Next Article in Journal
Application of Bioinspired Structural Ceramics with High-Temperature Electrical Insulation and High Adhesion in K-Type Coaxial Thermocouples
Previous Article in Journal
Effect of a Polyhexanide-Based Antiseptic Composition on Dentin Microhardness and Mechanical Properties: An In Vitro Study
Previous Article in Special Issue
Impact of NaHCO3/Na2CO3 Buffer Reagent on Mitigating the Corrosion of C110 Steel in Water-Based Annulus Protection Fluid at Ultrahigh Temperature
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Towards a Sustainable Material Protection: Olanzapine Drugs and Their Derivatives as Corrosion Inhibitors for C1018 Steel in 1 M Hydrochloric Acid

1
Chemistry Department, College of Science, Imam Abdulrahman Bin Faisal University, P.O. Box 76971, Dammam 31441, Saudi Arabia
2
Interdisciplinary Research Center for Advanced Materials (IRC-AM), King Fahd University of Petroleum & Minerals (KFUPM), Dhahran 31261, Saudi Arabia
3
Department of Pharmaceutical Chemistry, College of Clinical Pharmacy, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
4
Department of Chemistry, King Fahd University of Petroleum and Minerals, Dhahran 31261, Saudi Arabia
5
SPIMACO Group, Operations, Dammam Pharma Plant, Plant Technical, Dammam 31441, Saudi Arabia
6
Basic and Applied Scientific Research Center, Imam Abdulrahman Bin Faisal University, P.O. Box 1982, Dammam 31441, Saudi Arabia
7
Interdisciplinary Research Center for Construction and Building Materials, King Fahd University of Petroleum & Minerals, Dhahran 31261, Saudi Arabia
8
Department of Mechanical Engineering, King Fahd University of Petroleum and Minerals (KFUPM), Dhahran 31261, Saudi Arabia
*
Authors to whom correspondence should be addressed.
Materials 2025, 18(12), 2902; https://doi.org/10.3390/ma18122902
Submission received: 6 May 2025 / Revised: 10 June 2025 / Accepted: 13 June 2025 / Published: 19 June 2025
(This article belongs to the Special Issue Advances in Corrosion and Protection of Metallic Materials)

Abstract

:
This study investigates the synthesis process and characterization methods and evaluates the inhibition behavior of olanzapine (2-methyl-4-(4-methyl-1-piperazinyl)-10H-thieno-[2,3-b] 1,5]benzodiazepine (OLZ)) and its derivatives, such as 3-(2-methyl-4-(4-methylpiperazin-1-yl)-10H-benzo[b]thieno[2,3-e] [1,4]diazepin-10-yl) propenamide (OLZ1) and Ethyl 2-(2-methyl-4-(4-methylpiperazin-1-yl)-10H-benzo[b]thieno[2,3-e][1,4]diazepin-10 yl) acetate (OLZ2) for carbon steel (C1018) in a 1 M HCl acidic solution. Fourier Transform Infrared Spectroscopy (FTIR) and Nuclear Magnetic Resonance (NMR) were employed to verify their molecular structures and functional groups, which characterized the derivatives after synthesis. Their corrosion inhibition potential for C1018 steel in acidic media was estimated by weight loss (WL) and electrochemical techniques, such as electrochemical impedance spectroscopy (EIS), linear polarization resistance (LPR), and potentiodynamic polarization (PDP), accompanied by surface analysis methods. The findings revealed that all three derivatives demonstrated exceptional inhibition performance, achieving maximum efficiencies of 88.83%, 91.20%, and 91.82% for OLZ, OLZ1, and OLZ2 at 300 ppm, respectively. Weight loss experiments across different temperatures further explored their inhibitory behavior. Although inhibition efficiency decreased with a temperature increase to 318 K, the derivatives still displayed notable performance, with maximum efficiencies of 74.75% for OLZ, 81.63% for OLZ1, and 79.44% for OLZ2. Polarization studies identified the corrosion inhibition mechanisms as an anodic type. Surface characterization of the C1018 steel coupons, both with and without the inhibitors, was performed using FTIR and scanning electron microscopy (SEM) combined with energy-dispersive X-ray spectroscopy (EDX). These analyses indicated the creation of a protective inhibitor layer on the carbon steel surface, reducing corrosion in the acidic environment. Overall, this study underscores the potential of these drug derivatives as corrosion inhibitors, combining structural insights and performance assessments to support their industrial application.

1. Introduction

Corrosion has been identified as a major challenge in industry due to the harsh environmental conditions that prevail in Saudi Arabia. Acidic corrosion is the most common in industrial fields, such as oil and gas, chemical processing, and metal manufacturing. In 2022, for example, it accounted for 32.4% of Saudi Arabia’s GDP. However, corrosion remains a costly challenge, with global losses estimated at USD 2.5 trillion annually, or 3.4% of the world’s GDP, according to NACE International [1]. In the U.S. oil and gas industry alone, corrosion costs approximately USD 1.4 billion annually in exploration and production, with additional expenses of USD 7 billion for monitoring, maintenance, and replacement in gas and liquid transmission pipelines [1,2].
Developing effective corrosion inhibitors is fundamental to minimizing these major economic losses. Organic compounds with heteroatoms such as oxygen, nitrogen, and sulfur, together with π-electron systems, represent the main components of conventional corrosion inhibitors, which improve their adsorption onto metal surfaces to boost protective performance [3,4]. However, the environmental toxicity associated with many traditional inhibitors has necessitated a paradigm shift toward sustainable alternatives that maintain high inhibition efficiency while minimizing ecological impact.
Recent research has directed attention toward the potential repurposing of pharmaceutical substances as corrosion inhibitors, particularly those containing N-heterocyclic structures. These compounds demonstrate strong potential to replace traditional inhibitors because they can easily decompose in the environment while showing reduced toxicity and cost-effectiveness [5].
The inherent molecular architecture of pharmaceuticals contains multiple heteroatoms, aromatic rings, and functional groups that allow these substances to create protective films on metals through adsorption reactions and metal ion complexation [6,7]. This structural versatility allows them to provide comprehensive protection against various corrosion manifestations, including uniform, pitting, and crevice corrosion [7]. Several studies have reported the use of drug compounds as corrosion inhibitors containing heteroatoms for inhibiting metal corrosion in an acidic medium (Table 1).
Olanzapine, a typical antipsychotic medication widely prescribed for the treatment of schizophrenia and bipolar disorder, represents a promising candidate for repurposing as a corrosion inhibitor. As a selective antagonist of 5HT2A/D2 receptors, olanzapine has received FDA authorization for multiple psychiatric applications and is listed among the World Health Organization’s essential medicines [8]. The number of annual U.S. prescriptions for olanzapine exceeds 3 million, which positions the medication among the most prescribed drugs [9,10]. The two-year shelf life of olanzapine results in large amounts of expired medication that must be disposed of while also creating opportunities to find sustainable uses for the waste.
Table 1. Comparison of different N-heterocyclic drugs that are used as corrosion inhibitors with the results from our study.
Table 1. Comparison of different N-heterocyclic drugs that are used as corrosion inhibitors with the results from our study.
No.Drug NameMetal/AlloyCorrosive MediumUsed MethodsMax IE%Ref.
1Helicure drugCarbon Steel1 M HClWL + EIS + EFM85.8% (at 300 ppm)[11]
2Spironolactone drugn C38 Carbon Steel10% HClWL + EIS + PDP95.8% (at 7.2 × 10−3 M)[12]
3AtenololCarbon Steel1 M HClWL + EIS + PDP93.8% (at 300 ppm)[13]
4IrbesartanCarbon Steel1 M HClOCP + EIS + LPR + PDP + CV94% at 300 mg L−1[14]
5MetoprololSteel alloy (st 37)1 M HClEIS + PDP87% (300 ppm)[15]
6HydralazineCarbon Steel1 M HClWL + EIS + PDP91.44% (1250 ppm at 323 K)[16]
7CiprofloxacinCarbon Steel1 M HCl WL + EIS69.99% (at 0.7% v/v%)[17]
8DoxofyllineSoft Steel1 M HClWL + EIS72.6% (200 ppm)[18]
9IrnocamCarbon Steel1 M HClWL + EIS75.4% (at 600 ppm)[19]
10LumeraxCarbon Steel1 M HClEIS + PDP97.6% (100 mg/L)[20]
11MeloxicamCarbon Steel1 M HClEIS + PDP80.8% (30 ppm)[21]
12OrnidazoleCarbon Steel1 M HClWL + EIS + PDP71.61% (2 × 10−5 M)[22]
13TramadolCarbon Steel1 M HClEIS + PDP97.2% (100 ppm)[23]
14DulcolaxCarbon Steel1 M HClEIS + PDP + AAS92.5% (500 ppm)[24]
15GentamicinCarbon Steel1 M HClWL + EIS + PDP75.06% (0.9% v/v)[25]
16NorfloxacinCarbon Steel1 M HClWL + EIS + PDP89.8%[26]
17OlanzapineCarbon Steel1 M HClLPR92.47Our work
The molecular structure of olanzapine features multiple nitrogen atoms within heterocyclic rings, aromatic systems with delocalized π-electrons, and functional groups capable of interacting with metal surfaces. These characteristics align with established criteria for effective corrosion inhibitors.
Furthermore, the possibility of synthesizing derivatives with enhanced inhibition properties through strategic molecular modifications offers additional avenues for optimizing performance while maintaining environmental compatibility. The convergence of increasing industrial demand for corrosion inhibitors, growing environmental concerns regarding conventional inhibitors, and the global imperative for sustainable resource utilization underscores the significance of investigating pharmaceutical compounds as potential corrosion inhibitors. This research direction not only addresses critical industrial challenges but also contributes to circular economic principles by transforming pharmaceutical waste into valuable industrial resources.
The present study aims to comprehensively investigate the corrosion inhibition efficacy of olanzapine and two novel synthesized derivatives on C1018 carbon steel in 1 M HCl solution. The research methodology encompasses advanced electrochemical techniques, immersion tests, computational calculations, and surface characterization using FTIR spectroscopy and SEM/EDX analysis to elucidate the inhibition mechanisms and quantify protective performance. By establishing structure–activity relationships and optimizing inhibition parameters, this work seeks to develop environmentally benign corrosion inhibitors derived from pharmaceutical compounds, thereby addressing both industrial corrosion challenges and environmental sustainability imperatives.

