic hydrogen evolution, which occurs at an acidic pH. When a corrosion-preventing material is introduced to an acidic solution, it may undergo the following kind of protonation (Onyeachu et al. 2019):

(6)IhnneutralIhnprotonated

The protonated inhibitor chemical can adsorb at the anodic sites on the metal interface through electrostatic interactions involving the anions of the electrolyte via a bridge type of connection. Because of electrostatic interactions, the protonated inhibitor may immediately adsorb to the cathodic areas. Through π-interactions, polar functional groups, a single pair of heteroatoms, and electron back-donation from the partially full d-orbitals of the metallic surface, the inhibiting molecule’s neutral configuration can adsorb at the metal’s surface (Obot et al. 2019a). Theoretical and experimental methods are valuable for determining the inhibitor’s adsorption effect. Increased charge transfer resistance, decreased weight loss, corrosion speed, reduced corrosion current concentrations, and decreased currents as a function of frequency indicate the inhibitor’s adsorption at the metal/electrolyte interface. Anodic, cathodic, or mixed inhibitor action can also be reflected in changes in corrosion potential (Chauhan et al. 2020).

4 Results and discussion

4.1 Quantum chemical calculations

Calculations involving quantum chemicals have long been used to investigate reaction processes. They are a very effective tool for researching corrosion inhibition processes (Mamand et al. 2023). Theoretical estimation of corrosion inhibitor performance is gaining popularity, coinciding with advances in computer hardware and the establishment of efficient methods that aided the routine construction of molecular quantum mechanical simulations. All quantum chemical characteristics were derived following geometric optimization when compared to all nuclear coordinates at the DFT level employing the Kohn–Sham technique (Andzelm et al. 1995). A chemical species’ border orbital (highest occupied molecular orbital HOMO and lowest unoccupied molecular orbital LUMO) is crucial in determining its reactivity, which was observed for the first time by Fukui. A strong relationship involving corrosion rates and EHOMO, which is connected with the molecule’s propensity to donate electrons, has been discovered. Obot and Obi-Egbedi (2010) reported that the inhibitor can bind to the metal surface via donor–acceptor interactions between the π-electrons of the heterocyclic compound and the unoccupied d-orbitals of the metal surface atoms. According to the characterization of the frontier molecular orbital (FMO) hypothesis, the capacity of inhibitor compounds to donate and take electrons, respectively, is frequently connected with the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO). EHOMO is connected to a molecule’s capacity for electron donation. Several investigations have demonstrated that inhibitors with high EHOMO values are more likely to transfer electrons to a suitable acceptor with a low empty molecular orbital energy. Their ELUMO value determines molecules’ electron-accepting capacity. A higher ELUMO value indicates a decreased probability of electrons being received by the molecule. For example, the inhibitor can contribute an electron from the metal’s d orbital to the vacant d orbital on the metal surface. Figure 2A demonstrates that the HOMO position inhibitors are related to the benzene ring, C–H, and LUMO sites in both inhibitors located near the heteroatoms. These represent the inhibitors’ sensitive locations for interaction between metal surfaces and inhibitor molecules. According to the literary works (Shokry 2014), Mulliken population and HOMO population studies can be utilized to determine probable inhibitor adsorption sites Yadav et al. (2016) and ElBelghiti et al. (2016). Several writers agree that the greater the amount and quantity of negatively charged heteroatoms available in an inhibitor molecule, the greater its capacity to be adsorbed on the metal surface through a donor–acceptor type bond. The excellent coordination and chelation properties of (Q4) improved its inhibitory behavior. These corrosion inhibitors reacted significantly with metals (Fe, Al, and Cu) by forming powerful coordination interactions. The formation of coordination bonds involving 8-HQ and metals was predominantly attributed to heteroatoms (N, O, P), aromatic rings, and polar functional groups. These electron-rich places are known as adsorption or coordination bonding locations, and they transfer nonbonding and π-electrons to the outermost metal atom’s vacant d-orbitals to form coordination bonds. This type of charge transfer creates interelectronic repulsion due to the electron concentrations of metals. Consequently, metals are compelled to contribute electron densities to these molecules’ unoccupied orbitals by retro- (back-) donation. Aromatic rings are better for retro-donation since their molecules possess lower LUMO energies. These operations take place on the metal surface at the same time. In coordination bonding, this is known as the synergistic effect. The materials listed in this study with their respective prices are identified in Figure 2A, considering that the values differ by state. In nonprotonated states, Q3 > Q6 > Q5 > Q4 > Q2 > Q1 in gas and aqueous phases. In protonated species, Q3 ≅ Q4 > Q6 > Q5 > Q2 > Q1. As a result, we found that Q3 has the highest value for EHOMO and Q1 has the lowest value compared to other molecules. The lowest values for ELUMO obtained and calculated by the Gaussian program are Q1 and Q5, as shown in Figure 2B.

Figure 2: 
						Quantum chemical calculations in different states gas aqueous phases at protonated and nonprotonated species based on DFT at 6–311 ++ G(d,p) basis set. (A)–(D) represents quantum chemical calculations.
Figure 2:

Quantum chemical calculations in different states gas aqueous phases at protonated and nonprotonated species based on DFT at 6–311 ++ G(d,p) basis set. (A)–(D) represents quantum chemical calculations.

Because a large bandgap requires a significant amount of ionization energy to remove an electron from the final occupied orbital, the inhibitor with the largest bandgap is the least effective corrosion inhibitor in terms of efficiency. Given the impact of both tiny and large bandgaps on the reactivity of compounds, the energy gap is essential in determining the inhibitor material’s sensitivity to adsorption on the metallic surface. Strong chemical activity and poor kinetic stability are connected with the lowest band gap energy, maximum softness, and highest systems. Q6 is the chemical that inhibits bandgap energy, the lowest and softest. The most significant corrosion inhibitors are often chemical molecules, which absorb free electrons from the metal and supply electrons to the vacant orbital of the metal. The order of corrosion inhibition efficiencies is as Q6 > Q2 > Q3 > Q4 > Q5 > Q1. As shown in Figure 2C and Tables 16, the resulting bandgap energy calculations show that Q6 has the lowest and Q1 has the highest value in terms of bandgap energy. The dipole moments of protonated and nonprotonated species in the gas and aqueous phases were calculated (Tables 16). The dipole moment of the molecules is the most commonly used measure to characterize polarity. A polar covalent bond’s dipole moment is an indication of its polarity. It is the product of atom charge and distance between two bound atoms. The molecular dipole moment of a whole molecule can be described as the vector sum of individual bond dipole moments. Dipole moment (in Debye) is another significant electrical characteristic from the nonuniform distribution of charges on the different atoms in the molecule. The creation of heavily adsorbed layers on the metal surface is caused by the intermolecular interactions becoming stronger and more significant as the dipole moment increases. The total surface coverage of inhibitors on the Fe, Al, and Cu surfaces can be predicted using dipole moments. The strong dipole moment most likely improves adsorption with chemical compounds and metal surfaces. The energy of deformability increases, allowing the molecule to adsorb on the metal surface. The volume of the inhibitors increases and increases the contact area between the molecule and the metal surface, enhancing the inhibitors’ capacity to stop corrosion. The corrosion-inhibiting action of these compounds is increased when compounds having a higher dipole moment, which enhances their corrosion resistance, are adsorbed on a metal surface. Increasing polarization leads the electrons to be more readily moved from a high level to a suitable lower level in an empty orbital due to an increase in the dipole moment’s value. The inhibitor with a high dipole moment has a high softness value, and the large adsorption capacity occurs in areas of high softness, increasing the inhibition corrosion performance, the softness of organic molecules may be influenced by the dipole moment. The order of all inhibitors for corrosion levels is arranged as follows: Q6 > Q2 > Q3 > Q4 > Q5 > Q1, which is precisely the same as the practical outcomes. The ionization potential (I) can be considered as important factor for interpreting the interaction of atoms or molecules. Low values of ionization potential indicate that molecules are inactive, and high values of ionization potential indicate that compounds are highly reactive, potentially playing a significant impact on the activity or inactivity of chemical processes. Tables 14 show that the increasing ionizability is rising, which is consistent with an upward trend in EHOMO. A good corrosion inhibitor is one that has a low electronegativity value. According to the findings in Tables 16, the electronegative order follows a pattern of Q4 > Q3 > Q2 > Q6 > Q5 > Q1 rises. The greatest electronegativity is found in molecule Q4, which implies the highest inhibition efficiency, according to Sanderson’s electronegativity matching approach. Q1 is predicted by low electronegativity and a high differential in electronegativity. DFT has effectively estimated electronegativity, one of the crucial chemical parameters for determining inhibition. Because electronegativity is the capacity of a molecule to attract electrons, a molecule with a high electronegativity draws the electron with greater force from the metal surface (Çakir and Emregül 2022). Because of their higher electronegativity than neutral states, protonated Q4 and Q3 inhibitors demonstrated superior inhibitory activity.

