Review article

Quantitative and qualitative insights into the corrosion mitigation mechanism of N,N-dibutyl aniline for mild steel in sulfuric acid

  • Meenakshi Gupta a ,
  • Mansi Y. Chaudhary b ,
  • Neeta Azad a ,
  • Shramila Yadav , b, *
Expand
  • a Department of Chemistry, Atma Ram Sanatan Dharma College, University of Delhi, Delhi 110021, India
  • b Department of Chemistry, Rajdhani College, University of Delhi, New Delhi 110015, India
  • 3 Keywords: NNDBA, DFT, Double-layer capacitance, Gravimetric method, Adsorption Isotherm
* Correspondence to: Department of Chemistry, Rajdhani College, University of Delhi (North Campus), New Delhi 110015, India. E-mail address: (S. Yadav).

Publishing services by Elsevier B.V. on behalf of KeAi Communications Co. Ltd. This is an open access article under the CC BY license (http://creativecommons.org/licenses/by/4.0/).

Received date: 2025-06-20

  Revised date: 2025-07-20

  Accepted date: 2025-07-29

  Online published: 2025-07-29

Copyright

3050-628X/© 2025 INTERNATIONAL SCIENCE ACCELERATOR PTY LTD.

Highlights

• The corrosion inhibition performance of N,N-dibutylaniline (NNDBA) was comprehensively evaluated using qualitative observations with quantitative techniques.

• Adsorption studies revealed that NNDBA adsorption on the mild steel surface follows the Langmuir isotherm model, indicating monolayer adsorption behavior.

• NNDBA functions as a mixed-type corrosion inhibitor, influencing both anodic and cathodic reactions.

• Molecular dynamics simulations were employed to investigate the interaction mechanism and spatial configuration of NNDBA on the mild steel surface.

Abstract

This study explored the prevention of mild steel (MS) corrosion in sulfuric acid through deployment of N, Ndibutylaniline (NNDBA) as an organic inhibitor. The inhibition efficacy was meticulously scrutinized using a blend of qualitative and quantitative methodologies. The gravimetric method was executed across a spectrum of NNDBA concentrations (10-1M-10-7M ) and temperatures, 298 K-328 K (with 10 K increments), facilitating an intricate kinetic and thermodynamic exploration of the inhibition mechanism. Adsorption isotherm analyses affirmed NNDBA's adherence to Langmuir's model, signifying a monolayer adsorption paradigm. The adsorption process was found to be spontaneous and thermodynamically favorable, predominantly governed by physisorption. Empirical data delineated an inverse relationship between temperature and inhibition efficiency, whereas an augmentation in NNDBA concentration bolstered corrosion resistance. Potentiodynamic Polarisation (PDP) confirmed that NNDBA is a mixed-type inhibitor with a maximum efficiency of 92.4%. Electrochemical Impedance Spectroscopy (EIS) measurements revealed a marked decrement in the double-layer capacitance at the Fe/H2SO4 interface, corroborating inhibitor adsorption. Notably, the lower value of the phase shift exponent n for NNDBA suggests increased surface heterogeneity due to inhibitor film formation. Scanning Electron Microscopy (SEM-2D) and Atomic Force Microscopy (AFM-3D) unveiled distinct morphological alterations indicative of surface passivation. Density functional theory (DFT) calculations provided insights into the electronic structure of NNDBA, revealing a highly negative EHOMO , low ELUMO , and a small ΔE(5.24eV), all suggesting strong reactivity and the formation of a stable metal-inhibitor complex. The mechanistic pathway and spatial orientation of the interaction between the NNDBA molecule and MS surface were explored through molecular dynamics simulation to gain insights into its inhibitory behavior. Thus, the theoretical insights harmonize with experimental findings, substantiating its efficacy as a potent corrosion mitigant.

Cite this article

Meenakshi Gupta , Mansi Y. Chaudhary , Neeta Azad , Shramila Yadav . Quantitative and qualitative insights into the corrosion mitigation mechanism of N,N-dibutyl aniline for mild steel in sulfuric acid[J]. Extreme Materials, 2025 , 1(3) : 44 -60 . DOI: 10.1016/j.exm.2025.07.001

Introduction

Corrosion is a pervasive issue, significantly affects the longevity and performance of materials, particularly metals. Mild steel, valued for its superlative mechanical attributes and commendable economic viability, is ubiquitously employed in construction, infrastructure, and various industrial applications. However, it's high susceptibility to corrosion leads to material degradation, compromised structural integrity, and increased maintenance costs. Mild steel corrosion primarily occurs through electrochemical processes, [1,2] where the metal reacts with environmental factors such as moisture, oxygen, and electrolytes. This phenomenon may materialize in a plethora of morphologies, encompassing uniform degradation, pitting phenomena, and crevice-induced corrosion, each presenting unique challenges for mitigation. Understanding the mechanisms of corrosion in mild steel is essential for developing effective prevention strategies, such as protective coatings, alloying, and cathodic protection.
Use of inhibitors represent a promising approach to mitigating corrosion in mild steel, [3] providing effective and adaptable solutions across various industries, including oil and gas [4], marine, chemical processing [5], construction, automotive [6], mining, aerospace [7], etc. these industries utilize corrosion inhibitors to enhance safety, reduce maintenance costs, and extend the lifespan of metal components and structures. Corrosion of mild steel not only leads to material degradation but also results in profound pecuniary detriments on a global scale. According to the National Association of Corrosion Engineers (NACE), metal corrosion across various industries accounts for approximately 3.4% of the world GDP annually. The application of effective corrosion inhibitors has the potential to substantially mitigate these economic losses [8,9]. Ongoing research into new formulations and environmentally friendly alternatives continues to enhance their applicability, contributing to improved longevity and performance of mild steel structures in corrosive environments [10].
Effective corrosion inhibitors are typically organic compounds [11-15] featuring π bonds in conjugation and/or heteroatoms such as phosphorus (P), sulfur (S), nitrogen (N), and oxygen (O). These chemical entities augment corrosion resilience by adhering to metallic substrates [16,17], engendering protective stratifications that obstruct electrochemical processes. Among these, nitrogen-containing compounds, amine derivatives have shown significant potential because of their strong adsorption properties and ability to form stable complexes with the metal surface. Their structural characteristics allow for strong interaction with the metal, making them valuable in protecting against corrosion in various industrial settings. These inhibitors often exhibit high efficiency and low toxicity, making them a viable choice for industrial applications.
In recent years, a significant body of research has investigated the application of amine derivatives for corrosion inhibition in acidic environments. For example, Poly (aniline-formaldehyde) [18] demonstrated inhibition efficiency (IE) exceeding 90% at 1 ppm, reaching a maximum of 98.75% at 10 ppm. Impedance and temperature studies confirmed an adsorption-based inhibition mechanism, while potentiodynamic polarization indicated its mixed-type inhibitor behavior. 4-Chloro-N-(pyridin-2-ylmethyl) aniline (CPYA) [19] was synthesized and evaluated as a corrosion inhibitor for MS in acidic medium, exhibiting a maximum IE of 96.0% at a concentration of 4.59mmol/L. Comprehensive analyses of corrosion kinetics and thermodynamic parameters were conducted to elucidate the inhibition mechanism. In a different study by M.E. Belghiti and coauthors [20], Benzylidene-aniline derivatives (NCF and NCCM) were studied via polarization and EIS techniques. NCF was found to show an IE of 89.72% at 10-3M. Langmuir adsorption isotherm was followed, with thermodynamically favorable processes, and both inhibitors were identified as mixed-type.
Many organic compounds have been recommended as effective corrosion inhibitors in the oil and gas industry, particularly under highly acidic conditions. Abdelkarim Ait Mansour and coworkers investigated environmentally benign organic compounds from the isonicotinohydrazide family [21,22] as potential eco-friendly and nontoxic corrosion inhibitors for N80 carbon steel (N80CS) in a 15wt\%HCl solution. Notably, heterocyclic derivatives such as isatin-hy-drazones-including (E)-1-octyl-3-[2-(5-oxo-4,4-diphenyl)-4,5-di-hydro-1H-imidazol-2-yl)hydrazono]indolin-2-one (OPHIHI) and (E) - 3-[2-(5-oxo-4,4-diphenyl) - 4,5-dihydro-1H-imidazol-2-yl)hy-drazono]indolin-2-one (OIHIHI) [23] have demonstrated promising inhibitory properties. Similarly, Schiff base compounds such as N-(4-methoxyphenyl)-1-(1H-pyrrol-2-yl)methanimine (MPM), 1-(furan-2-yl)- N-(4-methoxyphenyl)methanimine (FMM), and N-(4-methox-yphenyl)-1-(thiophen-2-yl) methenamine (MTM) [24] have been studied for their corrosion inhibition performance on mild steel in 0.5 M H2SO4. Among these, MTM exhibited a maximum inhibition efficiency of 97.93% at a concentration of 250mgL-1, as measured by electrochemical methods. These studies underscore the relevance of nitrogencontaining heterocycles and their derivatives as potent corrosion inhibitors in acidic media, reinforcing the rationale for investigating NNDBA in the present work.
Aniline, p-toluidine, p-anisidine, p-phenitidine and N, N- dimethyly aniline have been found to exhibit inhibition efficiencies ranging from 72.7 % to 94.9 % as inhibitors for corrosion mitigation of mild steel in hydrofluoric acid (HF) [25]. Extensive research has been conducted on the corrosion behavior of mild steel in acidic environments using structurally analogous corrosion inhibitors. Notable examples include aniline [26], 4-acetamidoantipyrine [27], 3-(1,3-oxazol-5-yl) aniline [28], F-EN (ethyl (2Z)-3-[(4-fluorophenyl)-amino]-but-2-enoate) [29], as well as p-toluidine, o-aminophenol, anthranilic acid, and o-phenylenediamine [30].
Fig. 1. Structure of N, N Dibutyl aniline (NNDBA).
In this study, the corrosion IE of N, N-dibutyl aniline for mild steel in a sulfuric acid medium is evaluated. N, N-Dibutyl aniline, an aromatic amine derivative, its molecular structure (Fig. 1), featuring an electron-donating aromatic ring and nitrogen-based functional groups, is expected to enhance adsorption and protect the steel surface by blocking active corrosion sites. The work involves coupon studies and electrochemical technique to assess the inhibition performance. Furthermore, adsorption characteristics and thermodynamic variables are scrutinized to elucidate the fundamental inhibition paradigm, providing insights into the applicability of N,N-dibutyl aniline as an effective corrosion inhibitor. This study advances the development of advanced corrosion inhibitors and provides insights into sustainable solutions for industrial corrosion challenges.

Materials and experimental methods

Materials

In gravimetric analysis, mild steel cubes of 1 cm×1 cm×1 cm were used while for the EIS study, mild steel of active area of 1 cm2 was submitted for corrosion in 0.5 M sulfuric acid (AR Grade) alone and in the presence of various concentrations of compound (NNDBA) in 0.5H2SO4 solution. NNDBA of Sigma-Aldrich (assay 97 %), a colourless liquid, was used. Mild steel of composition C: 0.15%,Mn:1.02%, S:0.0255%,P:0.25%,Si:0.8%, and remainder Fe was used. Prior to submersion in the corrosive testing media, mild steel specimens underwent abrasion in accordance with the ASTM G1-03 standard, which outlines the prescribed procedures for the preparation and cleaning of corrosion test specimens [31,32]. The specimens were subsequently rendered free of grease using acetone and thereafter desiccated within a desiccator.

Inhibitor

To ensure effective interaction between the inhibitor and the mild steel surface, the solubility of N,N-dibutylaniline (NNDBA) in 1MH2SO4 was considered. NNDBA contains a nitrogen atom with a lone pair of electrons, which can readily undergo protonation in acidic media, forming a positively charged ammonium species. This protonated form is ionic in nature, thereby enhancing the solubility of NNDBA in the acidic solution. During solution preparation, no turbidity, precipitation, or phase separation was observed across all concentrations studied (10-1M to 10-7M ), indicating that NNDBA was fully soluble and remained homogeneously dispersed throughout the experiments.

Experimental methods

Gravimetric studies: Meticulously polished and desiccated mild steel coupons were precisely weighed using a Mettler Balance before being immersed in diverse solutions (acid, 10-1M,10-3M,10-5M,10-7M acid-compound solution) for a duration of six hours at controlled temperatures from 298 K-328 K. Upon completion of the six-hour immersion period, the coupons were extracted from their respective solutions and rigorously cleansed with distilled water to eliminate all corrosion residues, then dried in a desiccator for 24 h, and finally weighed to obtain weight loss of the specimens.
Electrochemical studies: Potentiodynamic polarization and electrochemical impedance studies were conducted using a CHI760C analyzer. A conventional three-electrode system was employed, consisting of a platinum electrode as the auxiliary electrode, a calomel electrode as the reference electrode, and a mild steel electrode as the working electrode.
Surface studies: SEM, 2-D surface morphology of pit-free polished coupon was done at 298 K on corrosion exposure to acid, at the highest concentration (10-1M ) and the lowest concentration (10-7M ) NNDBA solution for 24 h by JEOL-840. AFM, 3-D surface morphology of crack-free polished surface of mild steel after exposure to acid, 10-1 M NNDBA solution for 24 h was done by VEECO CP II scanning probe microscope.
Quantum chemical calculations: The geometry of the NNDBA molecule was refined via Density Functional Theory (DFT) computations executed using the DMol3 module within the Materials Studio 2020 suite. The hybrid B3LYP exchange-correlation functional, combined with the double numeric plus polarization (DNP) basis set, was employed to elucidate the frontier molecular orbital energies and achieve an energetically favorable structural conformation.
Molecular dynamic simulation (MD Simulation): The adsorption behavior of NNDBA on the mild steel surface was modeled using the Simulated Annealing method coupled with the SMART algorithm, implemented through the adsorption locator tool in the materials studio 2020 suite. Post-adsorption, a corrosive environment was simulated by introducing water, H+, and SO4 2- ions. MD simulations were performed in the Forcite module using the B3LYP forcefield with Ewald summation for electrostatics and van der Waals interactions. The system was equilibrated under an NVT ensemble at 298 K using a Nosé thermostat and random initial velocities, running for 500 ps. These parameters enabled precise evaluation of ligand-metal interactions in a simulated corrosive medium.