2. Experimental Section

2.1. Synthesis of the Corrosion Inhibitors

D-1 (OLZ 1) and D-2 (OLZ 2) were synthesized in neat conditions and under microwave irradiation (Figure 1). The compounds were synthesized through the reaction of olanzapine with acrylamide and ethyl chloroacetate to give D-1 and D-2, respectively, and were recrystallized from acetone. Microwave irradiations were conducted in an Anton Paar Monowave 300 single-mode microwave in 30 mL Pyrex [27,28]. The progress of the reactions was monitored using Merck silica gel 60F-254 thin-layer plates. The compounds were characterized using 1H NMR and FTIR. The 1HNMR spectra were recorded at 500 MHz (Bruker) and IR spectra were recorded on an FTIR (02) spectrophotometer (PerkinElmer, Bruker). The 1HNMR and IR spectra are presented in the Supplementary Materials (Figures S1 and S2) [29].

2.2. Weight Loss Measurements

The weight loss or immersion test is one of the primary tests used to estimate the inhibition efficiency of a substance. The sample was prepared, weighed, and then immersed in 250 mL of the solution in a jar with a lid, in the absence and presence of the optimum concentration for each inhibitor (300 ppm), for 24 h. After that, C1018 coupons were taken out and treated by (pickling) to remove oxide contaminants formed on the surface in atmospheric oxygen when the metal is exposed to air [30]. After that, the corrosion rate was calculated [31,32] as follows:
C o r r o s i o n   R a t e m m y = 87,600 W A T D
C o r r o s i o n   R a t e m p y = 3.45 × 10 6 × w A T D
I n h i b i t i o n   E f f i c i e n c y = C R ° C R i C R ° × 100
where w is the coupon weight loss in grams, A is the coupon surface area in cm2, T is the immersion period in 24 h, D is the coupon density in g/cm3, CR° is the corrosion rate of the solution without an inhibitor, and CRi is the corrosion rate of the solution with an inhibitor.

2.3. Electrochemical Measurements

The electrochemical evaluation was carried out using a standard three-electrode corrosion cell consisting of carbon steel (C1018) as the working electrode, silver/silver chloride electrode (Ag/AgCl) as the reference electrode, and graphite rod as the counter electrode. The electrodes were immersed in 300 mL of 1 M HCl solution. Gamry reference 3000 potentiostat/galvanostat/ZRA (Gamry Instrument, USA) was used to perform all the electrochemical measurements. Before testing, the working electrode was immersed in the test solution for 1 h to achieve a stable open circuit potential (OCP). All electrochemical experiments were performed in static, aerated solutions at room temperature. Electrochemical impedance spectroscopy (EIS), linear polarization resistance (LPR), and potentiodynamic polarization (PDP) tests were conducted to assess the corrosion behavior of carbon steel with and without inhibitors. EIS measurements were conducted over a frequency range of 100 kHz to 10 mHz with an AC amplitude of ±10 mV versus OCP. This was followed by LPR tests using a potential scan of ±10 mV relative to the OCP at a rate of 0.125 mV/s. Lastly, PDP tests were carried out by scanning the system at 0.5 mV/s from −250 mV in the cathodic direction to 250 mV in the anodic direction relative to the OCP. The Gamry Echem Analyst 7.9.0 software package was used for data analysis, fitting, and simulations. To evaluate the corrosion rate and inhibition efficiency, all experiments were performed three times to achieve high precision and reliability of the results.

2.4. Computational Methods

2.4.1. DFT Calculations

The DFT calculations for the four olanzapine molecules were performed using the Dmol3 module, Biovia Material Studio [33]. The GAA/BLYP functional, combined with the DNP basis sets to expand the Kohn−Sham orbitals with a confining cutoff set at 3.8 Å, was employed in the calculations. The COSMO method was applied to simulate the solvent effect of water [34]. The convergence tolerance was set to fine. Based on the frontier orbital theory, the electronic parameters in terms of the highest occupied molecular orbital (EHOMO), energy of the lowest unoccupied molecular orbital (ELUMO), energy gap (ΔE), electron affinity (A), ionization potential (I), electronegativity (χ), global hardness (η), and the fraction of electron transfer (ΔN) were calculated according to Equations (4)–(9):
Δ E = E L U M O E H O M O
I = E H O M O
A = E L U M O
χ = E H O M O + E L U M O 2 = I + A 2
η = E H O M O E L U M O 2 = I A 2
Δ N = χ F e χ I n η F e η I n
where χFe and ηFe represent the absolute electronegativity and hardness of iron, which have the theoretical values of 0 and 7 eV/mol, respectively [35].
The geometry optimization process was conducted with the following criteria: the convergence tolerance of the energy, maximum force, and maximum displacement were set at 1 × 10−5 Ha, 2 × 10−3 Ha Å−1, and 5 × 10−3 Å, respectively, while the SCF convergence tolerance was set at 1 × 10−6 and the real space cutoff radius at 4.0 Å.

2.4.2. MD Simulations

A molecular dynamics (MD) simulation was conducted to obtain the interaction energy between the Fe surface and the inhibitor molecules. The Fe(110) crystallographic surface was selected for its stability and abundance of active sites [36]. A supercell of the surface scaled to 10 × 10 times its original size was created and optimized. The interaction models comprised three layers: Fe(110), a solvent with the molecule, and a 30 Å vacuum. The solvent layer was designed to simulate 1 M HCl and included 491 H2O, 9 Cl, 9 H3O+, and 1 inhibitor molecule. All MD simulations were performed using the Forcite module and the COMPASS force field. The models underwent dynamic runs under the NVT ensemble at 298 K, controlled by the Andersen thermostat. The simulation time was 500 ps with a 1 fs time step. The interaction energy was then calculated based on Equation (10):
E i n t e r a c t i o n = E t o t a l E s u r f a c e + s o l u t i o n + E i n h i b i t o r
where E t o t a l is the total energy of the entire system; E s u r f a c e + s o l u t i o n is the energy of the surface with the solution; and E i n h i b i t o r is the energy of the inhibitor molecule only. The negative magnitude of interaction energy gives the values of binding energy ( E b i n d i n g ), as expressed in Equation (11).
E b i n d i n g = E i n t e a c t i o n
Several simulation models were constructed to study the diffusion of corrosive particles across the inhibitor films. These models include a blank model containing 180 H2O, 5 Cl, and 5 H3O+, as well as inhibited models with 10 molecules. The built models were optimized using the COMPASS forcefield. Subsequently, dynamics runs were performed (i.e., equilibrium and production runs) under the NVT ensemble at room temperature (298 K). The conformations generated from the MD simulations were used to track the corrosive particles (H2O, Cl, and H3O+) using the mean square displacement (MSD), as expressed in Equation (12). The diffusion coefficients (D) of the species were then calculated by fitting the MSD–time curves according to Einstein’s law of diffusion (Equation (13)) [37,38]:
r 2 t = r i t r i 0 2
D = 1 6 N lim t d d t i = 1 N r 2 t
where r 2 t is the MSD at time t; N represents the total number of particles; and r i t and r i 0 denote the position of a particle i at time t and at the initial time, respectively. The periodic boundary conditions were applied to all surface–molecule interaction models. The Ewald accuracy, cutoff distance, and spline width were set at 1 × 10−4 Kcal mol−1, 15.5 Å, and 1 Å, respectively. Prior to the production run, an equilibration run was executed at 200 ps NPT.

2.5. Surface Analyses

2.5.1. FTIR Analysis

FTIR is a spectroscopic technique that can provide critical information on the interactions between inhibitor molecules and the substrate by detecting changes in the vibrational frequencies of chemical bonds. This information helps us understand the corrosion inhibition mechanism [39]. A small amount of the powder sample was placed directly onto the ATR crystal, without the need for KBr. Gentle pressures were applied to ensure optimal contact between the sample and the crystal. The Shimadzu—IRAFFINITY-2 instrument was used for the FTIR analysis.

2.5.2. Scanning Electron Microscopy (SEM)

The morphology of the C1018 carbon steel surface exposed to uninhibited and inhibited 1 M hydrochloric acid was analyzed using scanning electron microscopy coupled with an energy-dispersive X-ray analyzer (VEGA 3 TESCAN coupled with AMETEK EDX) operated at 20 kV with a 10 mm working distance. SEM can provide information on the morphology and the type of corrosion on the metal surface, while EDX is used for elemental analysis [40].