Table 1:

Quantum chemical parameters for 1,3-bis (quinoline-8-yloxy) propane (Q1).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −5.516 −9.097 −5.84 −6.647
LUMO (eV) −1.169 −4.486 −1.36 −2.097
Ionization energy (eV) 5.516 9.097 5.84 6.647
Electron affinity (eV) 1.169 4.486 1.36 2.097
Energy gap (eV) 4.347 4.611 4.48 4.55
Hardness (eV) 2.1735 2.3055 2.24 2.275
Softness (eV)-1 0.230 0.217 0.223 0.220
Electronegativity (eV) 3.3425 6.7915 3.6 4.372
Chemical potential (eV) −3.3425 −6.7915 −3.6 −4.372
Electrophilicity (eV) 2.570 10.003 2.893 4.201
Nucleophilicity (eV)-1 0.389 0.100 0.346 0.238
Back-donation (eV) −0.543 −0.576 −0.560 −0.569
Electron transfer 1.538 2.946 1.607 1.922
Dipole moment 2.803 3.871 4.412 6.369
Total energy T.E (a.u) −581.018 −582.767 −581.035 −579.826
Table 2:

Quantum chemical parameters for acenaphtho[1,2-b]quinoxaline (Q2).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −5.997 −10.086 −6.127 −7.136
LUMO (eV) −2.024 −6.217 −2.149 −3.304
Ionization energy (eV) 5.997 10.086 6.127 7.136
Electron affinity (eV) 2.024 6.217 2.149 3.304
Energy gap (eV) 3.973 3.869 3.978 3.832
Hardness (eV) 1.9865 1.9345 1.989 1.916
Softness (eV)-1 0.252 0.258 0.251 0.261
Electronegativity (eV) 4.0105 8.1515 4.138 5.22
Chemical potential (eV) −4.0105 −8.1515 −4.138 −5.22
Electrophilicity (eV) 4.048 17.174 4.304 7.111
Nucleophilicity (eV)-1 0.247 0.058 0.232 0.141
Back-donation (eV) −0.497 −0.484 −0.497 −0.479
Electron transfer 2.019 4.214 2.080 2.724
Dipole moment 1.376 1.686 2.090 4.790
Total energy T.E (a.u) −801.472 −801.200 −801.479 −801.259
Table 3:

Quantum chemical parameters for acenaphtho[1,2-b]pyrazine (Q3).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −6.562 −10.894 −6.607 −7.579
LUMO (eV) −2.449 −7.087 −2.515 −3.37
Ionization energy (eV) 6.562 10.894 6.607 7.579
Electron affinity (eV) 2.449 7.087 2.515 3.37
Energy gap (eV) 4.113 3.807 4.092 4.209
Hardness (eV) 2.0565 1.9035 2.046 2.1045
Softness (eV)-1 0.243 0.263 0.244 0.238
Electronegativity (eV) 4.5055 8.9905 4.561 5.4745
Chemical potential (eV) −4.5055 −8.9905 −4.561 −5.4745
Electrophilicity (eV) 4.935 21.232 5.084 7.120
Nucleophilicity (eV)-1 0.203 0.047 0.197 0.140
Back-donation (eV) −0.514 −0.476 −0.512 −0.526
Electron transfer 2.191 4.723 2.229 2.601
Dipole moment 1.773 0.342 2.596 0.288
Total energy T.E (a.u) −647.843 −647.545 −647.851 −647.613
Table 4:

Quantum chemical parameters for 8-hydroxyquinoline (Q4).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −5.712 −11.291 −5.848 −7.531
LUMO (eV) −1.437 −6.763 −1.487 −3.011
Ionization energy (eV) 5.712 11.291 5.848 7.531
Electron affinity (eV) 1.437 6.763 1.487 3.011
Energy gap (eV) 4.275 4.528 4.361 4.52
Hardness (eV) 2.1375 2.264 2.1805 2.26
Softness (eV)-1 0.234 0.221 0.229 0.221
Electronegativity (eV) 3.5745 9.027 3.6675 5.271
Chemical potential (eV) −3.5745 −9.027 −3.6675 −5.271
Electrophilicity (eV) 2.989 17.996 3.084 6.147
Nucleophilicity (eV)-1 0.335 0.056 0.324 0.163
Back-donation (eV) −0.534 −0.566 −0.545 −0.565
Electron transfer 1.672 3.987 1.682 2.332
Dipole moment 2.595 2.212 3.551 2.811
Total energy T.E (a.u) −637.157 −636.881 −637.162 −636.953
Table 5:

Quantum chemical parameters for 1,4-dibutyl-6-methyl-1,4-dihydroquinoxaline-2,3-dione (Q5).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −5.712 −10.281 −5.803 −7.105
LUMO (eV) −1.16 −5.793 −1.348 −2.695
Ionization energy (eV) 5.712 10.281 5.803 7.105
Electron affinity (eV) 1.16 5.793 1.348 2.695
Energy gap (eV) 4.552 4.488 4.455 4.410
Hardness (eV) 2.276 2.244 2.2275 2.205
Softness (eV)-1 0.220 0.223 0.224 0.227
Electronegativity (eV) 3.436 8.037 3.576 4.900
Chemical potential (eV) −3.436 −8.037 −3.576 −4.900
Electrophilicity (eV) 2.594 14.392 2.870 5.444
Nucleophilicity (eV)-1 0.386 0.069 0.348 0.184
Back-donation (eV) −0.569 −0.561 −0.557 −0.551
Electron transfer 1.510 3.582 1.605 2.222
Dipole moment 5.990 7.426 8.778 10.298
Total energy T.E (a.u) −922.266 −922.003 −922.280 −922.073
Table 6:

Quantum chemical parameters for 1-[3-phenyl-5-quinoxalin-6-yl-4,5-dihydropyrazol-1-yl]butan-1-one (Q6).

Quantum chemical parameters Gas phase Aqueous phases
Nonprotonated Protonated Nonprotonated Protonated
HOMO (eV) −5.802 −9.385 −5.949 −7.136
LUMO (eV) −2.160 −5.903 −2.125 −2.957
Ionization energy (eV) 5.802 9.385 5.949 7.136
Electron affinity (eV) 2.160 5.903 2.125 2.957
Energy gap (eV) 3.642 3.482 3.824 4.179
Hardness (eV) 1.821 1.741 1.912 2.090
Softness (eV)-1 0.275 0.287 0.262 0.239
Electronegativity (eV) 3.981 7.644 4.037 5.047
Chemical potential (eV) −3.981 −7.644 −4.037 −5.047
Electrophilicity (eV) 4.352 16.781 4.262 6.094
Nucleophilicity (eV)-1 0.230 0.060 0.235 0.164
Back-donation (eV) −0.455 −0.435 −0.478 −0.522
Electron transfer 2.186 4.391 2.111 2.415
Dipole moment 4.596 7.827 6.850 10.260
Total energy T.E (a.u) −1,106.504 −1,106.245 −1,106.521 −1,106.314