Results and discussions

Gravimetric method

In an acidic medium, MS undergoes anodic dissolution, accompanied by a cathodic reaction involving the reduction of hydrogen ions (H+)to hydrogen gas (H2), as widely reported in the literature [33,34]:
Fe→Fe2++2e-
2H++2e-→H2
Thus, the weight loss of metal in an acidic medium is a direct consequence of these 2 reactions, along with the high solubility of corrosion products. The gravimetric method is a widely used approach for studying corrosion in which the weight loss of a metal specimen immersed in a corrosive medium is measured over a defined period. In the present investigation, MS coupons were subjected to immersion in sulfuric acid, in presence and absence of NNDBA, inhibitory agent for 6 h at studied temperatures: 298 K,308 K,318 K, and 328 K. The percentage IE, surface coverage (θ ), and corrosion rate (Cr ) were calculated using following equations [35]:
$\begin{array}{c}IE\%=\left(1-\frac{{w}_{i}}{{w}_{o}}\right)\times 100\end{array}$
$\begin{array}{c}\theta =\left(1-\frac{{w}_{i}}{{w}_{o}}\right)\end{array}$
wo and wi are the mass depletion of mild steel in the acidic solution and inhibitor-containing solution, respectively. The corrosion kinetics of mild steel in 0.5MH2SO4, both with and without diverse concentrations of the formulated additives, was determined from weight loss data using the standard calculation method [36,37]:
$\begin{array}{c}{C}_{r}=\frac{87.6{W}_{L}}{At\rho }\end{array}$
Cr is the corrosion rate of the mild steel (in mm/y), WL is the mass depletion of the mild steel (in mg ) and t is the immersion time (6 h ), ρ is the density of the mild steel (7.06 g/cm3 ), 87.6 is constant for corrosion process and A is exposed area of MS specimen (equals to 1 cm2 ).
The findings, listed in Table 1 indicated that the coupon experienced the highest weight loss in sulfuric acid in the absence of the inhibitor and at elevated temperatures, signifying an accelerated corrosion rate. In contrast, the presence of NNDBA led to a marked reduction in weight loss, underscoring its efficacy as a corrosion inhibitor. NNDBA demonstrated optimal inhibition performance at the lowest temperature (298 K ) and the highest concentration (10-1M ), where the weight loss was reduced to a mere 0.0041 g. However, an increase in temperature and a corresponding decrease in NNDBA concentration resulted in a subsequent rise in weight loss. The inhibition efficiency and surface coverage both exhibited a positive correlation with increasing NNDBA concentrations across all temperature conditions (Fig. 2), denoting that the inhibitor promotes the establishment of a continuous protective stratification upon the metallic substrate. This protective layer efficaciously sequesters the underlying substrate from pernicious anionic entities and attenuates the electrochemical redox phenomena that propagate the corrosion mechanism [38]. In contrast, the inhibition efficiency and surface coverage decreased as the temperature increased, signifying that at elevated temperatures, the dissolution kinetics of mild steel become predominant. This phenomenon can be ascribed to the attenuation of the adsorption process under heightened thermal conditions, suggesting that the inhibition mechanism is predominantly governed by physisorption [39].
Table 1 Temperature and concentration dependent corrosion rate and inhibition efficiency of mild steel in H2SO4 medium in the presence and absence of NNDBA determined via gravimetric method.
Temp. (K) Conc. of NNDBA (M) Initial weight Iw (g) Final weight Fw (g) Weight loss (g) IE (%) Θ K1⊕10-4 (hr-1 ) t1/2 (hr) Cr (mm/y)
298 10-1 5.2486 5.2445 0.0041 94.8 0.948 1.30 5319.81 8.48
10-3 5.9763 5.9632 0.0131 83.4 0.834 3.66 1894.48 27.09
10-5 5.1042 5.0854 0.0188 76.1 0.761 6.15 1126.61 38.88
10-7 6.9352 6.9099 0.0253 67.9 0.679 6.09 1137.50 52.32
Acid 7.4629 7.3841 0.0788 - - 17.69 391.64 162.96
308 10-1 5.8761 5.8645 0.0116 86.4 0.864 3.29 2103.82 23.99
10-3 6.4962 6.4819 0.0143 83.3 0.833 3.67 1886.48 29.57
10-5 6.6329 6.6097 0.0232 72.9 0.729 5.84 1186.48 47.98
10-7 5.9756 5.9436 0.032 62.6 0.626 8.95 774.23 66.18
Acid 7.3616 7.2761 0.0855 - - 19.47 355.86 176.81
318 10-1 5.9672 5.9295 0.0377 85.5 0.855 10.57 655.93 77.96
10-3 6.883 6.8332 0.0498 80.8 0.808 12.10 572.50 102.99
10-5 6.9231 6.8356 0.0875 66.2 0.662 21.20 326.84 180.95
10-7 6.463 6.3487 0.1143 55.9 0.559 29.74 232.98 236.37
Acid 7.0705 6.8113 0.2592 - - 62.26 111.31 536.02
328 10-1 6.8435 6.7953 0.0482 84.1 0.841 11.78 588.17 99.68
10-3 5.9748 5.8689 0.1059 65.1 0.651 29.81 232.46 219.00
10-5 5.8795 5.7479 0.1316 56.7 0.567 37.74 183.65 272.15
10-7 5.3298 5.1887 0.1411 53.5 0.535 44.73 154.94 291.79
Acid 7.6199 7.3162 0.3037 - - 67.80 0.0102 628.05
Fig. 2. Effect of NNDBA concentration on corrosion protection performance of mild steel at all studied temperatures.

Kinetic Study

Kinetic studies of the corrosion process exhibit characteristics akin to diffusion, where elevated temperatures reduce the concentration of inhibitor molecules on the metal surface, while lower temperatures enhance the adsorption of inhibitor molecules. This underscores the significance of kinetic analysis in comprehensively understanding and optimizing the corrosion inhibition mechanism. In this kinetic study, various kinetic parameters were examined, including activation energy, rate constant, and half-life data for mild steel in 0.5H2SO4.
The weight loss of MS in 0.5H2SO4 at different time intervals was investigated and tabulated in Table 2.
The mitigation of MS corrosion in H2SO4 is mainly driven by the dissolution of metal ions, exhibiting a first-order dependence on the unreacted metal surface area. For the kinetic study, the first-order rate equation is expressed as follows [40]:
$\begin{array}{c}ln\Delta W=-{k}_{1}t+ln{F}_{w}\end{array}$
where ΔW is weight loss, k1 is the first-order rate constant, Iw represents the initial weight of the coupon before immersion, Fw denotes the final weight of the coupon after immersion, and t is the duration for which the metal was exposed to the solution.
Table 2 Time-dependent mass depletion data for mild steel corrosion in H2SO4 solution.
Time (hrs) Initial weight Iw(g) Final weight Fw(g) Weight loss in H2SO4Δ W( g) ln(ΔW/g)
2 6.3940 6.3860 0.0080 -4.83
4 7.1260 7.1030 0.0230 -3.77
6 7.0490 7.0050 0.0440 -3.12
8 6.8450 6.7640 0.0810 - 2.51
Fig. 3. Time-dependent corrosion behavior of mild steel in H2SO4.
The first-order nature of MS corrosion in H2SO4 is confirmed by the ln (ΔW ) vs. time plot (Fig. 3), which supports the linearity of the data, with a correlation coefficient close to unity (R2=0.9808 ). Since MS follows first-order rate kinetics in 0.5H2SO4, the rate constant was calculated [41] using Eq. 5:
$\begin{array}{c}{k}_{1}=\frac{2.303}{t}log\frac{{I}_{w}}{{F}_{w}}\end{array}$
The half-life period (t1/2 ), defined as the time required for half of the reactant to deteriorate, is determined using Eq. 6.
$\begin{array}{c}{t}_{1/2}=\frac{0.693}{{k}_{1}}\end{array}$
The half-life is longer in the presence of an inhibitor, aligning with the findings of other researchers [42,43]. According to their studies, an effective inhibitor is one that extends the time required for the metal to convert into its corrosion products. Table 1 revealed that the rate constant declines with higher inhibitor concentrations but shows an increasing trend with temperature elevation. The elevated temperatures increase the kinetic energy of the inhibitor molecules adsorbed on the mild steel surface, thereby accelerating their desorption from the metal and promoting a faster corrosion rate and a higher rate constant [44].
Analogously, with an escalation in inhibitor concentration, a higher density of molecules becomes available for adsorption onto the mild steel substrate, thereby significantly suppressing the corrosion reaction and diminishing its rate constant. The half-life analysis further corroborates this observation [45].

Energy of Activation

The rates of MS corrosion were evaluated at various temperatures, both in absence and presence of different concentrations of NNDBA. These data were used to determine the activation energy (Ea ) for the metal dissolution process. The apparent activation energy of corrosion in the acidic medium was considered using the Arrhenius Equation, expressed as [46]:
$\begin{array}{c}log{C}_{r}=\frac{{E}_{a}}{2.303RT}+log\lambda \end{array}$
Cr is the corrosion rate in mm/y, Ea is the activation energy kJ/mol, λ is the Arrhenius pre-exponential, R is the gas constant expressed in J/K/mol and T is the absolute temperature in K.
Ea and λ values were extrapolated from the gradient and intercept of the Arrhenius plots (Fig. 4), which delineate the correlation between logCr and 1/T. These computed values are systematically presented in Table 3. The analytical data illustrate that Ea associated with the corrosion of mild steel in 0.5MH2SO4 exhibits a substantial increase in the presence of the NNDBA inhibitor relative to the uninhibited acidic medium. This pronounced elevation in activation energy signifies that NNDBA efficiently adheres to the metallic substrate, engendering a protective stratification that functions as a formidable barrier to both charge and mass transfer processes. Consequently, the overall corrosion kinetics experience a substantial diminution.
Fig. 4. Temperature dependence of corrosion rate: logCr vs. 1/T for mild steel in H2SO4 medium in the presence and absence of NNDBA.
Table 3 Energy of activation (Ea ) for mild steel corrosion in H2SO4 medium in the presence and absence of NNDBA.
S. No. Concentration (M) Ea(kJ/mol)
1 10-1 70.04
2 10-3 60.71
3 10-5 58.09
4 10-7 53.49
5 Acid 41.84
An increase in temperature leads to higher corrosion rates in both the presence and absence of the inhibitor. However, as the concentration of NNDBA increases, corrosion rates decrease. This is explained by the improved adsorption of the inhibitor on the MS surface at higher concentrations, which blocks corrosion-prone areas and increases the activation energy required for the corrosion process.
The augmented activation energy (Ea ) observed in the inhibitorcontaining medium, relative to the uninhibited solution, signifies that the inhibition mechanism involves physical adsorption. Activation energy (Ea ) values exceeding 80 kJ/mol generally suggest chemisorption, while those below this threshold point to physisorption. The observed Ea thus confirms the formation of a physically adsorbed protective film by the inhibitor on the mild steel surface [47]. The outcomes are shown in Table 3, where the Ea values remain below 80 kJ/mol, denoting a physisorption-dominated mechanism. Hence, empirical data substantiate that NNDBA undergoes physisorptive adherence to the mild steel substrate, effectively mitigating its corrosive degradation. Ea values exhibited a sequential augmentation, escalating from 41.84 kJ/mol in the uninhibited medium to 70.04 kJ/mol at the maximal inhibitor concentration (10-1M ) within the acidic environment. The discernible decline in inhibition efficacy with rising temperature in the presence of the inhibitor can be ascribed to the attenuation of physisorptive interactions under elevated thermal conditions [48].

Adsorption isotherm

Temperature exerts a profound influence on the kinetics of corrosion reactions. An elevation in temperature amplifies the corrosion rate due to enhanced ionic mobility, augmented diffusion coefficients, and the intensified activation of interfacial electrochemical reactions. Consequently, the efficacy of corrosion inhibitors is also temperature-contingent, as their adsorption dynamics and interactive affinity with the metallic substrate are inherently thermosensitive. A comprehensive evaluation of the most appropriate adsorption isotherm, alongside an analysis of the thermodynamic parameters governing the corrosion process, is imperative for elucidating the mechanistic pathways of inhibitor adsorption. These parameters serve to delineate the spontaneity of the inhibitory reaction, the energetics of the adsorption process, and the structural robustness of the protective inhibitor-derived film, thus offering critical insights into the stability and efficacy of the corrosion mitigation strategy.
The identification of the nature of adsorption usually relies on plotting various adsorption isotherms (Langmuir, Freundlich, Temkin, El-Awady, Flory Huggins and Frumkin) in order to identify the most appropriate model as illustrated in Fig. 5 [49].
Table 4 shows that the Langmuir Adsorption Isotherm follows as R2=0.9999, which presupposes that the adsorbed entities inhabit singular surface sites exclusively, devoid of any mutual interactions, for elucidating the adsorption behavior of the inhibitor on the metal surface.