3. Results and Discussion

3.1. Weight Loss Measurements

Weight loss measurement was conducted to evaluate the performance of the corrosion inhibitors at 298 K and 318 K. Table 2 and Figure 2 present the inhibition efficiency values for OLZ, OLZ1, and OLZ2 at these two temperatures. At 298 K, the inhibition efficiencies were 92.98%, 90.82%, and 96.23% for OLZ, OLZ1, and OLZ2, respectively, indicating a stronger interaction of the inhibitor molecules at lower temperatures [41,42]. However, at the elevated temperature of 318 K, the inhibition efficiencies decreased to 74.75%, 81.63%, and 79.44% for OLZ, OLZ1, and OLZ2, respectively. The decrease in inhibition efficiency with increasing temperature suggests the desorption of the inhibitor molecules from the metal surface. Interestingly, OLZ1 exhibited the smallest decrease in inhibition efficiency with temperature, suggesting that the acrylamide group enhances the thermal stability of the adsorbed inhibitor film. This could be attributed to stronger interactions between the acrylamide group and the metal surface, possibly through coordination bonds involving the amide nitrogen and oxygen atoms.

3.2. Open Circuit Potential Measurements

Figure 3 presents the OCP versus time plots for the blank solution and solutions containing 300 ppm of OLZ, OLZ1, and OLZ2. In the absence of inhibitors (blank), the OCP stabilized at approximately −0.47 V vs. Ag/AgCl, indicating active corrosion of the carbon steel substrate. Upon the addition of the inhibitors, a notable shift toward more positive (less negative) potential was observed. The OCP values stabilized at approximately −0.40 V, −0.42 V, and −0.44 V for OLZ, OLZ1, and OLZ2, respectively. This positive shift in OCP suggests that all three compounds predominantly affect the anodic reaction of the corrosion process [43]. The OCP measurements also revealed that potential stabilization occurred within the first 1000 s of immersion for all systems, after which minimal fluctuations were observed for the remainder of the 3500 s test period. This indicates the rapid formation of a protective film on the metal surface in the presence of inhibitors [44].

3.3. Electrochemical Impedance Spectroscopy (EIS)

EIS was employed to analyze the corrosion inhibition on C1018 in 1 M HCl solution, both with and without olanzapine and derivatives (OLZ, OLZ1, and OLZ2). This method reveals both the electrochemical reaction rates and the surface properties of the examined systems. Nyquist plots reveal the corrosion of steel in 1 M HCl solution that consists of semicircles with their centers below the real axis [16], as shown in Figure 4. These characteristics of charge transfer-controlled dissolution with significantly increased diameters in the presence of inhibitors indicate enhanced interfacial charge transfer resistance at the metal/electrolyte interface.
The Nyquist plots for the blank and OLZ, OLZ1, and OLZ2 inhibited solutions presented four semicircular capacitive arcs between high- and low-frequency regions; the high-frequency time constant (Rf//CPEf) represented a defective/porous protective OLZ and its derivatives film on the metal surface while the low-frequency time constant (Rct//CPEdl) represented the double layer formation at the metal/electrolyte interface [45,46]. The Bode plots (Figure 5) further support these findings, showing increased impedance magnitude |Z| and phase angle shifts toward more negative values in the presence of inhibitors. The phase angle plots exhibit a single time constant, which broadens in the presence of inhibitors, indicating the formation of a more compact and protective film on the metal surface [47,48].
The equivalent circuit model used to fit the EIS data is presented in Figure 4 (insert). The parameters include Rs, which is the solution resistance, CPEdl, which is the constant phase element for the double-layer, and Rct, which is the charge transfer resistance. The n values (ranging from 0.867 to 0.996) are less than unity, confirming the non-ideal capacitive behavior of the metal–solution interface. Moreover, higher n values indicate greater surface heterogeneity because of inhibitor adsorption. The observed inductive effects in the EIS data are related to the complex structure of the double layer at the metal–solution interface [49,50]. It is a key indicator of the interface’s electrical properties and inhibitor adsorption characteristics. These Cdl values were calculated from the impedance data by applying Equation (14) in Table 3.
C dl   =   ( Y º R c t 1 n ) 1 / n
The Cdl is the double-layer capacitance, Rct presents charge transfer, and Y0 is the CPE coefficient (reciprocal of impedance and also known as admittance).
The reduction in the electrical double layer value from 193.850 (blank) μF·cm−2 to (79.578, 65.471, 51.416) μF·cm−2 in the presence of OLZ and its derivatives (OLZ, OLZ1, and OLZ2) indicates an increase in the thickness of the electrical double layer, which retards the corrosion process as presented in Table 3. Also, this decrease in (Cdl) at the metal–solution interface with the presence of OLZ, OLZ1, and OLZ2 can result from a lower local dielectric constant, which leads to inhibitors adsorbed on the C1018 carbon steel. Quantitative analysis through non-linear least squares fitting to an equivalent electrical circuit revealed a huge increase in faradaic polarization resistance (Rp), from 140.897 Ω·cm2 for the uninhibited system to 1200.97, 1430.42, and 1646.42 Ω·cm2 for OLZ, OLZ1, and OLZ2, respectively, as shown in Table 3. Inhibition efficiency from the EIS technique was calculated utilizing the relationship defined in Equation (15). The experimental results consistently demonstrate that all three compounds exhibit excellent corrosion inhibition properties, with inhibition efficiencies exceeding 88% at 300 ppm. OLZ2 has the highest inhibition efficiency. This trend suggests that the structural modifications introduced in OLZ1 and OLZ2 enhance the corrosion inhibition properties of the parent molecule.
I E % = R p ( i n h ) R p ( b l a n k ) R p ( i n h ) × 100
IE is the corrosion efficiency, and Rp is the polarization resistance. The polarization resistance values without any inhibitor present are denoted by Rp (blank), while the values with an inhibitor present are denoted by Rp(inh).
The superior performance of OLZ2 is attributed to its ethyl acetate functionality, which provides additional oxygen donor atoms for coordinating covalent bonding with d-orbitals of iron and enhanced molecular hydrophobicity. The acrylamide group in OLZ1 showed intermediate performance because of its additional electron-rich centers and π-electron interactions with the metal substrate. The consistent inhibition hierarchy across multiple electrochemical techniques validates the reliability of these findings. This study created structure–activity relationships for olanzapine-based corrosion inhibitors and the results show their effectiveness as protective agents for carbon steel in aggressive acidic environments.

3.4. Potentiodynamic Polarization (PDP) Studies

PDP measurements were conducted to further elucidate the corrosion inhibition mechanism of the studied compounds. The polarization curves for carbon steel in 1 M HCl with and without 300 ppm of OLZ, OLZ1, and OLZ2 are shown in Figure 6. The electrochemical parameters of corrosion potential (Ecorr), corrosion current density (icorr), and inhibition efficiency (% IE) values were determined from the corresponding Tafel plots and the obtained data are shown in Table 4. The inhibition efficiencies can be calculated from the icorr values using the following Equation (16):
I E % = i c o r r ( b l a n k ) i c o r r ( i n h ) i c o r r ( b l a n k ) × 100
where icorr (blank) and icorr (inh) are the sample’s corrosion current density in 1 M HCl without and with inhibitors, respectively.
The corrosion current density (icorr) decreased significantly from 200.00 μA cm−2 for the blank solution to 20.100, 15.900, and 14.600 μA cm−2 for OLZ, OLZ1, and OLZ2, respectively. This reduction in icorr corresponds to inhibition efficiencies of 89.95%, 92.05%, and 92.70% for OLZ, OLZ1, and OLZ2, respectively, which is in agreement with the EIS results. The corrosion potential (Ecorr) shifted from −447.00 mV for the blank solution to −401.000, −397.00, and −422.000 mV for OLZ, OLZ1, and OLZ2, respectively. All three compounds act as anodic-type inhibitors, affecting the anodic metal dissolution.
The anodic (βa) and cathodic (βc) Tafel slopes were also modified in the presence of inhibitors. The anodic Tafel slope decreased from 160.80 mV/dec for the blank to 100.00, 89.000, and 91.400 mV/dec for OLZ, OLZ1, and OLZ2, respectively. Similarly, the cathodic Tafel slope increased from 120.00 mV/dec for the blank to 132.700, 218.600, and 241.600 mV/dec for OLZ, OLZ1, and OLZ2, respectively. These changes in Tafel slopes indicate that the inhibitors modify the mechanism of both anodic dissolution and cathodic hydrogen evolution, with a more pronounced effect on the anodic reaction. This anodic type of inhibition behavior is further confirmed by the potentiodynamic polarization results, which show that both the anodic branches of the polarization curves are affected by the presence of inhibitors. The changes in both the anodic (βa) and cathodic (βc) Tafel slopes indicate that the inhibitors modify the mechanism of both electrode reactions, resulting in a more pronounced increase in anodic Tafel slopes.