Kinetic, internal, and potential energy combine to form a system’s total energy. Hohenberg and Kohn proved that the total energy of a system is a unique effect of charge density (like atomic nuclei), including electron many-body implications (exchange and correlation) and the presence of a static external potential. The smallest value of the total energy functional represents the ground state energy of the system. This minimum is subsequently given by the electronic charge density, which is the actual single particle ground state energy. According to our analysis and ranking of the best inhibition efficiency, Q6 has the highest total energy of all the inhibitors, followed by Q2 > Q3 > Q4 > Q5 > Q1. Pearson and Songstad (1967) classified molecules as well as ions into four groups: hard acids, hard bases, soft acids, and soft bases and demonstrated that species from the same groups like to react with one another (the Hard and Soft Acid/Base (HSAB) Principle). Hard acids like to coordinate with hard bases, whereas soft acids prefer to interact with soft bases. Polarized chemical species are called hard, and unpolarized chemical species are called soft. Hard–hard interactions are mostly ionic, whereas primarily covalent are soft–soft interactions. Electrons are transferred between chemical species during chemical reactions. In comparison to hard molecules, soft molecules are more susceptible to charge transfer. Chemical hardness is directly related to chemical stability because hard compounds are less reactive. It can be observed that the molecular hardness values of chemical species in a reaction give vital hints to the anticipated process and assessment of reaction products. Chemical hardness is a key reactivity attribute of matter, and it is defined as resistance to electron cloud polarization or deformation of chemical species is an estimate of the stabilities and reactivities of molecules. In regard to chemical reactivity, soft molecules are more susceptible to reactivity than hard molecules in unimolecular processes like isomerization and dissociation. A molecule with a high softness value indicates low bandgap energy and increases the ability to donate and accept electrons. Whenever electron transfer occurs, it indicates a good corrosion inhibitor, and increasing hardness leads to decreased corrosion inhibition. As indicated in Tables 16, the ranking of the materials specified for corrosion prevention is as follows Q6 > Q2 > Q3 > Q4 > Q5 > Q1. The number of electrons transferred (∆Nmax) in a molecule is proportional to its capacity to donate electrons to the metal’s surface. The ∆Nmax values for the four compounds investigated vary from 1.538 to 4.790. A larger N value suggests a stronger proclivity to donate electrons to the electron-deficient site and a greater proclivity to interact and adsorb on the metal surface. N grows in the following order for molecules Q3 > Q6 > Q2 > Q4 > Q5 > Q1. The molecule Q3 (4.79) has the highest proportion of electrons transmitted (N), whereas Q1 and Q5 have the lowest proportions, which results in the lowest inhibitory efficiency. An electronic back donation process can occur as a result of the interaction between the inhibiting molecule and the surface of a metal in a basic model of charge transfer for donations and back donation of charges. Tables 16 reveal that electronic back-donation charges can occur during the process of interaction between the examined molecules and metal ions since their values are less than zero. When a zero and EBack-donation, the charges transferred to the molecule are energetically favorable, as results of the Q6 and Q3 are more energetically favored than others. The positive number of electrons transported (N) demonstrates that the molecules are electron acceptors, whereas the negative number of electrons transferred (N) demonstrates that the molecules are electron donors. As a result, as these inhibitors’ electron-donating capability to the metal surface increases, so does their inhibition effectiveness. The inhibitory efficiency increased as the metal surface’s electron-donating capacity increased (Lukovits et al. 2001). The quinolines studied in this study exhibit charge transfer characteristics toward mild steel. ∆Nmax < 3.6 eV represents a molecule’s inclination to donate electrons to a metal surface. In this study, all electron transfer values for all compounds and in all different states are positive. However, these values have changed. In some states, values are very high, such as nonprotonated Q6 and Q4, which have the highest value. Electrophilicity is a measure of a chemical species’ propensity to take electrons. The molecule with the greatest electrophilicity rating may accept the most electrons. A strong electrophile has a low electrophile and chemical potential, whereas a strong nucleophile has a significant electrophile and chemical potential. Lgaz et al. (2020) reported that a molecule increases its electron acceptance capacity if the electrophilic value is high, but conversely, if the electrophilic value is low, its electron acceptance capacity decreases. Moreover, molecules with a high nucleophile value have a high electron-donating ability. Therefore, it can be indicated which molecule has the highest electron-donating capacity or which molecule has the highest electron-accepting capacity, as shown in Tables 16. In general, Q3 and Q6 have the highest values in terms of electrophilicity. It implies that these two molecules have a higher ability to accept electrons than other molecules. Therefore, the best inhibitor against corrosion is a molecule with high electron acceptance.

Electrophilicity and nucleophilicity depend on several essential chemical factors. A molecule’s capacity to behave as an electrophile is determined by its ability to be an electron sink and receive a negative charge. Therefore, the molecule must be electron-deficient to accept the link to an electron source, and a positive molecule is an ideal electrophile. Because of its resonance, a neutral molecule can also be a good electrophile. If a resonance structure is viable, an electron sink at a bond site can certainly accept electrons to form a bond.

Furthermore, removing a good leaving group in a neutral molecule might cause an orbital to empty, resulting in a positive charge. An anion is an electron producer, and it cannot receive more electrons and, hence, cannot be a suitable electrophile. Weak polar or polarizable linkages are common in good electrophiles. It means that electrons might preferentially favor one side of a bond over the others, resulting in dipoles, and on one side of these bonds, there will be a partial positive charge.

4.2 Thermodynamic parameters

As seen in Figure 3, all compounds’ thermodynamic parameters – entropy (S), enthalpy (H), and Gibbs free energy (G) – were determined using DFT with the B3LYP/6–311 ++ G(d,p) basis set in the gas phase between 100 and 1,000 K. Such parameters characterize the progress and direction of a chemical process. Positive enthalpy values indicate an endothermic reaction, making dissolution difficult (Mamand et al. 2023; Rasul et al. 2023). It is known that an increase in temperature results in an increase in entropy, heat capacity, and enthalpy. Furthermore, Gibbs free energy decreases with increasing temperatures. The positive entropy value observed in this study confirms that the corrosion process exhibits a favorable entropic behavior, as shown in Figure 3. The phenomenon of spontaneous adsorption is indicated by the presence of negative values for the Gibbs free energy. From a comprehensive perspective, physisorption phenomena are characterized by Gibbs free energy values that are typically less than approximately (−40 kJ mol−1), and chemisorption is observed when the values are around or more negative than (−40 kJ mol−1). According to the calculations, the Gibbs free energy values for acenaphtho[1,2-b]pyrazine (Q3) and 8-hydroxyquinoline (Q4) are −14.402 kJ mol⁻1 and −21.362 kJ mol⁻1, respectively. These numerical values closely approximate the critical thresholds associated with physisorption.

Figure 3: 
						Thermodynamic parameters for all compounds.
Figure 3:

Thermodynamic parameters for all compounds.

Quinoxalines have been studied as corrosion inhibitors for different metals and alloys. Including a pyrazine ring next to the phenyl ring enhances the coverage of the target metals. Such an achievement by the two nitrogen atoms was adsorbed through sharing their lone pair electrons. 1,3-Bis(quinoline-8-yloxy) propane (Q1) is made composed of a central propane backbone with quinoline-8-yloxy groups connected to it. The presence of quinoline rings in the structure indicates protonation sites. Q1 comprises quinoline rings, which can serve as protonation sites. The nitrogen atoms in the quinoline rings have lone pairs of electrons, allowing them to receive protons in acidic situations. The protonation of Q1 would result in the formation of quinolinium cations, in which the nitrogen atoms carry positive charges. Acenaphtho[1,2-b]quinoxaline (Q2) comprises acenaphthene and quinoxaline rings fused. It lacks easily protonatable functional groups like amino or hydroxyl groups, which may limit its ability to undergo protonation. Although, under highly acidic conditions, protonation of Q2 could occur through an acid-aromatic ring reaction. However, compared to molecules with readily protonatable groups, the protonation of Q2 might be relatively less significant.

Q3, like Q2, is a polycyclic aromatic compound. However, it differs from Q2 because it incorporates a pyrazine ring instead of a quinoxaline ring. Similar to Q2, it also lacks easily protonatable functional groups. As a result, Q3 may exhibit relatively low reactivity toward protonation. 8-Hydroxyquinoline Q4, often known as oxine, is a quinolone hydroxy derivative. Q4 is a well-known chelating ligand commonly employed in qualitatively detecting metallic ions. Q4 will likely experience proton loss, forming Q4 chelate complexes with various metal ions. The coordination and chelation properties of Q4 are well recognized. As a result, Q4 has an extremely high potential for corrosion inhibition. Q4 strongly coordinates with metallic surfaces by using nonbonding electrons of oxygen and nitrogen, resulting in significant corrosion inhibition. When 8-hydroxyquinoline is dissolved in an acidic solution, the acidic nature of the solution provides protons that can interact with the basic nitrogen atom of Q4. The results indicated the formation of a protonated species often referred to as a quinolinium cation, where the nitrogen atom carries a positive charge. The protonation of 8-hydroxyquinoline can significantly affect its reactivity and chemical behavior. 1,4-Dibutyl-6-methyl-1,4-dihydroquinoxaline-2,3-dione Q5 and butyl and methyl substituents affect its protonation reactivity. Protonation is less likely because of the carbonyl groups and the dione functionality. As a result, compared to compounds having functional groups that may protonate more readily, Q5 may have comparatively low protonation reactivity. The electron-rich system of Q5 may interact with the acid under very acidic circumstances, resulting in protonation. However, protonation could be less advantageous due to the dione functionality’s electron-withdrawing properties. Q6 is 3-phenyl-5-quinoxalin-6-yl-4,5-dihydropyrazol-1-yl]butan-1-one that consists of a quinoxaline ring, a pyrazole ring, a phenyl group, and a butanone group. The nitrogen atoms in the quinoxaline and pyrazole rings can potentially undergo protonation. Due to electron-rich aromatic rings and nitrogen atoms with lone pairs of electrons, Q6 can accept protons under suitable conditions, displaying protonation reactivity. In acidic environments, the acid can interact with the lone pairs of electrons on the nitrogen atoms or other suitable sites within Q6. This interaction leads to protonation and the formation of a positively charged species.