Thermodynamic Parameters and their Implications

The Langmuir model was employed to calculate thermodynamic parameters, offering deeper insight into the nature of the adsorption process and its impact on corrosion inhibition. According to Equation 8, a plot of log(c/θ) versus log(c) yields a straight line, where the intercept corresponds to the adsorption equilibrium constant (Kads  ) at different temperatures.
The Kads  values (as presented in Table 5) exhibited a declining trend with increasing temperature, signifying robust adsorption of the NNDBA onto the MS interface at lower thermal conditions. However, as the temperature escalates, the adsorbed NNDBA molecules demonstrate a propensity to disengage from the MS surface, indicating a reduction in adsorption stability at elevated temperatures [50].
Gibb's free energy of adsorption ($\mathrm{\Delta }{G}_{\text{ads}\text{ }}^{\circ }$) value at temperature range 298 K-328 K was determined according to Eq. 14.
$\begin{array}{c}\Delta {G}_{ads}^{\circ }=-2.303RTlog\left(55.5\mathrm{ }{K}_{ads}\right)\end{array}$
R is the gas constant (8.314 J/mol ), T is the temperature (K ), and 55.5 is the standard molar concentration of water in the solution [51].
Broadly, four distinct modalities of adsorptive interactions may transpire at the metal-inhibitor interface [52]:
▪ Electrostatic attraction arises between the charged metallic substrate and ionized inhibitor molecules. ▪ Coordination via lone electron pairs residing on heteroatoms within the inhibitor molecule, facilitating interaction with the metal. ▪ π-electron delocalization, wherein the π electron cloud of the inhibitor engages with the metallic surface. ▪ A synergistic amalgamation of two or more of the aforementioned mechanisms.
The initial scenario typifies the physisorption mechanism, whereas the subsequent two correspond to chemisorption. The final case embodies a hybridized adsorption process, encompassing both physical and chemical adsorption phenomena. Thus, the type of adsorption can be decided by thermodynamic parameters.
The variation in Gibb's free energy of adsorption ( $\mathrm{\Delta }{G}_{\text{ads}\text{ }}^{\circ }$ ), assessed across multiple thermal conditions, is delineated in Table 5. The thermodynamic parameter $\mathrm{\Delta }{G}_{ads}^{\circ }$ with magnitudes equal to or below -20 kJ/mol signifies electrostatic interactions, indicative of physisorption, whereas values exceeding -40 kJ/mol denote the establishment of coordinate covalent bonding, characteristic of chemisorption [53]. The entirety of the obtained $\mathrm{\Delta }{G}_{\text{ads}\text{ }}^{\circ }$ values were negative and remained below 20 kJ/mol, signifying the thermodynamic viability and spontaneity of NNDBA adsorption onto the mild steel substrate, predominantly governed by a physisorption mechanism.
Fig. 5. Different adsorption isotherms for mild steel in H2SO4 medium in the presence and absence of NNDBA (a) Langmuir, (b) Freundlich, (c) Temkin, (d) ElAwady, (e) Flory-Huggins, and (f) Frumkin adsorption isotherm.
The standard Gibb's free energy, enthalpy and entropy are related as [54]:
$\begin{array}{c}\Delta {G}_{ads}^{\circ }=\Delta {H}_{ads}^{\circ }-T\Delta {S}_{ads}^{\circ }\end{array}$
By using the Van't Hoff equation, the enthalpy of adsorption ( $\mathrm{\Delta }{H}_{\text{ads}\text{ }}^{\circ }$ ) was determined as follows [55]:
$\begin{array}{c}log{\mathrm{K}}_{\mathrm{a}\mathrm{d}\mathrm{s}}=\frac{\mathrm{\Delta }{\mathrm{H}}_{\mathrm{a}\mathrm{d}\mathrm{s}}^{\circ }}{2.303\mathrm{R}\mathrm{T}}+D\end{array}$
The graphical representation in Fig. 6, depicting the correlation between 1/T and logKads , exhibits a linear trend, wherein the gradient corresponds to the enthalpy of adsorption ($\mathrm{\Delta }{H}_{\text{ads}\text{ }}^{\circ }$) value. In the present study, $\mathrm{\Delta }{H}_{\text{ads}\text{ }}^{\circ }$ was determined to be -3.4 kJ/mol, indicative of an exothermic process wherein thermal energy is dissipated as inhibitor molecules affix themselves onto the metallic substrate. This thermodynamic parameter substantiates the spontaneity and favorable nature of the adsorption phenomenon. The enthalpy value, being below 10 kJ/mol, further corroborates the predominance of physisorption as the governing adsorption mechanism for NNDBA on the mild steel interface [56]. Physisorption is characterized by the involvement of weak van der Waals interactions or electrostatic attractions between the inhibitor molecules and the metal surface, rendering the process reversible and susceptible to thermal perturbations. As the system's temperature elevates, the weakly adsorbed inhibitor species exhibit an enhanced propensity for desorption, thereby attenuating the overall inhibition efficacy [57].
Table 4 Analysis of adsorption behavior using different isotherm models.
Isotherm Isotherm Equations Temp. (K) Equation R2
Langmuir $\mathrm{l}\mathrm{o}\mathrm{g}\frac{c}{\theta }=\mathrm{l}\mathrm{o}\mathrm{g}c-\mathrm{l}\mathrm{o}\mathrm{g}{K}_{ads}$ (eq. 8) 298 y=0.9763x+0.0023 1
308 y=0.9761x+0.0253 1
318 y=0.9680x+0.0200 1
328 y=0.9675x+0.0651 0.9999
Freundlich $\mathrm{l}\mathrm{o}\mathrm{g}\theta =\mathrm{l}\mathrm{o}\mathrm{g}{K}_{ads}+\frac{1}{n}\mathrm{l}\mathrm{o}\mathrm{g}c$ (eq. 9) 298 y=0.0237x-0.0023 0.9963
308 y=0.0239x-0.0253 0.9430
318 y=0.0320x-0.0200 0.9588
328 y=0.0325x-0.0651 0.9188
Temkin $\theta =\frac{-2.303\mathrm{l}\mathrm{o}\mathrm{g}{K}_{ads}}{2a}-\frac{2.303\mathrm{l}\mathrm{o}\mathrm{g}c}{2a}\mathrm{ }$ (eq.10) 298 y=4.4x+98.15 0.9903
308 y=4.09x+92.66 0.9552
318 y=5.17x+92.78 0.9676
328 y=5.01x+84.89 0.8871
Flory-Huggins $\mathrm{l}\mathrm{o}\mathrm{g}\frac{\theta }{c}=\mathrm{l}\mathrm{o}\mathrm{g}{K}_{FH}+{x}_{FH}\mathrm{l}\mathrm{o}\mathrm{g}(1-\theta )$ (eq. 11) 298 y=6.8517x+9.3482 0.8874
308 y=12.501x+12.123 0.9779
318 y=11.166x+10.501 0.9769
328 y=10.479x+8.9185 0.8040
El-Awady $\mathrm{l}\mathrm{o}\mathrm{g}\frac{\theta }{1-\theta }=\mathrm{l}\mathrm{o}\mathrm{g}{K}^{\text{'}}+\mathrm{y}\mathrm{l}\mathrm{o}\mathrm{g}\mathrm{c}$ (eq. 12) 298 y=0.1502x+1.2984 0.9413
308 y=0.1003x+0.9398 0.9753
318 y=0.1168x+0.9144 0.9790
328 y=0.1071x+0.7213 0.8473
Frumkin $\mathrm{l}\mathrm{n}\frac{\theta }{(1-\theta )c}=\mathrm{l}\mathrm{n}{K}_{ads}+2a\theta \mathrm{ }$ (eq. 13) 298 y=-43.807x+46.103 0.9771
308 y=-48.207x+47.232 0.9480
318 y=-37.939x+37.595 0.9610
328 y=-35.748x+33.068 0.8531
Table 5 Thermodynamic insights from the Langmuir adsorption isotherm at all studied temperatures.
Temp (K) Kads(M-1) $\mathrm{\Delta }{\mathrm{G}}_{\text{ads}\text{ }}^{\circ }$ (kJ/mol ) $\mathrm{\Delta }{\mathrm{H}}_{\mathrm{a}\mathrm{d}\mathrm{s}}^{\circ }$ (kJ/mol ) $\mathrm{\Delta }{\mathrm{S}}_{\text{ads}\text{ }}^{\circ }(\mathrm{J}/$ K/mol)
298 0.995 -9.9 -3.4 21.95
308 0.943 -10.1 21.88
318 0.955 -10.5 22.33
328 0.861 -10.6 21.79
Fig. 6. Temperature dependence of adsorption equilibrium constant: logKadsvs.1/T.
The entropy of adsorption ($\mathrm{\Delta }{S}_{\text{ads}\text{ }}^{\circ }$ ), as ascertained through Eq. 15, is quantified at 21.99 J/K/mol. This parameter elucidates that the adherence of inhibitor molecules onto the metallic substrate engenders an augmentation in disorder or molecular randomness, signifying that the resultant adsorbed layer exhibits a diminished degree of structural organization relative to the uninhibited inhibitor species dispersed in solution. Such an entropy-driven phenomenon is generally perceived as thermodynamically advantageous in the realm of corrosion mitigation, as it implies the spontaneous and facile establishment of a protective molecular barrier over the metal surface, thereby enhancing the inhibitor's efficacy in curtailing corrosive degradation [58].
Fig. 7. Cathodic and anodic polarization run via PDP studies for various concentrations of H2SO4 in the presence and absence of NNDBA at 298 K.

Potentiodynamic polarization studies (PDP)

Potentiodynamic polarization studies were conducted with NNDBA concentrations of 10-1M,10-3M,10-5M, and 10-7M at 298 K. Potential values were plotted against the logarithm of current (Icorr  ). Before each cathodic and anodic polarization run, the open circuit potential (OCP) was measured against the saturated calomel electrode. Key electrochemical parameters, including corrosion potential (Ecorr  ), corrosion current (Icorr  ), cathodic slopes (bc ), and anodic slopes (ba ) were determined by the Tafel plots for various inhibitor concentrations, with acid plots overlaid at 298 K (shown in Fig. 7) and data summarized in Table 6. IE and θ of NNDBA were figure out by using Eqs. 17,18:
Table 6 Corrosion current and inhibition efficiency for mild steel in H2SO4 medium in the presence and absence of NNDBA determined via PDP studies at 298 K.
Concentration (M) - Ecorr  (mV) βc (mV/dec) βa (mV/dec) Icorr  (mA/cm2 ) IE (%) θ
10-1 500.0 172.8 100.0 0.0692 94.2 0.942
10-3 476.9 138.5 76.9 0.1514 87.4 0.874
10-5 492.3 138.5 100.0 0.2818 76.5 0.765
10-7 493.1 117.3 78.8 0.3311 72.5 0.725
Acid 465.8 106.7 120.0 1.2021 -
Fig. 8. Bode plot of mild steel in H2SO4 and in 10-1M NNDBA at 298 K.
$\begin{array}{c}IE\left(\%\right)=\left(\frac{{\mathrm{I}}_{\mathrm{H}2\mathrm{S}\mathrm{O}4}-{\mathrm{I}}_{\mathrm{N}\mathrm{N}\mathrm{D}\mathrm{B}\mathrm{A}}}{{\mathrm{I}}_{\mathrm{H}2\mathrm{S}\mathrm{O}4}}\right)\times 100\end{array}$
$\theta =\left(\frac{{\mathrm{I}}_{\mathrm{H}2\mathrm{S}\mathrm{O}4}-{\mathrm{I}}_{\mathrm{N}\mathrm{N}\mathrm{D}\mathrm{B}\mathrm{A}}}{{\mathrm{I}}_{\mathrm{H}2\mathrm{S}\mathrm{O}4}}\right) $
IH2SO4 and INNDBA  are corrosion currents in acid and inhibitor-acid solution respectively.
The potentiodynamic polarization plots (Fig. 7) in this study demonstrate that NNDBA molecules act as effective corrosion inhibitors, significantly influencing corrosion kinetics parameters. Table 6 indicates a significant reduction in Icorr  values for the corrosive solution containing inhibitor molecules compared to the uninhibited solution. This effect becomes more pronounced with increasing inhibitor concentration. The data further revealed that the introduction of NNDBA inhibitors into H2SO4 solution leads to a substantial decrease in Icorr  values, underscoring the impact of the inhibitor in minimizing corrosion [59].
The addition of NNDBA molecules induced a noticeable shift in both cathodic and anodic polarization plots. The analysis of shifts in the obtained corrosion potential (Ecorr  ) offered valuable insights into the anodic or cathodic characteristics of the electrochemical processes involved. This assessment enabled the classification of inhibitor molecules as anodic, cathodic, or mixed-type, based on their influence on Ecorr . Typically, an inhibitor is designated as anodic or cathodic when the displacement in Ecorr  between the inhibited and uninhibited systems exceeds 85 mV [60].
The highest shift in Ecorr  is found to be 34.2 mV, which categorized NNDBA a mixed type inhibitor. Furthermore, the presence of NNDBA leads to a greater increase in the cathodic slope compared to the anodic slope, suggesting that NNDBA more effectively suppresses hydrogen evolution than anodic dissolution. This indicates that, while NNDBA functions as a mixed-type inhibitor, it exhibits a slightly stronger cathodic inhibition effect [61]. The irregular variations in βc and βa values imply the possible involvement of additional species in the adsorption process.
Fig. 9. Nyquist plot of mild steel in H2SO4 and in 10-1 M NNDBA at 298 K.
At a concentration of 10-1M, the inhibition efficiency was recorded at 92.4%, whereas at the lowest concentration (10-7M ) of NNDBA at 298 K, it was 72.5%. This indicates that NNDBA effectively mitigates corrosion even at low concentration as well. Furthermore, the results are consistent with the gravimetric data.