3.5. Linear Polarization Resistance (LPR)

A linear polarization resistance technique was employed to evaluate the efficiency of OLZ and its derivatives in inhibiting corrosion at 1 M HCl. This measurement enables real-time nondestructive corrosion rate evaluation as an electrochemical assessment tool that researchers commonly use to measure material corrosion rates. The ±10 mV is relative to OCP, which produces a direct relationship between corrosion current and potential values. A small polarization selection maintained an accurate measurement of corrosion rates because it avoided permanent disruption of the corrosion process [51,52].
Table 4 presents both polarization resistance (Rp) and efficiency values. The LPR parameters derived from the curve fitting appear in Table 4. The Ecorr values of inhibited solutions show more positive values than the blank acid solution, while the Ecorr values do not follow any specific pattern when using different inhibitors. The inhibition efficiency (% IE) of the OLZ inhibitors was calculated using the following Equation (18) [53]:
R p = β c × β a 2.3 β c + β a i c o r r
I E % = R p ( i n h ) R p ( b l a n k ) R p ( i n h ) × 100
where Rp is the polarization resistance. The polarization resistance values without any inhibitor present are denoted by Rp (blank), while the values with an inhibitor present are denoted by Rp (inh) and icorr, which is the corrosion current density. βa is the anodic and βc cathodic Tafel slope.
The results in Table 4 demonstrate that polarization resistance values increase after adding inhibitors compared to the acid solution without inhibitors. The addition of OLZ together with its derivatives proves effective at blocking corrosion in this extremely acidic environment, as the addition of OLZ and its derivatives produced identical patterns of decreased corrosion rates together with enhanced inhibition efficiencies. The polarization resistance (Rp) increased from 132.20 Ω for the blank solution to 1135.66 Ω, 1473.00 Ω, and 1760.66 Ω for OLZ, OLZ1, and OLZ2, respectively. The corresponding inhibition efficiencies are 88.35%, 91.01%, and 92.47% for OLZ, OLZ1, and OLZ2, respectively, which are in excellent agreement with the values obtained from EIS and PDP.

3.6. Quantum Chemistry Calculations

This section discusses the results of the quantum chemistry-based calculations. The molecular orbitals play a crucial role in determining the electron transfer process of an organic molecule [54,55]. The frontier molecular orbitals of the olanzapine derivatives are shown in Figure 7 and Figure 8. The HOMOs and LUMOs are concentrated on the OLZ-based molecules, except for the piperazine rings. These locations likely denote and accept electrons on the iron surface during the adsorption process [56]. The reactivity parameters of OLZ and its derivatives are presented in Table 5. The energy gap (ΔE) values of the HOMO-LUMO spectrum for OLZ, OLZ1, and OLZ2 were 2.778, 2.785, and 2.790 eV, respectively. A smaller gap indicates higher chemical reactivity, suggesting good inhibition efficiency [57]. In this study, all molecules showed almost similar ΔE values. These theoretical results matched well with the experimental findings based on the electrochemical and weight loss measurements. The dipole moment (μ) is another vital parameter in determining a molecule’s effectiveness as a corrosion inhibitor, as it reflects bond polarity. The molecules studied exhibited dipole moments higher than water (1.88 Debye) [58], indicating their potential to outcompete water molecules and adsorb onto the substrate at the metal–solution interface.
Fukui function based on the Mulliken analysis was performed to characterize the local reactivity of the molecules [59]. The results of the analysis display the fk and fk+ indices of all molecules. The highest fk values are localized around the nitrogen and sulfur atoms, suggesting these sites are more prone to electrophilic attack. The fk+ values are significant in the sulfur atoms in the thiazepine ring, indicating these sites are more prone to nucleophilic attack. These heteroatoms are likely involved in the adsorption process on the steel surface, facilitating the formation of a protective layer as presented in Figure 9.

3.7. Adsorption Behavior

The interaction of the molecular surface system was investigated using MD simulations to show the adsorption configurations of the inhibitor molecules on the Fe(110) surface in the presence of 1 M HCl, and the results are presented in Figure 10. The adsorption models revealed that the olanzapine derivative molecules were adsorbed on the substrate in a nearly parallel orientation. Although the overall geometry of olanzapine and its derivatives is non-planar due to their three-dimensional structures, their large size allows for more adsorption sites on the metal substrate. This is seen with the OLZ2 molecule, which achieved a more optimal adsorption mode than the others. The adsorption position is crucial for maximizing surface coverage and thereby enhancing the inhibition ability of the molecules [60,61]. Based on the generated dynamic models, the binding energy of all the molecules was calculated, with the results providing values of 173.21, 190.90, and 191.81 kcal/mol for OLZ-Fe(110), OLZ1-Fe(110), and OLZ2-Fe(110), respectively. OLZ2 exhibited the highest binding energy due to its adsorption configuration. These results are consistent with the DFT calculations as well as the experimental findings, which showed that OLZ2 demonstrated excellent inhibition performance in all aspects.
Figure 11 shows a linear correlation between the IE% and the binding energy derived from the MD simulations. Despite the limited dataset, the trend indicated that molecules with higher binding energies exhibited better inhibition performance. This is consistent with the concept that stronger adsorption of the inhibitor on the steel surface enhances protective behavior. The determination coefficient (R2 = 0.9048) confirmed a strong linear relationship, which reinforced the relevance of binding energy as a predictive descriptor in structure–activity relationships. The validity of this correlation is further supported by the fact that MD simulations account for both the inhibitor and solvent interactions, thereby making the correlation more representative of the experimental system.

3.8. Diffusion Study

The diffusion models of the blank and inhibited systems are shown in Figure 12 and Figure 13 and Table 6. In principle, the inhibition film should be able to prevent the migration of the species to the metal surface [62,63]. Thus, the MSD analysis was used to track the mobility of the particles in the different inhibitors’ layers, showing the double-log plots of the MSD–time curves for corrosive particles across the blank and inhibited films. The data indicated an increasing trend of MSD values over time, implying the diffusion region of the particles. In addition, the MSD curves for the inhibited systems were significantly lower than those for the blank systems, indicating that the inhibitor molecules restricted particle movement. Based on the MSD data, the diffusion coefficients of H2O, H3O+, and Cl were calculated, as presented in Table 6. The inhibitor molecules significantly decreased the D-values of H2O, H3O+, and Cl by around 85%, 83%, and 84%, respectively, on average. It was found that OLZ2 is most effective in reducing the diffusion of H2O (0.35 × 10−9 m2·s−1), while OLZ2 and OLZ were more effective for H3O+ (0.17 × 10−9 m2·s−1) and Cl (0.07 × 10−9 m2·s−1), respectively.

3.9. Surface Analysis

3.9.1. SEM-EDX Analysis

Scanning electron microscopy (SEM) coupled with energy-dispersive X-ray spectroscopy (EDX) was used to examine the surface morphology and elemental composition of the carbon steel samples before and after immersion in the test solutions [64,65]. The SEM images and EDX spectra are presented in Figure 14a,b. The SEM image of the bare metal surface shows a relatively smooth morphology with some polishing marks. After immersion in 1 M HCl (blank solution), the surface appears severely corroded with numerous pits and cracks, indicating an aggressive attack by the acidic medium. In contrast, the surfaces of samples immersed in inhibitor-containing solutions exhibit significantly less damage, with smoother morphologies suggesting the formation of protective films. The surface protection appears to be most effective with OLZ2, followed by OLZ1 and OLZ. The EDX analysis presented in Table 7 provides quantitative information on the elemental composition of the sample surfaces. The bare metal surface primarily consists of iron (Fe, 85.56%) with small amounts of carbon (C, 9.69%), oxygen (O, 1.69%), and other elements. After immersion in 1 M HCl, the iron content decreases dramatically to 62.71%, while the oxygen content increases to 20.76%, indicating severe oxidation of the metal surface. In the presence of inhibitors, the iron content is largely preserved (83.90%, 84.19%, and 85.49% for OLZ, OLZ1, and OLZ2, respectively), suggesting effective protection against corrosion. Furthermore, the detection of nitrogen (N) on the inhibited surfaces (1.74%, 1.68%, and 1.68% for OLZ, OLZ1, and OLZ2, respectively) provides direct evidence for the adsorption of inhibitor molecules, as nitrogen is a key element in the molecular structure of these compounds. The carbon content is also higher on the inhibited surfaces (11.41%, 10.75%, and 10.81% for OLZ, OLZ1, and OLZ2, respectively) compared to the bare metal (9.69%), further confirming the presence of organic inhibitor molecules. The oxygen content is significantly lower on the inhibited surfaces (1.62%, 1.78%, and 2.24% for OLZ, OLZ1, and OLZ2, respectively) compared to the corroded surface in HCl (20.76%), indicating reduced oxidation of the metal in the presence of inhibitors.
The surface analysis techniques of SEM-EDX provide direct evidence for the adsorption of inhibitor molecules on the metal surface and the formation of protective films. The SEM images clearly show the protective effect of the inhibitors, with significantly less surface damage compared to the blank solution. The smoother morphology of the inhibited surfaces suggests the formation of a uniform protective film that shields the metal from the aggressive medium. The EDX analysis confirms the presence of inhibitor molecules on the metal surface through the detection of nitrogen, which is a key element in the molecular structure of these compounds. The higher carbon content and lower oxygen content on the inhibited surfaces compared to the corroded surface in HCl further support the formation of a protective organic film that prevents oxidation of the metal.

3.9.2. FTIR Spectroscopy

FTIR spectroscopy was employed to characterize the molecular structure of the inhibitors and to identify potential functional groups involved in the adsorption process on the metal surface. The results of FTIR analysis of pure OLZ, OLZ1, and OLZ2 are presented in Figure 15. The spectra reveal characteristic absorption bands corresponding to various functional groups present in the inhibitor molecules. All three compounds exhibit bands in the region of 3200–3500 cm−1 attributed to N-H stretching vibrations, bands around 2900–3000 cm−1 due to C-H stretching, and bands in the region of 1600–1700 cm−1 assigned to C=O stretching vibrations. The spectra of OLZ1 and OLZ2 show additional bands compared to the parent OLZ, confirming the successful incorporation of acrylamide and ethyl acetate groups, respectively. The blank solution showed a peak at 3300 cm−1 that was due to O-H stretching, but it was absent in the presence of the inhibitors. The -OH, -C=N, C=C, and aromatic ring on the surface of the metal serve as active sites for the interaction of the inhibitors.