4.3 Mulliken charge distribution

Predicting the active sites of inhibitor compounds is essential for studies on corrosion inhibition processes as shown in Figure 4. FMOs, Mulliken charge distribution, molecular electrostatic potential (MEP), and Fukui functions are frequently used to determine charge density distribution and possible binding sites. The FMOs are made up of HOMO and LUMO, and it is often believed that their energies are related to their ability to give and receive electrons. See Supplementary Material, it is feasible to distinguish between an inhibitor molecule’s active site – which is under attack from electrophile metal cations – and its site – which is prepared to take electrons (LUMO). The Mulliken charge distribution is used to identify the locations of an inhibitor molecule interacting with a metal surface. The HOMO energy density allocation is frequently used in conjunction with the Mulliken charge distribution to forecast the active site of the inhibitor molecule. A computer method called MEP indicates a molecule’s chemical reactivity. It explains how the electrical potential generated by the nuclei and electrons encircling molecules is distributed spatially. The electrostatic potential at each position r in the vicinity of a molecule is represented by Equation (7).

(7)V(r)=AZA|RAr|ρ(r)|rr|dr

where (ZA) is the charge of the nucleus A at position RA, and ρ(r) is the molecular electron density function. The sign of V(r), it depends on whether the electron or nucleus impact is noticeable at a particular place. Three-dimensional diagrams called MEP maps display the distribution of charges within molecules. Charges are related to their qualities. Color-graded mapping of the contours of the total electron density can be used to determine the active sites of nucleophiles and electrophiles in a molecular system. Different surface electrostatic potentials are indicated and depicted by the different colorings. The blue executive summary on most MEP maps shows the largest positive site. The red outline typically represents the largest negative site, which is easy to contribute electrons to. These are the preferred targets for nucleophilic and electrophilic attacks (Sanz et al. 1988). MEP maps to determine which regions of all inhibitors were the most active Sanz and others. The oxygen and nitrogen atoms in the ring are redder than those in other areas, based on color correction and the surrounding environment. Hence, the oxygen and nitrogen atoms in the rings are more active compared to other atoms. The change in a molecule’s electron density at a given position as the number of electrons changes is represented by the Fukui function, which is defined as the first derivative of the electron density (r) concerning the number of electrons N at a constant external potential v(r), as shown in Equation (8). The Fukui indices (fk+, fk, and fk0) may be calculated to evaluate the local reactivity of a molecule. Koumya et al. (Boussalah et al. 2012; Roy et al. 1999) reported that greater f(r) correlates to a more reactive active center of a molecule.

(8)f(r)=[ρ(r)N]v(r)
(9)fk+=[ρk(N+1)ρk(N)
(10)fk=ρk(N)ρk(N1)
fk0=12[ρk(N+1)ρk(N1)]
f(k)=fk+fk

where the corresponding electrical densities of site k for cationic, anionic, and neutral species are ρk(N), ρk(N+1), and ρk(N1). The highest values of fk+ and fk, respectively, indicate the locations most likely to experience nucleophilic and electrophilic attacks. The condensed Fukui functions can be used to investigate an inhibitor’s active site. Active sites in molecules with the most condensed Fukui functions encourage greater reactivity. These functions indicate which atoms in a molecule are more likely to give or take an electron or pair of electrons. The active site for nucleophilic assaults is the atom in the molecule with the greatest value of fk+. Meanwhile, the active site for electrophilic attacks is the atom in the molecule with the highest value of fk. In this study, we obtained active sites on heteroatoms Supplementary Material. In heteroatoms, electronic charges, Fukui functions, and proton affinities (PA) are critical in determining active sites on inhibitor compounds. Molecular electrostatic potential (MEP) is a descriptor for defining active areas for electrophilic and nucleophilic assaults that is related to electronic density. MEP of protonated and nonprotonated in the gas and aqueous phase inhibitors. The sections of the MEP of the selected inhibitors in this study are calculated by utilizing the optimized geometry at the 631 ++ G(d,p) to examine reactive sites for electrophilic and nucleophilic assault. The red and yellow colors represent electrophilic reactivity in the negative sections of the MEP, whereas blue colors suggest nucleophilic reactivity in the positive parts. The estimated findings suggest that the negative potential sites are on electronegative atoms (oxygen, nitrogen, and sulfur), while the positive potential sites are around the hydrogen atoms. In this work, the heteroatom plays an important role in determining the active sites.

Figure 4: 
						Chemical structures of selected inhibitors.
Figure 4:

Chemical structures of selected inhibitors.

4.4 Molecular dynamic simulation

The adsorption behavior and orientation of corrosion inhibitors on metal–electrolyte interfaces are studied using Monte Carlo simulations (MC) and molecular dynamics (MD) models. The results provide useful guidance. MD simulations are a computer-based modeling technique that may describe a molecule’s development as a trajectory or time function according to classical Newtonian physics principles (Guo et al. 2015). Since MC computations are the first ab initio force field that can efficiently, precisely, and concurrently estimate the condensed-phase characteristics for a wide range of chemical systems and gas-phase features, it is very beneficial to perform a fast simulation. In a specifically created simulation box, the computations model the interaction between the inhibitor and the metal surface when the inhibitor is in direct contact with the surface (Khaled 2009). The adsorption system or simulation results are composed of the optimized inhibitor molecules and the formed layers of iron atoms. The Fe, Cu, and Al (110) surface is widely utilized as the adsorption substrate because of its low surface energy and large coordination number of substrate atoms, resulting in more active interaction sites between the metal and the inhibitor molecules (Mamand and Qadr 2023b).

COMPASS is the first ab initio force field to enable a simultaneous and accurate prediction of condensed-phase and gas-phase features for a broad range of polymers and molecules. The COMPASS force field has been extended to span a more comprehensive range of molecules relevant to researchers researching ionic liquids. This is known as the COMPASSII force field (Sun et al. 2016). Copper and copper alloys are widely used in a wide range of manufacturing processes in the aquatic and nonaqueous. Copper is widely used in the electrotechnical field since it has excellent electrical conductivity.

Moreover, silver has higher electrical conductivity at room temperature than other metals. The significant energy differential allows electrons to flow between the valence and conduction bands. To prevent copper corrosion, several organic compounds are utilized. Adsorption at the metal–solution interface is the first stage in the mechanism of action of organic corrosion inhibitors in acidic environments. The electrical properties of the molecules (adsorbate), the chemical makeup of the solution, the characteristics of the metal’s outer layer, the temperature at which it happens of the reaction occurs, and the electrochemical potential at the metal–solution interface all influence the mechanism of adsorption (Dumont 1991). The attractive forces between the adsorbate and the metal are required for adsorption. Physisorption is the result of electrostatic interactions between the electrically charged surface of the metal and inhibitory organic ions or dipoles (El-Hajjaji et al. 2018).

The substrate–adsorbate configuration parameters consist of the total energy, which is the combined energies of the adsorbate components, rigid adsorption energy, and deformation energy as shown in Table 79. Adsorption energy is the energy required to adsorb the relaxed adsorbate component onto the substrate. The adsorption energy is the combined total of the rigidity and deformation energies of the adsorbate component. Rigid adsorption energy quantifies the energy produced or needed to adsorb the unrelaxed adsorbate component onto the substrate without any geometric adjustments. Deformation energy quantifies the energy generated during the relaxation of the adsorbate component on the substrate surface. The term (dEad/dNi) represents the energy of substrate–adsorbate combinations when one adsorbate component is eliminated. Monte Carlo simulations were employed to investigate the adsorption characteristics of quinoxaline derivatives (Q1–Q6) on various metal surfaces (Fe(110), Cu(111), and Al(110)) in neutral and their protonated states. The data in Tables 79 reveal that the adsorption energies varied significantly across the different inhibitors and metal surfaces. Monte Carlo simulations revealed that specific quinoxaline derivatives displayed remarkably stable low-energy adsorption arrangements on each investigated metal surface (Fe(110), Cu(111), and Al(110)) as shown in Figure 5. Our theoretical calculations identified promising adsorption behavior for quinoxaline derivatives on various metal surfaces. Notably, Q5 emerged as the frontrunner on Fe(110) with an exceptional adsorption energy (−1,277.413 kcal/mol), suggesting its potential as a highly effective iron corrosion inhibitor. Similarly, Q6 exhibited the highest favorable adsorption energies on Cu(111) (−723.728 kcal/mol) and Al(110) (−769.109 kcal/mol), indicating its potential for copper and aluminum applications. Interestingly, despite not having the highest adsorption energy on every surface, Q6 displayed consistently strong adsorption across different metals. This can be attributed to several factors: its larger conjugated electron system leading to enhanced π-backbonding, its smaller substituent allowing for closer interaction with the metal surface compared to Q5, and a potential additional interaction from the nitrogen atom in its pyrazole ring. These findings not only contribute to understanding the structure–property relationship between quinoxaline derivatives and their adsorption behavior but also pave the way for the development of novel and targeted corrosion inhibitors for specific metal compatibility.

Table 7:

The outputs and descriptors were calculated by the Monte Carlo simulation for adsorption on Fe (110) in protonated states.