Electrochemical impedance spectroscopy (EIS)

Electrochemical Impedance Spectroscopy (EIS) was conducted to evaluate the corrosion inhibition efficacy of NNDBA in H2SO4 medium on MS. The resulting Nyquist and Bode plots, depicted in Figs. 8 and 9, respectively, illustrate the impedance response of the system. Spectral acquisition was performed at an NNDBA concentration of 10-1M in H2SO4 at a controlled temperature of 298 K. Additionally, Fig. 10 presents the equivalent electrochemical circuit model, which elucidates the inhibition mechanism of NNDBA.
A Constant Phase Element (CPE) is used in EIS modeling to represent non-ideal capacitive behaviour of the double layer at the electrode-electrolyte interface. Instead of acting like a perfect capacitor, most real systems exhibit a distribution of time constants due to surface roughness, inhomogeneity, or porous films. The impedance of a CPE is given by ZCPE  as per Eq. 19:
$\begin{array}{c}{Z}_{CPE}=\frac{1}{{Y}_{O}(j\omega {)}^{n}}\end{array}$
Where, Yo is CPE constant (in μS⋅sn⋅cm-2 ), ω is angular frequency n : exponent (0≤n≤1 ), describing how far the system deviates from ideal capacitive behavior.
Fig. 10. Equivalent Circuit.
The impedance parameters namely, double-layer capacitance (Cd1 ), charge transfer resistance (Rct ), and inhibition efficiency (IE) were calculated using the following equations [62]:
$\begin{array}{c}{C}_{dl}=\frac{1}{2\pi \mathrm{f}{R}_{ct}}\end{array}$
f represents the frequency at which the semicircle reaches its maximum height along the imaginary axis, and Rct is determined from the diameter of the semicircle. IE % was evaluated using the equation [63]:
$\begin{array}{c}IE\left(\%\right)=\left(\frac{{R}_{ct,NNDBA}-{R}_{ct,H2SO4}}{{R}_{ct,NNDBA}}\right)\times 100\end{array}$
where Rct,NNDBA and Rct,H2SO4 correspond to the charge transfer resistance in the presence and absence of NNDBA, respectively.
The Nyquist plot (Fig. 9) exhibits a characteristic semi-circular arc, with its diameter serving as a direct measure of the corrosion inhibition effect exerted by NNDBA. A larger semicircle diameter in the presence of NNDBA, as compared to the system containing only acid, indicates a surge in Rct and a concomitant diminution in Cdl [64]. This enhancement in Rct signifies the protective film of NNDBA on the mild steel surface, effectively mitigating acid-induced degradation [65]. The diminution in Cdl further substantiates that the adsorption of NNDBA at the metal-solution interface mitigates the deleterious effects exerted by corrosive ions [66]. Moreover, the perturbations in phase angles, discernible in the Bode plots (Fig. 8), furnish corroborative validation of NNDBA adsorption onto the mild steel substrate [67].
The equivalent circuit, shown in Fig. 10, comprises the Rct in series with a Constant Phase Element (CPE), Q, which is substituted for Cdl and placed in parallel with the polarization resistance (Rp ). This electrical circuit topology, substantiated by an exceptionally reduced fitting error of χ2=1.527×10-3, encapsulates the intrinsic surface heterogeneity of the mild steel specimens and delineates deviations from quintessential capacitive behaviour [68,69].
The use of CPE instead of Cdl ensures a more accurate representation of real electrochemical systems, minimizing frequency-dependent deviations. Lower CPE for NNDBA indicates more compact and less reactive surface, suggesting strong adsorption of the inhibitor and fewer electroactive sites. The observed reduction in Cdl upon NNDBA addition indicates an increase in double-layer thickness and a consequent reduction in surface activity [70]. The observed increase in the Cdl value for the inhibitor suggests thick, compact inhibitor film reducing charge accumulation. This might appear counterintuitive at first because NNDBA has a slightly higher Cdl, yet it's a better inhibitor. This anomaly could be attributed to the complexity of adsorption layer, which might include contributions from both capacitive and resistive pathways. However, this increase is relatively small and the dominating influence is the drastic rise in Rct, suggesting that the inhibitor still forms a protective barrier. A pronounced augmentation in Rp further signifies that the inhibitor mitigates corrosion either by impeding the anodic dissolution kinetics or by orchestrating the formation of a robustly adsorbed protective layer on the mild steel interface. The lower value of n for NNDBA shows greater heterogeneity due to inhibitor film formation. This heterogeneity is beneficial in this context, as it reflects an irregular but protective barrier that impedes corrosion processes. The elevated IE provides compelling substantiation for this inference (Table 7).
Furthermore, the circuit model incorporates an inductive element (L), likely indicative of relaxation processes associated with the stabilization of the adsorbed layer. Such phenomena may originate from the intricate interplay between inhibitor molecules and transient corrosion intermediates residing on the electrode interface. The presence of inductance could thus reflect complex adsorption phenomena involving NNDBA and reaction by-products, further contributing to the overall corrosion inhibition mechanism [71].

Scanning electron microscopy (SEM)

To elucidate topographical intricacies of MS coupon alone, in acid and in inhibitor solution at 298 K, SEM analysis was meticulously conducted.
Fig. 11 (a) depicts the morphological transformations induced by immersion in H2SO4 under 3000 magnifications. In the corrosive medium of 0.5MH2SO4, pronounced surface degradation ensues, manifesting as intensified pitting and exacerbated fissuring [72].
Conversely, Fig. 11 (b) delineates the morphological characteristics of the mild steel surface subjected to 10-1 M NNDBA, while Fig. 11 (c) portrays the steel exposed to 10-7M NNDBA at 3000 magnifications, respectively. The SEM micrographs of the inhibitor-treated specimen exhibit a notable absence of corrosion artifacts, highlighting the potent protective efficacy of NNDBA. The inhibitor molecules facilitate the formation of a robust passivation layer, effectively shielding the metallic substrate from the aggressive acidic environment. Remarkably, even at an ultra-low concentration of 10-7M, NNDBA exhibits significant adsorption, as evidenced by modifications in interfacial electrochemical interactions. These findings underscore NNDBA's strong adsorption capability and its role in mitigating corrosive degradation through interfacial modifications [73].
Table 7 Electrochemical Impedance parameters for mild steel in H2SO4 medium in the presence and absence of NNDBA at 298 K.
Compound Rs (Ω/cm2 ) Rp(Ω/cm2) Rct (Ω/cm2 ) fmax (Hz) CPE (μS.sncm- 2 ) n Cdl(μF/cm2) IE (%)
NNDBA 2.218 2171 2168.78 1.74 74.92 0.7321 421.96 96.32
 24HSO 2.809 82.43 79.621 5.49 172.5 0.9515 364.28 -
Fig. 11. Scanning electron micrograph of mild steel in (a) 0.5H2SO4 (b) 10-1 M NNDBA in H2SO4 (c) 10-7 M NNDBA in H2SO4 at 3000 magnifications.
Fig. 12. AFM of (a) plain mild steel specimen, (b) mild steel specimen in the 0.5H2SO4 solution (c) mild steel specimen in presence of 10-1MNNDBAin H2SO4.

Atomic force microscopy (AFM)

Atomic Force Microscopy (AFM) was employed to analyze the surface topography of mild steel after six hours of immersion in 0.5 M H2SO4, both in the presence and absence of NNDBA, under ambient conditions, as SEM micrographs do not offer quantitative information on surface roughness.
Fig. 12 (a-c) presents the 3D images of MS under different conditions, providing insights into the topographical changes resulting from corrosion and inhibitor adsorption. Fig. 12(a) represents the plain, uncorroded MS surface, which appears relatively smooth. In contrast, Fig. 12(b) shows the MS surface after exposure to the blank acidic solution, where the average surface roughness increases significantly to 176.5 nm. This high roughness value indicates severe corrosion damage, characterized by an uneven and bumpy structure with noticeable pits and irregularities caused by the aggressive attack of sulfuric acid [74]. Conversely, in the presence of NNDBA, Fig. 12(c), the AFM image reveals a significantly smoother surface. The reduction in peak and valley heights to approximately 13.6 nm suggests a substantial decrease in surface roughness [75]. The smoother surface observed in the AFM analysis confirms the effectiveness of NNDBA in mitigating corrosion by forming a uniform protective film, thereby reducing surface deterioration. These findings support the role of NNDBA as an effective corrosion inhibitor for mild steel in 0.5MH2SO4 and are consistent with results obtained from other experimental techniques, further reinforcing its protective capabilities.

Quantum chemical studies

Quantum chemical computations enable a profound elucidation of the donor-acceptor dynamics between the frontier molecular orbitals (FMOs) of the inhibitor and the MS surface. This theoretical framework delves into the electronic structural paradigm within the HOMO-LUMO frontier orbital construct, meticulously analyzing the spatial electron density topology, Mulliken charge dispersions, and an array of quantum-mechanical descriptors pivotal for deciphering the mechanistic intricacies of corrosion inhibition.

Relation between inhibition efficiency and FMO energies

An in-depth analysis of HOMO identifies molecular regions with the highest electron-donating propensity toward electrophiles, corresponding to ionization potential (IP), while LUMO highlights areas with maximal electron-accepting capacity from nucleophiles, linked to electron affinity (EA). A highly negative EHOMO (-4.419eV) indicates weakened donor ability, favoring physisorption on the metal surface [76] i.e., reflects the tendency of the NNDBA molecules to donate electrons to vacant 3d-orbital of Fe. This result is in corroboration with experimental kinetic and thermodynamic parameters: Ea and $\mathrm{\Delta }{G}_{ads}^{\circ },\mathrm{\Delta }{H}_{ads}^{\circ },\mathrm{\Delta }{S}_{ads}^{\circ }$.
Diminished ELUMO (0.821eV) values signify an increased propensity of the molecule to function as an electron acceptor from metal surface thereby facilitating electrophilic interactions [77]. The low ΔE value (5.24 eV) signifies the enhanced stability of the metal-inhibitor complex on the MS surface, thereby augmenting the corrosion inhibition efficiency of NNDBA [78]. A reduced energy gap (ΔE ) enhances inhibition efficiency by lowering the excitation energy needed for electron abstraction from the highest occupied orbital; such species are more polarizable and exhibit heightened chemical reactivity [79].(Fig. 13)
When a molecule gets protonated, a molecule loses a unit of negative charge, making its conjugate acid electron-deficient. This electronic change results in a higher energy gap (ΔE ) in protonated NNDBA, which is a reflection of increased molecular stability. This stability supports the formation of a strong and persistent adsorptive film over the metal substrate. In addition, the presence of a positive charge at the nitrogen atom increases the electrostatic attraction of the molecule with the negatively charged mild steel surface and thus makes adsorption stronger [80]. Importantly, the high EHOMO  value for the protonated species (Fig. 14) indicates an increased tendency to donate electrons, which increases the inhibitor-metal interaction. In addition, the decreased ELUMO  in the protonated species increases its electron acceptance capability, promoting back-donation from the metal and hence encouraging more stable and long-lasting surface interactions [81].
Fig. 13. Optimized geometry of NNDBA with HOMO and LUMO distributions and their corresponding energy levels.
Fig. 14. Optimized geometry of protonated NNDBA with HOMO and LUMO distributions and their corresponding energy levels.

Relation between electronic parameters and type of adsorption

Chemical hardness (ηinhibitor  ) quantifies an atom, ion, or molecule's resistance to minor perturbations in chemical reactions, preserving its electronic structure with minimal polarization which is formulated [82] as:
$\begin{array}{c}{\eta }_{\text{inhibitor}\text{ }}=0.5+(I-A)\end{array}$
I= Ionization Energy =-EHOMO =4.419eV
A= Electron Affinity =-ELUMO =-0.821eV
A lower hardness value (eV) denotes higher reactivity and lower stability.
NNDBA, exhibiting moderate hardness (2.62 eV) predominantly adsorbs onto mild steel via physisorption, forming a protective barrier that impedes aggressive species (H+,SO4 2- ) from surface interaction. This aligns with corroborative analytical techniques [83].