3.10. Inhibition Mechanism

Olanzapine and its derivatives represent a promising pharmaceutical compound in the field of corrosion inhibition, combining high efficacy with environmental sustainability. The molecular structure of olanzapine features multiple heterocyclic systems containing nitrogen and sulfur atoms, along with delocalized π-electron systems, providing multiple adsorption sites on metallic surfaces. These combined features are hardly common in most organic molecules used as corrosion inhibitors.
The analysis of corrosion inhibition depends on how inhibitors adsorb onto steel surfaces. Organic compounds adsorb to surfaces through physical or chemical methods, but both mechanisms can happen at the same time. The adsorption process depends on three main factors, which include inhibitor charge, molecular structure, and metal surface properties. The inhibitor adsorption process relies on functional groups such as aromatic rings (N-H, C=O, C=C, and C-O) and N present in OLZ and its derivatives, which act as adsorption centers. The electron transfer capability of these functional groups to metal d-orbitals plays a crucial role in the adsorption process [66].
FTIR spectral data delivers essential information about how the studied compounds inhibit the corrosion process, as shown in Figure 15 and Table 8. It is clear in Figure 15 that the inhibited samples show diminished O-H stretching bands at 3300 cm−1, which indicates that inhibitor molecules move water molecules off the metal surface to block their involvement in corrosion reactions. This displacement is a crucial step in the inhibition process, as water molecules are essential for the electrochemical reactions that drive corrosion.
The decreased intensity of Fe-O peaks at 490 cm−1 in the presence of inhibitors proves that inhibitors block the formation of iron oxides, which act as primary corrosion products. Inhibitor molecules adsorb on metal surface active sites to block both anodic and cathodic reactions that occur during corrosion. By impeding these electrochemical reactions, the inhibitors effectively reduce the corrosion rate.
FTIR spectra of corrosion products show that inhibitor-specific peaks persist, which demonstrates that inhibitors maintain their adsorption to metal surfaces throughout the testing duration. Multiple adsorption mechanisms operate based on the various functional groups found in inhibitor molecules [67,68]. The proposed mechanism in OLZ is shown in Figure 15, where the immersion of steel coupons in the blank HCl resulted in metal dissolution, whereas the addition of OLZ resulted in inhibitor adsorption on the steel surface through chemisorption and physisorption. The adsorption process receives additional support from electrostatic forces between charged inhibitor molecules and metal surfaces, especially in acidic conditions where the metal surface develops a positive charge and inhibitor molecules become protonated. This type of interaction corresponds to physisorption and is generally weaker than chemisorption but can still contribute significantly to the overall inhibition effect. The spectra of OLZ1 and OLZ2 show additional bands compared to the parent OLZ, confirming the successful incorporation of acrylamide and ethyl acetate groups, respectively.

4. Conclusions

This study determined the inhibition of corrosion for carbon steel in 1 M HCl solution using a comprehensive suite of electrochemical and surface analytical techniques by olanzapine (OLZ) and two novel derivatives, OLZ1 and OLZ2. This study determined the following results from the experimental data obtained:
  • The three inhibition properties of OLZ and its derivatives on C1018 carbon steel in 1 M HCl solution were very high at 300 ppm, where inhibition efficiencies exceeded 88%. However, inhibition efficiency was temperature- and concentration-dependent.
  • Electrochemical studies indicate that the three inhibitors are predominantly anodic. PDP-derived inhibition efficiencies agreed with those obtained from EIS measurements.
  • At 298 K, the inhibition efficiency followed the order OLZ2 > OLZ1 > OLZ, while at 318 K, it followed the order OLZ1 > OLZ2 > OLZ, which confirms that the structural modifications enhanced the corrosion inhibition properties of the parent molecule.
  • The SEM-EDX analysis demonstrates that OLZ and its derivatives effectively inhibit steel corrosion in 1 M HCl compared to the uninhibited solution.
  • The FTIR spectroscopy results show that the metal surface active sites include -NH, -C=N, C=C, and aromatic ring structures.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/ma18122902/s1. Figure S1: Fourier Transformer Infrared Spectra of OLZ, OLZ1, and OLZ2. Figure S2: Nuclear Magnetic Resonance Spectra of (a) OLZ, (b) OLZ1, (c) OLZ2. Table S1: Chemical shifts of 1H NMR spectra of compounds OLZ, OLZ1, and OLZ2.

Author Contributions

Conceptualization, A.H.A., M.A. and M.S.G.; methodology, H.M.A.O., M.S.G., N.A., M.Y.A., N.M.A.O., A.R.A.-S. and I.B.; software, A.H.A., H.M.A.O. and A.A.B.; validation, H.M.A.O., A.A.B., N.A., M.Y.A., N.M.A.O., A.R.A.-S. and I.B.; formal analysis, A.H.A., H.M.A.O., A.A.B., N.A., M.Y.A., N.M.A.O., A.R.A.-S. and I.B.; investigation, A.H.A., M.A. and M.S.G.; resources, A.H.A., M.A., I.U.T., M.S.G. and I.A.; data curation, A.H.A., H.M.A.O., A.A.B., N.A. and I.A.; writing—original draft preparation, A.H.A., M.A., M.S.G. and A.A.B.; writing—review and editing, A.H.A., M.A. and I.U.T.; visualization, H.M.A.O. and A.A.B.; supervision, A.H.A. and M.A. All authors have read and agreed to the published version of the manuscript.

Funding

This research received no external funding.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Informed consent was obtained from all subjects involved in the study.