Inhibitors Total energy (kcal mol−1) Adsorption energy (kcal mol−1) Rigid adsorption energy (kcal mol−1) Deformation energy (kcal mol−1) dEad/dNi (kcal mol−1)
Q1 53.519 −121.202 −5.082 −123.119 −121.202
Q2 86.454 −269.419 −2.206 −267.213 −269.419
Q3 74.792 −218.305 −4.357 −213.948 −218.305
Q4 20.901 −146.1809 −2.062 −144.118 −146.180
Q5 45.065 −1,277.413 −2.194 −265.218 −127.413
Q6 55.383 −724.723 −3.1781 −721.545 −724.723
Table 8:

The outputs and descriptors of the Monte Carlo simulation were calculated for adsorption on Cu (111) in protonated states.

Inhibitors Total energy (kcal mol−1) Adsorption energy (kcal mol−1) Rigid adsorption energy (kcal mol−1) Deformation energy (kcal mol−1) dEad/dNi (kcal mol−1)
Q1 54.328 −120.392 −4.268 −123.124 −120.392
Q2 86.197 −269.676 −2.471 −267.204 −269.676
Q3 76.916 −216.181 −2.223 −213.958 −216.181
Q4 22.180 −144.901 −0.781 −144.120 −144.901
Q5 52.258 −127.219 −2.602 −127.616 −260.219
Q6 56.378 −723.728 −2.145 −721.582 −723.728
Table 9:

The outputs and descriptors calculated by the Monte Carlo simulation for adsorption on Al (110) in protonated states.

Inhibitors Total energy (kcal mol−1) Adsorption energy (kcal mol−1) Rigid adsorption energy (kcal mol−1) Deformation energy (kcal mol−1) dEad/dNi (kcal mol−1)
Q1 −3.353 −158.075 −62.052 −416.022 −158.075
Q2 52.967 −302.906 −35.881 −267.024 −302.906
Q3 40.907 −252.190 −38.334 −213.855 −252.190
Q4 3.094 −163.987 −19.881 −144.106 −163.987
Q5 1.763 −160.714 −56.198 −160.516 −310.714
Q6 10.997 −769.109 −52.773 −716.335 −769.109
Figure 5: 
						Molecular simulations for the most favorable adsorption modes were obtained for the selected molecules on the Al, Cu, and Fe (110) surface (side view).
Figure 5:

Molecular simulations for the most favorable adsorption modes were obtained for the selected molecules on the Al, Cu, and Fe (110) surface (side view).

The total energy of the substrate–adsorbate configuration is characterized as a combination of the energies of the adsorbate components, deformation energy, and rigid adsorption energy. In this study, we have observed that the highest value of deformation energy can determine the highest corrosion resistance of the material, as shown in Tables 79. Adsorption energy is the energy released during the adsorption of the relaxation adsorbate element on the substrate. The rigid adsorption energy and the deformation energy of the adsorbate component add up to the adsorption energy. The energy released (or required) each time the unrelaxed adsorbate component was adsorbed on the substrate before the geometry optimization stage is represented by the rigid adsorption energy (Aquino-Torres et al. 2020). The energy produced when the deposited adsorbate constituent was loosened on the substrate surface is reported as the deformation energy. All adsorption energies shown in Tables 79 are negative, implying that adsorption is carried out spontaneously. The inhibitor molecules’ significant adsorption can explain the metal surfaces’ higher negative interaction energy values. Recognizing the adsorption phenomenon is critical in corrosion issues. Theoretical investigations aid in identifying the most stable adsorption sites for a wide range of compounds. The obtained data may be used to learn more about the corrosion system, including the probable location of the attack on a surface, the most enduring location for inhibitor adsorption, and the adsorption energy of the adsorbed layer. In light of these calculations, we can point out that the ranking of these molecules for anticorrosion is according to this ranking Q6 > Q2 > Q3 > Q4 > Q5 > Q1.

4.5 Nonlinear optical properties (NLO)

Nonlinear optical (NLO) materials are important in nonlinear optics, especially for information technology and industrial applications. The first static calculation was done on the optimized molecular geometry using the B3LYP/6–311 ++ G(d,p) computational method. The mean polarizability (|α0|) and first hyperpolarizability (β0) can be defined in terms of their x, y, and z components (Al-Shamiri et al. 2022; Evangalin et al. 2018; Villemin et al. 2018).

μt=[μx2+μy2+μz2]12αt=(αxx+αyy+αzz)/3β0=(βx2+βy2+βz2)12βx=βxxx+βxyz+βxzzβy=βyy+βxxy+βyzzβz=βzzz+βxxz+βyzz

Large values of the polarizability and hyperpolarizability components in one direction imply significant delocalization of charge in that direction. Table 10 displays the calculated mean polarizability (α0) and initial hyperpolarizability (β0) for all compounds. The polarizability (α0) and first hyperpolarizability (β0) parameters of the GAUSSIAN 09 output are achieved in atomic units (a.u.), and the computed values have been converted into electrostatic units (e.s.u.) (for α;1 a.u = 0.1482 × 10−24 e.s.u., for β;1 a.u = 8.6393 × 10−33 e.s.u.). Urea is used to study molecular NLO properties. Since there were no experimental standards of NLO characteristics of the compounds, it is commonly utilized as a reference point for comparisons (Guerrab et al. 2020). The polarizability of Q4 is found to be the lowest, while Q1 exhibits the highest polarizability among the calculated values. Furthermore, the determination of the molecular hyperpolarizability magnitude (β0) holds significant importance within a nonlinear optical (NLO) system. Urea’s first hyperpolarizability (β) exceeds that of all other compounds in terms of their first hyperpolarizability. These results indicate that the entirety of the compounds may not possess the potential for nonlinear optical (NLO) applications, as presented in Table 10.

Table 10:

Nonlinear optical properties of all compounds.

Parameters Q1 Q2 Q3 Q4 Q5 Q6
α xx 333.9230 60.9300 261.6870 149.3240 232.6380 337.3770
α yy 0.0000 0.0000 0.0000 −7.3320 6.7970 −2.9050
α zz 301.7730 227.1390 207.7680 114.8970 244.4020 255.7670
α t 211.8987 96.0230 156.4850 85.6297 161.2790 196.7463
α (esu) × 10−24 31.4034 14.2306 23.1911 12.6903 23.9015 29.1578
β xxx 0.0000 0.0000 17.0718 10.7467 −3.9231 −27.9611
β xyy 0.0000 0.0000 −29.6781 7.7477 8.5690 51.6500
β xzz 0.0000 0.0000 7.5842 −1.2891 −0.8738 −22.9230
β x 0.0000 0.0000 −5.0221 17.2053 3.7721 0.7659
β yyy −12.3423 0.0000 0.0001 −0.9905 76.7717 −56.9751
β xxy −48.6777 0.0000 0.0000 6.8849 2.6983 −78.6809
β yzz −3.2509 0.0000 0.0000 −6.5005 −11.7598 19.2379
β y −64.2709 0.0000 0.0001 −0.6061 67.7102 −116.4181
β zzz 0.0007 −12.3822 0.0000 0.0000 −6.0436 −0.5928
β xxz −0.0037 −2.8253 0.0006 0.0000 −8.8260 26.7631
β yyz −0.0001 18.3761 0.0003 0.0000 18.4248 −35.6911
β z −0.0031 3.1686 0.0009 0.0000 3.5552 −9.5208
β 0(esu) × 10−33 64.271 3.167 5.022 17.216 67.908 116.809
β 0(esu) = 372.8 × 10−33 for urea

4.6 NBO analysis

NBO analysis describes intramolecular interactions and charge transfer. NBO analysis explains the donor–acceptor behavior of atoms in the molecule. Density Functional Theory (DFT) is used to study electron delocalization and the interactions between occupied and empty orbitals. Furthermore, the degree of stabilization resulting from orbital interaction is directly proportional to the disparity in energy levels between the orbitals involved. Therefore, the interaction that exhibits strong stabilization occurs between efficient donors and efficient acceptors.

Moreover, the E2 value represents the magnitude of the interaction energy between an electron acceptor and electron donors. A higher E2 value indicates a stronger propensity for electron donation from the donor to the acceptor (Evangalin et al. 2018; Kanmazalp 2017; Villemin et al. 2018). The NBO analysis results for the molecule were obtained using the B3LYP/6–311 ++ G(d,p) level of theory, as presented in Supplementary Material.

E(2)=ΔEij=qiF(i,j)2εjεi

where qi is the donor orbital occupancy, εj and εi are the diagonal elements, and F(i,j) is the off diagonal NBO Fock matrix element (İzzet et al. 2016; Khalid et al. 2020).