Relation between ΔN and inhibition mechanism

ΔN denotes the fractional electron transfer from a donor to the acceptor's vacant orbitals, quantifying electronic delocalization in do-nor-acceptor interactions, and is mathematically expressed as:
$\begin{array}{c}\Delta N=\frac{{\chi }_{Fe}-{\chi }_{NNDBA}}{2\left({\eta }_{Fe}+{\eta }_{NNDBA}\right)}\end{array}$
It's crucial to highlight that the electronegativity of iron (χFe ) is considered to be 4.82 eV and the theoretical value of hardness of metal surface is taken as zero (ηFe=0eV ), while the electronegativity of the inhibitor is calculated by:
$\begin{array}{c}{\chi }_{NNDBA}=0.5\times (I+A)\end{array}$
Thus, χNNDBA  from Eq. 24 is 1.799 eV and ηNNDBA  from Eq. 22 is 2.62 eV [84]. ΔN elucidates the vectorial nature of electron transfer between interacting systems; when the inhibitor (χNNDBA =1.799eV ) and iron (χFe=4.82eV ) are brought into contact, electrons migrate from the species with lower electronegativity to that with higher electronegativity until thermodynamic equilibrium of chemical potentials is established. A positive ΔN reflects electron donation from the inhibitor to the metal surface, whereas a negative ΔN signifies a reverse electron flow from the metal to the inhibitor. The computed electron transfer fraction (ΔN=0.577 ) aligns with Lukovit's criterion, remaining below 3.6 and positive, signifying that inhibition efficiency escalates with NNDBA's electron-donating propensity toward the mild steel (MS) surface. The positive ΔN values substantiate electron migration from neutral inhibitors to Fe, reinforcing the adsorption mechanism [85].

Relation between dipole moment & efficacy of NNDBA

NNDBA exhibits a dipole moment of 1.92 Debye exceeding that of water (1.88 Debye), implying potent dipole-dipole interactions with the metal surface. This suggests NNDBA possesses a superior adsorption affinity for mild steel, effectively displacing pre-adsorbed water molecules [86].

Condensed fukui function

The Fukui function provides insight into site-specific reactivity indices within a molecular framework, wherein the local reactivity of the selected inhibitors is ascertained via Fukui indices computed from Mulliken charge distributions of their anionic and cationic states.
The Fukui parameters (Eqs. 25-27) discern electrophilic and nucleophilic centers within a molecule through the application of Yao's dual descriptor:
f(k)+=[qk(N+1)-qk(N)] for nucleophilic attack
f(k)-=[qk(N)-qk(N+1)] for electrophilic attack
$\begin{array}{c}\Delta {\mathrm{f}}_{\mathrm{k}}=f(k{)}^{+}-f(k{)}^{-}={f}^{2}\end{array}$
Upon electron acceptance by a molecule, the atomic charges are represented as qk(N+1), whereas qk(N) corresponds to the charge distribution in its neutral state, and qk(N-1) reflects the atomic charges following electron loss. The Fukui functions f(k)+and f(k)- characterize the nucleophilic and electrophilic tendencies at a given site, respectively, while Δfk denotes the second-order Fukui function, capturing the differential reactivity between these two modes [87].
In general, the greatest f(k)-signifies the site for electrophilic assault whereas the highest f(k)+indicates the site for nucleophilic attack. The positive Fukui function at a particular site in a molecule indicates where the molecule is likely to experience an increase in electron density, which can be related to areas of high reactivity. In corrosion, this means that the molecule is more susceptible to undergoing reduction reactions, which is common in corrosion processes. If a molecule with high Fukui positive values is used as a corrosion inhibitor, it will be more effective in protecting the metal surface by interacting with electron rich sites.
As observed from Table 8, the phenyl carbons C9 and C10, along with hydrogens H28,H29,H36, and H 37, exhibit the highest f(k)+ values, marking them as prominent nucleophilic centers. Simultaneously, atoms N1, C9, and C10 also show elevated f(k)-values, indicating strong electrophilic behavior. A high Fukui negative index highlights region more prone to electron loss and potential oxidation. The dual role of C and H atoms in NNDBA as both electron donors and acceptors promotes strong interaction with the metal surface, aiding in corrosion inhibition. Notably, the convergence of nucleophilic and electrophilic activity at C9 and C10 points to their exceptional reactivity, making them key sites in the molecule's protective action against acid-induced corrosion of mild steel [88].
Table 8 Fukui functions (f*,f, and f2 ) for atoms of NNDBA.
Atom f+ f f2
N1 -0.001 0.15 -0.151
C2 -0.02 -0.047 0.027
C3 -0.027 -0.041 0.014
C4 -0.022 -0.005 -0.017
C5 -0.028 0 -0.028
C6 -0.006 0.021 -0.027
C7 -0.025 0.005 -0.03
C8 -0.024 0.005 -0.029
C9 0.096 0.047 0.049
C10 0.098 0.047 0.051
C11 -0.01 -0.006 -0.004
C12 -0.008 -0.008 0
C13 0.062 0.053 0.009
C14 0.068 0.045 0.023
C15 -0.006 0.061 -0.067
H16 0.022 0.038 -0.016
H17 0.035 0.058 -0.023
H18 0.017 0.041 -0.024
H19 0.042 0.051 -0.009
H20 0.02 -0.012 0.032
H21 0.026 0.029 -0.003
H22 0.023 -0.015 0.038
H23 0.031 0.023 0.008
H24 0.023 0.022 0.001
H25 0.017 0.005 0.012
H26 0.019 0.026 -0.007
H27 0.018 0.003 0.015
H28 0.109 0.032 0.077
H29 0.104 0.037 0.067
H30 0.001 0.011 -0.01
H31 0.02 0.034 -0.014
H32 0.016 0.015 0.001
H33 -0.003 0.016 -0.019
H34 0.021 0.034 -0.013
H35 0.017 0.014 0.003
H36 0.094 0.066 0.028
H37 0.096 0.064 0.032
H38 0.084 0.08 0.004
The dual descriptor (Δfk ) offers a refined perspective on atomic reactivity, where a positive value (Δfk>0 ) indicates dominance of nucleophilic attack, and a negative value (Δfk<0 ) reflects a preference for electrophilic interaction. As illustrated in Figure..., analysis of second-order Fukui function values for NNDBA reveals that 52.64% of its atoms exhibit nucleophilic character (Δfk>0 ), while 47.37% display electrophilic behavior (Δfk<0 ). This slight predominance of nucleophilic sites suggests that NNDBA has a marginally nucleophilic nature, enhancing its ability to effectively inhibit corrosion on the mild steel surface [89].

Molecular dynamics simulation

Molecular dynamics (MD) simulation, rooted in Monte Carlo methods, is an advanced computational approach used to investigate the molecular-level behavior of corrosion inhibitors. It enables the evaluation of adsorption energies and elucidates the inhibition mechanism on metal surfaces. MD simulations offer insights into the spatial orientation and interaction of inhibitor molecules with mild steel in acidic media [90]. At standard temperature and pressure, iron crystallizes in a body-centered cubic (bcc) structure, exhibiting facets such as (110), (100), (211), (311), (111), (321), and (210). Among these, the Fe (110) surface is the most thermodynamically stable, possessing the lowest surface energy and covering the largest area of the crystal. Hence, Fe(110) was selected as the representative adsorption site for inhibitor simulations in corrosive environments [91] (Fig. 15).
Fig. 15. Bar graph representing the Fukui functions (f*,f, and f2 ) for individual atoms of NNDBA.
Fig. 16 shows side and top views of the stable configurations of the inhibitor adsorbed on the mild steel surface in sulfuric acid, based on MD simulation. The NNDBA molecule strongly adheres to the Fe(110) surface. From the side view, the molecule lies nearly parallel to the surface, allowing full surface coverage. Adsorption occurs mainly through atoms N(1), planar phenylic carbons C(9),C(10), hydrogens H (28), H(29), and the alkyl chain hydrogens H(36),H(37), aligning well with Fukui function analysis. This enables the inhibitor to effectively block active sites and protect the metal surface.
The interaction between the inhibitor and the Fe(110) surface is represented by the interaction energy (Einteraction  ), calculated as:
$\begin{array}{c}{\mathrm{E}}_{\text{interaction}\text{ }}={\mathrm{E}}_{\text{total}\text{ }}-\left({\mathrm{E}}_{\text{surface}\text{ }+\text{ }\text{water}\text{ }+\text{ }\text{sulphuric acid}\text{ }}+{\mathrm{E}}_{\text{inhibitor}\text{ }}\right)\end{array}$
Etotal  represents the total energy of the complete system, while Esurface + water + sulphuric acid corresponds to the energy of the Fe(110) surface in combination with water and sulfuric acid, and Einhibitor  denotes the isolated energy of the inhibitor molecule. The binding energy (Ebinding  ) is defined as the negative of the interaction energy (Einteraction  ), establishing the relation:
$\begin{array}{c}{\mathrm{E}}_{\text{binding}\text{ }}=-{\mathrm{E}}_{\text{interaction}\text{ }}\end{array}$
A higher magnitude of binding energy, such as 123.11kcal/mol, implies a stronger affinity of the inhibitor towards the mild steel substrate, indicating robust adsorption. The negative sign of the binding energy further confirms the thermodynamic favorability of the interaction, validating the spontaneous nature of inhibitor adsorption on the Fe (110) surface. This computational insight is consistent with the experimental findings [92].(Table 9)
Table 9 Outputs calculated by the Monte Carlo simulation for adsorption of NNDBA, on Fe (110).
Parameter Value (kJ/mol)
Interaction energy -123.11
Binding energy 123.11
For a better assessment of NNDBA's inhibition performance, a comparison with structurally similar inhibitors has been presented. Table 10 outlines essential details such as the metal type, corrosive environment, inhibitor concentration, and recorded inhibition efficiencies from previous studies.

Proposed mechanism

The inhibitory mechanism of NNDBA on MS in 0.5MH2SO4 is elucidated through a confluence of experimental observations and theoretical computations. NNDBA mitigates corrosion through a triad of synergistic pathways:
Fig. 16. MD simulation captures in (a) top view of adsorbed NNDBA molecule on mild steel surface (b) front view of interactions between NNDBA and mild steel (c) top view of adsorbed NNDBA molecule on mild steel in presence of H2O and H2SO4 (d) top view of adsorbed NNDBA molecule on mild steel in presence of H2O and H2SO4.
1. The molecular architecture of NNDBA incorporates nitrogen heteroatoms endowed with lone pairs, while the electron-donating butyl substituents augment the electron density on both the nitrogen center and the adjacent phenylic carbon. This electronic enrichment promotes robust interactions between the lone pairs on nitrogen and the delocalized π-electrons of the aromatic moieties with the vacant d-orbitals of surface iron atoms [102].
Table 10 Comparative Overview of NNDBA and Structurally Related Corrosion Inhibitors.
Inhibitor IE (%) Metal Media Adsorption Isotherm Reference
N-Methyl Formanilide 83.1 Mild steel H2SO4 El-Awady adsorption [36]
(E)-N(2-Chlorobenzylidene)-2-Fluorobenzenamine and 89.7 Mild Steel HCl Langmuir [93]
(E)-N(2-Chlorobenzylidene) - 3-Chloro - 2-Methylbenzenamine 85.6
(Z) - 3-(1-(2-(4-amino - 5-mercapto - 4H-1,2,4-triazol - 3-yl) hydrazono)ethyl) - 2 H -chromen - 2-one 86.1 Carbon Steel 1 M HCl Langmuir [94]
5 -imino - 1,2,4-dithiazolidine-3-thione (IDTT) 86.5 Mild steel 1 M HCl Langmuir [95]
1-((8-hydroxyquinolin - 5-yl)methyl)urea 87.3 Carbon Steel 1 M HCl Langmuir [96]
(E) - 3-(2-(benzofuran-2-yl)vinyl)quinoxalin-2(1 H)-one 89.6 Mild Steel 1 M HCl Langmuir [97]
d toluene-2,4-diisocyanate-4-(1Himidazole-ly) aniline (TDIA) 89.99 Stainless Steel 304 L 1MHCl+3.5%NaCl Langmuir [98]
4-methyl-1-Phenyl-3-(p-tolyldiazenyl) -2,3-dihydro-1H-pyrrol-2-ol (MPTHP) 90.6 N - 80 Steel 10%HCl Langmuir [99]
(2-benzoyl-4-nitro-N-[(1H-pyrazol-1-yl)methyl]aniline (BNPMA)) 93.2 Carbon Steel 1 M HCl Langmuir [100]
4,6-bis(3,5-dimethyl-1 H-pyrazol - 1-yl)-N-(4-methoxyphenyl)-1,3,5-triazin-2-amine 93.6 Carbon Steel 0.25 M H2SO4 Frumkin [101]
N, N-dibutyl aniline (Our Result) 94.2 Mild Steel 0.5 M H2SO4 Langmuir Our result
Fig. 17. Visualization of interaction between NNDBA and mild steel (MS) surface.
Density Functional Theory (DFT) calculations, corroborated by Fukui function analysis, substantiate this electronic interplay.
Further support from Monte Carlo simulations and empirical findings confirms that NNDBA establishes a protective adsorptive interface on the Fe substrate via its polar functional groups. These groups orchestrate a physisorption mechanism wherein NNDBA molecules adopt a nearly parallel orientation relative to the steel surface, held predominantly by van der Waals forces. This spatial configuration facilitates the formation of a compact, uniform monolayer that maximizes surface coverage, thereby effectively shielding the MS substrate from aggressive ionic ingress. By occluding active anodic and cathodic sites, the inhibitor substantially curtails the kinetics of electron transfer processes and impedes the diffusion pathways of corrosive species, culminating in significant corrosion attenuation.
2. The molecular framework of NNDBA encompasses an extended hydrophobic alkyl moiety, which imparts a significant degree of water-repellency upon adsorption onto the metal surface. This hydrophobic barrier impedes the ingress of aqueous and ionic species to the underlying substrate, thereby disrupting the electrolyte-metal interactions essential for corrosion initiation and propagation. Consequently, the corrosion kinetics are markedly suppressed due to restricted access of water and aggressive ions to the active sites.
This multifaceted inhibition mechanism as shown in Fig. 17, underlines the potent protective efficacy of NNDBA, rooted in both molecular-level electronic interactions and surface adsorption dynamics.