Data Availability Statement

The data presented in this study are not publicly available due to privacy restrictions related to the personal information of the participants. However, data can be made available upon reasonable request from the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Jacobson, G.A. NACE international’s IMPACT study breaks new ground in corrosion management research and practice. Bridge 2016, 46, 28–32. [Google Scholar] [CrossRef]
  2. Fayomi, O.; Akande, I.; Odigie, S. Economic Impact of Corrosion in Oil Sectors and Prevention: An Overview. J. Phys. Conf. Ser. 2019, 1378, 022037. [Google Scholar] [CrossRef]
  3. Al-Moubaraki, A.H.; Obot, I.B. Corrosion challenges in petroleum refinery operations: Sources, mechanisms, mitigation, and future outlook. J. Saudi Chem. Soc. 2021, 25, 101370. [Google Scholar] [CrossRef]
  4. Raghavendra, N. Green compounds to attenuate aluminum corrosion in HCl activation: A necessity review. Chem. Afr. 2020, 3, 21–34. [Google Scholar] [CrossRef]
  5. Baari, M.J.; Sabandar, C.W. A Review on expired drug-based corrosion inhibitors: Chemical composition, structural effects, inhibition mechanism, current challenges, and future prospects. Indones. J. Chem. 2021, 21, 1316–1336. [Google Scholar] [CrossRef]
  6. Narang, R.; Vashishth, P.; Bairagi, H.; Sehrawat, R.; Shukla, S.; Mangla, B. Applicability of drugs as sustainable corrosion inhibitors. Adv. Mater. Lett. 2023, 14, 2304-1732. [Google Scholar] [CrossRef]
  7. Tanwer, S.; Shukla, S.K. Recent advances in the applicability of drugs as corrosion inhibitor on metal surface: A review. Curr. Res. Green Sustain. Chem. 2022, 5, 100227. [Google Scholar] [CrossRef]
  8. Piggott, T.; Moja, L.; Huttner, B.; Okwen, P.; Raviglione, M.C.B.; Kredo, T.; Schünemann, H.J. WHO Model list of essential medicines: Visions for the future. Bull. World Health Organ. 2024, 102, 722. [Google Scholar] [CrossRef]
  9. Rehan, S.T.; Siddiqui, A.H.; Khan, Z.; Imran, L.; Syed, A.A.; Tahir, M.J.; Jassani, Z.; Singh, M.; Asghar, M.S.; Ahmed, A. Samidorphan/olanzapine combination therapy for schizophrenia: Efficacy, tolerance and adverse outcomes of regimen, evidence-based review of clinical trials. Ann. Med. Surg. 2022, 79, 104115. [Google Scholar] [CrossRef]
  10. ClinCalc DrugStats Database. The Top 300 Drugs of 2022 in the United States by Prescription; ClinCalc LLC: Rockville, MD, USA, 2022. [Google Scholar]
  11. Al-Nami, S.Y. Investigation of Adsorption and Inhibitive Effect of Expired Helicure Drug on Mild Steel Corrosion in Hydrochloric Acid Solution. Int. J. Electrochem. Sci. 2020, 15, 2685–2699. [Google Scholar] [CrossRef]
  12. Maduelosi, N.J.; Iroha, N.B. Insight into the adsorption and inhibitive effect of spironolactone drug on C38 carbon steel corrosion in hydrochloric acid environment. J. Bio-Tribo-Corros. 2021, 7, 6. [Google Scholar] [CrossRef]
  13. Karthik, G.; Sundaravadivelu, M. Studies on the inhibition of mild steel corrosion in hydrochloric acid solution by atenolol drug. Egypt. J. Pet. 2016, 25, 183–191. [Google Scholar] [CrossRef]
  14. Srivastava, M.; Tiwari, P.; Srivastava, S.K.; Prakash, R.; Ji, G. Electrochemical investigation of Irbesartan drug molecules as an inhibitor of mild steel corrosion in 1 M HCl and 0.5 M H2SO4 solutions. J. Mol. Liq. 2017, 236, 184–197. [Google Scholar] [CrossRef]
  15. Mohammadinejad, F.; Zandi, M.S.; Bahrami, M.J.; Golshani, Z. Metoprolol: New and efficient corrosion inhibitor for mild steel in hydrochloric and sulfuric acid solutions. Acta Chim. Slov. 2020, 67, 710–719. [Google Scholar] [CrossRef]
  16. Prasanna, B.; Praveen, B.; Hebbar, N.; Venkatesha, T. Anticorrosion potential of hydralazine for corrosion of mild steel in 1 m hydrochloric acid solution. J. Fundam. Appl. Sci. 2015, 7, 222–243. [Google Scholar] [CrossRef]
  17. Geethamani, P.; Kasthuri, P.; Aejitha, S. An expired non-toxic drug acts as corrosion inhibitor for mild steel in hydrochloric acid medium. Int. J. Chem. Pharm. Sci. 2015, 3, 1442. [Google Scholar]
  18. Kalkhambkar, A.G.; Rajappa, S.; Manjanna, J.; Malimath, G. Effect of expired doxofylline drug on corrosion protection of soft steel in 1 M HCl: Electrochemical, quantum chemical and synergistic effect studies. J. Indian Chem. Soc. 2022, 99, 100639. [Google Scholar] [CrossRef]
  19. Sharma, S.; Ganjoo, R.; Thakur, A.; Kumar, A. Investigation of corrosion performance of expired Irnocam on the mild steel in acidic medium. Mater. Today Proc. 2022, 66, 540–543. [Google Scholar] [CrossRef]
  20. Dohare, P.; Chauhan, D.S.; Hammouti, B.; Quraishi, M.A. Experimental and DFT investigation on the corrosion inhibition behavior of expired drug lumerax on mild steel in hydrochloric acid. Bioanal. Electrochem. 2017, 9, 762. [Google Scholar]
  21. Abdollahi, F.; Foroughi, M.; Shahidi Zandi, M.; Kazemipour, M. Electrochemical investigation of meloxicam drug as a corrosion inhibitor for mild steel in hydrochloric and sulfuric acid solutions. Prog. Color Color. Coat. 2020, 13, 155–165. [Google Scholar]
  22. Shi, Y.; Bai, W.; Guo, J.; Gao, L.; Chen, Y.; Wu, Y.; Liang, L. Inhibition of three kinds of expired nitroimidazole antibiotics. Anti-Corros. Methods Mater. 2018, 65, 398–407. [Google Scholar] [CrossRef]
  23. Dohare, P.; Chauhan, D.; Sorour, A.; Quraishi, M. DFT and experimental studies on the inhibition potentials of expired Tramadol drug on mild steel corrosion in hydrochloric acid. Mater. Discov. 2017, 9, 30–41. [Google Scholar] [CrossRef]
  24. Hameed, R.A.; Aleid, G.M.; Khaled, A.; Mohammad, D.; Aljuhani, E.H.; Al-Mhyawi, S.R.; Alshammary, F.; Abdallah, M. Expired dulcolax drug as corrosion inhibitor for low carbon steel in acidic environment. Int. J. Electrochem. Sci. 2022, 17, 220655. [Google Scholar] [CrossRef]
  25. Srinivasulu, A.; Kasthuri, P. Study of inhibition and adsorption properties of mild steel corrosion by expired pharmaceutical Gentamicin drug in hydrochloric acid media. Orient. J. Chem. 2017, 33, 2616–2624. [Google Scholar] [CrossRef]
  26. Xuehui, P.; Xiangbin, R.; Kuang, F.; Jiandong, X.; Baorong, H. Inhibiting effect of ciprofloxacin, norfloxacin and ofloxacin on corrosion of mild steel in hydrochloric acid. Chin. J. Chem. Eng. 2010, 18, 337–345. [Google Scholar]
  27. Obermayer, D.; Glasnov, T.N.; Kappe, C.O. Microwave-assisted and continuous flow multistep synthesis of 4-(pyrazol-1-yl) carboxanilides. J. Org. Chem. 2011, 76, 6657–6669. [Google Scholar] [CrossRef]
  28. Obermayer, D.; Kappe, C.O. On the importance of simultaneous infrared/fiber-optic temperature monitoring in the microwave-assisted synthesis of ionic liquids. Org. Biomol. Chem. 2010, 8, 114–121. [Google Scholar] [CrossRef]
  29. Gunawan, R.; Nandiyanto, A.B.D. How to read and interpret 1H-NMR and 13C-NMR spectrums. Indones. J. Sci. Technol. 2021, 6, 267–298. [Google Scholar] [CrossRef]
  30. G31-72; Standard Practice for Laboratory Immersion Corrosion Testing of Metals. Annual Book of ASTM Standards; ASTM: West Conshohocken, PA, USA, 2004.
  31. Satapathy, A.; Gunasekaran, G.; Sahoo, S.; Amit, K.; Rodrigues, P. Corrosion inhibition by Justicia gendarussa plant extract in hydrochloric acid solution. Corros. Sci. 2009, 51, 2848–2856. [Google Scholar] [CrossRef]
  32. ASTM G1-03; Standard Practice for Preparing, Cleaning, and Evaluating Corrosion Test Specimens. ASTM: West Conshohocken, PA, USA, 2004.
  33. Shankar, U.; Gogoi, R.; Sethi, S.K.; Verma, A. Introduction to materials studio software for the atomistic-scale simulations. In Forcefields for Atomistic-Scale Simulations: Materials and Applications; Springer: Singapore, 2022; pp. 299–313. [Google Scholar]
  34. Klamt, A.; Schüürmann, G. COSMO: A new approach to dielectric screening in solvents with explicit expressions for the screening energy and its gradient. J. Chem. Soc. Perkin Trans. 2 1993, 799–805. [Google Scholar] [CrossRef]
  35. Qiang, Y.; Zhi, H.; Guo, L.; Fu, A.; Xiang, T.; Jin, Y. Experimental and molecular modeling studies of multi-active tetrazole derivative bearing sulfur linker for protecting steel from corrosion. J. Mol. Liq. 2022, 351, 118638. [Google Scholar] [CrossRef]
  36. Obot, I.; Bahraq, A.A.; Alamri, A.H. Density functional theory and molecular dynamics simulation of the corrosive particle diffusion in pyrimidine and its derivatives films. Comput. Mater. Sci. 2022, 210, 111428. [Google Scholar] [CrossRef]
  37. Sokolov, I.M.; Klafter, J. From diffusion to anomalous diffusion: A century after Einstein’s Brownian motion. Chaos: Interdiscip. J. Nonlinear Sci. 2005, 15, 026103. [Google Scholar] [CrossRef] [PubMed]
  38. Hu, S.-Q.; Guo, A.-L.; Yan, Y.-G.; Jia, X.-L.; Geng, Y.-F.; Guo, W.-Y. Computer simulation of diffusion of corrosive particle in corrosion inhibitor membrane. Comput. Theor. Chem. 2011, 964, 176–181. [Google Scholar] [CrossRef]
  39. Dutta, A. Fourier transform infrared spectroscopy. In Spectroscopic Methods for Nanomaterials Characterization; Elsevier: Amsterdam, The Netherlands, 2017; pp. 73–93. [Google Scholar]
  40. Mohammed, A.; Abdullah, A. Scanning electron microscopy (SEM): A review. In Proceedings of the 2018 International Conference on Hydraulics and Pneumatics—HERVEX, Băile Govora, Romania, 7–9 November 2018; pp. 77–85. [Google Scholar]
  41. Nahlé, A.; Abu-Abdoun, I.I.; Abdel-Rahman, I. Effect of Temperature on the Corrosion Inhibition of Trans-4-Hydroxy-4′-Stilbazole on Mild Steel in HCl Solution. Int. J. Corros. 2012, 2012, 380329. [Google Scholar] [CrossRef]
  42. Obot, I.; Obi-Egbedi, N.; Umoren, S. Adsorption characteristics and corrosion inhibitive properties of clotrimazole for aluminium corrosion in hydrochloric acid. Int. J. Electrochem. Sci. 2009, 4, 863–877. [Google Scholar] [CrossRef]
  43. Sarkar, T.K.; Saraswat, V.; Mitra, R.K.; Obot, I.; Yadav, M. Mitigation of corrosion in petroleum oil well/tubing steel using pyrimidines as efficient corrosion inhibitor: Experimental and theoretical investigation. Mater. Today Commun. 2021, 26, 101862. [Google Scholar] [CrossRef]
  44. Tang, J.; Wang, H.; Jiang, X.; Zhu, Z.; Xie, J.; Tang, J.; Wang, Y.; Chamas, M.; Zhu, Y.; Tian, H. Electrochemical behavior of jasmine tea extract as corrosion inhibitor for carbon steel in hydrochloric acid solution. Int. J. Electrochem. Sci. 2018, 13, 3625–3642. [Google Scholar] [CrossRef]
  45. Oguzie, E.; Li, Y.; Wang, F. Corrosion inhibition and adsorption behavior of methionine on mild steel in sulfuric acid and synergistic effect of iodide ion. J. Colloid Interface Sci. 2007, 310, 90–98. [Google Scholar] [CrossRef]
  46. Haruna, K.; Saleh, T.A.; Quraishi, M. Expired metformin drug as green corrosion inhibitor for simulated oil/gas well acidizing environment. J. Mol. Liq. 2020, 315, 113716. [Google Scholar] [CrossRef]
  47. Mobin, M.; Zamindar, S.; Banerjee, P. Mechanistic insight into adsorption and anti-corrosion capability of a novel surfactant-derived ionic liquid for mild steel in HCl medium. J. Mol. Liq. 2023, 385, 122403. [Google Scholar] [CrossRef]
  48. Touir, R.; Errahmany, N.; Rbaa, M.; Benhiba, F.; Doubi, M.; Kafssaoui, E.E.; Lakhrissi, B. Experimental and computational chemistry investigation of the molecular structures of new synthetic quinazolinone derivatives as acid corrosion inhibitors for mild steel. J. Mol. Struct. 2024, 1303, 137499. [Google Scholar] [CrossRef]
  49. Fernandes, C.M.; Costa, A.R.; Leite, M.C.; Martins, V.; Lee, H.-S.; da CS Boechat, F.; de Souza, M.C.; Batalha, P.N.; Lgaz, H.; Ponzio, E.A. A detailed experimental performance of 4-quinolone derivatives as corrosion inhibitors for mild steel in acid media combined with first-principles DFT simulations of bond breaking upon adsorption. J. Mol. Liq. 2023, 375, 121299. [Google Scholar] [CrossRef]
  50. Acidi, A.; Sedik, A.; Rizi, A.; Bouasla, R.; Rachedi, K.O.; Berredjem, M.; Delimi, A.; Abdennouri, A.; Ferkous, H.; Yadav, K.K. Examination of the main chemical components of essential oil of Syzygium aromaticum as a corrosion inhibitor on the mild steel in 0.5 M HCl medium. J. Mol. Liq. 2023, 391, 123423. [Google Scholar] [CrossRef]
  51. Millard, S.; Law, D.; Bungey, J.; Cairns, J. Environmental influences on linear polarisation corrosion rate measurement in reinforced concrete. NDT E Int. 2001, 34, 409–417. [Google Scholar] [CrossRef]
  52. Olasunkanmi, L.O.; Obot, I.B.; Kabanda, M.M.; Ebenso, E.E. Some quinoxalin-6-yl derivatives as corrosion inhibitors for mild steel in hydrochloric acid: Experimental and theoretical studies. J. Phys. Chem. C 2015, 119, 16004–16019. [Google Scholar] [CrossRef]