In this investigation, several types of electronic transitions were commonly observed: σ → σ*, π → π*, n → σ*, n → π*, and π* → π* as tabulated. In Q4 and Q3, the greatest interaction is between π*(C4–C5) → π*(C1–C6) and π*(C5–C6) → π*(C1–C2) with the stabilization energy of 252.09 kcal/mol and 249.09 kcal/mol, respectively. In Q6 and Q4, the electron donating from LP(2) (O26) and LP(1) (N11) to the antibonding acceptor π*(N18–C22) and π*(C1–C6) with stabilization energy of 26.99, and 34.41 kcal/mol is the most important interactions. Besides, LP(1)O13 to the antibonding acceptor π*(C8–C9) stabilization energy of 22.68 kcal/mol is also a significant interaction in Q5.

5 Conclusions

A theoretical investigation was carried out utilizing the B3LYP functionals and a 6-311G ++ (d,p) basis set to learn about the reactive behavior of some quinoxaline and their potential applicability as corrosion inhibitors. Using the Fukui functions, the HOMO, and the Mulliken population analysis, we determined that the reactive sites of these derivatives are primarily located in the hetero atoms. Based on the discovery of paramedics deciding on the rust ability, we found that the ranking is as follows: Q6 > Q2 > Q3 > Q4 > Q5 > Q1. Some parameters don’t participate in the ranking, but based on most parameters such as HOMO, LUMO, bandgap energy, softness, hardness, and Monte Carlo simulation concluded. The positive entropy value observed in this study confirms that the corrosion process exhibits a favorable entropic behavior. In addition, the Gibbs free energy values for acenaphtho[1,2-b] pyrazine (AP) and 8-hydroxyquinoline are −14.402 kJ mol−1 and 6.092, −21.362 kJ mol−1, respectively. These numerical values closely approximate the critical thresholds associated with physisorption. These findings suggest that not all chemicals may be suitable for NLO applications. According to the calculations carried out by NBO, the most important interaction in Q4 and Q3 takes place between *(C4–C5) and *(C1–C6), whereas *(C5–C6) and *(C1–C2) have stabilization energy of 252.09 and 249.09 kcal/mol, respectively.


Corresponding author: Karzan A. Omar, Department of Chemistry, Faculty of Science & Health, Koya University, KoyaKOY45, Kurdistan Region – F.R., Iraq; and Department of Medical Laboratory Science, College of Science, Cihan University, Erbil, Iraq, E-mail:

Acknowledgments

The authors would like to thank the Chemistry Department of Koya University.

  1. Research ethics: Not applicable.

  2. Author contributions: The authors have accepted responsibility for the entire content of this manuscript and approved its submission.

  3. Competing interests: The authors state no conflict of interest.

  4. Research funding: None declared.

  5. Data availability: Not applicable.

References

Ahmed, L., Bulut, N., Kaygili, O., and Rebaz, O. (2023). Quantum chemical study of some basic organic compounds as the corrosion inhibitors. J. Phys. Chem. Funct. Mater. 23: 634–642, https://doi.org/10.54565/jphcfum.1263803.Search in Google Scholar

Al-Shamiri, H.A., Sakr, M.E., Abdel-Latif, S.A., Negm, N.A., Abou Kana, M.T., El-Daly, S.A., and Elwahy, A.H. (2022). Experimental and theoretical studies of linear and non-linear optical properties of novel fused-triazine derivatives for advanced technological applications. Sci. Rep. 12: 9937, https://doi.org/10.1038/s41598-022-22311-z.Search in Google Scholar PubMed PubMed Central

Andzelm, J., Kölmel, C., and Klamt, A. (1995). Incorporation of solvent effects into density functional calculations of molecular energies and geometries. J. Chem. Phys. 10: 9312–9320, https://doi.org/10.1063/1.469990.Search in Google Scholar

Anwar Omar, R., Koparir, P., Koparir, M., and Safin, D.A. (2023). A novel cyclobutane-derived thiazole–thiourea hybrid with a potency against COVID-19 and tick-borne encephalitis: synthesis, characterization, and computational analysis. J. Sulfur Chem. 45: 1–18, https://doi.org/10.1080/17415993.2023.2260918.Search in Google Scholar

Aquino-Torres, E., Camacho-Mendoza, R.L., Gutierrez, E., Rodriguez, J.A., Feria, L., Thangarasu, P., and Cruz-Borbolla, J. (2020). The influence of iodide in corrosion inhibition by organic compounds on carbon steel: theoretical and experimental studies. Appl. Surf. Sci. 514: 145928, https://doi.org/10.1016/j.apsusc.2020.145928.Search in Google Scholar

Boussalah, N., Ghalem, S., El Kadiri, S., Hammouti, B., and Touzani, R. (2012). Theoretical study of the corrosion inhibition of some bipyrazolic derivatives: a conceptual DFT investigation. Rev. Chem. Intermed. 38: 2009–2023, https://doi.org/10.1007/s11164-012-0522-0.Search in Google Scholar

Çakir, B. and Emregül, K.C. (2022). A theoretical approach to the inhibitive effect of two new Schiff base compounds. Commun. Fac. Sci. Univ. Ankara, Ser. B: Chem. Chem. Eng. 64: 1–19.Search in Google Scholar

Chauhan, D.S., Singh, P., and Quraishi, M. (2020). Quinoxaline derivatives as efficient corrosion inhibitors: current status, challenges and future perspectives. J. Mol. Liq. 320: 114387, https://doi.org/10.1016/j.molliq.2020.114387.Search in Google Scholar

Chen, S., Zhu, B., and Liang, X. (2020). Corrosion inhibition performance of coconut leaf extract as a green corrosion inhibitor for X65 steel in hydrochloric acid solution. Int. J. Electrochem. Sci. 15: 1–15, https://doi.org/10.20964/2020.01.39.Search in Google Scholar

Dhanaraj, C.J. and Johnson, J. (2015). Quinoxaline based bio-active mixed ligand transition metal complexes: synthesis, characterization, electrochemical, antimicrobial, DNA binding, cleavage, antioxidant and molecular docking studies. J. Photochem. Photobiol., B 151: 100–109, https://doi.org/10.1016/j.jphotobiol.2015.07.010.Search in Google Scholar PubMed

Dumont, R.S. (1991). A metropolis Monte Carlo method for computing microcanonical statistical rate constants. J. Comput. Chem. 12: 391–401, https://doi.org/10.1002/jcc.540120313.Search in Google Scholar

ElBelghiti, M., Karzazi, Y., Dafali, A., Hammouti, B., Bentiss, F., Obot, I., Bahadur, I., and Ebenso, E.-E. (2016). Experimental, quantum chemical and Monte Carlo simulation studies of 3, 5-disubstituted-4-amino-1, 2, 4-triazoles as corrosion inhibitors on mild steel in acidic medium. J. Mol. Liq. 218: 281–293, https://doi.org/10.1016/j.molliq.2016.01.076.Search in Google Scholar

El Ghayati, L., Sert, Y., Sebbar, N.K., Ramli, Y., Ahabchane, N.H., Talbaoui, A., Mague, J.T., El Ibrahimi, B., Taha, M.L., Essassi, E.M., et al.. (2021). Syntheses of novel 1, 5-benzodiazepine derivatives: crystal structures, spectroscopic characterizations, Hirshfeld surface analyses, molecular docking studies, DFT calculations, corrosion inhibition anticipation, and antibacterial activities. J. Heterocycl. Chem. 58: 270–289, https://doi.org/10.1002/jhet.4167.Search in Google Scholar

El-Hajjaji, F., Belghiti, M., Hammouti, B., Jodeh, S., Hamed, O., Lgaz, H., and Salghi, R. (2018). Adsorption and corrosion inhibition effect of 2-mercaptobenzimidazole (surfactant) on a carbon steel surface in an acidic medium: experimental and Monte Carlo simulations. Port. Electrochim. Acta 36: 197–212, https://doi.org/10.4152/pea.201803197.Search in Google Scholar

El Sayed, H., El Nemr, A., Esawy, S.A., and Ragab, S. (2006). Corrosion inhibitors. Part II: quantum chemical studies on the corrosion inhibitions of steel in acidic medium by some triazole, oxadiazole and thiadiazole derivatives. Electrochim. Acta 51: 3957–3968.10.1016/j.electacta.2005.11.010Search in Google Scholar

Erdoğan, Ş., Safi, Z.S., Kaya, S., Işın, D.Ö., Guo, L., and Kaya, C. (2017). A computational study on corrosion inhibition performances of novel quinoline derivatives against the corrosion of iron. J. Mol. Struct. 1134: 751–761, https://doi.org/10.1016/j.molstruc.2017.01.037.Search in Google Scholar

Evangalin, J., Bharanidharan, S., and Dhandapani, A. (2018). Molecular spectroscopic investigations of (E)-1-(4 methylbenzylidene) urea using DFT method. Arch. Pharmacal Res. 2: 10–21.Search in Google Scholar

Feng, L., Yang, H., Cui, X., Chen, D., and Li, G. (2018). Experimental and theoretical investigation on corrosion inhibitive properties of steel rebar by a newly designed environmentally friendly inhibitor formula. RSC Adv. 8: 6507–6518, https://doi.org/10.1039/c7ra13045g.Search in Google Scholar PubMed PubMed Central

Fergachi, O., Benhiba, F., Rbaa, M., Ouakki, M., Galai, M., Touir, R., Lakhrissi, B., Oudda, H., and Touhami, M.E. (2019). Corrosion inhibition of ordinary steel in 5.0 M HCl medium by benzimidazole derivatives: electrochemical, UV–visible spectrometry, and DFT calculations. J. Bio- Tribo-Corros. 5: 1–13.10.1007/s40735-018-0215-3Search in Google Scholar

Fu, J.-J., Li, S.-N., Wang, Y., Cao, L.-H., and Lu, L.-D. (2010). Computational and electrochemical studies of some amino acid compounds as corrosion inhibitors for mild steel in hydrochloric acid solution. J. Mater. Sci. 45: 6255–6265, https://doi.org/10.1007/s10853-010-4720-0.Search in Google Scholar

Guerrab, W., Lgaz, H., Kansiz, S., Mague, J.T., Dege, N., Ansar, M., Marzouki, R., Taoufik, J., Ali, I.H., Chung, I.-M., et al.. (2020). Synthesis of a novel phenytoin derivative: crystal structure, Hirshfeld surface analysis and DFT calculations. J. Mol. Struct. 1205: 127630, https://doi.org/10.1016/j.molstruc.2019.127630.Search in Google Scholar

Guo, L., Ren, X., Zhou, Y., Xu, S., and Gong, Y. (2015) Monte Carlo simulations of corrosion inhibition of copper by two Schiff bases. In: 2015 International Conference on Materials, Environmental and Biological Engineering. Atlantis Press.10.2991/mebe-15.2015.143Search in Google Scholar

Guo, L., Obot, I.B., Zheng, X., Shen, X., Qiang, Y., Kaya, S., and Kaya, C. (2017). Theoretical insight into an empirical rule about organic corrosion inhibitors containing nitrogen, oxygen, and sulfur atoms. Appl. Surf. Sci. 406: 301–306, https://doi.org/10.1016/j.apsusc.2017.02.134.Search in Google Scholar

Haque, J., Srivastava, V., Chauhan, D.S., Quraishi, M., Kumar, A.M., and Lgaz, H. (2020). Electrochemical and surface studies on chemically modified glucose derivatives as environmentally benign corrosion inhibitors. Sustain. Chem. Pharm. 16: 100260, https://doi.org/10.1016/j.scp.2020.100260.Search in Google Scholar

Hussein, Y.T. and Azeez, Y.H. (2023). DFT analysis and in silico exploration of drug-likeness, toxicity prediction, bioactivity score, and chemical reactivity properties of the urolithins. J. Biomol. Struct. Dyn. 41: 1168–1177, https://doi.org/10.1080/07391102.2021.2017350.Search in Google Scholar PubMed

Izadi, M., Shahrabi, T., and Ramezanzadeh, B. (2018). Active corrosion protection performance of an epoxy coating applied on the mild steel modified with an eco-friendly sol-gel film impregnated with green corrosion inhibitor loaded nanocontainers. Appl. Surf. Sci. 440: 491–505, https://doi.org/10.1016/j.apsusc.2018.01.185.Search in Google Scholar

İzzet, K., KİRAZ, A.Ö., Cetişli, H., Donat, R., and Kolsuz, N. (2016). Theoretical calculation on a compound formed by methyl alcohol and simmondsin. Int. J. Second. Metab. 3: 39–348.10.21448/ijsm.240701Search in Google Scholar

Kanmazalp, S.D.r. (2017). Investigation of theoretical calculations of 2-(1-phenylethylideneamino) guanidine compound: NBO, NLO, HOMO-LUMO and MEP analysis by DFT method. Karaelmas Sci. Eng. J. 7: 491–496.Search in Google Scholar

Khaled, K. (2009). Monte Carlo simulations of corrosion inhibition of mild steel in 0.5 M sulphuric acid by some green corrosion inhibitors. J. Solid State Chem. 13: 1743–1756, https://doi.org/10.1007/s10008-009-0845-y.Search in Google Scholar

Khalid, M., Ali, A., Jawaria, R., Asghar, M.A., Asim, S., Khan, M.U., Hussain, R., ur Rehman, M.F., Ennis, C.J., and Akram, M.S. (2020). First principles study of electronic and nonlinear optical properties of A–D–π–A and D–A–D–π–A configured compounds containing novel quinoline–carbazole derivatives. RSC Adv. 10: 22273–22283, https://doi.org/10.1039/d0ra02857f.Search in Google Scholar PubMed PubMed Central

Koparir, P., Omar, R.A., Sarac, K., Ahmed, L.O., Karatepe, A., Taskin-Tok, T., and Safin, D.A. (2022). Synthesis, characterization and computational analysis of thiophene-2, 5-diylbis ((3-mesityl-3-methylcyclobutyl) methanone). Polycycl. Aromat. Comp. 43: 1–19, https://doi.org/10.1080/10406638.2022.2112712.Search in Google Scholar

Lgaz, H., Salghi, R., Jodeh, S., Ramli, Y., Larouj, M., Toumiat, K., Quraishi, M., Oudda, H., and Jodeh, W. (2016). Understanding the adsorption of quinoxaline derivatives as corrosion inhibitors for mild steel in acidic medium: experimental, theoretical and molecular dynamic simulation studies. J. Steel Struct. Construct. 2: 100011.10.4172/2472-0437.1000111Search in Google Scholar

Lgaz, H., Salghi, R., Masroor, S., Kim, S.-H., Kwon, C., Kim, S.Y., Yang, Y.-J., and Chung, I.-M. (2020). Assessing corrosion inhibition characteristics of hydrazone derivatives on mild steel in HCl: insights from electronic-scale DFT and atomic-scale molecular dynamics. J. Mol. Liq. 308: 112998, https://doi.org/10.1016/j.molliq.2020.112998.Search in Google Scholar

Liu, Y. and Shi, X. (2009). Electrochemical chloride extraction and electrochemical injection of corrosion inhibitor in concrete: state of the knowledge. Corros. Rev. 27: 53–82, https://doi.org/10.1515/corrrev.2009.27.1-2.53.Search in Google Scholar

Lukovits, I., Kalman, E., and Zucchi, F. (2001). Corrosion inhibitors—correlation between electronic structure and efficiency. Corrosion 57: 3–8, https://doi.org/10.5006/1.3290328.Search in Google Scholar

Maldonado, N. and Amo-Ochoa, P. (2021). Advances and novel perspectives on colloids, hydrogels, and aerogels based on coordination bonds with biological interest ligands. Nanomaterials 11: 1865, https://doi.org/10.3390/nano11071865.Search in Google Scholar PubMed PubMed Central

Mamad, D.M., Omer, R.A., and Othman, K.A. (2023a). Quantum chemical analysis of amino acids as anti-corrosion agents. Corros. Rev. 41: 703–717, https://doi.org/10.1515/corrrev-2023-0031.Search in Google Scholar

Mamad, D.M., Rasul, H.H., Awla, A.H., and Omer, R.A. (2023b). Insight into corrosion inhibition efficiency of imidazole-based molecules: a quantum chemical study. Dokl. Phys. Chem. 511: 125–133, https://doi.org/10.1134/s0012501623600043.Search in Google Scholar

Mamand, D.M. and Qadr, H.M. (2021). Comprehensive spectroscopic and optoelectronic properties of bbl organic semiconductor. Prot. Met. Phys. Chem. Surf. 57: 943–953, https://doi.org/10.1134/s207020512105018x.Search in Google Scholar

Mamand, D. and Qadr, H. (2023a). Quantum computations and density functional theory on corrosion inhibition efficiency of BIA, HBT, MBI and PIZ compounds. CPST. 14: 159–172.10.15407/hftp14.02.159Search in Google Scholar

Mamand, D.M. and Qadr, H.M. (2023b). Corrosion inhibition efficiency and quantum chemical studies of some organic compounds: theoretical evaluation. Corros. Rev. 41: 427–441, https://doi.org/10.1515/corrrev-2022-0085.Search in Google Scholar

Mamand, D.M., Azeez, Y.H., and Qadr, H.M. (2023). Monte Carlo and DFT calculations on the corrosion inhibition efficiency of some benzimide molecules. Mong. J. Chem. 24: 1–10, https://doi.org/10.5564/mjc.v24i50.2435.Search in Google Scholar

Mattsson, A.E., Schultz, P.A., Desjarlais, M.P., Mattsson, T.R., and Leung, K. (2004). Designing meaningful density functional theory calculations in materials science: a primer. Model. Simulat. Mater. Sci. Eng. 13: R1, https://doi.org/10.1088/0965-0393/13/1/r01.Search in Google Scholar