Conclusion

The heterocyclic N-substituted aniline derivative, N,N Dibutylaniline (NNDBA), comprising a planar aniline core substituted with two butyl chains on nitrogen, was systematically investigated as a corrosion inhibitor for MS in 0.5MH2SO4 using both experimental and theoretical methodologies. Gravimetric analysis confirmed that MS corrosion followed first-order kinetics, with NNDBA significantly reducing the corrosion rate and rate constant, while prolonging the halflife of the reaction. The inhibition efficiency (IE) increased with NNDBA concentration and decreased with temperature, indicating a physisorption mechanism. The calculated activation energy (Ea ) in presence of NNDBA was 70.04 kJ/mol which was higher than that of the uninhibited system but remained below 80 kJ/mol, supporting physical adsorption. Thermodynamic analysis revealed that the standard free energy of adsorption (ΔG ads  ) was approximately -10 kJ/mol, further confirming electrostatic interactions consistent with physisorption.
Adsorption followed the Langmuir isotherm, with favorable Kads values indicating monolayer coverage and spontaneous adsorption. Potentiodynamic polarization (PDP) measurements identified NNDBA as a mixed-type inhibitor, with a maximum shift in Ecorr  of 34.2 mV. At a concentration of 10-1M, NNDBA achieved an inhibition efficiency of 92.4%, while even at the lowest concentration of 10-7M at 298 K, the efficiency remained substantial at 72.5%. Electrochemical impedance spectroscopy (EIS) supported these findings, with increased Rct and reduced Cdl, indicating the formation of a compact inhibitor layer at the metal interface.
Surface characterization via SEM and AFM confirmed the formation of a protective film on the inhibited MS surface. Quantum chemical calculations yielded a highly negative EHOMO (-4.419eV), low ELUMO  (0.821 eV), and a small energy gap (ΔE=5.24eV ), suggesting strong reactivity and enhanced metal-inhibitor complex stability. Fukui indices identified key active sites contributing to adsorption. Molecular dynamics simulations revealed that NNDBA adopts a stable, parallel alignment on Fe(110), interacting via the nitrogen atom, aromatic ring, and adjacent hydrogens. Monte Carlo simulations supported these findings with strongly negative interaction energies, indicating thermodynamically favorable and exothermic adsorption.
In conclusion, NNDBA demonstrates excellent corrosion inhibition performance in acidic media through strong surface adsorption, electrostatic interaction, and the formation of a stable protective film. Its performance is well-supported by kinetic, thermodynamic, electrochemical, surface, and theoretical data, confirming its potential as an effective and reliable corrosion inhibitor for mild steel.
Authors contribution
Meenakshi Gupta: Concept, experimentation, analysis of the data, and writing the first draft. Mansi Yuvender Choudhary: Plotting of graphs and drafting of manuscript. Neeta Azad: Software, Quantum chemical research and result interpretation. Shramila Yadav: Writing, Interpretation of results, data analysis, figure generation, and finalizing the manuscript.
CRediT authorship contribution statement
Neeta Azad: Writing - review & editing, Software. Shramila Yadav: Writing - review & editing, Writing - original draft, Formal analysis. Meenakshi Gupta: Writing - original draft, Formal analysis, Data curation. Mansi Y. Chaudhary: Writing - review & editing, Software.
Declaration of Competing Interest
The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.
[1]
N. Perez, Electrochemical corrosion, Materials Science: Theory and Engineering, Springer Nature Switzerland, Cham, 2024, pp. 835-898.

[2]
R. Rodrigues, S. Gaboreau, J. Gance, I. Ignatiadis, S. Betelu, Reinforced concrete structures: a review of corrosion mechanisms and advances in electrical methods for corrosion monitoring, Constr. Build. Mater. 269 ( 2021) 121240, https://doi.org/10.1016/j.conbuildmat.2020.121240.

[3]
A. Kadhim, A.A. Al-Amiery, R. Alazawi, M.K.S. Al-Ghezi, R.H. Abass, Corrosion inhibitors. A review, Int. J. Corros. Scale Inhib. 10 (1) ( 2021) 54-67

[4]
V. S, S.A. Umoren, Corrosion inhibitors in the oil and gas industry, John Wiley & Sons, ( 2020), https://doi.org/10.1002/9783527822140.ch18.

[5]
A. Kumar, J. Singh, Overview on corrosion in automotive industry and thermal power plant, Proc. Eng. Sci. 4 ( 2022) 13-22, https://doi.org/10.24874/PES04.01. 003.

[6]
P. Visser, H. Terryn, J.M. Mol, Aerosp. Coat. Act. Prot. Coat. N. Gener. Coat. Met. ( 2016) 315-372.

[7]
S.A. El-Enin, A. Amin, Review of corrosion inhibitors for industrial applications, Int. J. Eng. Res. Rev. 3 (2) ( 2015) 127-145.

[8]
A. Ashwathareddy, S. Rao, S.S. Subramaniyam, P.G. Krishna, K.M. Rama, S. Kodange, Corrosion mitigation of mild steel in acidic medium using liquid crystals: a comprehensive review, Inorg. Chem. Commun. ( 2024) 113071, https://doi.org/10.1016/j.inoche.2024.113071.

[9]
F.M. Galleguillos Madrid, A. Soliz, L. Cáceres, M. Bergendahl, S. Leiva-Guajardo, C. Portillo, M. Páez, Green corrosion inhibitors for metal and alloys protection in contact with aqueous saline, Materials 17 (16) ( 2024) 3996, https://doi.org/10.3390/ma17163996.

[10]
N. Hossain, M. Asaduzzaman Chowdhury, M. Kchaou, An overview of Green corrosion inhibitors for sustainable and environment friendly industrial development, J. Adhes. Sci. Technol. 35 (7) ( 2021) 673-690, https://doi.org/10.1080/01694243.2020.1816793.

[11]
H. Assad, A. Kumar, Understanding functional group effect on corrosion inhibition efficiency of selected organic compounds, J. Mol. Liq. 344 ( 2021) 117755, https://doi.org/10.1016/j.molliq.2021.117755.

[12]
L. Chen, D. Lu, Y. Zhang, Organic compounds as corrosion inhibitors for carbon steel in HCl solution: a comprehensive review, Materials 15 (6) ( 2022) 2023, https://doi.org/10.3390/ma15062023.

[13]
M.A. Ahmed, S. Amin, A.A. Mohamed, Current and emerging trends of inorganic, organic and eco-friendly corrosion inhibitors, RSC Adv. 14 (43) ( 2024) 31877-31920, https://doi.org/10.1039/D4RA05662K.

[14]
L.T. Popoola, Organic Green corrosion inhibitors (OGCIs): a critical review, Corros. Rev. 37 (2) ( 2019) 71-102, https://doi.org/10.1515/corrrev-2018-0058.

[15]
S. Dewangan, N. Vaidya, A.K. Bhatia, Oxygen-containing heterocyclic compounds as Green corrosion inhibitors, In Computational Modelling and Simulations for Designing of Corrosion Inhibitors, Elsevier, 2023, pp. 395-414, https://doi.org/ 10.1016/B978-0-323-95161-6.00012-6.

[16]
M.L. Zheludkevich, J. Tedim, M.G.S. Ferreira, Smart" coatings for active corrosion protection based on multi-functional micro and nanocontainers, Electrochim. Acta 82 ( 2012) 314-323, https://doi.org/10.1016/j.electacta.2012.04.095.

[17]
X.H. To, N. Pebere, N. Pelaprat, B. Boutevin, Y. Hervaud, A corrosion-protective film formed on a carbon steel by an organic phosphonate, Corros. Sci. 39 (10-11) ( 1997) 1925-1934, https://doi.org/10.1016/S0010-938X(97)00086-3.

[18]
M.A. Quraishi, S.K. Shukla, Poly (aniline-formaldehyde): a new and effective corrosion inhibitor for mild steel in hydrochloric acid, Mater. Chem. Phys. 113 (23) ( 2009) 685-689, https://doi.org/10.1016/j.matchemphys.2008.08.028.

[19]
C.M. Fernandes, M.V.P. de Mello, N.E. dos Santos, A.M.T. de Souza, M. Lanznaster, E.A. Ponzio, Theoretical and experimental studies of a new aniline derivative corrosion inhibitor for mild steel in acid medium, Mater. Corros. 71 (2) ( 2020) 280-291, https://doi.org/10.1002/maco.201911065.

[20]
M.E. Belghiti, S. Bouazama, S. Echihi, A. Mahsoune, A. Elmelouky, A. Dafali,... M. Tabyaoui, Understanding the adsorption of newly Benzylidene-aniline derivatives as a corrosion inhibitor for carbon steel in hydrochloric acid solution: experimental, DFT and molecular dynamic simulation studies, Arab. J. Chem. 13 (1) ( 2020) 1499-1519, https://doi.org/10.1016/j.arabjc.2017.12.003.

[21]
A. Ait Mansour, K. Subbiah, H. Lgaz, M.R. Al-Hadeethi, M. Messali, T. Park,... R. Salghi, Investigating corrosion failure in N80 carbon steel: experimental and theoretical insights into isonicotinohydrazide derivatives as inhibitors in acidic conditions, Inorg. Chem. Commun. 161 ( 2024) 112007, https://doi.org/10.1016/j.inoche.2023.112007.

[22]
A. Ait Mansour, M.R. Al-hadeethi, H. Lgaz, K. Subbiah, M. Messali, H.S. Lee, R. Salghi, Exploring the potential of isonicotinohydrazide derivatives in N80 steel corrosion control: an integrated approach through synthesis, modeling, and experimentation in acidic environments, Colloids Surf. A Physicochem. Eng. Asp. 679 ( 2023) 132542, https://doi.org/10.1016/j.colsurfa.2023.132542.

[23]
A. Ait Mansour, A.E.A. Allah, H. Lgaz, M. Messali, H.S. Lee, L. Bazzi, B. Hammouti, Evaluation of N80 carbon steel corrosion in 15wt\%HCl using isatin-hydrazones: a comprehensive approach with chemical, electrochemical techniques, and DFTB calculations, J. Mol. Struct. 1321 ( 2025) 139910, https://doi.org/10.1016/j.molstruc.2024.139910.

[24]
Sheetal A.K. Singh A. Ait Mansour S. Thakur B. Pani M. Singh R. Salghi, Aromaticity of heterocyclic compounds and their corrosion inhibition property: experimental and theoretical analysis, Langmuir 40 (41) ( 2024) 21675-21692, https://doi.org/10.1021/acs.langmuir.4c02707.

[25]
S. Vishwanatham, Inhib. Eff. some Anilin Compd. Corros. mild Steel 3%HF ( 1998).

[26]
R.T. Loto, C.A. Loto, T. Fedotova, Electrochemical studies of mild steel corrosion inhibition in sulfuric acid chloride by aniline, Res Chem. Inter. 40 ( 2014) 1501-1516, https://doi.org/10.1007/s11164-013-1055-x.

[27]
N.S. Abtan, M.A.I. Al-Hamid, L.A. Kadhim, F.F. Sayyid, F.T.M. Noori, A. Kadum, W.K. Al-Azzawi, Unlocking the power of 4-Acetamidoantipyrine: a promising corrosion inhibitor for preserving mild steel in harsh hydrochloric acid environments, Prog. Color. Color. Coat. 17 (1) ( 2024) 85-96, https://doi.org/10.30509/pccc.2023.167147.1223.

[28]
A. Alamiery, W.K. Al-Azzawi, Investigation of 3-(1, 3-oxazol-5-yl) aniline as a highly efficient corrosion inhibitor for mild steel in 1 m HCl solution, Int. J. Low. Carbon Technol. 18 ( 2023) 850-862, https://doi.org/10.1093/ijlct/ctad069.

[29]
M.F. Carlos, G.K. Barboza, A. Echevarria, Anticorrosive effect of halogenated aniline enaminoesters on carbon steel in HCl, Int. J. Corros. 2022 (1) ( 2022) 7218063, https://doi.org/10.1155/2022/7218063.

[30]
M.M. El-Naggar, A.S. Amin, S.M. Syam, S.M. Refaat, B.N. Ahmed, Inhibition mechanism of mild steel corrosion in acidic media by some amine compounds, Benha J. Appl. Sci. 7 (4) ( 2022) 231-237, https://doi.org/10.21608/bjas.2022.257801.

[31]
G. ASTM, Standard practice for preparing, cleaning, and evaluating corrosion test specimens, Am. Soc. Test. Mater. ( 2003).

[32]
Ali N., Fonna S., Saputra Y., Ariffin A.K., & Supardi J. Efficiency of Syzygium cumini Fruit Extract as a Green Corrosion Inhibitor for Low Carbon Steel in Hydrochloric Acid Solution. Indonesian Journal of Chemistry. DOI: 10.22146/ijc.103598.

[33]
A. Rahman, M. Ismail, M. Hussain, Inhibition of corrosion of mild steel in hydrochloric acid by bambusa arundinacea, Int. Rev. Mech. Eng. 5 ( 2011) 59-63.

[34]
M.M. Solomon, S.A. Umoren, I.I. Udosoro, A.P. Udoh, Inhibitive and adsorption behavior of carboxymethyl cellulose on mild steel corrosion in sulphuric acid solution, Corros. Sci. 52 (4) ( 2009) 1317-1325, https://doi.org/10.1016/j.corsci.2009.11.041.

[35]
C. Zeng, Z.-Y. Zhou, W.-J. Mai, Q.-H. Chen, J.-B. He, B.-K. Liao, Exploration on the corrosion inhibition performance of salvia miltiorrhiza extract as a Green corrosion inhibitor for Q235 steel in HCl environment, J. Mater. Res. Technol. 32 ( 2024) 3857-3870, https://doi.org/10.1016/j.jmrt.2024.09.003.

[36]
M. Gupta, N. Azad, M.Y. Chaudhary, Y.S. Sharma, S. Yadav, Inhibitory potential of N-Methyl formanilide on mild steel corrosion in sulfuric acid: quantum chemical and experimental perspectives, Asian J. Chem. 36 (11) ( 2024) 2521-2529, https://doi.org/10.14233/ajchem.2024.32292.

[37]
R.T. Loto, C.A. Loto, Data on the comparative evaluation of the corrosion inhibition of vanillin and vanillin admixed with rosmarinus officinalis on mild steel in dilute acid media, Chem. Data Collect. 24 ( 2019) 100290, https://doi.org/10.1016/j.cdc.2019.100290.

[38]
M. Sabiha, Y. Kerroum, M. El Hawary, M. Boudalia, A. Bellaouchou, O. Hammani, H.M.A. Amin, Investigating the adsorption and corrosion protection efficacy and mechanism of marjoram extract on mild steel in HCl medium, Molecules 30 (2) ( 2025) 272, https://doi.org/10.3390/molecules30020272.

[39]
N.I. Kairi, J. Kassi, The effect of temperature on the corrosion inhibition of mild steel in 1 m HCl solution by curcuma longa extract, Int. J. Electrochem. Sci. 8 ( 2013) 7138-7155, https://doi.org/10.1016/S1452-3981(23)14836-X.

[40]
O.D. Ofuyekpone, O.G. Utu, B.O. Onyekpe, Corrosion inhibition for alloy 304 L (UNS S30403) in 1 mH2SO4 solution by centrosema pubescens leaves extract, Appl. Surf. Sci. Adv. 3 ( 2021) 100061, https://doi.org/10.1016/j.apsadv.2021. 100061.

[41]
M. Beniken, R. Salim, E. Ech-chihbi, M. Sfaira, B. Hammouti, M.E. Touhami, M. Taleb, Adsorption behavior and corrosion inhibition mechanism of a polyacrylamide on c-steel in 0.5 mH2SO4 : electrochemical assessments and molecular dynamic simulation, J. Mol. Liq. 348 ( 2022) 118022, https://doi.org/10.1016/j.molliq.2021.118022.

[42]
O. Dickson, O.G. Utu, B.O. Onyekpe, A.A. Adediran, M. Oki, Data on corrosion inhibition effect of stylosanthes gracilis extract on UNS S30403 austenitic stainless steel in dilute acid solution, Chem. Data Collect. 35 ( 2021) 100763, https://doi.org/10.1016/j.cdc.2021.100763.

[43]
N.C. Ngobiri, E.E. Oguzie, N.C. Oforka, O. Akaranta, Comparative study on the inhibitive effect of sulfadoxine-pyrimethamine and an industrial inhibitor on the corrosion of pipeline steel in petroleum pipeline water, Arab. J. Chem. 12 (7) ( 2019) 1024-1034, https://doi.org/10.1016/j.arabjc.2015.04.004.

[44]
I. Mu'azu, A.A.M. Ayuba, M. Hussein, I. Fater, Corrosion inhibition of mild steel in 0.3 m hydrochloric acid solution using urena lobata leaves extract. Applied Journal of Environmental Engineering Science, 10 (4) ( 2024) 206-223, https://doi.org/10.48422/IMIST.PRSM/ajees-v10i4.49700.

[45]
F.E. Abeng, B.I. Ita, V.C. Anadebe, V.I. Chukwuike, K.M. Etiowo, P.Y. Nkom, O.O. Ekerenam, N.B. Iroha, I.J. Ikot, Multidimensional insight into the corrosion mitigation of clonazepam drug molecule on mild steel in chloride environment: empirical and computer simulation explorations, Results Eng. 17 ( 2023) 100924, https://doi.org/10.1016/j.rineng.2023.100924.

[46]
F. Sabirneeza, A. Rahiman, S. Sethumanickam, Corrosion inhibition, adsorption, and thermodynamic properties of poly(vinyl alcohol-cysteine) in molar HCl, Arab. J. Chem. 10 (2) ( 2017) S3358-S3366, https://doi.org/10.1016/j.arabjc.2014.01.016.

[47]
S. Minjibir, Abubakar, M. Ladan, Corrosion inhibition potential of prosopis juliflora leaves extract on mild steel in H2SO4 solutions, Adv. J. Chem. Sect. A 6 (3) ( 2023) 311-323, https://doi.org/10.22034/ajca.2023.398774.1372.

[48]
A.A. Khadom, Effect of temperature on corrosion inhibition of copper-nickel alloy by tetraethylenepentamine under flow conditions, J. Chil. Chem. Soc. 59 (3) ( 2014), https://doi.org/10.4067/s0717-97072014000300004.

[49]
A.I. Abbas, S.A. Ahmed, W.K. Al-Azzawi, M.M. Hanoon, A. Alamiery, A. Isahak, K. Wan Nor Roslam, A.A.H. Kadhum, Corrosion inhibition of mild steel in hydrochloric acid environment using thiadiazole derivative: weight loss, thermodynamics, adsorption, and computational investigations, South Afr. J. Chem. Eng. 41 (1) ( 2022), https://doi.org/10.1016/j.sajce.2022.06.011.

[50]
A. Fawzy, M. Abdallah, I.A. Zaafarany, S.A. Ahmed, I.I. Althagafi, Thermodynamic, kinetic and mechanistic approach to the corrosion inhibition of carbon steel by new synthesized amino acids-based surfactants as Green inhibitors in neutral and alkaline aqueous media, J. Mol. Liq. 265 ( 2018) 276-291, https://doi.org/10.1016/j.molliq.2018.05.140.

[51]
A.O. Okewale, O.A. Adesina, Kinetics and thermodynamic study of corrosion inhibition of mild steel in 1.5 m HCl medium using cocoa leaf extract as inhibitor, J. Appl. Sci. Environ. Manag. 24 (1) ( 2020) 37-47, https://doi.org/10.4314/jasem.v24i1.6.

[52]
E. Ituen, O. Akaranta, A. James, Evaluation of performance of corrosion inhibitors using adsorption isotherm models: an overview, Chem. Sci. Int. J. 18 (1) ( 2017) 1-34, https://doi.org/10.9734/CSIJ/2017/28976.

[53]
S.C. Ikpeseni, G.O. Odu, H.I. Owamah, P.U. Onochie, D.C. Ukala, Thermodynamic parameters and adsorption mechanism of corrosion inhibition in mild steel using jatropha leaf extract in hydrochloric acid, Arab. J. Sci. Eng. 46 ( 2021) 7789-7799, https://doi.org/10.1007/s13369-021-05488-9.

[54]
M.A. Enad, Synthesis of new aromatic azo-schiff sulfonamide derivatives as carbon steel anticorrosive in 1 m hcl, Eur. Sci. Method. J. 2 (3) ( 2024) 22-37.

[55]
S. Yadav, S. Kaushik, N. Dheer, S. Kumar, G. Singh, M. Chaudhary, M. Gupta, Experimental investigation of anti-corrosive behaviour of beta vulgaris: a Green approach, J. Appl. Nat. Sci. 15 (3) ( 2023) 1315-1325, https://doi.org/10.31018/jans.v15i3.4969.

[56]
G. Pandimuthu, K. Muthupandi, T.W. Chen, S.M. Chen, A. Sankar, P. Muthukrishnan, S.P. Rwei, Inhibitor effect of N-(5-((4-chlorophenyl) diazenyl)-2-hydroxy benzylidene)-2-hydroxy benzohydrazide for mild steel corrosion in chloride and sulphate acidic solutions, Int. J. Electrochem. Sci. 16 (11) ( 2021) 211135, https://doi.org/10.20964/2021.11.51.

[57]
L. Emembolu, C. Igwegbe, Investigation of temperature correlations on corrosion inhibition of carbon steel in acid media by flower extract, Eur. J. Eng. Appl. Sci. 5 (1) ( 2022) 29-36, https://doi.org/10.31018/jans.v15i3.4969.

[58]
P.P. Kumari, P. Shetty, S.A. Rao, Electrochemical measurements for the corrosion inhibition of mild steel in 1 M hydrochloric acid by using an aromatic hydrazide derivative, Arab. J. Chem. 10 (5) ( 2017) 653-663, https://doi.org/10.1016/j.arabjc.2014.09.005.

[59]
S.K. Saha, M. Murmu, N.C. Murmu, P. Banerjee, Benzothiazolylhydrazine azomethine derivatives for efficient corrosion inhibition of mild steel in acidic environment: integrated experimental and density functional theory cum molecular dynamics simulation approach, J. Mol. Liq. 364 ( 2022) 120033, https://doi.org/10.1016/j.molliq.2022.120033.

[60]
P. Vashishth, H. Bairagi, R. Narang, S.K. Shukla, B. Mangla, Thermodynamic and electrochemical investigation of inhibition efficiency of Green corrosion inhibitor and its comparison with synthetic dyes on MS in acidic medium, J. Mol. Liq. 365 ( 2022) 120042, https://doi.org/10.1016/j.molliq.2022.120042.

[61]
S.A. Al Kiey, A.A. El-Sayed, A.M. Khalil, Controlling corrosion protection of mild steel in acidic environment via environmentally benign organic inhibitor, Colloids Surf. A Physicochem. Eng. Asp. 683 ( 2024) 133089, https://doi.org/10.1016/j.colsurfa.2023.133089.

[62]
V. Saraswat, M. Yadav, Improved corrosion resistant performance of mild steel under acid environment by novel carbon dots as Green corrosion inhibitor, Colloids Surf. A Physicochem. Eng. Asp. 627 ( 2021) 127172, https://doi.org/10.1016/j.colsurfa.2021.127172.

[63]
K. Dahmani, M. Khattabi, I. Saber, O. Kharbouch, M. Galai, S.M. Alharbi, F. Benhiba, A. Shaim, Z.S. Safi, M. Ebn Touhami, M. Cherkaoui, Study on the corrosion inhibition properties of quinoxaline derivatives as acidizing corrosion inhibitors for mild steel: synthesis, experimental analysis, and theoretical insights, Chem. Afr. 7 (10) ( 2024) 5461-5483, https://doi.org/10.1007/s42250-024-01135-6.

[64]
H.M. Elabbasy, H.S. Gadow, Study the effect of expired tenoxicam on the inhibition of carbon steel corrosion in a solution of hydrochloric acid, J. Mol. Liq. 321 ( 2021) 114918, https://doi.org/10.1016/j.molliq.2020.114918.

[65]
E. Kamali Ardakani, E. Kowsari, A. Ehsani, Imidazolium-derived polymeric ionic liquid as a Green inhibitor for corrosion inhibition of mild steel in 1.0 m HCl : experimental and computational study, Colloids Surf. A Physicochem. Eng. Asp. 586 ( 2020) 124195, https://doi.org/10.1016/j.colsurfa.2019.124195.

[66]
H. Rahmani, K.I. Alaoui, K.M. Emran, A. El Hallaoui, M. Taleb, S. El Hajji, B. Labriti, E. Ech-chihbi, B. Hammouti, F. El-Hajjaji, Experimental and DFT investigation on the corrosion inhibition of mild steel by 1, 2, 3- triazole regioisomers in 1M hydrochloric acid solution, Int. J. Electrochem. Sci. 14 (1) ( 2019) 985-998, https://doi.org/10.20964/2019.01.80.

[67]
S. Haddou, K. Zaidi, O. Dagdag, A. Hbika, M. Adil Mahraz, M. Bouhrim, A.S. Alqahtani, O.M. Noman, H. Kim, A. Aouniti, B. Hammouti, A. Chahine, Theoretical and electrochemical evaluation of Cannabis sativa L. extracts as corrosion inhibitors for mild steel in acidic medium, ChemistryOpen ( 2024) e202400273, https://doi.org/10.1002/open.202400273.

[68]
M.Y. Chaudhary, S. Yadav, P. Bansal, Y.S. Sharma, M. Gautam, C. Chandra, M. Gupta, Towards sustainable corrosion inhibition: a combined experimental and computational study of ethyl triphenyl phosphonium iodide on aluminium in acidic medium, Sustain. Chem. Environ. 9 ( 2025) 100221, https://doi.org/10.1016/j.scenv.2025.100221.

[69]
M. Zhang, L. Guo, M. Zhu, K. Wang, R. Zhang, Z. He, Y. Lin, S. Leng, V. Chikaodili Anadebe, X. Zheng, Akebia trifoliate koiaz peels extract as environmentally benign corrosion inhibitor for mild steel in HCl solutions: integrated experimental and theoretical investigations, J. Ind. Eng. Chem. 101 ( 2021) 227-236, https://doi.org/10.1016/j.jiec.2021.06.009.

[70]
K.G. Prajapati, P.S. Desai, B.B. Parmar, A.M. Patel, Comprehensive study on the corrosion inhibition of aluminum in HCl by N1, N1 '-(ethane-1,2-diyl)di(ethane-1,2-diamine): experimental and theoretical approaches, Results Surf. Interfaces 17 ( 2024) 100347, https://doi.org/10.1016/j.rsurfi.2024.100347.

[71]
C. Verma, H. Lgaz, D.K. Verma, E.E. Ebenso, I. Bahadur, M.A. Quraishi, Molecular dynamics and Monte Carlo simulations as powerful tools for study of interfacial adsorption behavior of corrosion inhibitors in aqueous phase: a review, J. Mol. Liq. 260 ( 2018) 99-120, https://doi.org/10.1016/j.molliq.2018.03.045.

[72]
W. Daoudi, A. El Aatiaoui, O. Dagdag, K. Zaidi, R. Haldhar, S.-C. Kim, A. Oussaid, A. Aouinti, A. Berisha, F. Benhiba, E.E. Ebenso, A. Oussaid, Anti-Corrosion coating formation by a biopolymeric extract of artemisia herba-alba plant: experimental and theoretical investigations, Coatings 13 (3) ( 2023) 611, https://doi.org/10.3390/coatings13030611.

[73]
S. Nabatipour, S. Mohammadi, A. Mohammadi, Synthesis and comparison of two chromone based schiff bases containing methoxy and acetamido substitutes as highly sustainable corrosion inhibitors for steel in hydrochloric acid, J. Mol. Struct. 1217 ( 2020) 128367, https://doi.org/10.1016/j.molstruc. 2020. 128367.

[74]
N. Benzbiria, A. Thoume, Z.A. El Caid, S. Echihi, A. Elmakssoudi, A. Zarrouk, M. Zertoubi, An investigation on the utilization of a synthesized benzodiazepine derivative as a corrosion inhibitor for carbon steel in sulfuric solution: chemical and electrochemical synthesis, surface analysis (SEM/AFM), DFT and MC simulation, Colloids Surf. A Physicochem. Eng. Asp. 681 ( 2024) 132744, https://doi.org/10.1016/j.colsurfa.2023.132744.

[75]
M.Y. Chaudhary, M. Gupta, P. Bansal, Y.S. Sharma, N. Dheer, A. Kant,... S. Yadav, Allyl triphenyl phosphonium bromide, an ionic liquid as an eco-friendly and Green inhibitor for corrosion of aluminium in hydrochloric acid: mechanistic insights and experimental validation, Sustain. Chem. Environ. ( 2025) 100206, https://doi.org/10.1016/j.scenv.2025.100206.

[76]
V.V. Mehmeti, A.R. Berisha, Corrosion study of mild steel in aqueous sulfuric acid solution using 4-methyl-4H-1,2,4-triazole-3-thiol and 2-mercaptonicotinic acid—an experimental and theoretical study, Front. Chem. 5 ( 2017) 61, https://doi.org/10.3389/fchem.2017.00061.

[77]
Y. Fernine, N. Arrousse, R. Haldhar, C.J. Raorane, E. Ech-Chihbi, S.C. Kim,... M. Taleb, Novel thiophene derivatives as eco-friendly corrosion inhibitors for mild steel in 1 m HCl solution: characterization, electrochemical and computational (DFT and MC simulations) methods, J. Environ. Chem. Eng. 10 (6) ( 2022) 108891, https://doi.org/10.1016/j.jece.2022.108891.

[78]
E.B. Caldona, M. Zhang, G. Liang, T.K. Hollis, C.E. Webster, D.W. Smith Jr, D.O. Wipf, Corrosion inhibition of mild steel in acidic medium by simple azolebased aromatic compounds, J. Electroanal. Chem. 880 ( 2021) 114858, https://doi.org/10.1016/j.jelechem.2020.114858.

[79]
Gupta M., Bhrara K., & Singh G. Artigo Corros. Prot. Mater., Vol. 31, No1 ( 2015).

[80]
O. Dagdag, Z. Safi, H. Erramli, O. Cherkaoui, N. Wazzan, L. Guo, A. El Harfi, Adsorption and anticorrosive behavior of aromatic epoxy monomers on carbon steel corrosion in acidic solution: computational studies and sustained experimental studies, RSC Adv. 9 (26) ( 2019) 14782-14796, https://doi.org/10.1039/C9RA01672D.

[81]
D.M. Mamand, H.M. Qadr, Corrosion inhibition efficiency and quantum chemical studies of some organic compounds: theoretical evaluation, Corros. Rev. 41 (4) ( 2023) 427-441, https://doi.org/10.1515/corrrev-2022-0034.

[82]
H.H. Rasul, D.M. Mamad, Y.H. Azeez, R.A. Omer, K.A. Omer, Theoretical investigation on corrosion inhibition efficiency of some amino acid compounds, Comput. Theor. Chem. 1225 ( 2023) 114177, https://doi.org/10.1016/j.comptc.2023.114177.

[83]
A. Thakur, S. Kaya, A.S. Abousalem, A. Kumar, Experimental, DFT and MC simulation analysis of vicia sativa weed aerial extract as sustainable and eco-benign corrosion inhibitor for mild steel in acidic environment, Sustain. Chem. Pharm. 29 ( 2022) 100785, https://doi.org/10.1016/j.scp.2022.100785.

[84]
K. Zaidi, N. Bouroumane, C. Merimi, A. Aouiniti, R. Touzani, A. Oussaid, S.M. Ibrahim, Iron-ligand complex, an efficient inhibitor of steel corrosion in hydrochloric acid media, J. Mol. Struct. 1284 ( 2023) 135434, https://doi.org/10.1016/j.molstruc.2023.135434.

[85]
M.Y. Chaudhary, M. Gupta, Y.S. Sharma, P. Bansal, S. Kaushik, R. Kanojia,... S. Yadav, Benzyl triphenyl phosphonium bromide as a corrosion inhibitor: a multifaceted study on aluminium protection in acidic environment, J. Ion. Liq. ( 2025) 100152, https://doi.org/10.1016/j.jil.2025.100152.

[86]
S. Pour-Ali, S. Hejazi, Tiazofurin drug as a new and non-toxic corrosion inhibitor for mild steel in HCl solution: experimental and quantum chemical investigations, J. Mol. Liq. 354 ( 2022) 118886, https://doi.org/10.1016/j.molliq.2022.118886.

[87]
O. Oyeneyin, D. Akerele, N. Ojo, O. Oderinlo, Corrosion inhibitive potentials of some 2 H -1-benzopyran-2-one derivatives-DFT calculations, Biointerface Res. Appl. Chem. 11 (6) ( 2021) 13968-13981, https://doi.org/10.33263/BRIAC116.1396813981.

[88]
D.M. Iamand, Y.H. Azeez, H.M. Qadr, Monte Carlo and DFT calculations on the corrosion inhibition efficiency of some benzamide molecules, Mong. J. Chem. 24 (50) ( 2023) 1-10, https://doi.org/10.5564/mjc.v24i50.2435.

[89]
F. Iorhuna, S.M. Adulfatah, A.M. Ayuba, Quinazoline derivatives as corrosion inhibitors on aluminum metal surface: a theoretical study, Adv. J. Chem. Sect. A 6 (1) ( 2023) 71-84.

[90]
O. Moumeni, M. Mehri, R. Kerkour, A. Boublia, F. Mihoub, K. Rebai, Y. Benguerba, Experimental and detailed DFT/MD simulation of α-aminophosphonates as promising corrosion inhibitor for XC48 carbon steel in HCl environment, J. Taiwan Inst. Chem. Eng. 147 ( 2023) 104918, https://doi.org/10.1016/j.jtice.2023.104918.

[91]
S.K. Saha, P. Ghosh, A. Hens, N.C. Murmu, P. Banerjee, Density functional theory and molecular dynamics simulation study on corrosion inhibition performance of mild steel by mercapto-quinoline schiff base corrosion inhibitor, Phys. E Lowdimensional syst. Nanost. 66 ( 2015) 332-341, https://doi.org/10.1016/j.molliq.2016.09.110.

[92]
N.N. Hau, D.Q. Huong, Effect of aromatic rings on mild steel corrosion inhibition ability of nitrogen heteroatom-containing compounds: experimental and theoretical investigation, J. Mol. Struct. 1277 ( 2023) 134884, https://doi.org/10.1016/j. molstruc.2022.134884.

[93]
M.E. Belghiti, S. Bouazama, S. Echihi, A. Mahsoune, A. Elmelouky, A. Dafali,... M. Tabyaoui, Understanding the adsorption of newly Benzylidene-aniline derivatives as a corrosion inhibitor for carbon steel in hydrochloric acid solution: experimental, DFT and molecular dynamic simulation studies, Arab. J. Chem. 13 (1) ( 2020) 1499-1519, https://doi.org/10.1016/j.arabjc.2017.12.003.

[94]
K. Belal, A.H. El-Askalany, E.A. Ghaith, et al., Novel synthesized triazole derivatives as effective corrosion inhibitors for carbon steel in 1 M HCl solution: experimental and computational studies, Sci. Rep. 13 ( 2023) 22180, https://doi.org/10.1038/s41598-023-49468-5.

[95]
Ahmed Al-Amiery, Comprehensive evaluation of 5-imino-1,2,4-dithiazolidine-3thione as a corrosion inhibitor for mild steel in hydrochloric acid solution, Sci. Rep. ( 2025), https://doi.org/10.1038/s41598-025-95104-9.

[96]
M. El Faydy, A. Barrahi, N. Timoudan, I. Warad, Z. Safi, N. Wazzan,... A. Zarrouk, Corrosion inhibition and adsorption behavior of two novel quinolin-8-ols on carbon steel surface in HCl : synthesis, electrochemical, surface characterization, and quantum chemical approaches, Can. Metall. Q. ( 2025) 1-23, https://doi.org/10.1080/00084433.2025.2508104.

[97]
A. Ech-chebab, M. Missioui, Y. Zaoui, O. Dagdag, H. Kim, A. Berisha, M. Ebn Touhami, Experimental and theoretical evaluation of a novel quinoxaline derivative as a corrosion inhibitor for mild steel in 1.0 M HCl, Can. Metall. Q. ( 2025) 1-16, https://doi.org/10.1080/00084433.2025.2522556.

[98]
Kabiru Haruna, Othman Hamouz, Tawfik Saleh, The corrosion inhibition performance of a diissocyanate-imidazole based organic compound during acid cleaning of MSF desalination plant, Heliyon 10 ( 2024) e38116, https://doi.org/10.1016/j.heliyon.2024.e38116.

[99]
Reham Wahba, Adel El-Sonbati, Mostafa Diab, Esam Gomaa, Marwa El-Nahass, Y. Abdallah, Electrochemical corrosion performance of N80 steel in acidized 10% HCl medium using 4-Methyl-1-Phenyl-3-(p-tolyldiazenyl) -2,3-Dihydro-1H-Pyrrol-2-ol, Heliyon 11 ( 2025) e42317, https://doi.org/10.1016/j.heliyon.2025.e42317.

[100]
L. Adlani, Nisrine Benzbiria, Abderrahim Titi, N. Timoudan, Fouad Benhiba, Ismail Warad, G. Kaichouh, Rachid Touzani, H. Zarrok,Burak Dikici, H. Oudda,

Abdelkader A. Zarrouk, Contact Zarrouk, Performance of a new pyrazole derivative in 1 m HCl on the corrosion of carbon steel: experimental, quantum chemical and molecular dynamics simulation studies, J. Dispers. Sci. Technol. ( 2024), https://doi.org/10.1080/01932691.2024.2304641.

[101]
Hassan Hammud, Nadeem Sheikh, Ihab Shawish, Hawra Bukhamsin, Dolayl AlHudairi, Angelina Wee, Malai Hamid, Sarah Maache, Hessa Al-Rasheed, Assem Barakat, Ayman El-Faham, Hany El-Lateef, Bis(dimethylpyrazolyl)-ani-line-s-triazine derivatives as efficient corrosion inhibitors for C-steel and computational studies, R. Soc. Open Sci. ( 2024) 11, https://doi.org/10.1098/rsos.231229.

[102]
N. Mohanapriya, M. Kumaravel, B. Lalithamani, Theoretical and experimental studies on the adsorption of n-[(E)-pyridin-2-ylmethylidene] aniline, a schiff base, on mild steel surface in acid media, J. Electrochem. Sci. Technol. 11 (2) ( 2020) 117-131, https://doi.org/10.33961/jecst.2019.00430.

Outlines

/