  53. Umoren, S.A. Polypropylene glycol: A novel corrosion inhibitor for ×60 pipeline steel in 15% HCl solution. J. Mol. Liq. 2016, 219, 946–958. [Google Scholar] [CrossRef]
  54. Zhang, J.; Qiao, G.; Hu, S.; Yan, Y.; Ren, Z.; Yu, L. Theoretical evaluation of corrosion inhibition performance of imidazoline compounds with different hydrophilic groups. Corros. Sci. 2011, 53, 147–152. [Google Scholar] [CrossRef]
  55. Guo, L.; Kaya, S.; Obot, I.B.; Zheng, X.; Qiang, Y. Toward understanding the anticorrosive mechanism of some thiourea derivatives for carbon steel corrosion: A combined DFT and molecular dynamics investigation. J. Colloid Interface Sci. 2017, 506, 478–485. [Google Scholar] [CrossRef]
  56. Zhang, X.; Kang, Q.; Wang, Y. Theoretical study of N-thiazolyl-2-cyanoacetamide derivatives as corrosion inhibitor for aluminum in alkaline environments. Comput. Theor. Chem. 2018, 1131, 25–32. [Google Scholar] [CrossRef]
  57. Obi-Egbedi, N.; Obot, I.; El-Khaiary, M.I. Quantum chemical investigation and statistical analysis of the relationship between corrosion inhibition efficiency and molecular structure of xanthene and its derivatives on mild steel in sulphuric acid. J. Mol. Struct. 2011, 1002, 86–96. [Google Scholar] [CrossRef]
  58. Verma, C.; Quraishi, M.; Singh, A. 5-Substituted 1H-tetrazoles as effective corrosion inhibitors for mild steel in 1 M hydrochloric acid. J. Taibah Univ. Sci. 2016, 10, 718–733. [Google Scholar] [CrossRef]
  59. Ouakki, M.; Galai, M.; Rbaa, M.; Abousalem, A.S.; Lakhrissi, B.; Rifi, E.; Cherkaoui, M. Investigation of imidazole derivatives as corrosion inhibitors for mild steel in sulfuric acidic environment: Experimental and theoretical studies. Ionics 2020, 26, 5251–5272. [Google Scholar] [CrossRef]
  60. Verma, C.; Quraishi, M.; Rhee, K.Y. Electronic effect vs. molecular size effect: Experimental and computational based designing of potential corrosion inhibitors. Chem. Eng. J. 2022, 430, 132645. [Google Scholar] [CrossRef]
  61. Hou, B.; Zhang, Q.; Li, Y.; Zhu, G.; Zhang, G. Influence of corrosion products on the inhibition effect of pyrimidine derivative for the corrosion of carbon steel under supercritical CO2 conditions. Corros. Sci. 2020, 166, 108442. [Google Scholar] [CrossRef]
  62. Wang, Y.; Wang, Z.; Zhang, L.; Lu, M. Theoretical insights into the inhibition performance of three neonicotine derivatives as novel type of inhibitors on carbon steel. J. Renew. Mater. 2020, 8, 819–832. [Google Scholar] [CrossRef]
  63. Yan, Y.; Wang, X.; Zhang, Y.; Wang, P.; Cao, X.; Zhang, J. Molecular dynamics simulation of corrosive species diffusion in imidazoline inhibitor films with different alkyl chain length. Corros. Sci. 2013, 73, 123–129. [Google Scholar] [CrossRef]
  64. Aziz, I.A.A.; Abdulkareem, M.H.; Annon, I.A.; Hanoon, M.M.; Al-Kaabi, M.H.; Shaker, L.M.; Alamiery, A.A.; Isahak, W.N.R.W.; Takriff, M.S. Weight loss, thermodynamics, SEM, and electrochemical studies on N-2-methylbenzylidene-4-antipyrineamine as an inhibitor for mild steel corrosion in hydrochloric acid. Lubricants 2022, 10, 23. [Google Scholar] [CrossRef]
  65. Obot, I.; Umoren, S.; Ankah, N. Pyrazine derivatives as green oil field corrosion inhibitors for steel. J. Mol. Liq. 2019, 277, 749–761. [Google Scholar] [CrossRef]
  66. Dao, D.Q.; Hieu, T.D.; Le Minh Pham, T.; Tuan, D.; Nam, P.C.; Obot, I.B. DFT study of the interactions between thiophene-based corrosion inhibitors and an Fe 4 cluster. J. Mol. Model. 2017, 23, 260. [Google Scholar] [CrossRef]
  67. Saranya, J.; Sounthari, P.; Parameswari, K.; Chitra, S. Acenaphtho [1,2-b] quinoxaline and acenaphtho [1,2-b] pyrazine as corrosion inhibitors for mild steel in acid medium. Measurement 2016, 77, 175–186. [Google Scholar] [CrossRef]
  68. Singh, P.; Srivastava, V.; Quraishi, M. Novel quinoline derivatives as green corrosion inhibitors for mild steel in acidic medium: Electrochemical, SEM, AFM, and XPS studies. J. Mol. Liq. 2016, 216, 164–173. [Google Scholar] [CrossRef]
Figure 1. Synthesis of derivatives (a) OLZ1 and (b) OLZ2.
Figure 1. Synthesis of derivatives (a) OLZ1 and (b) OLZ2.
Materials 18 02902 g001
Figure 2. Variation in inhibition efficiency values by weight loss of carbon steel in the absence and presence of inhibitors in 1 M HCl at 298 K and 318 K.
Figure 2. Variation in inhibition efficiency values by weight loss of carbon steel in the absence and presence of inhibitors in 1 M HCl at 298 K and 318 K.
Materials 18 02902 g002
Figure 3. Open circuit potential plots of OLZ, OLZ1, and OLZ2.
Figure 3. Open circuit potential plots of OLZ, OLZ1, and OLZ2.
Materials 18 02902 g003
Figure 4. Nyquist plots for the corrosion control of carbon steel in 1 M HCl in the presence and absence of OLZ, OLZ1, and OLZ2.
Figure 4. Nyquist plots for the corrosion control of carbon steel in 1 M HCl in the presence and absence of OLZ, OLZ1, and OLZ2.
Materials 18 02902 g004
Figure 5. Bode plots for the corrosion control of carbon steel in 1 M HCl in the presence and absence of OLZ, OLZ1, and OLZ2.
Figure 5. Bode plots for the corrosion control of carbon steel in 1 M HCl in the presence and absence of OLZ, OLZ1, and OLZ2.
Materials 18 02902 g005
Figure 6. Potentiodynamic polarization plot for the corrosion control of carbon steel in 1 M HCl by different concentrations of inhibitors: OLZ, OLZ1, and OLZ2.
Figure 6. Potentiodynamic polarization plot for the corrosion control of carbon steel in 1 M HCl by different concentrations of inhibitors: OLZ, OLZ1, and OLZ2.
Materials 18 02902 g006
Figure 7. Optimized structures and orbital distributions of the olanzapine derivative molecules.
Figure 7. Optimized structures and orbital distributions of the olanzapine derivative molecules.
Materials 18 02902 g007
Figure 8. EHOMO, ELUMO, and the energy gap of olanzapine derivative molecules.
Figure 8. EHOMO, ELUMO, and the energy gap of olanzapine derivative molecules.
Materials 18 02902 g008
Figure 9. Fukui parameters of the olanzapine-based molecules.
Figure 9. Fukui parameters of the olanzapine-based molecules.
Materials 18 02902 g009
Figure 10. Adsorption configurations of the olanzapine derivative molecules on the Fe(110) surface and their corresponding binding energy.
Figure 10. Adsorption configurations of the olanzapine derivative molecules on the Fe(110) surface and their corresponding binding energy.
Materials 18 02902 g010
Figure 11. Correlation between MD-derived binding energy and experimental inhibition efficiency for OLZ, OLZ1, and OLZ2.
Figure 11. Correlation between MD-derived binding energy and experimental inhibition efficiency for OLZ, OLZ1, and OLZ2.
Materials 18 02902 g011
Figure 12. Diffusion models of the blank and inhibited systems.
Figure 12. Diffusion models of the blank and inhibited systems.
Materials 18 02902 g012aMaterials 18 02902 g012b
Figure 13. MSD curves of the corrosive particles across the blank and inhibited films: (a) H2O; (b) H3O+; and (c) Cl.
Figure 13. MSD curves of the corrosive particles across the blank and inhibited films: (a) H2O; (b) H3O+; and (c) Cl.
Materials 18 02902 g013aMaterials 18 02902 g013b
Figure 14. (a) Scanning electron micrographs (first row) of C1018 steel samples: polished fresh metal, immersed in 1 M HCl solution and immersed in 1 M HCl solution containing OLZ, OLZ1, and OLZ2. (b) EDX graphs (second row) of C1018 steel samples: polished fresh metal, immersed in 1 M HCl so-lution and immersed in 1 M HCl solution containing OLZ, OLZ1, and OLZ2.
Figure 14. (a) Scanning electron micrographs (first row) of C1018 steel samples: polished fresh metal, immersed in 1 M HCl solution and immersed in 1 M HCl solution containing OLZ, OLZ1, and OLZ2. (b) EDX graphs (second row) of C1018 steel samples: polished fresh metal, immersed in 1 M HCl so-lution and immersed in 1 M HCl solution containing OLZ, OLZ1, and OLZ2.
Materials 18 02902 g014aMaterials 18 02902 g014b
Figure 15. FTIR spectra of C1018 and corrosion products after 6 h of immersion with and without an inhibitor: (a) OLZ, (b) OLZ1, and (c) OLZ2.
Figure 15. FTIR spectra of C1018 and corrosion products after 6 h of immersion with and without an inhibitor: (a) OLZ, (b) OLZ1, and (c) OLZ2.
Materials 18 02902 g015aMaterials 18 02902 g015b
Table 2. Weight loss data of carbon steel showing the corrosion rate and inhibition efficiency in the absence and presence of the inhibitors in 1 M HCl at 298 K and 318 K.
Table 2. Weight loss data of carbon steel showing the corrosion rate and inhibition efficiency in the absence and presence of the inhibitors in 1 M HCl at 298 K and 318 K.
InhibitorAt 298 KAt 318 K
△ wC.RIE%△ wC.RIE%
mpymmpympymmpy
Blank1.0341876.0847.63-2.8755361.45136.13-
OLZ0.07131.73.3492.98%0.6881353.434.3674.75%
OLZ10.092172.054.3690.82%0.495984.82581.63%
OLZ20.03570.511.7996.23%0.5411101.8527.9779.44%
Table 3. The data obtained from fitted EIS curves for corrosion control of carbon steel with the presence and absence of OLZ, OLZ1, and OLZ2 at 300 ppm.
Table 3. The data obtained from fitted EIS curves for corrosion control of carbon steel with the presence and absence of OLZ, OLZ1, and OLZ2 at 300 ppm.
Inh.Rs
Ω·cm2
CPEdl Rct
Ω·cm2
CPEf Rf
Ω·cm2
Cdl μF·cm−2Rp
Ω·cm2
IE
%
STD
IE
χ2 × 10−3
Y01 (mΩ sn cm−2)n1Y02 (mΩ sn cm−2)n2
Blank131.1589.3800.8671.7470.4650.9588.005193.850140.897--9.94
OLZ0.7463.6570.96710.8992.7360.8821189.33379.5781200.9788.280.09404.14
OLZ11.6447.530.9721416.320.8200.98412.44965.4711430.4290.150.19563.21
OLZ21.3250.410.996898.0629.4130.904747.03351.4161646.4291.450.20435.14
Table 4. Results for the inhibitory effect of OLZ, OLZ1, and OLZ2 on carbon steel by potentiodynamic polarization and linear polarization resistance experiments.
Table 4. Results for the inhibitory effect of OLZ, OLZ1, and OLZ2 on carbon steel by potentiodynamic polarization and linear polarization resistance experiments.
InhibitorPDP LPR
Ecorr
(mV/SCE)
icorr
(μA cm−2)
βa
(mV/dec)
βc
(mV/dec)
IE
%
χ2 × 10−3Rp
Ω·cm2
IE
%
STD
CR
Blank−447.00200.00160.80120.00-23.93132.20--
OLZ−401.00020.100100.00132.70089.95191.61135.6688.350.0655
OLZ1−397.00015.90089.000218.60092.0575.341473.0091.010.2311
OLZ2−422.00014.60091.400241.60092.70821.11760.6692.470.3597
Table 5. Quantum chemical indices of the molecules.
Table 5. Quantum chemical indices of the molecules.
ParametersInhibitor Molecules
OLZOLZ1OLZ2
EHOMO (eV)−4.643−4.637−4.669
ELUMO (eV)−1.865−1.852−1.879
I (eV)4.6434.6374.669
A (eV)1.8651.8521.879
ΔE (eV)2.7782.7852.790
χ (eV)3.2543.2453.274
η (eV)1.3891.3931.395
ΔN (eV)−2.697−2.697−2.671
μ (Debye)4.6036.5105.160
Table 6. Diffusion coefficients of the corrosive particles in films.
Table 6. Diffusion coefficients of the corrosive particles in films.
Inhibitor FilmD (×10−9 m2·s−1)
H2OH3O+Cl
Blank2.792.571.24
OLZ0.390.350.07
OLZ10.660.870.25
OLZ20.350.170.19
Table 7. Atomic percentage of metal in the presence and absence of OLZ, OLZ1, and OLZ2 by EDX.
Table 7. Atomic percentage of metal in the presence and absence of OLZ, OLZ1, and OLZ2 by EDX.
ElementMetalMetal with HClMetal with OLZMetal with OLZ1Metal with OLZ2
N1.570.911.741.681.68
O1.6920.761.621.782.24
C9.6914.4211.4110.7510.81
S0.460.360.430.430.39
P0.650.510.590.730.67
Mn0.380.330.310.450.40
Fe85.5662.7183.9084.1985.49
Table 8. FTIR band assignments for corrosion products and inhibitor effects.
Table 8. FTIR band assignments for corrosion products and inhibitor effects.
NO.BandAbsorbance PeakReason
1Fe-O450The presence of a sharp, intense peak indicates the formation of ferrous and ferric oxides or hydroxides
2O-H3300The presence of a hydroxyl group is an indication of the formation of ferric oxides or hydroxides
3C-Cl1000Presence of chloride ions as corrosion products
4Fe-O490Reduction in peak intensity due to the presence of the inhibitor
5C-Cl1000Reduction in peak intensity due to the presence of the inhibitor
6O-H3300Disappearance of the hydroxyl group in the presence of the inhibitor
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Omar, H.M.A.; Ankah, N.; Gomaa, M.S.; Alkhaldi, M.Y.; Osman, N.M.A.; Al-Subaie, A.R.; Aldossary, I.; Baig, I.; Bahraq, A.A.; Aljohani, M.; et al. Towards a Sustainable Material Protection: Olanzapine Drugs and Their Derivatives as Corrosion Inhibitors for C1018 Steel in 1 M Hydrochloric Acid. Materials 2025, 18, 2902. https://doi.org/10.3390/ma18122902