Mousavi, M., Mohammadalizadeh, M., and Khosravan, A. (2011). Theoretical investigation of corrosion inhibition effect of imidazole and its derivatives on mild steel using cluster model. Corros. Sci. 53: 3086–3091, https://doi.org/10.1016/j.corsci.2011.05.034.Search in Google Scholar

Obot, I. and Obi-Egbedi, N. (2010). Adsorption properties and inhibition of mild steel corrosion in sulphuric acid solution by ketoconazole: experimental and theoretical investigation. Corros. Sci. 52: 198–204, https://doi.org/10.1016/j.corsci.2009.09.002.Search in Google Scholar

Obot, I., Kaya, S., Kaya, C., and Tüzün, B. (2016). Theoretical evaluation of triazine derivatives as steel corrosion inhibitors: DFT and Monte Carlo simulation approaches. Rev. Chem. Intermed. 42: 4963–4983, https://doi.org/10.1007/s11164-015-2339-0.Search in Google Scholar

Obot, I., Haruna, K., and Saleh, T. (2019a). Atomistic simulation: a unique and powerful computational tool for corrosion inhibition research. Arab. J. Sci. Eng. 44: 1–32, https://doi.org/10.1007/s13369-018-3605-4.Search in Google Scholar

Obot, I., Meroufel, A., Onyeachu, I.B., Alenazi, A., and Sorour, A.A. (2019b). Corrosion inhibitors for acid cleaning of desalination heat exchangers: progress, challenges and future perspectives. J. Mol. Liq. 296: 111760, https://doi.org/10.1016/j.molliq.2019.111760.Search in Google Scholar

Olasunkanmi, L.O., Kabanda, M.M., and Ebenso, E.E. (2016). Quinoxaline derivatives as corrosion inhibitors for mild steel in hydrochloric acid medium: electrochemical and quantum chemical studies. Phys. E Low-dimens. Syst. Nanostruct. 76: 109–126, https://doi.org/10.1016/j.physe.2015.10.005.Search in Google Scholar

Omar, R.A., Koparir, P., Sarac, K., Koparir, M., and Safin, D.A. (2023). A novel coumarin-triazole-thiophene hybrid: synthesis, characterization, ADMET prediction, molecular docking and molecular dynamics studies with a series of SARS-CoV-2 proteins. Chem. Sci. J. 1356: 6, https://doi.org/10.1007/s12039-022-02127-0.Search in Google Scholar PubMed PubMed Central

Omer, R.A., Koparir, P., and Ahmed, L. (2022a). Theoretical determination of corrosion inhibitor activities of 4-allyl-5-(pyridin-4-yl)-4H-1, 2, 4-triazole-3-thiol-thione tautomerism. Indian J. Chem. Technol. 29: 75–81.Search in Google Scholar

Omer, R.A., Koparir, P., and Ahmed, L.O. (2022b). Characterization and inhibitor activity of two newly synthesized thiazole. J. Bio- Tribo-Corros. 8: 28, https://doi.org/10.1007/s40735-021-00625-1.Search in Google Scholar

Omer, R.A., Koparir, P., and Koparir, M. (2022c). Synthesis, experimental and theoretical characterization with inhibitor activity for 1, 2, 4-traizaol derivatives. Indian J. Chem. 61: 1278–1287.10.56042/ijc.v61i12.69443Search in Google Scholar

Omer, R., Koparir, P., Koparir, M., Rashid, R., Ahmed, L., and Hama, J. (2022d). Synthesis, characterization and DFT study of 1-(3-mesityl-3-methylcyclobutyl)-2-((4-phenyl-5-(thiophen-2-yl)-4H-1, 2, 4-triazol-3-yl) thio) ethan-1-one. Prot. Met. Phys. Chem. Surf. 58: 1077–1089, https://doi.org/10.1134/s2070205122050185.Search in Google Scholar

Omer, R.A., Ahmed, K.M., Kamal, S., Omer, K.A., Hamad, W.M., and Mamad, D.M. (2023a). N, N-bis (2, 4-dihydroxy benzaldehyde) benzidine: synthesis, characterization, DFT, and theoretical corrosion study. J. Mol. Struct. 1300: 137279, https://doi.org/10.1016/j.molstruc.2023.137279.Search in Google Scholar

Omer, R.A., Mustafa, R.A., Salih, S., Hamad, W., and Taher, S. (2023b). Synthesis of a new series of benzothiazole compounds and study of their liquid crystal properties. Passer J. Basic Appl. Sci. 5: 78–84.Search in Google Scholar

Onyeachu, I., Chauhan, D., Ansari, K., Obot, I., Quraishi, M., and Alamri, A.H. (2019). Hexamethylene-1, 6-bis (N-d-glucopyranosylamine) as a novel corrosion inhibitor for oil and gas industry: electrochemical and computational analysis. New J. Chem. 43: 7282–7293, https://doi.org/10.1039/c9nj00023b.Search in Google Scholar

Parlak, A.E., Omar, R.A., Koparir, P., and Salih, M.I. (2022). Experimental, DFT and theoretical corrosion study for 4-(((4-ethyl-5-(thiophen-2-yl)-4H-1, 2, 4-triazole-3-yl) thio) methyl)-7, 8-dimethyl-2H-chromen-2-one. Arab. J. Chem. 15: 104088, https://doi.org/10.1016/j.arabjc.2022.104088.Search in Google Scholar

Pearson, R.G. (1988). Absolute electronegativity and hardness: application to inorganic chemistry. Inorg. Chem. 27: 734–740, https://doi.org/10.1021/ic00277a030.Search in Google Scholar

Pearson, R.G. and Songstad, J. (1967). Application of the principle of hard and soft acids and bases to organic chemistry. J. Am. Chem. Soc. 89: 1827–1836, https://doi.org/10.1021/ja00984a014.Search in Google Scholar

Rasul, H.H., Mamad, D.M., Azeez, Y.H., Omer, R.A., and Omer, K.A. (2023). Theoretical investigation on corrosion inhibition efficiency of some amino acid compounds. Comput. Theor. Chem. 1225: 114177, https://doi.org/10.1016/j.comptc.2023.114177.Search in Google Scholar

Roy, R.K., Pal, S., and Hirao, K. (1999). On non-negativity of Fukui function indices. J. Chem. Phys. 110: 8236–8245, https://doi.org/10.1063/1.478792.Search in Google Scholar

Sanz, F., Manaut, F., José, J., Segura, J., Carbó, M., and De la Torre, R. (1988). Automatic determination of MEP patterns of molecules and its application to caffeine metabolism inhibitors. J. Mol. Struct. 170: 171–180, https://doi.org/10.1016/0166-1280(88)80060-5.Search in Google Scholar

Shokry, H. (2014). Molecular dynamics simulation and quantum chemical calculations for the adsorption of some azo-azomethine derivatives on mild steel. J. Mol. Struct. 1060: 80–87, https://doi.org/10.1016/j.molstruc.2013.12.030.Search in Google Scholar

Sun, H., Jin, Z., Yang, C., Akkermans, R.L., Robertson, S.H., Spenley, N.A., Miller, S., and Todd, S.M. (2016). COMPASS II: extended coverage for polymer and drug-like molecule databases. J. Mol. Model. 22: 1–10, https://doi.org/10.1007/s00894-016-2909-0.Search in Google Scholar PubMed

Umoren, S.A. and Solomon, M.M. (2017). Synergistic corrosion inhibition effect of metal cations and mixtures of organic compounds: a review. J. Environ. Chem. Eng. 5: 246–273, https://doi.org/10.1016/j.jece.2016.12.001.Search in Google Scholar

Villemin, D., Abbaz, T., and Bendjeddou, A. (2018). Molecular structure, HOMO, LUMO, MEP, natural bond orbital analysis of benzo and anthraquinodimethane derivatives. Pharm. Biol. Eval. 5: 1–15, https://doi.org/10.26510/2394-0859.pbe.2018.04.Search in Google Scholar

Wang, D., Gao, L., Zhang, D., Yang, D., Wang, H., and Lin, T. (2016). Experimental and theoretical investigation on corrosion inhibition of AA5052 aluminium alloy by l-cysteine in alkaline solution. Mater. Chem. Phys. 169: 142–151, https://doi.org/10.1016/j.matchemphys.2015.11.041.Search in Google Scholar

Yadav, M., Gope, L., Kumari, N., and Yadav, P. (2016). Corrosion inhibition performance of pyranopyrazole derivatives for mild steel in HCl solution: gravimetric, electrochemical and DFT studies. J. Mol. Liq. 216: 78–86, https://doi.org/10.1016/j.molliq.2015.12.106.Search in Google Scholar


Supplementary Material

This article contains supplementary material (https://doi.org/10.1515/corrrev-2024-0007).


Received: 2024-01-26
Accepted: 2024-04-21
Published Online: 2024-07-19
Published in Print: 2024-12-17

© 2024 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 21.9.2025 from https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2024-0007/html?lang=en
Scroll to top button