AMA Style

Omar HMA, Ankah N, Gomaa MS, Alkhaldi MY, Osman NMA, Al-Subaie AR, Aldossary I, Baig I, Bahraq AA, Aljohani M, et al. Towards a Sustainable Material Protection: Olanzapine Drugs and Their Derivatives as Corrosion Inhibitors for C1018 Steel in 1 M Hydrochloric Acid. Materials. 2025; 18(12):2902. https://doi.org/10.3390/ma18122902

Chicago/Turabian Style

Omar, Habibah M. A., Nestor Ankah, Mohamed S. Gomaa, Malak Y. Alkhaldi, Nadir M. A. Osman, Abdullah R. Al-Subaie, Ibrahim Aldossary, Irshad Baig, Ashraf A. Bahraq, Marwah Aljohani, and et al. 2025. "Towards a Sustainable Material Protection: Olanzapine Drugs and Their Derivatives as Corrosion Inhibitors for C1018 Steel in 1 M Hydrochloric Acid" Materials 18, no. 12: 2902. https://doi.org/10.3390/ma18122902

APA Style

Omar, H. M. A., Ankah, N., Gomaa, M. S., Alkhaldi, M. Y., Osman, N. M. A., Al-Subaie, A. R., Aldossary, I., Baig, I., Bahraq, A. A., Aljohani, M., Toor, I. U., & Alamri, A. H. (2025). Towards a Sustainable Material Protection: Olanzapine Drugs and Their Derivatives as Corrosion Inhibitors for C1018 Steel in 1 M Hydrochloric Acid. Materials, 18(12), 2902. https://doi.org/10.3390/ma18122902

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop