Next Article in Journal
A Hybrid Solar–Thermoelectric System Incorporating Molten Salt for Sustainable Energy Storage Solutions
Previous Article in Journal
Robust Intrusion Detection System Using an Improved Hybrid Deep Learning Model for Binary and Multi-Class Classification in IoT Networks
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Recent Development of Corrosion Inhibitors: Types, Mechanisms, Electrochemical Behavior, Efficiency, and Environmental Impact

by
Denisa-Ioana (Gheorghe) Răuță
1,
Ecaterina Matei
2 and
Sorin-Marius Avramescu
3,*
1
Biotechnical Systems Engineering Doctoral School, National University of Science and Technology Politehnica Bucharest, 060042 Bucharest, Romania
2
Department of Metallic Materials Processing and Eco-Metallurgy, National University of Science and Technology Politehnica Bucharest, 313 Spl. Independentei, 060042 Bucharest, Romania
3
Faculty of Animal Productions Engineering and Management, University of Agronomic, Sciences and Veterinary Medicine of Bucharest, 011464 Bucharest, Romania
*
Author to whom correspondence should be addressed.
Technologies 2025, 13(3), 103; https://doi.org/10.3390/technologies13030103
Submission received: 1 December 2024 / Revised: 29 January 2025 / Accepted: 31 January 2025 / Published: 5 March 2025
(This article belongs to the Section Environmental Technology)

Abstract

:
This review examines recent advances in corrosion inhibitor technologies, with a focus on sustainable and environmentally friendly solutions that address both industrial efficiency and environmental safety. Corrosion is a ubiquitous problem, contributing to massive economic losses globally, with costs estimated between 1 and 5% of GDP in different countries. Traditional inorganic corrosion inhibitors, while effective, are often based on toxic compounds, necessitating the development of more environmentally friendly and non-toxic alternatives. The present work highlights innovative eco-friendly corrosion inhibitors derived from natural sources, including plant extracts and oils, biopolymers, etc., being biodegradable substances that provide effective corrosion resistance with minimal environmental impact. In addition, this review explores organic–inorganic hybrid inhibitors and nanotechnology-enhanced coatings that demonstrate improved efficiency, durability, and adaptability across industries. Key considerations, such as application techniques, mechanisms of action, and the impact of environmental factors on inhibitor performance, are discussed. This comprehensive presentation aims to contribute to updating the data on the development of advanced corrosion inhibitors capable of meeting the requirements of modern industries while promoting sustainable and safe practices in corrosion management.

1. Introduction

The estimated global cost of corrosion was USD 2.5 trillion/year in 2013, equivalent to 3.8% of the global GDP (Gross Domestic Product), based on NACE (National Statistical Classification of Economic) International in 2016. In the European Region, the total cost of corrosion (for 2013) was USD 701.5 billion. The distribution of corrosion costs in the United States is presented in Figure 1 [1]. Most countries anticipate corrosion costs to range between 1% and 5% of their GDP annually [2,3].
According to NACE International [4], applying efficient corrosion management measures already available may cut these expenses by 15–35%, resulting in worldwide savings of USD 375–875 trillion annually.
Corrosion management is crucial for cost savings and asset life, ensuring individual safety and environmental impacts. Implementing a corrosion management system can maximize these savings [2].
The present cost of corrosion is calculated by combining design, construction, repairs, and replacement costs. The actual price of corrosion is the difference between a strategy that does not address corrosion and control and the existing approach [5]. Proper corrosion prevention strategies can reduce economic losses, improve the environment, and enhance public safety, while addressing various forms of corrosion, according to Yang [3].

2. Corrosion Basics

Corrosion stands as an irreversible deterioration of metal, alloy, ceramic, and polymer surfaces due to the action of different chemical compounds that forces the transformation of these materials into more thermodynamically and kinetic stable forms [6]. This process requires a complex set of conditions (Figure 2) that favor specific electrochemical reactions that lead to the degradation of solids and the compromise of their properties and further use [7,8,9].

3. Corrosion Types

Redox reactions occur at the material interface, because corrosion processes are naturally electrochemical in nature [10]. Corrosion mechanisms can vary based on material type, the type of corrosive environment, and the presence of additional substances. Rusting, pitting, cracking, and other types of damage may arise from development of these reactions. This increase the risk of equipment breakdown, the failure of certain methods, and safety risks. Understanding these mechanisms is crucial for preventing and minimizing corrosion, maintaining material integrity, and extending service life, thus enhancing overall performance [11]. Regarding corrosive environments, corrosion can be grouped as dry or wet [6].
Electrolytes are materials that facilitate the formation of anodic and cathodic processes and ion transfer, completing the corrosion circuit. Corrosion often occurs in pipeline steels when carbon dioxide or hydrogen sulfide are present; these compounds are primarily found in oil and gas but are sometimes combined with water, sand, microorganisms, and other organic compounds [12]. The metal will oxidize at the anode, and the metal itself will absorb the electrons provided at the cathode. As the ions generated at the anode move through the solution to the cathode, they will combine with the oxygen reduction product to form an oxide layer. The anode and cathode can move at any time, which leads to corrosion of the metal surface [13,14].
Corrosion occurs in several forms (Figure 3).

3.1. Uniform Corrosion

It can be considered general surface corrosion when the thickness of the metal is slowly reduced in a homogeneous manner because the agent causing corrosion is evenly distributed (this often occurs in the atmosphere and aqueous media) [15].

3.2. Localized Attack

Isolated areas of the material are affected (crevice and pitting corrosion) when a metal surface corrodes more quickly in tiny sections than it does throughout in a corrosive medium. Because of the corroding medium’s presence, these tiny portions of the metal surface are largely corroded at a greater speed. The region with a restricted oxygen source will become the anode, while the region with an unlimited oxygen source will become the cathode [13].

3.2.1. Crevice Corrosion

Corrosion is caused by the presence of water (moisture) in the surrounding environment. If the material has cracks, water penetrates the cracks, stagnates there, and causes corrosion. The anodic reaction can either be limited to a particular area or occur evenly over the metal surface. At the anodic site, corrosion products may accumulate if the soluble metal ions can react with the solution to generate an insoluble compound (Figure 4a) [14].

3.2.2. Pitting Corrosion

Corrosion on a material begins in a limited area, forming pits or holes on the surface as it grows deeper and larger over time. This localized corrosion occurs in areas with lower oxygen levels, often where contaminants or water are present, creating regions where one area acts as an anode and another as a cathode. It penetrates deeply into the material, even with a protective coating, causing structural weaknesses and potential failure. Passive materials are particularly susceptible to this type of corrosion, which is influenced by pH and cathode/anode ratio [16,17,18].
Within the pit, at the anode, an oxidation reaction occurs, producing metal ions and electrons (Figure 5).
After that, these metal ions then enter the solution, while the electrons pass through the metal to the cathode. The presence of water initiates a different process called hydrolysis:
Mn+ + H2O→MO + H+ + ne
The acidic conditions in pitting corrosion result from metal ion hydrolysis, which prevents the pit walls from re-passivating. This autocatalytic process is limited, causing tiny holes and increased metal loss. Since the anodic site is relatively small, pit penetration is influenced by cathodic control, resulting in a high degree of depth.

3.3. Grain Boundaries

The material’s resistance to corrosive environments can vary along boundaries or other lines of weakness.

3.3.1. Galvanic Corrosion

This type of corrosion occurs when two distinct metals are present in an electrolyte. The less noble metal acts as the anode and corrodes, while the more noble metal functions as the cathode, providing protection. The metal with a higher reduction potential or higher position in the electrochemical series is more susceptible to rust [19].
For example, in an HCl solution, a conducting electrolyte, iron (Fe) and zinc (Zn) exhibit different behaviors. Fe serves as the cathode r, because it is positioned ahead of Zn in the electrochemical series, making it more noble. Consequently, Fe, being less reactive, remains protected while Zn undergoes corrosion.
Cathode—reduction reaction: Fe2+ + 2e→Fe
Anode—oxidation reaction: Zn − 2e→Zn2+
General reaction of galvanic corrosion:
MetalA + MetalB + Electrolyte (aq)→MetalA+ + MetalB− + Electrolyte (aq)
This type of corrosion is a frequent problem in maritime settings, where ships and offshore constructions are made from various metallic alloys.

3.3.2. Stress Corrosion

Mechanical stress, such as static or applied tensile stress, combined with a corrosive environment causes stress corrosion cracking (SCC) [18]. SCC refers to the cracking of an alloy caused by the interaction of applied or residual stress, during forming and manufacturing, with specific environmental conditions [20]. Stress can be caused by pressures exerted during manufacturing, processing, or heat treatment, as well as from residual locked-in stress. As a result, the electrode potential fluctuates from one spot to another. Corrosion occurs as a mechanism to reduce stress, with anodes forming in highly stressed areas and cathodes in stress-free regions [21]. This form of corrosion is frequently observed in sectors such as nuclear power generation, where metals are subjected to high stress and harsh environmental conditions [22]. In low-pH environments, corrosion primarily occurs through a cathodic process. The hydrogen evolution reaction can be particularly problematic, as atomic hydrogen may penetrate the metal and induce internal stresses. In neutral and alkaline media, another significant cathodic reaction often takes place, with its rate depending on the amount of oxygen present on the metal surface. The following equations describe these reactions:
Cathodic reaction in acidic medium: 2H+ + 2e→H2
Cathodic reaction in neutral medium: O2 + H2O + ne→nOH
Cathodic reaction in basic medium: O2 + 4H+ + 4e→2 H2O

3.3.3. Intergranular Corrosion

A form of localized corrosion known as intergranular corrosion occurs when cohesive stress is reduced, starting and progressing into the grain boundaries, thereby weakening the material. This type of corrosion is associated with impurity segregation and/or the depletion of one of the alloying elements at the grain boundaries [23].

3.3.4. Erosion Corrosion

The hydrocarbon and mineral processing industries are concerned about erosion in flow-changing devices caused by sand movement. Increased impact velocity can cause catastrophic damage, mass transfer, and an accelerated erosion–corrosion rate in solid particle erosion–corrosion. Factors such as pipe designs, material properties, hydrodynamics, phase interactions, and the characteristics of scattered phases contribute to the complexity of erosion in pipelines. In hydrocarbon production pipelines, turbulent slug flow regimes, including a liquid film under the gas phase and a high-flow rate gas phase above the pipeline, can further accelerate erosion-induced damage [24,25,26].

3.3.5. Cavitation Corrosion

In the context of fluid dynamics, cavitation is an unavoidable form of failure in components that are flow-handling such as pipelines, water turbines, and propellers. It is characterized by the rapid deterioration of material surfaces caused by the constant impact of shock waves and microjets generated when bubbles collapse. Due to safety concerns and significant financial costs, the issue of cavitation erosion in flow-handling elements of maritime engineering has driven the development of improved materials to mitigate its effects [26,27].

3.3.6. Fretting Corrosion

Fretting corrosion is a significant practical concern, as it can considerably reduce the fatigue properties of structural components [26]. The main cause of fretting corrosion is small cyclic movements between two materials under cyclic stress, leading to surface degradation. In ambient conditions, corrosion actively contributes to fretting corrosion damage [27]. The fretting process accelerates corrosion by continuously breaking down and reforming the protective oxide layer or passive coating on metallic surfaces, a cycle known as the re-passivation process. Furthermore, relative motion within the crevice intensifies corrosion processes, compounded by local oxygen depletion [28].

3.4. Biological Corrosion

Biological corrosion is influenced by the activity of living organisms, such as bacteria and algae (Figure 6). This type of corrosion is also known as microbial corrosion, microbiologically influenced corrosion (MIC), or microbially induced corrosion [29]. It is a unique form of corrosion that occurs under various conditions due to the presence of living organisms and their metabolic processes. These organisms directly impact anodic and cathodic reactions, degrade protective coatings, and create corrosive environments, making biological corrosion distinct from other forms of corrosion [30].
According to conservative estimates, microbially induced corrosion accounts for up to 20% of corrosion in aquatic systems [5]. As the global population ages, implantation procedures are becoming increasingly common. Implant alloys, primarily composed of stainless steels, cobalt–chromium alloys, and titanium alloys, are critical components of biomedical implant devices. Corrosion presents a significant challenge in this context, and the properties of passive oxide coatings, along with material characteristics, are crucial factors in material selection and development. Improved material selection, design, and quality control could significantly reduce or even eliminate corrosion in implant devices [31]. Examples of bacteria known to accelerate iron corrosion include Bacillus licheniformis, Acetobacter aceti, Acetobacterium sp., Chlorella vulgaris, Shewanella putrefaciens, and Pseudomonas aeruginosa, among others [30,32,33,34,35,36,37].

4. Techniques for Corrosion Prevention

Traditional corrosion protection techniques, despite their limitations, have been used for decades to combat corrosion caused by wear, UV rays, and chemical attacks (Figure 7). These methods require periodic maintenance and reapplication, which can be costly and logistically challenging for industries. One common approach is the use of sacrificial anodes, which are materials designed to corrode preferentially over the metal they protect. However, such methods are often inflexible and not adaptable, making them less suitable for modern industries. As new materials and components are introduced into increasingly demanding environments, advanced corrosion protection techniques are become more pressing necessary [38].
Corrosion is the process through which metals return to their original oxide state, creating significant barriers that prevent them from reaching thermodynamic equilibrium. The primary cause of corrosion is a reduction in Gibbs free energy, which causes metals to revert to their reduced oxide state. This complex issue has plagued humanity for years, affecting metals and alloys across various industries and applications worldwide [39].
∆G = ∆H − T∆S
∆G0 = −RTlnK (at equilibrium)
ΔG = ΔG0 + RTlnQ (not in the standard circumstances)
∆G—Gibbs free energy (indicates whether reaction is exothermic or endothermic); ∆G0—standard change in Gibbs free energy; ∆H—change in enthalpy; ∆S—change in entropy; T—temperature in Kelvin; R—gas constant (8.314 J/K·mol); K—natural logarithm of K, equilibrium constant; Q—reaction quotient (represents the reaction’s initial conditions).
Corrosion control techniques (Figure 8) include galvanization, painting, dealloying, and the application of inhibitors [40]. The purpose of corrosion control methods is to reduce material corrosion to a manageable level, so that the materials can reach their intended lifecycle. Some corrosion management techniques aim to eliminate corrosion entirely. Over time, significant advancements have been made in the field of corrosion prevention and protection [41]. The most practical and cost-effective technique for combating corrosion is the use of corrosion inhibitors. Corrosion inhibitors prevent corrosion by adsorbing on the metal surface and inhibiting one or more electrochemical processes at the solution/metal interface, particularly on low-carbon steel surfaces.

4.1. Coatings

The best method for protecting metallic surfaces is the application of coatings. To prevent corrosion, coatings must form a strong physical barrier that prevents aggressive species from reaching the metallic interface [42].
Coatings play a critical role in industrial applications because they improve the tribological performance of materials, reducing wear, friction, and surface damage. Coatings designed to suppress corrosion must provide an effective physical barrier to block corrosive agents from contacting the metal surface [43]. The use of toxic substances such as volatile organic compounds (VOCs), hazardous air pollutants, and chromium (VI) has been significantly restricted due to environmental and human health regulations. The goal is to develop eco-friendly, non-toxic coatings to address these concerns. Industrial coatings can be categorized into two types:
Organic;
Inorganic.
Organic coatings, such as polymer-based paints and coatings, are widely used due to their versatility, ease of application, and visually appealing results [44]. These coatings offer excellent properties in terms of corrosion resistance, UV protection, chemical resistance, and aesthetic properties [45], making them popular in industries such as construction, automotive, and consumer goods.
Inhibitors are effective in preventing mild steel corrosion, but some organic inhibitors can be harmful due to their poor biodegradability. This has led to a search for green corrosion inhibitors, which are non-toxic, renewable, and affordable. These inhibitors are derived from plant-based sources, making them widely available and environmentally friendly [46].
Alternatively, inorganic coatings include ceramic and metallic coatings. Metallic coatings, also known as alloy coatings, provide enhanced surface hardness, wear protection, and higher corrosion resistance [47].
These coatings can be applied through methods such as thermal spraying and electroplating. Zinc-rich primers are the most popular and effective anticorrosion coatings within the range of sacrificial coatings [48]. Zinc-rich primers provide cathodic protection to steel substrates by enabling zinc particles to interact with each other. When zinc interacts with water, oxygen, and carbon dioxide, it produces corrosion products like zinc oxide, hydroxide, hydrozincites, and carbonates. These corrosion products form a stable barrier over primer film flaws, delaying water absorption.
Other coatings that are commonly used include titanium (Ti), nickel alloys (Ni alloy), copper (Cu), silicon carbide (SiC), etc. [49]. Magnesium alloys, known as “green engineering materials of the 21st century”, are gaining popularity due to their high specific strength, low density, and excellent damping capabilities. These properties make them ideal for applications in electronic communications, transportation, and aerospace industries [50]. Common methods for applying coatings include spraying, dipping, and electrostatic deposition. Spraying is efficient, reliable, and easy, while dipping or immersion provides uniform application for complex geometries. The process typically involves dispersing coating ingredients in a solvent, followed by application via brushing, dipping, or spraying. Electroplating is a popular method for applying metal coatings, using an electrochemical process to deposit a metal layer onto a substrate [49].

4.2. Cathodic Protection

The two forms of cathodic protection are shown in Figure 9 and are as follows [51]:
Impressed current— this method uses a direct power supply feeding systems, where the positive pole is coupled to an anode. The anode can be either inert (such as silicon-cast iron, graphite, or platinized and activated titanium) or soluble (such as iron).
Galvanic or sacrificial anodes—galvanic or sacrificial anodes are created by coupling a less noble metal, such as iron or magnesium, with the substrate to provide cathodic protection. This process is used to protect various materials, including copper alloys, stainless steels, steel in seawater, and structures in soil or freshwater environments [46]. More reactive metals, like magnesium (Mg), aluminum (Al), and zinc (Zn), are often used as sacrificial anodes, because they corrode faster than the protected metal, thereby offering cathodic protection for metals such as steel and aluminum alloys [52].
Figure 9. Cathodic protection general scheme: (a) galvanic anodes and (b) impressed currents [53].
Figure 9. Cathodic protection general scheme: (a) galvanic anodes and (b) impressed currents [53].
Technologies 13 00103 g009
Cathodic protection employs two distinct techniques that differ in their current circulation mechanisms. It mitigates or prevents corrosion through two primary effects: thermodynamic, which reduces the driving voltage for the corrosion process; and kinetic, which increases the electrical resistance of the system [51].
In impressed current systems, local acidification occurs due to hydrolysis, which dissolves galvanic anodes such as iron, magnesium, zinc, and aluminum. For insoluble anodes, the anodic reaction involves oxygen evolution and chlorine evolution, making these anodes inert. The cathodic protection capability of sacrificial anticorrosion coatings makes them one of the most widely used and recommended coatings for preventing corrosion on metal surfaces in a harsh environment, such as industrial and marine settings [48]. The sacrificial metal acts as an anode with the lowest potential, forming a galvanic couple with the base metal. This process protects the base metal from corrosion due to the sacrificial metal’s lower potential [54].

4.3. Anodic Passivation

Another effective method for preventing corrosion on metallic surfaces is to transform the metal into an electrochemical cell and manage its potential in the passive (anodic protection). Unlike cathodic protection, anodic protection applies current directly to the structure that needs to be protected [55].
Anodic protection works by increasing the metal’s potential to form a thin passive layer on its the surface. This oxide layer, composed of the metal’s own oxides, acts as a protective barrier by isolating the substrate from the electrolyte [56]. Metals and alloys, such as iron, carbon steel, stainless steel, titanium, aluminum, chromium, and nickel, can be effectively protected against highly corrosive conditions, including exposure to caustic soda and sulfuric and phosphoric acids, through anodic protection [55].
The foundation of anodic corrosion prevention is the impressed anodic current supplied by a potentiostat, which keeps the metal within the passive potential range [57]. Even in unfavorable circumstances, anodic protection allows materials that are capable of passivation to remain in a “passive” state or be brought into one [58]. The main challenge of anodic protection is the breakdown of the passive oxide layer by aggressive anions, like chloride and bromide ions, in the solution. The effectiveness of this protection depends significantly on the passivity of the steel and the integrity of the passive oxide layer, making their maintenance crucial for long-term corrosion resistance [59].

4.4. Pre-Treatment

Thoroughly cleaning the metal surface completely is a crucial step before applying any protective measures [6]. Any surface contamination can hinder the ability of coating elements or precursors to reach the substrate, reducing the coating’s adhesion and overall effectiveness.
The primary objectives of pre-treatment procedures are to enhance the corrosion resistance of the coated metal system [60]:
-
Improving the metallic substrate’s resistance to corrosion;
-
Creation of a suitable surface for the subsequent layers of the coating system.
The three primary pre-treatment techniques used to remove contamination are [60] as follows:
  • Mechanical cleaning: This process rapidly removes the exterior layers of a material, including methods such as grinding, blasting, and machining. It is effective regardless of the surface’s microstructure or chemical composition.
  • Chemical pre-treatment: Chemical solutions used for pre-treatment are categorized into alkaline and acidic solutions. These chemicals help can eliminate grease, oil residues, and other contaminants left from manufacturing or mechanical pre-treatment. Organic solvents, such as ethanol and acetone, are commonly used in scientific research for this purpose.
  • Laser technology: Due to its high automation, selectivity, efficiency, and environmental friendliness, laser surface texturing has gained popularity in recent years for preparing functional surfaces [61]. Laser processing combines surface texturing with laser chemical modification, making it a popular method for material surface treatment [62]. When a solid surface is irradiated with a laser, it develops nano- and microscale structures, which reduce the surface energy of the substrate or enable specific interactions between the surface and the fouling materials [63]. Additionally, laser surface modification, involving melting and annealing materials, is being explored for improving corrosion resistance in metallic alloys. This process homogenizes microstructures, dissolves intermetallic particles, and extends solid solubility through fast cooling rates, all without altering the surface’s chemical composition [64]. Laser cladding is another advanced process where particles are heated using a defocused laser beam, creating a metallurgical bond between the substrate and clad layer. Factors such as laser power, beam size, scanning velocity, cladding geometry, dilution rate, melt pool profile, aspect ratio, and clad layer thickness influence the coating’s quality [65]. The technique selective laser melting (SLM) creates metal objects with high surface roughness, which can limit their direct application without further processing [66]. Laser shock peening (LSP) is a new surface treatment method designed to reduce metal corrosion. It involves applying high-energy-density laser pulses to create a cavity on the target surface, inducing deep mechanical residual compressive stresses that enhance corrosion resistance [67].

4.5. Relevant Design and Sustainable Material Selection

A thorough understanding of corrosion mechanisms is essential for optimizing material selection [68]. Choosing appropriate materials for specific applications is crucial in industrial design and product development, as improper selection can lead to high costs and product failures [69].
High-fatality catastrophic incidents in the aviation industry have driven safety improvements caused by corrosion-related cracking, which can severely compromise vehicle performance and endanger human lives. Many corrosion-related incidents remain unreported due to liability concerns or insufficient evidence. A notable case is the 1988 Boeing 737 structural failure, which emphasized the risks associated with aging aircraft, where corrosion-accelerated fatigue was identified as the primary cause of failure [70].
Another tragic incident occurred on 19 November 1984, at a large liquefied petroleum gas (LPG) storage and distribution center in San Juan Ixhuatepec, 20 km north of Mexico City. The disaster began with an LPG leak, likely caused by a pipe break or rupture due to excessive pressure, ultimately resulting in 650 fatalities [71].

5. Corrosion Inhibitors

An inhibitor is a substance that minimizes metal loss, reduces hydrogen embrittlement, protects against pitting, and reduces acid fumes. It acts as a barrier, reducing the oxidation rate of metals, and is easy to apply without significant disruption, making it an effective method for corrosion prevention [72].
In recent years, a growing number of eco-friendly or “green” corrosion inhibitors have been discovered, including plant extracts, expired drugs, ionic liquids, etc. [73].
By covering the surface and delaying metal corrosion, inhibitors protect the metals. Good inhibitors should slow down chemical reactions. When there is an adequate amount of inhibitor present, a thin film is formed around the metal during the reaction, and this protective layer is created to prevent aggressive components in the environment from damaging the material [74].
Shielding is a fundamental inhibitor protection mechanism that creates an insulating barrier over metal surfaces to prevent corrosive substances from entering and shielding them from the outside world. This mechanism is commonly used in organic coatings, where iron and other metals interact with inhibitors like oxygen and nitrogen to build complex coatings, preventing water contact and preventing electron, water, and oxygen transport [75].
The right inhibitors can lower corrosion rates and extend metal life by coating surfaces, delaying corrosion, and acting as a protective layer by reducing chemical reactions and forming a thin film [40].
The corrosion of metals and alloys leads to high restoration costs for the repair and replacement of various equipment pieces. Moreover, it poses a significant safety risk to people. Therefore, developing effective corrosion inhibitors is essential. When selecting corrosion inhibitors, the presence of π-electrons and electron-rich molecular structures are important factors to consider [76].
The high toxicity to chromate, phosphate, and arsenic compounds has led to strict international laws, reducing their use and increasing the demand for alternative inhibitors with similar anticorrosive properties that also address environmental and health concerns. Molybdate compounds and rare-earth metal salts, like cerium chloride, are examples of these compounds [72].
The corrosion inhibitor market is projected to grow from an estimated USD 7.9 billion in 2021 to USD 10.1 billion in 2026, at a compound annual growth rate of 4.9% [77].

5.1. Classification of Corrosion Inhibitors

Corrosion inhibitors can be classified as anodic, cathodic, or mixed, depending on their mechanism of action on metal surface reactions and their interactions with ions. They provide optimal protection by simultaneously affecting both cathodic and anodic reactions. Their efficiency depends on factors such as concentration, molecular structure, functional groups, conjugation, and type of bonding atom [78].
They are categorized based on their influence on the anodic, cathodic, or both reactions in the corrosion process.

5.1.1. Anodic Corrosion Inhibitors or Chemical Passivators

When precipitation blocks the anodic sites, the anodic reaction slows down, and the corrosion potential shifts in the cathodic direction. This type of corrosion inhibitor functions by forming a protective oxide layer on the metal surface, leading to a significant anodic shift. As a result, passivation occurs, effectively reducing the corrosion rate [79].

5.1.2. Cathodic Corrosion Inhibitors

Cathodic inhibitors slow down the cathodic reaction and shift the corrosion potential in the anodic direction by blocking cathodic sites. They facilitate the movement of positive ions toward the cathode, forming a protective layer that reduces or stops the cathodic reaction. This process decreases the diffusion rate of reducing elements to the metal surface, increasing its susceptibility of hydrogen-induced cracking while effectively blocking cathodic reduction processes. As a result, these inhibitors are widely used for corrosion protection [72].
Cathodic inhibitors are frequently used due to their environmental acceptability and ability to prevent hydrogen from diffusing into metal surfaces, thereby reducing environmental harm. Examples include polyphosphates [80], zinc salts [81], and cerium salts [82,83]. However, certain cathodic inhibitors, such as sulfur, can promote hydrogen diffusion into metal surfaces, leading to contamination by forming harmful hydrogen sulfides upon reacting with hydrogen [84,85].
Other variants used as cathodic corrosion inhibitors include ionic liquid monocationic and dicationic compounds [86], as well as hexylmeythylimidazolium tetrafluoroborate ionic liquid [40]. In deaerated HCl solutions, 2-aminomethylbenzimidazole and bis(benzimidazol-2-ylethyl) sulfide can act as cathodic inhibitors in the initial stage for steel [87]. Additionally, 1,12-bis(1,2,4-triazolyl) dodecane is also an excellent choice of cathodic inhibitor for steel in HCl solution [88].
The presence of a protective layer between electrolyte and substrate significantly reduces the average polarization current. This layer acts as a barrier on the metal surface, facilitating the formation of insoluble compounds.

5.1.3. Mixed Corrosion Inhibitors

Substances that form films and produce surface precipitates can obstruct both anodic and cathodic sites. By inhibiting reactions at these sites, mixed-type inhibitors, commonly film-forming compounds, facilitate the establishment of protective processes on the surface. These inhibitors have several advantages, as they control both reactions.
Example include silicates and the phosphate-based corrosion inhibitors [89]. In solution, a combination of imidazoline and phosphate inhibitors forms a lamellar structure [90]. Compounds derived from imidazolium, pyridinium, and dimethyl-ethylbenzylammonium serve as mixed corrosion inhibitors [91]. The corrosion rate decreases when the active anode and cathode areas are reduced.
Inhibitor films can be classified into several types:
(a)
Chemisorbed films, where lone pair of electrons is donated to a central adsorption atom in a functional group;
(b)
Electrostatic adsorption films, which rely on electrostatic interactions;
(c)
Precipitation and/or complex films, which result from reactions between dissolved metal ion and organic inhibitor molecules.
Based on their chemical structure and functional groups, corrosion inhibitors can also be classified as organic, inorganic, or hybrid inhibitors (Figure 10).

5.1.4. Organic Inhibitors

Organic inhibitors contain substances such as amines, alcohols, and amino acids. These compounds work by adhering to the metal surface, creating a protective barrier that shields the metal from corrosive environments. Due to their effectiveness, organic inhibitors are largely used in industries, including oil and gas, petrochemical, and marine applications [78].
Organic compounds containing nitrogen (N), sulfur (S), and oxygen (O) play a crucial role in preventing corrosion reactions. They are particularly effective on metallic materials like copper, iron, and aluminum alloys, as they inhibit the formation of insoluble deposits on the metal surface.
Organic inhibitors adsorb inhibitor molecules onto the electrolyte/metal interface, preventing both cathodic and anodic reactions. This process can be either chemical or physical, depending on factors such as chemical structure, concentration, electrolyte type, external charge, and molecule circulation. Physisorption and chemisorption are two primary methods through which organic inhibitors function. The mechanism of action of organic corrosion inhibitors is based on the adsorption of inhibitor molecules, which slow down both cathodic and anodic reactions. Physical adsorption involves electrostatic interactions at the interface, while chemical adsorption involves chemical bonding between the inhibitor molecules and the metal substrate [92,93].
There are various types of organic inhibitors, classified according to their active functional groups, as follows:
Amines, phenols, nitrite, phosphoric acid, carboxylic acid inhibitors, sulfonic acid, triphosphoric acid, and ethylene diamine inhibitors [94].
The durability of corrosion inhibitors is influenced by various environmental factors, such as pH, temperature, concentration, and the presence of other chemical species, including impurities, suspended particles in the atmosphere, etc. [95].
To prevent corrosion in unfavorable electrolytes, corrosion inhibitors must maintain stable functional groups. These inhibitors can interact with aggressive electrolytes, which often contain ions such as nitrate, sulfate, or chloride, reducing their effectiveness. In severe conditions, functional groups susceptible to nucleophilic or electrophilic attacks, such as organic fractions like -NH2 or -OH, may degrade or undergo chemical modifications. Depending on the nature of the electrolyte, polar functional groups like -OH, -NH2, -CN, -COOC2H5, and -COOH can be transformed into charged species [79,96].
Physiochemisorption refers to the simultaneous occurrence of physisorption and chemisorption in aqueous electrolytes, with protonation allowing charged metal surfaces and protonated molecules to interact through physisorption or electrostatic bonding [96]. A key indicator of adsorption type is the Gibbs free energy (ΔGads). Pure chemisorption is characterized by ΔGads values of −40 kJ/mol or higher (negative) and −20 kJ/mol or higher (positive) [97], indicating that the adsorption primarily follows physiochemisorption [98]. Intermediate ΔGads values suggest a mixture of physical and chemical adsorption, reflecting metal–inhibitor interaction. Although establishing a direct correlation between inhibitory efficiency and ΔGads is challenging, these limits provide a useful indication [99].
By nature, organic corrosion inhibitors can be classified:
  • Synthetic organic corrosion inhibitors:
    -
    1,2,3-triazole derivatives serve as corrosion inhibitors for metal surfaces. Unprotonated molecules, such as lone pairs from oxygen and nitrogen atoms and π-electrons from the benzene ring, can occupy the metal surface’s empty iron orbitals, resulting in the chemisorption phenomenon. The triazole derivatives interact with the metal surface through donor–acceptor interactions between the -N=N-, -C=N-, and (-C=C-) π-electrons of the benzene rings and the empty orbitals of the metal atoms [100].
    -
    Quinoline and its derivatives act as corrosion inhibitors in acidic media for mild steel. Through coordination bonding, quinoline derivatives with polar substituents, such as hydroxyl (-OH), methoxy (-OMe), amino -NH2), and nitro (-NO2), among others, efficiently adsorb and form highly stable chelating complexes with surface metallic atoms. Quinoline is expected to interact strongly with metallic surfaces due to its high electron density (10 π- and 2 non-bonding electrons). Using a donor–acceptor mechanism, organic compounds interact with metal surfaces throughs the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO), which relate to the compounds’ ability to donate and accept electrons [101,102].
    -
    Oxalamide (Carbamothiol) derivatives act as corrosion inhibitors for copper in 3.5% NaCl [103].
    -
    Imidazoline derivatives act as corrosion inhibitors and surface-active agents for steel alloy in 1 M H2SO4 and HCl [104].
    -
    Imidazolone, an eco-friendly corrosion inhibitor for mild steel [105], is a mixed-type inhibitor, and follows the Langmuir adsorption model, forming heterogeneous films on the mild steel surface.
    -
    Quinazolinone derivatives act as mixed-type inhibitors for mild steel in a 1.0 M HCl medium. The Langmuir isotherm suggests that the dissolution of mild steel is an endothermic process and that these inhibitors function through physical adsorption on metallic surfaces [106].
    -
    Amino acid derivatives—novel benzimidazole derivatives incorporating three amino acids (tryptophan, tyrosine, and histidine) were designed for acidic solutions. The Langmuir adsorption model was followed, showing the presence of both chemisorption and physisorption processes [107].
    -
    Surfactants. Surfactants are surface-active substances with inherent amphiphilicity, containing both hydrophobic and hydrophilic groups. They serve as effective organic corrosion inhibitors, especially in protecting metals from corrosion. This protection is achieved through adsorption, driven the coordination or electrostatic electron-donating functions of the hydrophilic group [108].
Wang et al. [109] investigated the corrosion inhibition capabilities of a morpholine gemini surfactant and its monomer in a CO2-saturated 5% NaCl solution using surface analysis, electrochemical measurements, and weight-loss methods. El-Monem et al. [110] synthesized two novel cationic surfactants: 1-dodecyl-3-(hydroxymethyl) pyridin-1-ium bromide and N-(2-hydroxyethyl)-N, N-dimethyldodecan-1-aminium bromide.
An important class of amphiphilic compounds with antibacterial properties, pyridinium gemini surfactants are typically chosen for use in industrial and medicinal settings. For the first time, Yumnam et al. [111] synthesized, characterized, and tested three pyridinium gemini surfactants as corrosion inhibitors for carbon steel in HCl. These surfactants feature varying alkyl chain lengths and two amide spacers: 3,3′-(malonylbis(azanediyl)) bis(1-dodecylpyridin-1-ium) dibromide, 3,3′-(malonylbis(azanediyl)) bis(1-tetradecylpyridin-1-ium) dibromide, and 3,3′-(,alonylbis(azanediyl)) bis(1-octadecylpyridin-1-ium) dibromide.
In this study, Wang et al. [112] thoroughly examined the mechanisms and inhibitory effects of three anionic surfactants: potassium laurate, potassium oleate, and dodecyl benzenesulfonic acid, on the corrosion of cobalt metal in a weakly alkaline glycine solution.
Four new non-ionic polymeric surfactants were developed through green synthesis using recycled poly(ethylene terephthalate) plastic waste to prevent corrosion on carbon steel in a 3.5% NaCl solution, which simulates artificial seawater. These surfactant inhibitors are of the mixed type [113].
  • Ionic liquids
Salts composed of ions, such as anions and cations, that are liquids at room temperature and up to 100 °C are known as ionic liquids. Their desirable properties, including low volatility, non-flammability, non-toxicity, chemical stability, high solubility in polar solvents, and the ability to adsorb onto metallic surfaces, have made them popular in various industries. Common classes of ionic liquids include imidazolium, triazolium, thiazolium, phosphonium, pyridinium, ammonium, pyrrolidinium, and pyridinium-based ionic liquids. More details and classifications of liquid ions can be found in the review by Ikeuba et al. [114]. Ionic liquids exhibit a mixed inhibitory effect [115].
A new polymeric ionic liquid was created using imidazolium ionic liquids with short alkyl chains as monomers, along with two control ionic liquids: 1-hydroxypropyl-3-vinylimidazolium chloride, 2,2′-diallylbisphenol A, and 1-aminopropyl-3-vinylimidazolium chloride [116].
(E)-3-ethyl-2-(4-hydroxystyryl) benzo [117] thiazol-3-ium iodide and (E)-3-ethyl-2-(4 hydroxy styryl) benzo [117] thiazol-3-ium bromide are two benzothiazole ionic liquids. These compounds provide exceptional protection for steel against aggressive acid solutions at varying temperatures when used as green corrosion inhibitors [118].
Three distinct ionic liquids—bis(2-hydroxyethyl) ammonium butanoate, bis(2-hydroxyethyl) ammonium octanoate, and bis(2-hydroxyethyl) ammonium decanoate—were tested by Wanees et al. [119] for their ability to inhibit the corrosion behavior of C-steel submerged in 10% formation water solutions at different temperatures and concentrations.
  • Drugs
As a result of their molecular structures, which promote adsorption at the metal–aqueous solution interface, researchers are exploring alternative green inhibitors for pharmaceuticals. The first antibiotics studied—ampicillin, cloxacillin, flucloxacillin, and amoxicillin—demonstrated inhibitory efficiencies of up to 90% [120]. Penicillin G and streptomycin, two antibacterial and antitubercular medications, also showed excellent results, inhibiting up to 90% of the corrosion process.
Pharmaceuticals typically contain key structural components—heteroatoms and a balance between hydrophobic and hydrophilic groups—that contribute to their high corrosion inhibition potential [121]. With their well-organized heteroatoms, such as oxygen and nitrogen, macrocyclic compounds improve electronic interactions and sp2 hybridization with metal surfaces, forming effective hydrophobic films that improve metal–inhibitor interactions and corrosion inhibition [122]. Oubahou et al. [123] have scientifically proven, through electrochemical, thermodynamic, and computational investigation, that expired Fenoprofen effectively reduces copper corrosion in H2SO4 solution. OS Michael et al. [124] demonstrated that in 0.5 M HCl, expired paracetamol inhibited mild steel corrosion in a mixed-type manner, and the inhibition process occurred independently. One such drug with potential that extends beyond its intended medical use is Itraconazole, a triazole antifungal medication, investigated by Iorhuna et al. [125]. Its potential as a corrosion inhibitor is suggested by its distinct molecular structure, composed of several heterocyclic rings with nitrogen atoms. This study aimed to determine, through weight loss measurement analyses, whether Itraconazole can inhibit aluminum corrosion in 0.7 M HCl. Additionally, Furosemide was investigated as a corrosion inhibitor for carbon steel in 1.0 M HCl, [126].
  • Natural organic corrosion inhibitors
The use of naturally occurring mild steel corrosion inhibitors derived from plants dates back to the 1960s, when tannins and their derivatives were employed to prevent steel, iron, and other utensils from corroding. Research indicates that plant-based extracts have been widely and effectively used in acidic media to reduce corrosion [127].
Natural resources have garnered significant attention, including gums, plant extracts/oils, biopolymers, sugars, amino acids, chemical medications (drugs), ionic liquids, and surfactants [128,129]. Green corrosion inhibitors can mitigate the negative impacts of corrosion on the economy, environment, health, communities, biodiversity, and infrastructure, while also helping to conserve biodiversity.
Due to human health and safety concerns, including the use of inexpensive, readily available, eco-friendly, and biodegradable substances, green corrosion inhibitors are the ideal choice. Examples include plant extracts, expired non-toxic medications, and various organic chemicals. These inhibitors vary in cost and performance. The types of natural inhibitors are as follows:
  • Biopolymers
In their natural form, gums are a blend of long-chain polysaccharides typically used in pharmaceutical, cosmetic, and food processing applications. The unique compositions of natural gums make them suitable for use as corrosion inhibitors [130].
In a neutral setting, Danyliak et al. [131] investigated the corrosion-resistance-enhancing capabilities of an environmentally friendly composition based on zinc acetate and polysaccharide (gum Arabic) for low-alloy carbon steel. Gum Arabic nanoparticles have also been studied by Assad et al. [132] for their ability to prevent corrosion in reinforced concrete exposed to a carbon dioxide environment for 180 days.
Using weight loss and Tafel polarization techniques, Abdallah [133] tested Guar gum as a corrosion inhibitor for carbon steel in a 1 M H2SO4 solution. According to the findings of the current study by Venkatesh et al. [134], Guar gum acted both as a corrosion inhibitor and a natural admixture.
Palumbo et al. [135] employed gravimetric and electrochemical methods to examine Guar gum as a potential environmentally friendly corrosion inhibitor for pure aluminum in a 1M HCl solution at various temperatures and immersion times. The findings indicated that Guar gum effectively prevented the corrosion of pure aluminum in the investigated environment.
Biswas et al. [136] demonstrated how Xanthan gum and its graft copolymer can inhibit mild steel corrosion in 15% HCl through both theoretical and experimental investigations. Additionally, in Cao et al.’s study [137], a novel natural polymer inhibitor based on Xanthan gum and β-Cyclodextrin was developed to inhibit the corrosion of L80 steel in 1 mol/L HCl. The corrosion inhibition effect of β-cyclodextrin modified Xanthan gum on L80 steel was assessed using surface analysis technology and electrochemical methods.
Pokhmurs’kyi et al. [138] investigated how Xanthan gum, along with its combination with a biogenic surfactant, prevented carbon steel from corroding in corrosive media containing chloride. The study demonstrated that Xanthan biopolymer effectively inhibits corrosion in steel by adsorbing iron ions and forming complexes with them.
Other gums that have been used as natural corrosion inhibitors have been reported by Vaidya et al. [130], including Rosin (commonly extracted from Pinus trees), Alginate (a biopolymer obtained from brown seaweed such as Laminaria or Fucus or produced by bacteria, Pachylobus edulis gum, Cashew tree gum, Raphia hookeri gum, Locust bean gum, Azadirachta indica gum, etc. Timothy et al. [139] provided a more detailed selection of natural gums for use as corrosion inhibitors.
Plant or animal cells produce macromolecular structures called biopolymers, which can be copolymerized to create grafted biopolymers, enhancing their properties. Polysaccharides such as starch, cellulose derivatives, chitosan, dextrin, pectin, and Guar gum are common examples of corrosion inhibitors. The functional groups in these polysaccharides are chemically modified to produce new biopolymers [140].
Due to their superior barrier properties, ease of customization, and potential for mass production, biopolymers have long been recognized as effective anticorrosion coating materials in both industrial and scientific applications [141].
A study conducted by Akachar et al. [142] investigated the synthesis and properties of a novel class of corrosion inhibitors based on lignin biopolymer and clays. The dispersed clay mineral in the lignin biopolymer, extracted from Alfa fibers, forms a barrier layer that protects against corrosive species.
For the first time, Toghan et al. [143] investigated the effectiveness of two linear biopolymers, keratan and chitosan, as environmentally safe metallic corrosion inhibitors for aluminum (Al) in 1.0 M NaCl solution at various temperatures. Additionally, Hao et al. [144], from Al–Li alloys, developed a green inhibitor based on chitosan (derived from marine biomass: crustaceans) and sodium alginate (from marine biomass: algae), to process waste dust from Al–Li alloys, serving as a modified hydrogen evolution inhibitor.
Sakakihara et al. [145] conducted experiments using a variety of sugars, including reducing sugars aldose (sugar with an aldehyde group), ketose (sugar with a ketone group), and syrup with an expiration date. The skeletal structures of these sugars are strikingly similar, with identical hydroxyl groups and hydrogen atoms at carbon positions 4–6.
To prevent carbon steel from corroding in an acidic environment, Rbaa et al. [146] developed two new sugar-based glucose derivatives: 5,6-anhydro-3-O-dodecyl-1,2-O-isopropylidene-α-d-glucofuranose and 5,6-anhydro-3-O-hexadecyl-1,2-O-isopropylidene-α-d-glucofuranose. Through a chemisorption process, these derivates form a strong interaction that prevents metal dissolution. In another study, Rabaa et al. [147] developed new glucose derivatives based on 8-hydroxyquinoline derived from D-glucose. These substances were evaluated for their ability to inhibit corrosion of mild steel in a 1.0 M HCl medium. According to the polarization study, the two inhibitors function as mixed-type inhibitors.
In this study [148], two thiocarbohydrazide-modified glucose derivatives were synthesized using N-glycosylic linkage. These derivatives serve as environmentally friendly corrosion inhibitors to prevent carbon corrosion of steel pipelines in the oil and gas sector.
  • Amino Acids
Amino acids and their derivatives have made significant progress in corrosion prevention due to their favorable environmental effects and their ability to form effective chelating complexes with metallic substrates. Amino acids such as proline, histidine, and tryptophan, which can bind to metal surfaces, are frequently used as corrosion inhibitors [77]. After reviewing the use of amino acids as corrosion inhibitors, Ibrahimi et al. [129] concluded that nearly all tested amino acid compounds showed strong potential as environmentally friendly inhibitors against metal corrosion in various media.
Using a density functional theory approach, Kumar et al. [149] investigated the corrosion inhibitory effects of cysteine, glutamic acid, glycine, and their derivative glutathione on copper.
A study conducted by Kasprzhitskii et al. [150] assessed the inhibitory effect of L-amino acids with varying side chain lengths on Fe surfaces in the gas phase. Using potentiodynamic polarization and impedance measurements, Sedikk et al. [151] investigated the corrosion inhibition capabilities of several amino acids, including L-cysteine, L-methionine, L-tyrosine, and L-histidine, as environmentally friendly corrosion inhibitors for brass in aerated and stirred 3% NaCl solution.
Yousif et al. [107] showed that under highly acidic conditions, three benzimidazole derivatives containing amino acid units, like tryptophan, tyrosine, and histidine, exhibit exceptional protective properties as sustainable corrosion inhibitors for steel.
Moura et al. [152] reported that, based on the efficacy of inhibitors containing imidazole, the amino acid L-histidine, which features an imidazole ring, was studied as a corrosion inhibitor for AISI 1018 carbon steel in chloride solution.
  • Plant extracts
For several reasons, plant-based extracts have gained significant attention recently [153,154,155,156]. These plants are abundant in many countries, and, due to the simplicity of the extraction process, they have become increasingly valuable [157,158].
Plant-based corrosion inhibitors are not only environmentally friendly, but also highly effective in preventing corrosion. Oils and extracts from plants, which are rich in complex organic compounds, are becoming more popular [159].
A suitable solvent is essential for plant extraction. As the extraction solvent permeates the plant tissue, it dissolves the target component, which is then extracted. The solvent must also maintain a stable redox state, be easy to handle, and not be toxic or flammable [160].
Rached et al. [161] studied Mentha pulegium from the Zaër region, with pulegone as the primary compound. In 0.5 M H2SO4, Mentha pulegium essential oil acted as a corrosion inhibitor for coppersurfaces, achieving an inhibitory efficiency of 91.0%. In another study, the same researchers [162] evaluated the green corrosion inhibition performance of Marrubium vulgare L. essential oil for copper surfaces, also in H2SO4 medium, with 95.51% effectiveness in preventing corrosion. The optimum essential oil concentration was found to be 1 g/L oil. Moreover, Mizoud et al. [159] investigated the essential oil of Urginea maritima for its ability to inhibit copper corrosion in a 0.5 M H2SO4 environment. At a low concentration of 0.5 g/L, the oil achieved an impressive inhibition rate of 94.5%, demonstrating its exceptional efficacy in corrosion control. Additionally, the researchers in ref. [163] explored the corrosion inhibition performance of Rosmarinus officinalis L. essential oil in mild steel exposed to 1 M HCl. The essential oil acted as a cathodic-type inhibitor, reducing corrosion activity by over 90%, at a dose of 1.5 g/L.
Other recent studies on the use of essential oils as environmental corrosion inhibitors have highlighted the use of the Inula Viscosa essential oil for XC48 steel in acidic conditions [164], the Warionia saharea essential oil for mild steel in HCl [165], and the Juniperus oxycedrus essential oil for mild steel in 1.0 M HCl medium [166].
To centralize the use of plant extracts as green corrosion inhibitors, a review table summarizing recent data from the literature was compiled (Table 1).
  • Inorganic inhibitors
Metals and their derived compounds are essential for living organisms, both in beneficial ways and with potential drawbacks, particularly in terms of toxicity, even at lower concentrations. Many of these conventional inhibitors are hazardous and pose serious risks to human health and the environment [182]. For example, some substances form a protective film or barrier on the metal surface, thereby reducing the corrosion rate. Examples include (ortho)phosphates, silicates (SiO3)2 [72], Schiff base molecules [183], chromate (CrO4−2) [184], nitrite (NO2) [185], and molybdate (MoO4−2) [186], and oxides such as aluminum oxide and titanium oxide.
To reduce or stop contact with corrosive agents, the barrier mechanism forms a protective layer on a metal surface. This layer is created either through the adsorption of inhibitor molecules or by the inhibitor’s reaction with metal ions via the passivation mechanism, leading to the formation of a passive film. The inhibitor also functions as a sacrificial anode in the cathodic protection system, which is established through the cathodic protection mechanism [187]. Despite the strong activity of chromium derivatives in corrosion reduction, they are highly toxic and are, therefore, less commonly used in such applications. Lanthanide salts (CeCl3, La(NO3)3, Sm(NO3)3, LaCl3, and SmCl3) present similar toxicity to NaCl [188]. As a result, lanthanide salts are considered efficient and green corrosion inhibitors. CeCl3, for example, has been used as an inhibitor for aluminum alloys and galvanized steel in NaCl solutions [189]. In this study, the authors demonstrated the formation of a protective layer on the surfaces of both type of metals.
  • Hybrid inhibitors (inorganic–organic) corrosion inhibitors
Finding an effective corrosion inhibitor formulation can be challenging, especially when combining different types of chemicals. Typically, the corrosion inhibition effect is assessed by testing individual compounds. However, combining compatible compounds can enhance the performance of the inhibitors. For example, because organic inhibitors have lower conductivity in an electrolyte compared to inorganic inhibitors, their inhibition effect tends to be weaker. As a result, significant research has been conducted on the combination of organic and inorganic inhibitors, aiming to leverage the benefits of both types of compounds.
For improved performance and synergistic corrosion protection, the combination of inorganic and organic components is highly effective. Hybrid coatings, which integrate inorganic materials such as metal oxides or silica nanoparticles with organic polymers, like epoxy, exhibit superior mechanical strength, enhanced adhesion, and increased resistance to environmental threats, including corrosion, due to their complex interactions [190].
Wu et al. [191] studied the combination of zinc oxide, titanium dioxide, and fluoroethylene vinyl ether polymer to develop a multifunctional coating. This organic–inorganic hybrid was sprayed and self-assembled onto the surface of corroded bronze. The aim of the project was to create an effective, low-cost coating technology for protecting bronze. With cross-linked interpenetrating network structures, the hybrid coating prevents corrosive media from directly contacting the bronze substrate, thereby delaying the diffusion of ions. The coating was applied via spraying, and the performance of coated samples was compared to uncoated samples to evaluate its anticorrosion effectiveness.
In this study, Wang et al. [192] developed a simple, one-step spray technique that uses only water to create a modified phosphate-based organic–inorganic composite coating. Alumina and polytetrafluoroethylene nanofillers were incorporated with aluminum phosphate, an inorganic binder, modified with silica sol.
In this study, Kabeb et al. [193] demonstrated how the amount of ammonium polyphosphate influenced the corrosion resistance and flame-retardant properties of hybrid graphene oxide/halloysite/intumescent flame-retardant coatings.
For magnesium alloys, Yu et al. [194] developed a superhydrophobic coating that provides proactive corrosion prevention. Molybdate, an inorganic corrosion inhibitor, and dopamine, an organic corrosion inhibitor, were encapsulated into graphene oxide, which served as a nanocontainer. The incorporation of octadecylamine enhanced the hydrophobic effect. This intelligent nanocontainer was uniformly distributed throughout an organic coating, enabling the spread of corrosion. The resulting superhydrophobic coating exhibited an exceptional hydrophobic effect.
Epoxy–silica organic–inorganic hybrid coatings are engineered with controlled molecular structures to enhance their permeability barrier properties. After 45 days in a 10-weight-percent H2SO4 solution, experiments showed that the 8% Si–O–Si structure provided the best resistance to acid penetration. The organic–inorganic hybrid structure also contributed to the densification of the coating, forming a highly cross-linked network that effectively prevented damage and the diffusion of corrosive media [195].
To control microbially induced corrosion of pipeline materials, Balakrishnan et al. [196] focused on developing a silane-based epoxy–biocide hybrid coating for carbon steels. The study determined the optimal inhibitory concentrations of biocides such as isothiazoline, benzalkonium chloride, and bronopol for the coatings. This highlighted the antimicrobial and anticorrosive properties of these non-oxidizing biocides.
  • Film-forming corrosion inhibitors work by creating a protective barrier or blocking layer made of a substance other than the inhibitor itself. Calcium and zinc salts are known to inhibit cathodic film formation, while benzoate is a typical example of an anodic film-forming inhibitor. These inhibitors act specifically at the cathode or anode [197].
  • Vapor-phase/volatile-phase corrosion inhibitors function by releasing protective vapors that form a thin film over the metal surface to prevent corrosion.
One of the modern technologies used to control corrosion, benefiting the global economy, is volatile corrosion inhibitors. A review of the literature revealed that alkanethiols have been used to prevent corrosion of copper and iron. For copper coated with dopamine in 3.5 weight percent NaCl, 11-mercaptoundecanoic acid serves as a corrosion inhibitor. Additionally, it is used in a phosphate buffer with 0.16 M NaCl to protect stainless steel from corrosion [198].
Volatile-phase corrosion inhibitors effectively prevent the corrosion of metals such as iron, zinc, and aluminum by creating a protective film or neutralizing surrounding reagents. Due to their high volatility, these inhibitors readily evaporate and condense on the metallic surface, providing comprehensive protection. They offer an all-encompassing solution for metal corrosion and can be incorporated into various materials, including coatings, foams, adhesives, powders, sprays, and plastics [199].
Gas-phase deposition holds promise for controlling inhibitor release and enhancing coating properties. However, it requires further refinement to address challenges related to process complexity, cost-efficiency, and scalability, especially when compared to traditional wet chemistry methods.
Vapor corrosion inhibitors, organic substances with vapor pressures ranging from approximately 10−7 to 10−2 mm Hg, are adsorbed onto the metal surface in a few monolayers. They are transported via diffusion through the gas phase, effectively preventing corrosion. Vapor-phase inhibitors are particularly useful for protecting metals in cavities and other hard-to-reach areas [200].
Volatile corrosion inhibitors are typically inorganic compounds with low, yet significant, vapor pressures. These inhibitors form a protective coating on metal surfaces by releasing molecules that adhere to the metal, creating a thin layer. When the corrosive product is soluble in the medium, corrosion accelerates. Volatile corrosive products evaporate, leaving the metal surface exposed to further attack, thereby accelerating the corrosion process [199].
Examples include dicyclohexylamine chromate and benzotriazole for protecting copper, and phenyl thiourea and cyclohexylamine chromate for brass [72].
Shen et al. [201] aimed to enhance the anticorrosion properties of cerium-salt-modified layered double-hydroxide coating through vapor-phase assembly of octanoic acid. The modified coating, created on an aluminum alloy, successfully incorporated octanoic acid into the cerium-salt-modified layered double-hydroxide structure. Electrochemical measurements were used to assess the coating’s corrosion resistance.
A new method for embedding unreacted corrosion inhibitor microparticles in organic coatings using gas deposition in a fluidized bed reactor has been developed. This technique ensures better dispersion of the organic particles and prevents unwanted reactions with the surrounding matrix. Researchers suggest that this method can release substantial amounts of organic corrosion inhibitor at damaged areas. Pulsed gas-phase deposition techniques create nanometer-thick oxide layers on flat, porous substrates. A TiOx nanoscale layer is applied to a model organic inhibitor, and modified inhibitor particles are then coated with a solvent-borne epoxy-amine [202].
Vapor-phase inhibitors have gained attention for their potential application in the marine field. Imidazole compounds, which are inexpensive and derived from vitamin B6 byproducts, serve as stable, high-efficiency, and low-toxicity vapor-phase corrosion inhibitors for iron.
These substances can be biodegraded by bacteria and do not have any negative physiological effects on humans [203]. Vorobyova et al. [204] aimed to investigate the effectiveness of grape pomace extract as an environmentally friendly vapor-phase green inhibitor for steel corrosion. These inhibitors are designed to protect metal surfaces from corrosion during storage and transportation of equipment.
Five amino-benzothiazole derivatives were synthesized and studied using weight loss and potential polarization techniques to evaluate their inhibitory effects on ozone depletion, caprification, and relative humidity at 100%. The aim was to develop corrosion inhibitors for multimetal vapor sputtering. Among the compounds tested, 6-methyl-benzothiazole cinnamate showed the highest performance, achieving an inhibition efficiency of over 90% for all metals under investigation [199]. Mild steel can be protected from atmospheric corrosion using a polyamine compound as a volatile corrosion inhibitor. Specifically, bis-piperidinium methyl-urea was developed as a volatile corrosion inhibitor for mild steel [199].
For example, molybdenum oxide (MoO3) is used as an inhibitor for shipping containers, along with volatile solids such as hexamethylene amine, cyclohexylamine, dicyclohexylamine salts, carboxylates, amides, and amines [8].
  • Nanotechnology corrosion inhibitors
The most effective method for controlled inhibitor delivery, tailored to specific requirement, involves using micro/nanocapsules filled with corrosion inhibitor and featuring a core–shell structure. Nanoinhibitors are recognized as one of the most effective and precise approaches to combat corrosion, due to their unique nanoscale properties, including distinct electrical, optical, magnetic, catalytic, physical, and chemical characteristics [205].
Furthermore, nanoparticles often act as microelectrodes in the presence of electrolytes, accelerating a range of redox reactions through electron transfer processes [206].
Nanotechnology enables new formulas to self-clean and self-heal. If the film is damaged or separates due to flow conditions, it reforms. Additionally, tiny nanoparticles possess a larger surface area, enhancing corrosion resistance [207].
Mesoporous nanosilica particles have been found to be compatible with ceramic, polymeric, and metal materials. Additionally, they serve effectively as a filler to enhance functional properties such as chemical, physical, sensing, electrochemical, drug delivery, selective growth, etc. [208]. The inclusion of nano-ZnO particles resulted in improved surface coverage and a more uniform coating, which further reduced the corrosion rate of the mild steel substrate, ZnTiO3 [40], and nanolignin [73]. While these materials have a high cost, they offer significant performance benefits [85].
Smart coatings corrosion inhibitors:
  • Self-healing coatings
Corrosion-protective coatings can self-heal through intrinsic, extrinsic, or combined healing mechanisms. Intrinsic self-healing relies on the coating’s unique molecular structure and does not require an external component. Damaged coatings can recover through reversible reactions triggered by external stimuli like light, chemicals, temperature, or humidity changes. These coatings offer significant advantages, including the ability to restore barrier properties even after multiple damage–healing cycles. One of their key benefits is their capacity to re-establish protective qualities after repeated damage–healing events [209]. Micro/nanocarriers are frequently designed to actively restore the coating’s protective performance by releasing the encapsulated healing agents in response to environmental changes [42].
Examples include polymer layers, silica–organic layers, conversion layers, metallic layers, and ceramic layers [210].
  • Antifouling corrosion inhibitors
Better corrosion protection is essential. Utilizing carriers loaded with biocides is highly appealing because they allow intelligent release. There are two main approaches for application:
(a)
Adding polymerizable agents to repair defects in the polymeric coating matrix;
(b)
Using corrosion inhibitors to prevent areas from corroding.
Corrosion inhibitors like hydroxyquinoline or triazoles, along with benzalkonium chloride, were encapsulated in mesoporous silica nanoparticles with a pH/sulfide ion-controlled release mechanism [211]. Antifouling agents, such as zinc pyrithione, were encapsulated in gel particles. Additionally, natural and environmentally friendly species with antifouling properties were encapsulated in coatings, microspheres, or polymer particles loaded with silver. Hydrogen-peroxide-producing enzymes were also encapsulated in silica particles using co-precipitation techniques with a polyethyleneimine template [212].
  • Super-hydrophobic coatings
Super-hydrophobicity and hydrophobicity, which make water and aqueous electrolytes repellent, are crucial surface characteristics for corrosion protection. These properties can be achieved by encapsulating beneficial species or by modifying the composition, structure, or morphology of the exterior layers of the coating system [213].
Examples include porous silica capsules and calcium hydroxide inside microcapsules [212]. To modify epoxy coatings, Garcia et al. [214] synthesized poly(urea-formaldehyde) capsules and encapsulated octyldimethylsilyloleate, a silyl ester. Upon reacting with moisture, the active agent forms a hydrophobic layer that enhances corrosion protection.

6. Mechanism of Action of Organic Corrosion Inhibitors

Among the more than 40 adsorption isotherm models, only a few are commonly used for electrochemical adsorption and, consequently, for evaluating the adsorption of corrosion inhibitors. According to data from the literature, the Freundlich, Temkin, Flory–Huggins, El-Awady, Bockris–Swinkels, Volmer, Hill de Boer, Parsons, and Dubinin–Radushkevich isotherms, along with their variations, are frequently applied [215].
The literature extensively describes various mathematical techniques for modeling adsorption mechanisms, including adsorption isotherms, as well as experimental methods for evaluating corrosion inhibition potential, such as weight loss measurement, electrochemical impedance spectroscopy, and scanning electron microscopy [121].
The adsorption isotherm model studies describe the nature of the adsorption or interaction of organic corrosion inhibitors on metallic surfaces. By fitting the degree of surface coverage and corrosion rate to various adsorption isotherm models, like Langmuir, El Awady, Frumkin, Temkin, Freundlich, Flory–Huggins, etc., the optimal adsorption isotherm model can be determined [10,216].
Adsorption isotherms define the relationship between surface coverage and inhibitor concentration based on various hypotheses, linking the equilibrium adsorption constant to fractional surface coverage [217]. These models provide essential insights into the fundamental mechanics of inhibitor adsorption onto metal surfaces [218].
The adsorption equilibrium constant and standard adsorption Gibbs energy are used to establish the relationship between inhibitor surface coverage and concentration in bulk solution. The Langmuir adsorption isotherm is commonly applied in corrosion inhibition research to determine the standard adsorption Gibbs energy based on reported inhibition efficiency [219].
The Langmuir isotherm, introduced in 1916, is the most widely used adsorption model [220]. According to this isotherm, the inhibitor adsorbs onto a surface with a fixed number of adsorption sites, and the adsorption process forms a monolayer [221]. The Langmuir isotherm serves as a fundamental equation for describing the low-concentration adsorption of organic chemicals, especially organic inhibitor molecules, on homogeneous surfaces. It assumes that the Gibbs free energy of adsorption remains constant and that all adsorption sites are equivalent, maintaining uniformity due to the surface cleaning and polishing process [215].
The equilibrium equation for a metal–electrolyte solution containing adsorbed molecules at a constant temperature is represented by the adsorption isotherm. This equation is straightforward for a given metal, electrolyte solution, and inhibitor, as it involves two independent variables: the inhibitor concentration in the solution, cinh(aq), and the degree of surface coverage by inhibitor molecules, θ [175].
c i n h θ = 1 K a d s + c i n h
The El-Awady isotherm assumes multilayer coverage or the binding of multiple molecules at a single active site [222]. This model can provide insight into the presence of multilayer adsorption behavior.
log θ 1 θ = y log c i n h + log K
K a d s = K 1 y
where y represents the number of inhibitor molecules present at a specific active site. When y > 1, it indicates multilayer protection or multibinding occupation [223].
The Temkin isotherm may further support the theory of chemisorptive interaction between inhibitor molecules and an energetically inhomogeneous metal surface [224].
θ = ln c i n h + K a d s
The Freundlich model can be used to analyze the degree of energetic inhomogeneity on the surface [225].
log θ = n   log c i n h + log K a d s
The Frumkin adsorption isotherm model:
log c inh θ 1 θ = 2 α   θ + 2.303 log K a d s
The Flory–Huggins adsorption isotherm model:
log θ c i n h = log K a d s + a log K a d s
Here, Kads—adsorption–desorption constant; θ—surface coverage degree; cinh—concentration of the inhibitor; α = 1/n, where n > 0 is an integer or real number.
Factors such as active sites, aromatic rings, steric hindrance, electron densities, the type of acidic solution, and the interaction between the p-orbital and the d-orbital of iron atoms all influence the stability of organic inhibitor molecules on metallic surfaces [67].
Surface (in)homogeneity for a pair of adsorbate/adsorbent can be determined through advanced analysis of fitting results. The surface can exhibit homogeneity (as described by the Langmuir model), exponential inhomogeneity (as described by the Freundlich model), or uniform inhomogeneity (as described by the Temkin model) [226].
Organic inhibitors can function through physisorption or chemisorption. Physisorption involves the formation of a protective blocking film on the metal’s surface, while chemisorption occurs when inhibitor molecules form direct chemical bonds with the metal’s exterior. Most organic inhibitors influence both cathodic and anodic reactions, but their effectiveness may vary between these processes. The effectiveness of the inhibitor on the metal’s surface is due to the formation of chemisorbed bonds, which create an insoluble layer of inhibitor molecules on the surface, reducing corrosion rates by impeding the metal’s interaction with the environment [227,228].
Organic molecules are generally preferred over the inorganic ones due to their lower toxicity. The physiochemisorption mechanism is linked to the inhibitory activity of organic compounds [228]. The inhibitory effect on corrosion activity is influenced by the presence of polar groups, heterocyclic systems, π-conjugation, and nucleophilicity, all of which significantly influence the chemical structure of the organic molecule [229].

7. Interactions of Corrosion Inhibitors with Metal Surface; Molecular Modeling

Using different organic compounds as corrosion inhibitors can lead to varying degrees of beneficial outcomes in terms of surface protection. Macroscopic effects are unquestionably the result of molecular interactions between chemical compounds and metal atoms, making a thorough evaluation of bond formation essential. This is a complex process that encompasses several aspects (Figure 11):
  • Chemical and electronic key configuration of corrosion inhibitors: This includes factors such as acidic and basic properties, molecular volume, lone pair electrons, dipole effects, etc. All these properties can influence the availability of inhibitor molecules to interact with the surface [230].
  • Bonding of inhibitor with the metal surface.
  • Surface configuration in terms of rugosity and chemistry.
  • Solvent influence: This mainly concerns the competition between inhibitor molecules and solvent molecules for surface sites. Also, the pH of the solvent plays a significant role as it can influence the inhibitor’s structure (whether protonated or non-protonated) and the surface charge based on the point of zero charge (PZC).
  • Effect of anodic and cathodic areas: Various types of reactions and phenomena can occur at the solid surface (Figure 11).
  • Effects of electrode potential: Polarization of the dual system, the inhibitor metal surface, can occur when the solid surface is exposed to different chemical species.
In earlier studies, the assessment of corrosion inhibitor efficiency was related only with electronic properties of the molecules. This approach is known as the molecular electronic-properties to inhibition-efficiency correlation (MEPTIC). Some authors present in their studies the shortcomings [231,232] of this approach in terms of its absence in accounting for other important parameters (Table 2). The correlation of these parameters with state-of-the-art density functional theory (DFT) molecular simulations provides valuable insights into the interactions between inhibitor molecules and metal surfaces. Moreover, the complexity of these phenomena requires the use of machine learning methods like neural networks [233,234,235,236]. Thus, MEPTIC evolved into a more comprehensive manner, like the self-assembled monolayer (SAM) and the evaluation of phase diagrams that account for competitive adsorption on solid surfaces. In a comprehensive study [230], several techniques (X-ray photoelectron spectroscopy, atomic force microscopy, corrosion tests, computational DFT modeling) were used to highlight the influence of molecule nature (2-mercapto-5-methoxybenzimidazole; SH-BimH-5OMe and 5-amino-2-mercaptobenzimidazole; SH-BimH-5NH2) on binding to a copper plate surface. Results from DFT and corrosion tests concur with microscopy and spectroscopy results, showing that the behavior of the molecules differs in terms of formation of ordered or less ordered structures on the surface, depending on the types of bonds formed. SH-BimH-5NH2 can form bonds using both S and NH2 groups, while SH-BimH-5OMe binds using only the S group.

8. Conclusions

The development of corrosion inhibitors, particularly through environmentally friendly and hybrid solutions, has opened new avenues for sustainable corrosion management, with significant benefits for the environment, economy, and society. As mentioned, inhibitors commonly used in the past, although effective, often present toxicity and environmental risks, which has led to a reorientation towards natural and hybrid alternatives. The review highlights recent advances, illustrating the potential of organic, synthetic, and natural inhibitors, as well as nanotechnologies and self-healing systems, which not only demonstrate high efficacy but also align with modern environmental regulations and industrial needs.
The exploration of organic corrosion inhibitors has proven essential in developing sustainable solutions for corrosion protection. This review emphasizes the essential role of organic compounds, particularly those derived from natural sources such as plant extracts, biopolymers, sugars, and gums, in the effective and environmentally friendly inhibition of corrosion. Organic inhibitors, with their abundant π-electrons and electron-dense functional groups, can efficiently adsorb on metal surfaces, forming protective barriers that mitigate both anodic and cathodic reactions. Compared to traditional inorganic inhibitors, organic options offer a distinct advantage in terms of environmental compatibility and reduced toxicity.
Innovations in organic inhibitor formulations, including hybrid systems that combine organic and inorganic components, extend the performance capabilities of these inhibitors. Such advances enable customized solutions for various industries, from petrochemicals to transportation, where corrosion poses significant operational challenges. Although organic inhibitors show promising results, further research is needed to optimize their durability and performance under extreme conditions. Ultimately, the mainstreaming of organic corrosion inhibitors marks a key shift towards safer, more sustainable corrosion management practices that align with global environmental and health standards.
The impetus for these advanced inhibitors comes from both economic pressures—given the global costs of corrosion—and the need to protect public health and the environment. Future research should continue to explore other natural sources and focus on scalability, particularly for industries heavily dependent on metallic infrastructure, such as transportation, energy, and construction.
Scaling up laboratory results to industrial applications involves difficulties, such as providing repeatability and consistency in complex and dynamic environments. Temperature fluctuations, pressure changes, and impurities can all have an impact on performance, making large-scale replication of lab results difficult. Another barrier to overcome is process optimization, which necessitates changes to reactor design, flow rates, and mixing efficiency. Material compatibility is particularly important, as industrial equipment materials may interact differently with chemicals studied at a lab scale. Economic challenges include the cost of raw materials, production, and implementation, while meeting regulatory and safety standards adds another degree of complexity. Long-term impact and durability assessments are required for industrial systems.
There are various challenges to the general adoption of green corrosion inhibitors. High production costs are a major concern, as many green inhibitors are obtained from natural sources, and the extraction and purification processes can be expensive, especially when scaled up for commercial applications. Additionally, their performance may not always match that of synthetic inhibitors, particularly under harsh industrial settings, limiting their application in certain industries. Green inhibitors also have stability and shelf-life issues, as natural substances tend to degrade faster and have a shorter lifespan. Scaling up and adjusting to industrial environments can be difficult. Furthermore, the absence of standardization in testing techniques impacts comparisons between green inhibitors and conventional alternatives.
Market opposition and supply chain boundaries limit the widespread of green inhibitors. Industries frequently hesitate to implement newer technologies because of perceived dangers and disruptions. Green inhibitors may potentially have negative environmental and ecological repercussions, such as diminution of resources or harmful impacts on non-target species. To overcome these obstacles, increasing investment in research and development is critical. Cost-effective extraction and synthesis technologies must be developed, and collaboration among academics, industry, and regulatory organizations can help standardize testing and certification processes. Raising awareness of the environmental benefits of green inhibitors could encourage industries to adopt sustainable alternatives, leading to a greener future.
While significant progress has been made, the journey toward fully sustainable and universally adaptable corrosion inhibitors is still ongoing. Bridging the gap between laboratory-scale efficacy and industrial-scale application will require interdisciplinary research and collaboration. The vision for corrosion inhibition moves beyond protection; it is about proactively preserving global assets through green, high-performance solutions that will set the standard for future infrastructure.

Author Contributions

Conceptualization, D.-I.R., S.-M.A. and E.M.; methodology, D.-I.R. and S.-M.A.; investigation, D.-I.R. and S.-M.A.; writing—review and editing, D.-I.R., S.-M.A. and E.M.; supervision, S.-M.A. and E.M. All authors have read and agreed to the published version of the manuscript.

Funding

The APC amount was provided from internal funds not pertaining any specific project. S.-M.A. gratefully acknowledges the support obtained from the Faculty of Animal Productions Engineering and Management, University of Agronomic, Sciences and Veterinary Medicine of Bucharest.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. NACE Study Estimates Global Cost of Corrosion. Available online: https://www.rustbullet.com/cost-of-corrosion/?srsltid=AfmBOoor60Cy-USkAyUAObxGazC-ThrGmbPpliLgqczvovNQZCWmvT-Q (accessed on 1 January 2025).
  2. Verma, C.; Ebenso, E.; Abdellaziz, B.; Hussain, C.M. Recent developments in Sustainable Corrosion Inhibitors: Design, Performance, and industrial scale applications. Mater. Adv. 2021, 2, 3806–3850. [Google Scholar] [CrossRef]
  3. Yang, L. 1—Introduction. In Techniques for Corrosion Monitoring; Yang, L., Ed.; Woodhead Publishing: Sawston, UK, 2008; pp. 1–5. [Google Scholar]
  4. NACE International. International Measures of Prevention, Application, and Economics of Corrosion Technologies Study (IMPACT); NACE International: Huston, TX, USA, 2016. [Google Scholar]
  5. Koch, G. Cost of corrosion. In Trends in Oil and Gas Corrosion Research and Technologies; Woodhead Publishing: Sawston, UK, 2017; pp. 3–30. [Google Scholar]
  6. Balangao, J.K. Corrosion of Metals: Factors, Types and Prevention Strategies. J. Chem. Health Risks 2024, 14, 79–87. [Google Scholar]
  7. Alkadir Aziz, I.A.; Annon, I.A.; Abdulkareem, M.H.; Hanoon, M.M.; Alkaabi, M.H.; Shaker, L.M.; Alamiery, A.A.; Wan Isahak, W.N.; Takriff, M.S. Insights into Corrosion Inhibition Behavior of a 5-Mercapto-1,2,4-triazole Derivative for Mild Steel in Hydrochloric Acid Solution: Experimental and DFT Studies. Lubricants 2021, 9, 122. [Google Scholar] [CrossRef]
  8. Al-Amiery, A.A.; Yousif, E.; Isahak, W.N.R.W.; Al-Azzawi, W.K. A Review of Inorganic Corrosion Inhibitors: Types, Mechanisms, and Applications. Tribol. Ind. 2023, 45, 313–339. [Google Scholar] [CrossRef]
  9. Dawood, M.A.; Alasady, Z.M.K.; Abdulazeez, M.S.; Ahmed, D.S.; Sulaiman, G.M.; Kadhum, A.A.H.; Shaker, L.M.; Alamiery, A.A. The corrosion inhibition effect of a pyridine derivative for low carbon steel in 1 M HCl medium: Complemented with antibacterial studies. Int. J. Corros. Scale Inhib. 2021, 10, 1766–1782. [Google Scholar]
  10. 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. 2021, 430, 132645. [Google Scholar] [CrossRef]
  11. Abtan, N.; Al-Hamid, M.; Kadhim, L.; Sayyid, F.; Noori, F.; Kadum, A.; Al-Amiery, A.; Al-Azzawi, W. Unlocking the Power of 4-Acetamidoantipyrine: A Promising Corrosion Inhibitor for Preserving Mild Steel in Harsh Hydrochloric Acid Environments. Prog. Color Color. Coat. 2023, 17, 85–96. [Google Scholar] [CrossRef]
  12. Nnoka, M.; Alaso Jack, T.; Szpunar, J. Effects of different microstructural parameters on the corrosion and cracking resistance of pipeline steels: A review. Eng. Fail. Anal. 2024, 159, 108065. [Google Scholar] [CrossRef]
  13. Tait, W.S. Chapter 27—Controlling Corrosion of Chemical Processing Equipment. In Handbook of Environmental Degradation of Materials, 3rd ed.; Kutz, M., Ed.; William Andrew Publishing: Norwich, NY, USA, 2018; pp. 583–600. [Google Scholar]
  14. Kim, Y.-J.; Bahn, C.B.; Baek, S.H.; Choi, W.; Song, G.D. Crevice chemistry and corrosion in high temperature water: A review. Nucl. Eng. Technol. 2024, 56, 3112–3122. [Google Scholar] [CrossRef]
  15. Qasim, A.; Khan, M.S.; Lal, B.; Shariff, A.M. A perspective on dual purpose gas hydrate and corrosion inhibitors for flow assurance. J. Pet. Sci. Eng. 2019, 183, 106418. [Google Scholar] [CrossRef]
  16. Aljibori, H.; Al-Amiery, A.; Isahak, W. Advancements in Corrosion Prevention Techniques. J. Bio- Tribo-Corros. 2024, 10, 78. [Google Scholar] [CrossRef]
  17. Mahmoodian, D.M. Reliability and Maintainability of In-Service Pipelines; Gulf Professional Publishing: Huston, TX, USA, 2018. [Google Scholar]
  18. Ahmad, Z. Principles of Corrosion Engineering and Corrosion Control; Elsevier: Burlington, MA, USA, 2006. [Google Scholar] [CrossRef]
  19. Harsimran, S.; Santosh, K.; Rakesh, K. Overview of Corrosion and its Control: A Critical Review. Proc. Eng. Sci. 2021, 3, 13–24. [Google Scholar] [CrossRef]
  20. Nogara, J.; Zarrouk, S.J. Corrosion in geothermal environment: Part 1: Fluids and their impact. Renew. Sustain. Energy Rev. 2018, 82, 1333–1346. [Google Scholar] [CrossRef]
  21. Fontana, M.G.; Greene, N.D.; Klerer, J. Corrosion Engineering. J. Electrochem. Soc. 1968, 115, 142C. [Google Scholar] [CrossRef]
  22. Zade, G.S.; Patil, K.D. Advances in Corrosion-Resistant Coatings. In Functional Coatings; John Wiley & Sons: Hoboken, NJ, USA, 2024; pp. 110–152. [Google Scholar]
  23. Liu, Q.; Qian, J.; Barker, R.; Wang, C.; Neville, A.; Pessu, F. Application of Double Loop Electrochemical Potentio-kinetic Reactivation for characterizing the intergranular corrosion susceptibility of stainless steels and Ni-based alloys in solar nitrate salts used in CSP systems. Eng. Fail. Anal. 2021, 129, 105717. [Google Scholar] [CrossRef]
  24. Yang, Z.; Li, L.; Qiao, Y.; Li, C.; Zhang, L.; Cui, J.; Ren, D.; Ji, H.; Zheng, Y. Cavitation erosion-corrosion properties of as-cast TC4 and LPBF TC4 in 0.6 mol/L NaCl solution: A comparison investigation. Ultrason. Sonochem. 2024, 108, 106947. [Google Scholar] [CrossRef]
  25. Li, L.; Nie, S.; Li, C.; Chen, X.; Qiao, Y.; Ma, R.; Chen, Z.; Zhang, L.; Cui, J. Study on cavitation erosion-corrosion behavior of CoCrFeNiMoCu0.1 high entropy alloy in 3.5 wt% NaCl solution. Ultrason. Sonochem. 2024, 110, 107021. [Google Scholar] [CrossRef]
  26. Reclaru, L.; Brooks, R.A.; Zuberbühler, M.; Eschler, P.Y.; Constantin, F.; Tomoaia, G. Evaluation of taper joints with combined fatigue and crevice corrosion testing: Comparison to human explanted modular prostheses. Mater. Sci. Eng. C 2014, 34, 69–77. [Google Scholar] [CrossRef]
  27. Schijve, J. Fretting Corrosion. In Fatigue of Structures and Materials; Springer: Dodrecht, The Netherlands, 2001. [Google Scholar]
  28. Royhman, D.; Pourzal, R.; Hall, D.; Lundberg, H.J.; Wimmer, M.A.; Jacobs, J.; Hallab, N.J.; Mathew, M.T. Fretting-corrosion in hip taper modular junctions: The influence of topography and pH levels—An in-vitro study. J. Mech. Behav. Biomed. Mater. 2021, 118, 104443. [Google Scholar] [CrossRef]
  29. Pal, M.; Lavanya, M. Microbial Influenced Corrosion: Understanding Bioadhesion and Biofilm Formation. J. Bio- Tribo-Corros. 2022, 8, 76. [Google Scholar] [CrossRef]
  30. Gonçalves, M.C.; Margarido, F. Materials for Construction and Civil Engineering: Science, Processing, and Design. In Corrosion; Springer International Publishing: Cham, Switzerland, 2015; pp. 679–716. [Google Scholar]
  31. Hansen, D. Metal Corrosion in the Human Body: The Ultimate Bio-Corrosion Scenario. Electrochem. Soc. Interface 2008, 17, 31–34. [Google Scholar] [CrossRef]
  32. Liu, D.; Jia, R.; Xu, D.; Yang, H.; Zhao, Y.; Khan, M.s.; Huang, S.; Wen, J.; Yang, K.; Gu, T. Biofilm inhibition and corrosion resistance of 2205-Cu duplex stainless steel against acid producing bacterium Acetobacter aceti. J. Mater. Sci. Technol. 2019, 35, 2494–2502. [Google Scholar] [CrossRef]
  33. Lou, Y.; Chang, W.; Cui, T.; Qian, H.; Huang, L.; Ma, L.; Hao, X.; Zhang, D. Microbiologically influenced corrosion inhibition of carbon steel via biomineralization induced by Shewanella putrefaciens. npj Mater. Degrad. 2021, 5, 59. [Google Scholar] [CrossRef]
  34. Lekbach, Y.; Liu, T.; Li, Y.; Moradi, M.; Dou, W.; Xu, D.; Smith, J.A.; Lovley, D.R. Microbial corrosion of metals: The corrosion microbiome. Adv. Microb. Physiol. 2021, 78, 317–390. [Google Scholar] [CrossRef]
  35. Lekbach, Y.; Li, Z.; Xu, D.; El Abed, S.; Dong, Y.; Liu, D.; Gu, T.; Koraichi, S.I.; Yang, K.; Wang, F. Salvia officinalis extract mitigates the microbiologically influenced corrosion of 304L stainless steel by Pseudomonas aeruginosa biofilm. Bioelectrochemistry 2019, 128, 193–203. [Google Scholar] [CrossRef]
  36. Zhou, E.; Li, H.; Yang, C.; Wang, J.; Xu, D.; Zhang, D.; Gu, T. Accelerated corrosion of 2304 duplex stainless steel by marine Pseudomonas aeruginosa biofilm. Int. Biodeterior. Biodegrad. 2018, 127, 1–9. [Google Scholar] [CrossRef]
  37. Jia, R.; Yang, D.; Abd Rahman, H.B.; Gu, T. Laboratory testing of enhanced biocide mitigation of an oilfield biofilm and its microbiologically influenced corrosion of carbon steel in the presence of oilfield chemicals. Int. Biodeterior. Biodegrad. 2017, 125, 116–124. [Google Scholar] [CrossRef]
  38. Thakur, A.; Kumar, A. Hybrid Nanomaterials as Next-Generation Corrosion Inhibitors for Metals and Alloys. In Innovations and Applications of Hybrid Nanomaterials; Khanna, V., Sharma, P., Mahajan, P., Eds.; IGI Global: Hershey, PA, USA, 2024; pp. 91–117. [Google Scholar]
  39. Nazeer, A.A.; Madkour, M. Potential use of smart coatings for corrosion protection of metals and alloys: A review. J. Mol. Liq. 2018, 253, 11–22. [Google Scholar] [CrossRef]
  40. Kumar, P.; Soni, I.; Jayaprakash, G.; Kumar, S.; Rao, S.; Flores-Moreno, R.; Swamirayachar, S. Experimental and theoretical studies of hexylmeythylimidazolium tetrafluoroborate ionic liquid as cathodic corrosion inhibitor for mild steel. Inorg. Chem. Commun. 2022, 146, 110110. [Google Scholar] [CrossRef]
  41. Zehra, S.; Mobin, M.; Aslam, R. Chapter 2—Corrosion prevention and protection methods. In Eco-Friendly Corrosion Inhibitors; Guo, L., Verma, C., Zhang, D., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 13–26. [Google Scholar]
  42. Auepattana-Aumrung, K.; Crespy, D. Self-healing and anticorrosion coatings based on responsive polymers with metal coordination bonds. Chem. Eng. J. 2023, 452, 139055. [Google Scholar] [CrossRef]
  43. Shaw, B.A.; Kelly, R. What is Corrosion? Electrochem. Soc. Interface 2006, 15, 24–26. [Google Scholar] [CrossRef]
  44. Olajire, A. Recent advances on organic coating system technologies for corrosion protection of offshore metallic structures. J. Mol. Liq. 2018, 269, 572–606. [Google Scholar] [CrossRef]
  45. Hayatdavoudi, H.; Rahsepar, M. A mechanistic study of the enhanced cathodic protection performance of graphene-reinforced zinc rich nanocomposite coating for corrosion protection of carbon steel substrate. J. Alloys Compd. 2017, 727, 1148–1156. [Google Scholar] [CrossRef]
  46. Presuel-Moreno, F.; Jakab, M.A.; Tailleart, N.; Goldman, M.; Scully, J.R. Corrosion-resistant metallic coatings. Mater. Today 2008, 11, 14–23. [Google Scholar] [CrossRef]
  47. Mondal, K.; Nunez, L.; Downey, C.; Van Rooyen, I. Recent Advances in the Thermal Barrier Coatings for Extreme Environments. Mater. Sci. Energy Technol. 2021, 4, 208–210. [Google Scholar] [CrossRef]
  48. Hussain, A.; Nuthalapati, S.; Pichumani, M.; Chakra, C. Research progress in organic zinc rich primer coatings for cathodic protection of metals—A comprehensive review. Prog. Org. Coat. 2021, 153, 106040. [Google Scholar] [CrossRef]
  49. Okokpujie, I.; Tartibu, L.; Musa-Basheer, H.; Adeoye, A. Effect of Coatings on Mechanical, Corrosion and Tribological Properties of Industrial Materials: A Comprehensive Review. J. Bio- Tribo-Corros. 2023, 10, 2. [Google Scholar] [CrossRef]
  50. Xie, J.; Zhang, J.; You, Z.; Liu, S.; Guan, K.; Wu, R.; Wang, J.; Feng, J. Towards developing Mg alloys with simultaneously improved strength and corrosion resistance via RE alloying. J. Magnes. Alloys 2021, 9, 41–56. [Google Scholar] [CrossRef]
  51. Pedeferri, P. Cathodic and Anodic Protection. In Corrosion Science and Engineering; Springer: Cham, Switzerland, 2018; pp. 383–422. [Google Scholar]
  52. Kadhum, A.; Betti, N.; Al-Adili, A.; Shaker, L.; Al-Amiery, A. Limits and developments in organic inhibitors for corrosion of mild steel: A critical review (Part two: 4-aminoantipyrine). Int. J. Corros. Scale Inhib. 2022, 11, 43–63. [Google Scholar] [CrossRef]
  53. Clematis, D.; Marroccu, A.; Panizza, M.; Barbucci, A. A Critical Analysis on the Current Design Criteria for Cathodic Protection of Ships and Superyachts. Materials 2022, 15, 2645. [Google Scholar] [CrossRef]
  54. Cheng, L.; Luo, Y.; Ma, S.; Guo, W.; Xiaohui, W. Corrosion resistance of inorganic zinc-rich coating reinforced by Ni-coated coal fly ash. J. Alloys Compd. 2019, 786, 791–797. [Google Scholar] [CrossRef]
  55. Silva, R.; Alemán, C.; Ferreira, C.A.; Armelin, E.; Ferreira, J.; Meneguzzi, Á. Smart Paint for anodic protection of steel. Prog. Org. Coat. 2014, 78, 116–123. [Google Scholar] [CrossRef]
  56. Edeleanu, C. Corrosion Control by Anodic Protection. Platin. Met. Rev. 1960, 4, 86–91. [Google Scholar] [CrossRef]
  57. Cowley, W.E.; Robinson, F.P.A.; Kerrich, J.E. Anodic Protection for the Control of Corrosion Fatigue. Br. Corros. J. 1968, 3, 223–237. [Google Scholar] [CrossRef]
  58. Paulekat, F.; Gräfen, H.; Kuron, D. Anodic corrosion protection of sulphuric acid plants with regard to the recovery of heat. Mater. Corros. 1982, 33, 254–262. [Google Scholar] [CrossRef]
  59. Ohtsuka, T. Corrosion Protection of Steels by Conducting Polymer Coating. Int. J. Corros. 2012, 2012, 915090. [Google Scholar] [CrossRef]
  60. Vaghefinazari, B.; Wierzbicka, E.; Visser, P.; Posner, R.; Arrabal, R.; Matykina, E.; Mohedano, M.; Blawert, C.; Zheludkevich, M.L.; Lamaka, S.V. Chromate-Free Corrosion Protection Strategies for Magnesium Alloys—A Review: Part III—Corrosion Inhibitors and Combining Them with Other Protection Strategies. Materials 2022, 15, 8489. [Google Scholar] [CrossRef]
  61. Liu, C.; Tong, S.; Yue, Y.; Wang, H.; Song, J.; Li, Y.; Wang, Q.; Wang, Z. Laser-based fabrication of superwetting titanium alloy with enhanced corrosion and erosion-corrosion resistance. Colloids Surf. A Physicochem. Eng. Asp. 2024, 688, 133648. [Google Scholar] [CrossRef]
  62. Emelyanenko, K.A.; Emelyanenko, A.M.; Boinovich, L.B. Laser Obtained Superhydrophobic State for Stainless Steel Corrosion Protection, A Review. Coatings 2023, 13, 194. [Google Scholar] [CrossRef]
  63. Cholkar, A.; Chatterjee, S.; Richards, C.; McCarthy, É.; Perumal, G.; Regan, F.; Kinahan, D.; Brabazon, D. Biofouling and Corrosion Protection of Aluminum Alloys Through Ultrafast Laser Surface Texturing for Marine Applications. Adv. Mater. Interfaces 2024, 11, 2300835. [Google Scholar] [CrossRef]
  64. Liu, Z.; Liu, H.; Viejo, F.; Aburas, Z.; Rakhes, M. Laser-induced microstructural modification for corrosion protection. Proc. Inst. Mech. Eng. Part C J. Mech. Eng. Sci. 2010, 224, 1073–1085. [Google Scholar] [CrossRef]
  65. Krishna, L.R.; Madhavi, Y.; Babu, P.S.; Rao, D.S.; Padmanabham, G. Strategies for corrosion protection of non-ferrous metals and alloys through surface engineering. Mater. Today Proc. 2019, 15, 145–154. [Google Scholar] [CrossRef]
  66. Macera, L.; Pullini, D.; Boschetto, A.; Bottini, L.; Mingazzini, C.; Falleti, G.L. Sol–Gel Silica Coatings for Corrosion Protection of Aluminum Parts Manufactured by Selective Laser Melting (SLM) Technology. Coatings 2023, 13, 1081. [Google Scholar] [CrossRef]
  67. Al-Ghezi, M.; Gaaz, T. A mini review on corrosion, inhibitors and mechanism types of mild steel inhibition in an acidic environment. Int. J. Corros. Scale Inhib. 2021, 10, 861–884. [Google Scholar] [CrossRef]
  68. Hill, R.T.; Ramirez, F.A.; Perez, A.L.; Monty, B.A. Material Selection and Corrosion Control for Topside Process and Utility Piping and Equipment. In Proceedings of the NACE CORROSION 2012, Salt Lake City, UT, USA, 11–15 March 2012. [Google Scholar]
  69. İpek, M.; Selvi, İ.H.; Findik, F.; Torkul, O.; Cedimoğlu, I.H. An expert system based material selection approach to manufacturing. Mater. Des. 2013, 47, 331–340. [Google Scholar] [CrossRef]
  70. Vachtsevanos, G.; Natarajan, K.; Rajamani, R.; Sandborn, P. Corrosion Processes: Sensing, Monitoring, Data Analytics, Prevention/Protection, Diagnosis/Prognosis and Maintenance Strategies: Sensing; Springer: Cham, Switzerland, 2020. [Google Scholar]
  71. Arturson, G. The tragedy of San Juanico—The most severe LPG disaster in history. Burns 1987, 13, 87–102. [Google Scholar] [CrossRef]
  72. Palanisamy, G. Corrosion Inhibitors; IntechOpen: London, UK, 2019. [Google Scholar]
  73. Luo, Z.-G.; Zhang, Y.; Wang, H.; Wan, S.; Song, L.-F.; Liao, B.-K.; Guo, X.-P. Modified Nano-Lignin as a Novel Biomass-Derived Corrosion Inhibitor for Enhanced Corrosion Resistance of Carbon Steel. Corros. Sci. 2023, 227, 111705. [Google Scholar] [CrossRef]
  74. Sowmyashree, A.S.; Somya, A.; Kumar, C.P.; Rao, S. Novel nano corrosion inhibitor, integrated zinc titanate nano particles: Synthesis, characterization, thermodynamic and electrochemical studies. Surf. Interfaces 2021, 22, 100812. [Google Scholar] [CrossRef]
  75. Zhu, X.; Wang, C. Corrosion inhibition performance of an environmentally friendly water-based nano-inhibitor. Alex. Eng. J. 2023, 70, 495–501. [Google Scholar] [CrossRef]
  76. Aljibori, H.; Abdulzahra, O.; Adily, A.; Al-Azzawi, W.; Al-Amiery, A.; Kadhum, A. Corrosion inhibition effects of concentration of 2-oxo-3- hydrazonoindoline in acidic solution, exposure period, and temperature. Int. J. Corros. Scale Inhib. 2023, 12, 438–457. [Google Scholar] [CrossRef]
  77. Verma, C.; Al-Moubaraki, A.; Alfantazi, A.; Rhee, K.Y. Heterocyclic amino acids-based green and sustainable corrosion inhibitors: Adsorption, bonding and corrosion control. J. Clean. Prod. 2024, 446, 141186. [Google Scholar] [CrossRef]
  78. Al-Amiery, A.A.; Isahak, W.N.R.W.; Al-Azzawi, W.K. Corrosion Inhibitors: Natural and Synthetic Organic Inhibitors. Lubricants 2023, 11, 174. [Google Scholar] [CrossRef]
  79. Wonnie Ma, I.A.; Ammar, S.; Kumar, S.S.A.; Ramesh, K.; Ramesh, S. A concise review on corrosion inhibitors: Types, mechanisms and electrochemical evaluation studies. J. Coat. Technol. Res. 2021, 19, 241–268. [Google Scholar] [CrossRef]
  80. Bastidas, D.M.; Martin, U.; Bastidas, J.M.; Ress, J. Corrosion Inhibition Mechanism of Steel Reinforcements in Mortar Using Soluble Phosphates: A Critical Review. Materials 2021, 14, 6168. [Google Scholar] [CrossRef]
  81. Mohamed, K.E.M.; Ibrahim, O.H.; El-Bedawy, M.E.; Ali, A.H. Synergistic effect of different Zn salts with sodium octanoate on the corrosion inhibition of carbon steel in cooling water. J. Radiat. Res. Appl. Sci. 2020, 13, 276–287. [Google Scholar] [CrossRef]
  82. Nishimoto, M.; Muto, I.; Sugawara, Y.; Hara, N. Role of Cerium Ions for Improving Pitting Corrosion Resistance of Sulfide Inclusions in Stainless Steels. In Electrochemical Society Meeting Abstracts; IOP Publishing: Bristol, UK, 2017; Volume MA2017-02, p. 698. [Google Scholar] [CrossRef]
  83. Aliofkhazraei, M.; Gharabagh, R.S.; Teimouri, M.; Ahmadzadeh, M.; Hasannejad, H.; Seyfoori, A. Effect of cerium ion addition on corrosion and wear characteristics of plasma electrolytic oxidation coating of CP-Ti. Prot. Met. Phys. Chem. Surf. 2016, 52, 1093–1099. [Google Scholar] [CrossRef]
  84. Özcan, M.; Dehri, İ.; Erbil, M. Organic sulphur-containing compounds as corrosion inhibitors for mild steel in acidic media: Correlation between inhibition efficiency and chemical structure. Appl. Surf. Sci. 2004, 236, 155–164. [Google Scholar] [CrossRef]
  85. Ahmed, A.A.; Al-mashhadani, M.H.; Hussain, Z.; Mohammed, S.A.; Yusop, R.M.; Yousif, E. Inhibition of Corrosion: Mechanisms and Classifications an Overview. Al-Qadisiyah J. Pure Sci. 2020, 25, 1–9. [Google Scholar] [CrossRef]
  86. HosseinpourRokni, M.; Naderi, R.; Soleimani, M.; Kowsari, E.; Pourfath, M. Indirect interactions between the ionic liquid and Cu surface in 0.5 M HCl: A novel mechanism explaining cathodic corrosion inhibition. Corros. Sci. 2023, 216, 111100. [Google Scholar] [CrossRef]
  87. Nam, N.; Hien, P.; Hoai, N.; Thu, V. A study on the mixed corrosion inhibitor with a dominant cathodic inhibitor for mild steel in aqueous chloride solution. J. Taiwan Inst. Chem. Eng. 2018, 91, 556–569. [Google Scholar] [CrossRef]
  88. Borbolla, J.; Thangarasu, P.; García-Ochoa, E. A new inhibitor for mild carbon steel: Electrochemical and DFT studies. J. Electroanal. Chem. 2005, 583, 8–16. [Google Scholar] [CrossRef]
  89. Coelho, L.B.; Lukaczynska-Anderson, M.; Clerick, S.; Buytaert, G.; Lievens, S.; Terryn, H.A. Corrosion inhibition of AA6060 by silicate and phosphate in automotive organic additive technology coolants. Corros. Sci. 2022, 199, 110188. [Google Scholar] [CrossRef]
  90. Wang, X.; Yang, J.; Chen, X.; Liu, C.; Zhao, J. Synergistic inhibition properties and microstructures of self-assembled imidazoline and phosphate ester mixture for carbon steel corrosion in the CO2 brine solution. J. Mol. Liq. 2022, 357, 119140. [Google Scholar] [CrossRef]
  91. Palomar, M.E.; Olivares-Xometl, C.O.; Likhanova, N.V.; Pérez-Navarrete, J.-B. Imidazolium, Pyridinium and Dimethyl-Ethylbenzyl Ammonium Derived Compounds as Mixed Corrosion Inhibitors in Acidic Medium. J. Surfactants Deterg. 2011, 14, 211–220. [Google Scholar] [CrossRef]
  92. Richards, C.A.J.; McMurray, H.N.; Williams, G. Smart-release inhibition of corrosion driven organic coating failure on zinc by cationic benzotriazole based pigments. Corros. Sci. 2019, 154, 101–110. [Google Scholar] [CrossRef]
  93. Guo, L.; Qi, C.; Zheng, X.; Zhang, R.; Shen, X.; Kaya, S. Toward understanding the adsorption mechanism of large size organic corrosion inhibitors on an Fe(110) surface using the DFTB method. RSC Adv. 2017, 7, 29042–29050. [Google Scholar] [CrossRef]
  94. Talat, R.; Asghar, M.A.; Tariq, I.; Akhter, Z.; Liaqat, F.; Nadeem, L.; Haider, A.; Ali, S. Evaluating the Corrosion Inhibition Efficiency of Pyridinium-Based Cationic Surfactants for EN3B Mild Steel in Acidic-Chloride Media. Coatings 2022, 12, 1701. [Google Scholar] [CrossRef]
  95. Fawzy, A.; Abdallah, M.; Zaafarany, I.; Ahmed, S.; Althagafi, I. 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. 2018, 265, 276–291. [Google Scholar] [CrossRef]
  96. Pour-Ali, S.; Hejazi, S. Tiazofurin drug as a new and non-toxic corrosion inhibitor for mild steel in HCl solution: Experimental and quantum chemical investigations. J. Mol. Liq. 2022, 354, 118886. [Google Scholar] [CrossRef]
  97. Al-Gorair, A.; Aboubaker Saleh, M.; Alotaibi, M.; Al-Juaid, S.; Abdallah, M.; Wanees, S. Potentiometric and polarization studies on the oxide film repair and retardation of pitting corrosion on indium in an alkaline aqueous solution utilizing some triazole compounds. Inorg. Chem. Commun. 2023, 158, 111497. [Google Scholar] [CrossRef]
  98. Mehta, R.K.; Yadav, M. Corrosion inhibition properties of expired Broclear medicine and its carbon dot as eco-friendly inhibitors for mild steel in 15% HCl. Mater. Sci. Eng. B 2023, 295, 116566. [Google Scholar] [CrossRef]
  99. Kokalj, A. Corrosion inhibitors: Physisorbed or chemisorbed? Corros. Sci. 2022, 196, 109939. [Google Scholar] [CrossRef]
  100. Hrimla, M.; Bahsis, L.; Laamari, M.R.; Julve, M.; Stiriba, S.-E. An Overview on the Performance of 1,2,3-Triazole Derivatives as Corrosion Inhibitors for Metal Surfaces. Int. J. Mol. Sci. 2022, 23, 16. [Google Scholar] [CrossRef]
  101. Verma, C.; Quraishi, M.A.; Ebenso, E.E. Quinoline and its derivatives as corrosion inhibitors: A review. Surf. Interfaces 2020, 21, 100634. [Google Scholar] [CrossRef]
  102. Dahmani, K.; Galai, M.; Rbaa, M.; Ech-Chebab, A.; Errahmany, N.; Guo, L.; AlObaid, A.A.; Hmada, A.; Warad, I.; Lakhrissi, B.; et al. Evaluating the efficacy of synthesized quinoline derivatives as Corrosion inhibitors for mild steel in acidic environments: An analysis using electrochemical, computational, and surface techniques. J. Mol. Struct. 2024, 1295, 136514. [Google Scholar] [CrossRef]
  103. Abdel-karim, A.M.; Hussien, H.M.; Shahen, S.; El-Shamy, O.A.A.; Ghayad, I.M.; Saleh, N.M.; El-Sattar, N.E.A.A. Green synthesis of novel bis structure of (Carbamothioyl) oxalamide derivatives as corrosion inhibitors for copper in 3.5% NaCl; experimental and theoretical investigation. J. Mol. Struct. 2024, 1295, 136597. [Google Scholar] [CrossRef]
  104. Aiad, I.A.; Hafiz, A.A.; El-Awady, M.Y.; Habib, A.O. Some Imidazoline Derivatives as Corrosion Inhibitors. J. Surfactants Deterg. 2010, 13, 247–254. [Google Scholar] [CrossRef]
  105. Ettahiri, W.; El Moutaouakil Ala Allah, A.; Lazrak, J.; Safir, E.H.; Yadav, K.K.; Hammouti, B.; Obaidullah, A.J.; Rais, Z.; Ramli, Y.; Taleb, M. Synthesis, characterization, theoretical, and experimental evaluation of novel imidazolone—Based compounds as eco-friendly corrosion inhibitors for mild steel. J. Ind. Eng. Chem. 2024, 140, 631–646. [Google Scholar] [CrossRef]
  106. Touir, R.; Errahmany, N.; Rbaa, M.; Benhiba, F.; Doubi, M.; Kafssaoui, E.H.; 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]
  107. Yousif, Q.A.; Bedair, M.A.; Fadel, Z.; Al-Odail, F.; Abuelela, A.M. Evaluating the efficacy of newly synthesized amino acid derivatives as corrosion inhibitors in acidic solutions. Inorg. Chem. Commun. 2024, 164, 112454. [Google Scholar] [CrossRef]
  108. Xiong, Y.; Cao, M. Application of surfactants in corrosion inhibition of metals. Curr. Opin. Colloid Interface Sci. 2024, 73, 101830. [Google Scholar] [CrossRef]
  109. Wang, T.; Dai, S.; Xiong, Y.; Yan, H.; Zhu, Y. The morpholine surfactants with corrosion inhibition and antibacterial activity: Experiments and theoretical calculations. Colloids Surf. A Physicochem. Eng. Asp. 2024, 700, 134784. [Google Scholar] [CrossRef]
  110. El-Monem, M.; Farag, A.; Khalil, M.; Migahed, M. Hydroxyethyl cationic surfactants as corrosion inhibitors for S90 steel in produced water: Electrochemical, and computational studies. J. Mol. Struct. 2024, 1314, 138702. [Google Scholar] [CrossRef]
  111. Yumnam, G.; Adhikari, S.; Pulikkal, A.K.; Rajaraman, P. Impacts of pyridinium gemini surfactants on corrosion inhibition of carbon steel. Surf. Interfaces 2023, 45, 103796. [Google Scholar] [CrossRef]
  112. Wang, F.; Zhang, S.; Ma, T.; Shi, Y.; Wang, X.; Du, H.; Liu, R.; Han, X. Three anionic surfactants for corrosion inhibition in cobalt CMP: Research on validity and mechanism. Surf. Interfaces 2024, 47, 104202. [Google Scholar] [CrossRef]
  113. Hameed, R.S.; Aleid, G.; Ragab, H.; Alshafey, H.; Aljuhani, E.; Almhyawi, S. Green synthesis of polymeric surfactants from recycling of plastic waste for applications on steel protection in the petroleum industry. Green Chem. Lett. Rev. 2024, 17, 2379442. [Google Scholar] [CrossRef]
  114. Ikeuba, A.I.; Essiet, N.; Echem, O.C.; Ezenobi, N.; Okon, E.; Okafor, P.C. A review of the application of ionic liquids as eco-friendly corrosion inhibitors for steel, aluminum, copper and magnesium alloys. J. Ion. Liq. 2024, 4, 100098. [Google Scholar] [CrossRef]
  115. El-Nagar, R.A.; Khalil, N.A.; Atef, Y.; Nessim, M.I.; Ghanem, A. Evaluation of ionic liquids based imidazolium salts as an environmentally friendly corrosion inhibitors for carbon steel in HCl solutions. Sci. Rep. 2024, 14, 1889. [Google Scholar] [CrossRef]
  116. Jia, H.; Jia, H.; Wang, Q.; Xu, Y.; Wang, B.; Wang, Q.; Li, X.; Wang, Z.; Lv, K.; Huang, P. Imidazolium-Based Polymeric Ionic Liquids with Short Alkyl Chains as Green Corrosion Inhibitors for Mild Steel in 1 M HCl: Experimental and Theoretical Investigations. Langmuir 2024, 40, 14141–14152. [Google Scholar] [CrossRef]
  117. Zunita, M.; Kevin, Y.J. Ionic liquids as corrosion inhibitor: From research and development to commercialization. Results Eng. 2022, 15, 100562. [Google Scholar] [CrossRef]
  118. Cui, L.; Lv, Y.; Dong, Y.; Liao, H.; Wu, S.; Li, X. Benzothiazole-based ionic liquids as environment-friendly and high-efficiency corrosion inhibitors for mild steel in HCl: Experimental and theoretical studies. J. Mol. Liq. 2024, 394, 123769. [Google Scholar] [CrossRef]
  119. Wanees, S.; Kamel, M.; Ibrahim, M.; Rashwan, S.; Atef, Y.; Gamal Abd Elsadek Ghunime, M. Corrosion inhibition and synergistic effect of ionic liquids and iodide ions on the corrosion of C-steel in formation water associated with crude oil. J. Umm Al-Qura Univ. Appl. Sci. 2023, 10, 107–119. [Google Scholar] [CrossRef]
  120. Vaszilcsin, N.; Kellenberger, A.; Dan, M.L.; Duca, D.A.; Ordodi, V.L. Efficiency of Expired Drugs Used as Corrosion Inhibitors: A Review. Materials 2023, 16, 5555. [Google Scholar] [CrossRef]
  121. Lessa, R.C.D.S. Synthetic Organic Molecules as Metallic Corrosion Inhibitors: General Aspects and Trends. Organics 2023, 4, 232–250. [Google Scholar] [CrossRef]
  122. Chauhan, D.S.; Verma, C.; Quraishi, M.A. Molecular structural aspects of organic corrosion inhibitors: Experimental and computational insights. J. Mol. Struct. 2021, 1227, 129374. [Google Scholar] [CrossRef]
  123. Oubahou, M.; el Aloua, A.; Benzbiria, N.; Harrari, S.; Takky, D.; Youssef, N.; Zeroual, A.; Wang, S.-F.; Syed, A.; Belghiti, M.E. Electrochemical, Thermodynamic and Computational Investigation of the Use of an Expired Drug as a Sustainable Corrosion Inhibitor for Copper in 0.5 M H2SO4. Mater. Chem. Phys. 2024, 323, 129642. [Google Scholar] [CrossRef]
  124. Michael, O.; Thompson, A. Experimental Evaluation of Unpreprocessed-Expired Paracetamol Drugs as Corrosion Inhibitor for Mild Steel in Hydrochloric Acid. In Proceedings of the 2023 FUTA SEET Conference, Akure, Nigeria, 7–9 November 2023. [Google Scholar]
  125. Iorhuna, F. Itraconazole Drug as Corrosion Inhibitor for Aluminium in 0.7 M Hydrochloric Acid Medium. J. Mater. Environ. Sci. 2024, 15, 501–511. [Google Scholar]
  126. Abd El Maksoud, S.; Fouda, A.E.A.; Badawy, H. Furosemide drug as a corrosion inhibitor for carbon steel in 1.0 M hydrochloric acid. Sci. Rep. 2024, 14, 9052. [Google Scholar] [CrossRef]
  127. Emmanuel, J. Corrosion protection of mild steel in corrosive media, a shift from synthetic to natural corrosion inhibitors: A review. Bull. Natl. Res. Cent. 2024, 48, 26. [Google Scholar] [CrossRef]
  128. Salleh, S.Z.; Yusoff, A.H.; Zakaria, S.K.; Taib, M.A.A.; Abu Seman, A.; Masri, M.N.; Mohamad, M.; Mamat, S.; Ahmad Sobri, S.; Ali, A.; et al. Plant extracts as green corrosion inhibitor for ferrous metal alloys: A review. J. Clean. Prod. 2021, 304, 127030. [Google Scholar] [CrossRef]
  129. El Ibrahimi, B.; Jmiai, A.; Bazzi, L.; El Issami, S. Amino acids and their derivatives as corrosion inhibitors for metals and alloys. Arab. J. Chem. 2020, 13, 740–771. [Google Scholar] [CrossRef]
  130. Vaidya, N.R.; Aklujkar, P.; Rao, A. Modification of natural gums for application as corrosion inhibitor: A review. J. Coat. Technol. Res. 2021, 19, 223–239. [Google Scholar] [CrossRef]
  131. Danyliak, M.O.; Zin, I.; Korniy, S. Corrosion inhibition of low-alloy carbon steel by gum Arabic and zinc acetate in neutral chloride-containing environment. J. Ind. Eng. Chem. 2023, 129, 267–277. [Google Scholar] [CrossRef]
  132. Asaad, M.A.; Huseien, G.F.; Baghban, M.H.; Raja, P.B.; Fediuk, R.; Faridmehr, I.; Alrshoudi, F. Gum Arabic Nanoparticles as Green Corrosion Inhibitor for Reinforced Concrete Exposed to Carbon Dioxide Environment. Materials 2021, 14, 7867. [Google Scholar] [CrossRef]
  133. Abdallah, M. Guar Gum as Corrosion Inhibitor for Carbon Steel in Sulfuric Acid Solutions. Port. Electrochim. Acta 2004, 22, 161–175. [Google Scholar] [CrossRef]
  134. Venkatesh, G.; Kamal, C.; Vennila, P.; Kaya, S.; Annaamalai, M.G.L.; Ibrahimi, B.E. Sustainable corrosion inhibitor for steel embedded in concrete by Guar Gum: Electrochemical and theoretical analyses. Appl. Surf. Sci. Adv. 2022, 12, 100328. [Google Scholar] [CrossRef]
  135. Palumbo, G.; Berent, K.; Proniewicz, E.; Banaś, J. Guar Gum as an Eco-Friendly Corrosion Inhibitor for Pure Aluminium in 1-M HCl Solution. Materials 2019, 12, 2620. [Google Scholar] [CrossRef]
  136. Biswas, A.; Pal, S.; Udayabhanu, G. Experimental and theoretical studies of xanthan gum and its graft co-polymer as corrosion inhibitor for mild steel in 15% HCl. Appl. Surf. Sci. 2015, 353, 173–183. [Google Scholar] [CrossRef]
  137. Cao, Y.; Zou, C.; Wang, C.; Chen, W.; Liang, H.; Lin, S. Green corrosion inhibitor of β-cyclodextrin modified xanthan gum for X80 steel in 1 M H2SO4 at different temperature. J. Mol. Liq. 2021, 341, 117391. [Google Scholar] [CrossRef]
  138. Pokhmurs’kyi, V.; Zin, I.; Tymus’, M.; Korniy, S.; Karpenko, O.; Khlopyk, O.; Korets’ka, N. Inhibition of the Corrosion of Carbon Steel by Xanthan Biopolymer. Mater. Sci. 2020, 55, 522–528. [Google Scholar] [CrossRef]
  139. Timothy, U.; Umoren, P.; Solomon, M.; Igwe, I.; Umoren, S. An appraisal of the utilization of natural gums as corrosion inhibitors: Prospects, challenges, and future perspectives. Int. J. Biol. Macromol. 2023, 253, 126904. [Google Scholar] [CrossRef]
  140. Nascimento, R.; Furtado, L.; Guimarães, M.J. Biopolymers as Corrosion Inhibitors: Relative Inhibition Potential of Biopolymers and Grafted Biopolymers. In Grafted Biopolymers as Corrosion Inhibitors: Safety, Sustainability, and Efficiency; John Wiley & Sons: Hoboken, NJ, USA, 2023; pp. 21–55. [Google Scholar]
  141. Sundar, S.K.; Solanki, J. Anticorrosion Coating Using Natural Biopolymer. In Functional Coatings: Innovations and Challenges; John Wiley & Sons: Hoboken, NJ, USA, 2024; pp. 153–167. [Google Scholar]
  142. Akachar, S.; AitAghzzaf, A.; Zarki, Y.; Idouhli, R.; Azaryouh, L.; El Achaby, M.; Draoui, K. Novel sustainable corrosion inhibitors for enhancing carbon steel corrosion resistance based on clay and lignin biopolymer extracted from raw Alfa fibers. J. Mater. Sci. 2024, 59, 17927–17944. [Google Scholar] [CrossRef]
  143. Toghan, A.; Alayyafi, A.; Alhussain, H.; Zaki, E.; Khodari, M.; Alqarni, N.; Masoud, E.M.; Eldesoky, A.; Eldesoky, A.; Fawzy, A. Effect of adsorption of two green biopolymers on the corrosion of aluminum in 1.0 M NaCl solution: Physicochemical, spectroscopic, surface and quantum chemical exploration. Int. J. Electrochem. Sci. 2024, 19, 100791. [Google Scholar] [CrossRef]
  144. Hao, T.; Xu, K.; Zheng, X.; Yao, X.; Li, J.; Yu, Y.; Liu, Z. Hydrogen inhibition of wet Al Li alloy dust collector systems using a composite green biopolymer inhibitor based on chitosan/sodium alginate: Experimental and theoretical studies. Int. J. Biol. Macromol. 2024, 278, 134708. [Google Scholar] [CrossRef]
  145. Yoko, S.; Yoshishige, M.; Tomoko, T.; Shinichiro, O.; Tomohiko, I.; Hidenori, Y. Comparison of Reducing Sugars as Agent to Retard Corrosion Rates. GEOMATE J. 2024, 26, 133–140. [Google Scholar]
  146. Rbaa, M.; Dohare, P.; Berisha, A.; Dagdag, O.; Lakhrissi, L.; Galai, M.; Lakhrissi, B.; Touhami, M.E.; Warad, I.; Zarrouk, A. New Epoxy sugar based glucose derivatives as eco friendly corrosion inhibitors for the carbon steel in 1.0 M HCl: Experimental and theoretical investigations. J. Alloys Compd. 2020, 833, 154949. [Google Scholar] [CrossRef]
  147. Rbaa, M.; Abousalem, A.S.; Rouifi, Z.; Lakhrissi, L.; Galai, M.; Zarrouk, A.; Lakhrissi, B.; Lakhrissi, Y. Selective synthesis of new sugars based on 8-hydroxyquinoline as corrosion inhibitors for mild steel in HCl solution-effect of the saturated hydrocarbon chain: Theoretical and experimental studies. Inorg. Chem. Commun. 2020, 118, 108019. [Google Scholar] [CrossRef]
  148. Wang, X.; Lei, Y.; Jiang, Z.N.; Zhang, Q.H.; Li, Y.; Liu, H.; Zhang, G. Developing two thiocarbohydrazide modified glucose derivatives as high-efficiency green corrosion inhibitors for carbon steel. Ind. Crops Prod. 2022, 188, 115680. [Google Scholar] [CrossRef]
  149. Kumar, D.; Jain, N.; Jain, V.; Rai, B. Amino Acids as Copper Corrosion Inhibitors: A Density Functional Theory Approach. Appl. Surf. Sci. 2020, 514, 145905. [Google Scholar] [CrossRef]
  150. Kasprzhitskii, A.; Lazorenko, G.; Nazdracheva, T.; Yavna, V. Comparative Computational Study of L-Amino Acids as Green Corrosion Inhibitors for Mild Steel. Computation 2021, 9, 1. [Google Scholar] [CrossRef]
  151. Ben Seddik, N.; Achache, M.; Zarki, Y.; Chraka, A.; Bouchta, D.; Raissouni, I. Computational, theoretical and experimental studies of four amino acids as corrosion inhibitors for brass in 3% NaCl medium. J. Mol. Liq. 2024, 397, 124113. [Google Scholar] [CrossRef]
  152. Moura, M.J.D.S.; Vasques, R.B.; Magalhães, S.J.D.M.; Almeida Neto, F.W.D.Q.; de Lima Neto, P.; dos Santos, L.P.M.; Florez, M.A.C.; Ribas, G.F.; Medeiros, S.L.S.; Salomão, F.C.C.S.; et al. Assessment of the Amino Acid L-Histidine as a Corrosion Inhibitor for a 1018 Carbon Steel in Aqueous Sodium Chloride Solution. Crystals 2024, 14, 703. [Google Scholar] [CrossRef]
  153. Toma, D.-I.; Baroi, A.M.; Din, A.; Vizitiu, D.E.; Fierascu, I.; Fierascu, R.C. Grapevine Plant Waste Utilization in Nanotechnology. AgroLife Sci. J. 2024, 13, 203–216. [Google Scholar] [CrossRef]
  154. Safieh, V.; Gholamhossein, D.; Ali, T. Impact of Kaolin Particle Film on Light Extinction Coefficient and Radiation Use Efficiency of Pistachio (Pistachia vera). AgroLife Sci. J. 2017, 6, 125–131. [Google Scholar]
  155. Petkova, N.; Denev, P. Evaluation of Fructan Contents in the Taproots of Plants Lactuca serriola L. and Sonchus oleraceus L. Sci. Pap. Ser. D Anim. Sci. 2013, 17, 117–122. [Google Scholar]
  156. Stoica, F.; Ratu, R.N.; Cara, I.G.; Topa, D.; Jitareanu, G. Exploring the potential of grape pomace powder as a functional ingredient in yogurt. Sci. Pap. Ser. D Anim. Sci. 2024, 67, 558–565. [Google Scholar]
  157. Dehghani, A.; Ghahremani, P.; Mostafatabar, A.H.; Ramezanzadeh, B. Plant extracts: Probable alternatives for traditional inhibitors for controlling alloys corrosion against acidic media—A review. Biomass Convers. Biorefinery 2022, 14, 7467–7486. [Google Scholar] [CrossRef]
  158. Sharma, S.; Solanki, A.S.; Thakur, A.; Sharma, A.; Kumar, A.; Sharma, S.K. Phytochemicals as eco-friendly corrosion inhibitors for mild steel in sulfuric acid solutions: A review. Corros. Rev. 2024, 42, 727–742. [Google Scholar] [CrossRef]
  159. Mzioud, K.; Habsaoui, A.; Rached, S.; Ech-Chihbi, E.; Ouakki, M.; Salghi, R.; Touhami, M. Experimental Investigation and Theoretical Modeling of Copper Corrosion Inhibition by Urginea maritima Essential Oil. Mater. Today Sustain. 2024, 27, 100906. [Google Scholar] [CrossRef]
  160. Zunita, M.; Rahmi, V.A. Advancement of Plant Extract/Ionic Liquid-Based Green Corrosion Inhibitor. Chem. Afr. 2023, 7, 505–538. [Google Scholar] [CrossRef]
  161. Rached, S.; Mzioud, K.; Habsaoui, A.; Galai, K.; Dahmani, K.; Ouakki, M.; Fartah, S.E.; Dkhireche , N.; Touhami, M.E. Inhibition of Copper Corrosion in Sulfuric Acid by Mentha pulegium L. Port. Electrochim. Acta 2024, 42, 137–153. [Google Scholar] [CrossRef]
  162. Rached, S.; Habsaoui, A.; Mzioud, K.; Lachhab, R.; Haida, S.; Errahmany, N.; Galai, M.; Touhami, M. Valorization of the green corrosion inhibitor Marrubium vulgare L.: Electrochemical, thermodynamic, theoretical & surface studies. Chem. Data Collect. 2023, 48, 101099. [Google Scholar] [CrossRef]
  163. Daoudi, W.; Dagdag, O.; Verma, C.; Berdimurodov, E.; Oussaid, A.; Berisha, A.; Oussaid, A.; Abboud, M.; El Aatiaoui, A. Rosmarinus officinalis L. Oil as an Eco-Friendly corrosion inhibitor for mild steel in acidic Solution: Experimental and computational studies. Inorg. Chem. Commun. 2024, 161, 112030. [Google Scholar] [CrossRef]
  164. Santana, C.; Nogueira da Cunha, J.; Rodrigues, J.G.; Duarte, J.; Freire, D.; D’Elia, E. Aqueous Extracts of the Castor Beans as a Corrosion Inhibitor of Mild Steel in HCl Media. J. Braz. Chem. Soc. 2020, 31, 1225–1238. [Google Scholar] [CrossRef]
  165. Ansari, A.; Youssefi, Y.; Tanghourte, M.; Ouassou, N.; Asoufar, N.; Znini, M.; Lgaz, H.; Mabrouk, E.H.; Azrour, M.; Lee, H.-S.; et al. Assessment of Warionia saharea Essential Oil as a Green Corrosion Inhibitor for Mild Steel in HCl: Experimental and Computational Studies. Coatings 2024, 14, 1164. [Google Scholar] [CrossRef]
  166. Mortadi, K.; El Amri, A.; Ouakki, M.; Hsissou, R.; Jebli, A.; Lebkiri, A.; Safi, Z.; Wazzan, N.; Berisha, A.; Cherkaoui, M.; et al. Electrochemical and theoretical studies on a bioactive Juniperus oxycedrus essential oil as a potential and ecofriendly corrosion inhibitor for mild steel in 1.0 M HCl environment. Inorg. Chem. Commun. 2024, 162, 112196. [Google Scholar] [CrossRef]
  167. Abdullah, H.; Anaee, R.; Khadom, A.; Ali, A.; Malik, A.; Kadhim, M. Experimental and theoretical assessments of the chamomile flower extract as a green corrosion inhibitor for aluminum in artificial seawater. Results Chem. 2023, 6, 101035. [Google Scholar] [CrossRef]
  168. Zhou, Z.; Min, X.; Wan, S.; Liu, J.; Liao, B.; Guo, X. A novel green corrosion inhibitor extracted from waste feverfew root for carbon steel in H2SO4 solution. Results Eng. 2023, 17, 100971. [Google Scholar] [CrossRef]
  169. Chowdhury, M.; Ahmed, M.M.; Hossain, N.; Islam, M.; Islam, S.; Rana, M. Tulsi and green tea extracts as efficient green corrosion inhibitor for the corrosion of aluminum alloy in acidic medium. Results Eng. 2023, 17, 100996. [Google Scholar] [CrossRef]
  170. Barboza, G.; Oliveira, M.; Neves, M.; Echevarria, A. Justicia brandegeeana as a green corrosion inhibitor for carbon steel in sulfuric acid. Green Chem. Lett. Rev. 2024, 17, 2320254. [Google Scholar] [CrossRef]
  171. Ezugha, S.; Chiedozie, A.; Eze, V. Inhibition Potentials of Solanum macrocarpon Leave Extract as a Green Corrosion Inhibitor of Mild Steel in an Acidic Solution. J. Bio- Tribo-Corros. 2024, 10, 81. [Google Scholar] [CrossRef]
  172. Malik, B.; Kaur, M.; Singh, K.; Kaur, J.; Saxena, A.; Sohal, H.S.; Husain, F.M.; Arshad, M.; Adil, M. An environmentally sustainable approach for the utilization of Erigeron bonariensis as a green inhibitor against weathering steel corrosion in acidic media. RSC Adv. 2024, 14, 16432–16444. [Google Scholar] [CrossRef] [PubMed]
  173. Kaur, J.; Saxena, A. Synergistic mixture of the Dactyloctenium aegyptium extract and KI as an environment friendly and highly efficient corrosion inhibitor for steel in 0.5 M HCl. J. Indian Chem. Soc. 2024, 101, 101317. [Google Scholar] [CrossRef]
  174. Hussein, S.A.; Khadom, A.A. Okra leaves extract as green corrosion inhibitor for steel in sulfuric acid: Gravimetric, electrochemical, and surface morphological investigations. Results Chem. 2024, 8, 101566. [Google Scholar] [CrossRef]
  175. Sajadi, G.S.; Salmanian, F.; Naghizade, R.; Hosseini, S.M.A. The inhibitive action of lemon verbena plant extract as an economical and eco-friendly corrosion inhibitor for mild steel in acidic solutions. Int. J. Electrochem. Sci. 2024, 19, 100699. [Google Scholar] [CrossRef]
  176. Mamudu, U.; Santos, J.H.; Umoren, S.A.; Alnarabiji, M.S.; Lim, R.C. Investigations of corrosion inhibition of ethanolic extract of Dillenia suffruticosa leaves as a green corrosion inhibitor of mild steel in hydrochloric acid medium. Corros. Commun. 2024, 15, 52–62. [Google Scholar] [CrossRef]
  177. Tan, B.; Liu, Y.; Gong, Z.; Zhang, X.; Chen, J.; Guo, L.; Xiong, J.; Liu, J.; Marzouki, R.; Li, W. Pyracantha fortuneana alcohol extracts as biodegradable corrosion inhibitors for copper in H2SO4 media. J. Mol. Liq. 2024, 397, 124117. [Google Scholar] [CrossRef]
  178. Fouda, A.; Molouk, A.; Atia, M.; El-Hossiany, A.; Almahdy, M. Verbena officinalis (VO) leaf extract as an anti-corrosion inhibitor for carbon steel in acidic environment. Sci. Rep. 2024, 14, 16112. [Google Scholar] [CrossRef]
  179. Nour El Houda, S.; Amel, B.; Malika, F. Trifolium repens extracts as a green corrosion inhibitor for carbon steel in a 3.5% NaCl solution. J. Taiwan Inst. Chem. Eng. 2024, 165, 105771. [Google Scholar] [CrossRef]
  180. Abeng, F.E.; Ita, B.I.; Ikpi, M.E.; Chukwuike, V.I.; Ikeuba, A.I.; Edim, M.M.; Chidiebere, M.A.; Thakur, A.; Anadebe, V.C. Millettia aboensis leaves extract as eco-friendly corrosion inhibitor for mild steel in acidizing solution: From experimental to molecular level prediction. Results Eng. 2024, 24, 102950. [Google Scholar] [CrossRef]
  181. Zhu, J.-H.; Lin, B.-L.; Duan, T.-H.; Lin, H.-Q.; Zhang, G.-Y.; Zhou, X.-X.; Xu, Y.-Y. Zea mays bracts extract as an eco-friendly corrosion inhibitor for steel in HCl pickling solution: Experimental and simulation studies. Arab. J. Chem. 2024, 17, 105895. [Google Scholar] [CrossRef]
  182. Raja, B.; Ismail, M.; Ghoreishiamiri, S.; Mirza, J.; Ismail, M.; Kakooei, S.; Rahim, A. Reviews on Corrosion Inhibitors—A Short View. Chem. Eng. Commun. 2016, 203, 1145–1156. [Google Scholar] [CrossRef]
  183. Shang, Z.; Zhu, J. Overview on plant extracts as green corrosion inhibitors in the oil and gas fields. J. Mater. Res. Technol. 2021, 15, 5078–5094. [Google Scholar] [CrossRef]
  184. Zehra, S.; Mobin, M.; Aslam, J. Chapter 13—Chromates as corrosion inhibitors. In Inorganic Anticorrosive Materials; Verma, C., Aslam, J., Hussain, C.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 251–268. [Google Scholar]
  185. Murmu, M.; Saha, S.K.; Murmu, N.C.; Banerjee, P. Chapter 14—Nitrate as corrosion inhibitor. In Inorganic Anticorrosive Materials; Verma, C., Aslam, J., Hussain, C.M., Eds.; Elsevier: Amsterdam, The Netherlands, 2022; pp. 269–296. [Google Scholar]
  186. Refaey, S.; El-Rehim, S.; Taha, F.; Saleh, M.; Ahmed, R. Inhibition of chloride localized corrosion of mild steel by PO43−, CrO42−, MoO42−, and NO2 anions. Appl. Surf. Sci. 2000, 158, 190–196. [Google Scholar] [CrossRef]
  187. Eltmimi, A.; Al-Amiery, A.; Allami, A.J.; Yusop, R.M.; Kadhum, A.; Allami, T. Inhibitive effects of a novel efficient Schiff base on mild steel in hydrochloric acid environment. Int. J. Corros. Scale Inhib. 2021, 10, 634–648. [Google Scholar] [CrossRef]
  188. Bernal, S.; Botana, F.J.; Calvino, J.J.; Marcos, M.; Pérez-Omil, J.A.; Vidal, H. Lanthanide salts as alternative corrosion inhibitors. J. Alloys Compd. 1995, 225, 638–641. [Google Scholar] [CrossRef]
  189. Arenas, M.A.; Bethencourt, M.; Botana, F.J.; de Damborenea, J.; Marcos, M. Inhibition of 5083 aluminium alloy and galvanised steel by lanthanide salts. Corros. Sci. 2001, 43, 157–170. [Google Scholar] [CrossRef]
  190. Aljibori, H.; Al-Amiery, A.; Kadhum, A. Advances in corrosion protection coatings: A comprehensive review. Int. J. Corros. Scale Inhib. 2023, 12, 1476–1520. [Google Scholar] [CrossRef]
  191. Wu, W.; Zhang, X.; Chen, Y.; Ji, J.; Zhang, F.; Guo, J.; Zhao, T.; Zhu, J.; Luo, H. Multifunctional organic-inorganic hybrid coating for enhanced bronze corrosion protection. J. Cult. Herit. 2024, 69, 113–125. [Google Scholar] [CrossRef]
  192. Wang, Z.; Li, X.; Zheng, H.; Deng, S.; Hou, Y.; Wu, Y.; Liu, M. A waterborne coating system for constructing inorganic-organic composite anti-corrosion and wear-resistant coating. Colloids Surf. A Physicochem. Eng. Asp. 2024, 694, 134120. [Google Scholar] [CrossRef]
  193. Kabeb, S.; Hassan, A.; Ahmad, F. Synergistic enhancement: Ammonium Polyphosphate’s impact on hybrid coating performance against corrosion and fire. Polym. Adv. Technol. 2024, 35, e6398. [Google Scholar] [CrossRef]
  194. Yu, X.; Zhang, M.; Chen, H. Superhydrophobic anticorrosion coating with active protection effect: Graphene oxide-loaded inorganic/organic corrosion inhibitor for magnesium alloys. Surf. Coat. Technol. 2024, 480, 130586. [Google Scholar] [CrossRef]
  195. Sun, J.; Sun, W.; Wang, L.; Xu, K.; Yang, Z.; Ma, Y.; Zhao, L.; Ma, S.; Xing, W.; Liu, G. Boosting the acid penetration barrier of epoxy-silica organic-inorganic hybrid coating via adjusting its molecular structures: Experimental and molecular dynamics simulation study. Colloids Surf. A Physicochem. Eng. Asp. 2024, 688, 133634. [Google Scholar] [CrossRef]
  196. Balakrishnan, A.; Govindaraj, S.; Gopala Krishna, N.; Thirumalaisamy, N.; Anne, R.; Sublime, N.; Philip, J. Enhancing microbiologically influenced corrosion protection of carbon steels with silanized epoxy-biocide hybrid coatings. Environ. Sci. Pollut. Res. 2024, 31, 13302–13326. [Google Scholar] [CrossRef]
  197. Zhang, Y.; Niu, X.; Zhou, J.; Wang, J.; Yang, C.; Hou, Z.; Zhu, Y.; Huang, L. Surface corrosion inhibition mechanism of sarcosine as a green novel inhibitor on a novel barrier layer material of cobalt in copper film CMP for GLSI. Mater. Sci. Semicond. Process. 2022, 140, 106402. [Google Scholar] [CrossRef]
  198. Belarbi, Z.; Singer, M.; Farelas, F.; Vu, N.; Young, D.; Nesic, S. Thiols as Volatile Corrosion Inhibitors for Top of the Line Corrosion. In Proceedings of the NACE CORROSION 2017, New Orleans, LA, USA, 26–30 March 2017. [Google Scholar]
  199. Ansari, F.A.; Verma, C.; Siddiqui, Y.S.; Ebenso, E.; Abdellaziz, B. Volatile corrosion inhibitors for ferrous and non-ferrous metals and alloys: A review. Int. J. Corros. Scale Inhib. 2018, 7, 126–150. [Google Scholar] [CrossRef]
  200. Furman, A.; Kharshan, M.; Chandler, C. Performance and Testing of Vapor Phase Corrosion Inhibitors. In Proceedings of the NACE CORROSION 2004, New Orleans, LA, USA, 28 March–1 April 2004. [Google Scholar]
  201. Shen, Y.; Zhang, D.; Zhang, Z.; Li, C.; Wu, W. High corrosion-resistant octanoic acid vapor-phase modified Ce-LDH surface film of aluminum alloy. Colloids Surf. A Physicochem. Eng. Asp. 2024, 686, 133370. [Google Scholar] [CrossRef]
  202. Zhao, J.; van Ommen, J.R.; Garcia, S.J. Gas-phase deposited nanolayers guard organic microparticles in polymer matrices for active corrosion protection at damages. Prog. Org. Coat. 2024, 192, 108522. [Google Scholar] [CrossRef]
  203. Wang, X.; Ren, J.; Li, Z.; Li, Y. Research progress of vapor phase corrosion inhibitors in marine environment. Environ. Sci. Pollut. Res. 2022, 29, 88432–88439. [Google Scholar] [CrossRef]
  204. Vorobyova, V.; Skiba, M.I.; Chygyrynets, O.J.P. A novel eco-friendly vapor phase corrosion inhibitor of mild steel. Pigment. Resin Technol. 2019, 48, 137–147. [Google Scholar] [CrossRef]
  205. Mahvidi, S.; Gharagozlou, M.; Mahdavian, M.; Naghibi, S. Potency of ZnFe2O4 Nanoparticles as Corrosion Inhibitor for Stainless Steel; the Pigment Extract Study. Mater. Res. 2017, 20, 1492–1502. [Google Scholar] [CrossRef]
  206. Shi, H.; Bi, H.; Yao, B.; Zhang, L. Dissolution of Au nanoparticles in hydrochloric acid solution as studied by optical absorption. Appl. Surf. Sci. 2000, 161, 276–278. [Google Scholar] [CrossRef]
  207. Roy, P.; Karfa, P.; Adhikari, U.; Sukul, D.D. Corrosion inhibition of mild steel in acidic medium by polyacrylamide grafted Guar gum with various grafting percentage: Effect of intramolecular synergism. Corros. Sci. 2014, 88, 246–253. [Google Scholar] [CrossRef]
  208. Liu, Y.; Chen, L.; Tang, Y.; Zhang, X.; Qiu, Z. Synthesis and characterization of nano-SiO2@octadecylbisimidazoline quaternary ammonium salt used as acidizing corrosion inhibitor. Rev. Adv. Mater. Sci. 2022, 61, 186–194. [Google Scholar] [CrossRef]
  209. Liu, T.; Ma, L.; Wang, X.; Wang, J.; Qian, H.; Zhang, D.; Li, X. Self-healing corrosion protective coatings based on micro/nanocarriers: A review. Corros. Commun. 2021, 1, 18–25. [Google Scholar] [CrossRef]
  210. Stankiewicz, A.; Szczygieł, I.; Szczygieł, B. Self-healing coatings in anti-corrosion applications. J. Mater. Sci. 2013, 48, 8041–8051. [Google Scholar] [CrossRef]
  211. Zheng, Z.; Huang, X.; Schenderlein, M.; Borisova, D.; Cao, R.; Moehwald, H.; Shchukin, D. Self-Healing and Antifouling Multifunctional Coatings Based on pH and Sulfide Ion Sensitive Nanocontainers. Adv. Funct. Mater. 2013, 23, 3307–3314. [Google Scholar] [CrossRef]
  212. Montemor, M. Functional and smart coatings for corrosion protection: A review of recent advances. Surf. Coat. Technol. 2014, 258, 17–37. [Google Scholar] [CrossRef]
  213. Wang, Q.; Li, J.; Zhang, C.; Qu, X.; Liu, J.; Yang, Z. Regenerative superhydrophobic coating from microcapsules. J. Mater. Chem. 2010, 20, 3211–3215. [Google Scholar] [CrossRef]
  214. García, S.; Fischer, H.; White, P.A.; Mardel, J.; Gonzalez-Garcia, Y.; Mol, J.M.C.; Hughes, A.E. Self-healing anticorrosive organic coating based on an encapsulated water reactive silyl ester: Synthesis and proof of concept. Prog. Org. Coat. 2011, 70, 142–149. [Google Scholar] [CrossRef]
  215. Vaszilcsin, C.; Putz, M.; Kellenberger, A.; Dan, M. On the evaluation of metal-corrosion inhibitor interactions by adsorption isotherms. J. Mol. Struct. 2023, 1286, 135643. [Google Scholar] [CrossRef]
  216. Kouakou, V.; Niamien, P.M.; Yapo, A.; Diaby, S.; Trokourey, A. Experimental and DFT Studies on the Behavior of Caffeine as Effective Corrosion Inhibitor of Copper in 1M HNO3. Orbital Electron. J. Chem. 2016, 8, 66–79. [Google Scholar] [CrossRef]
  217. He, Y.; Ren, S.; Wang, X.; Young, D.; Mohamed-Saïd, M.; Santos, B.A.; Serenario, M.; Singer, M. Temperature Dependence of Adsorption and Effectiveness for a Pyrimidinium-Type Corrosion Inhibitor on Mild Steel. Corrosion 2023, 80, 177–186. [Google Scholar] [CrossRef] [PubMed]
  218. Ech-Chihbi, E.; Es-Sounni, B.; Kerdoune, C.; Mouhib, A.; Bakhouch, M.; Salim, R.; Salghi, R.; Hammouti, B.; Mazoir, N.; Chafiq, M.; et al. Corrosion inhibition performance and adsorption mechanism of new synthesized symmetrical diarylidenacetone-based inhibitor on C38 steel in 15 % HCl medium: Theoretical insight and experimental validation. Colloids Surf. A Physicochem. Eng. Asp. 2024, 702, 135073. [Google Scholar] [CrossRef]
  219. Kokalj, A. On the use of the Langmuir and other adsorption isotherms in corrosion inhibition. Corros. Sci. 2023, 217, 111112. [Google Scholar] [CrossRef]
  220. Al-Amiery, A.; Shaker, L.; Betti, N. Corrosion inhibition effect and adsorption behaviour of nicotinic acid derivative on mild steel in HCl media. Mater. Today Proc. 2021, 56, 2204–2208. [Google Scholar] [CrossRef]
  221. Feng, C.; E, J.; Han, W.; Deng, Y.; Zhang, B.; Zhao, X.; Han, D. Key technology and application analysis of zeolite adsorption for energy storage and heat-mass transfer process: A review. Renew. Sustain. Energy Rev. 2021, 144, 110954. [Google Scholar] [CrossRef]
  222. Beda, R.H.B.; Niamien, P.M.; Avo Bilé, E.B.; Trokourey, A. Inhibition of Aluminium Corrosion in 1.0 M HCl by Caffeine: Experimental and DFT Studies. Adv. Chem. 2017, 2017, 6975248. [Google Scholar] [CrossRef]
  223. Gobara, M.; Baraka, A.; Zaghloul, B. Inhibition of mild steel corrosion in sulfuric acid solution using collagen. Res. Chem. Intermed. 2014, 41, 7245–7261. [Google Scholar] [CrossRef]
  224. Noor, E.; Al-Moubaraki, A.; Alghamdi, A. Continuous Studies on Using Camel’s Urine as Nontoxic Corrosion Inhibitor–Corrosion Inhibition of Al–Cu Alloy in Alkaline Solutions. Arab. J. Sci. Eng. 2018, 44, 237–250. [Google Scholar] [CrossRef]
  225. Vigdorowitsch, M.; Pchelintsev, A.; Tsygankova, L.; Tanygina, E. Freundlich Isotherm: An Adsorption Model Complete Framework. Appl. Sci. 2021, 11, 8078. [Google Scholar] [CrossRef]
  226. Vigdorowitsch, M.; Tsygankova, L.Y.; Shel, N.V. Polylogarithmic Adsorption Isotherm at Surface Linear Energetic Heterogeneity. Bull. Natl. Res. Nucl. Univ. MEPhI 2020, 9, 389–395. [Google Scholar] [CrossRef]
  227. Kovačević, N.; Milošev, I.; Kokalj, A. How relevant is the adsorption bonding of imidazoles and triazoles for their corrosion inhibition of copper? Corros. Sci. 2017, 124, 25–34. [Google Scholar] [CrossRef]
  228. Quraishi, M.; Chauhan, D. Heterocyclic Organic Corrosion Inhibitors: Principles and Applications; Elsevier: Amsterdam, The Netherlands, 2020. [Google Scholar]
  229. Goyal, M.; Kumar, S.; Bahadur, I.; Verma, C.; Ebenso, E.E. Organic corrosion inhibitors for industrial cleaning of ferrous and non-ferrous metals in acidic solutions: A review. J. Mol. Liq. 2018, 256, 565–573. [Google Scholar] [CrossRef]
  230. Neupane, S.; Losada-Pérez, P.; Tiringer, U.; Taheri, P.; Desta, D.; Xie, C.; Crespo, D.; Mol, J.M.C.; Milošev, I.; Kokalj, A.; et al. Study of Mercaptobenzimidazoles as Inhibitors for Copper Corrosion: Down to the Molecular Scale. J. Electrochem. Soc. 2021, 168, 051504. [Google Scholar] [CrossRef]
  231. Kokalj, A.; Lozinšek, M.; Kapun, B.; Taheri, P.; Neupane, S.; Losada-Pérez, P.; Xie, C.; Stavber, S.; Crespo, D.; Renner, F.U.; et al. Simplistic correlations between molecular electronic properties and inhibition efficiencies: Do they really exist? Corros. Sci. 2021, 179, 108856. [Google Scholar] [CrossRef]
  232. Winkler, D.A.; Breedon, M.; Hughes, A.E.; Burden, F.R.; Barnard, A.S.; Harvey, T.G.; Cole, I. Towards chromate-free corrosion inhibitors: Structure–property models for organic alternatives. Green Chem. 2014, 16, 3349–3357. [Google Scholar] [CrossRef]
  233. Kokalj, A. Molecular modeling of organic corrosion inhibitors: Calculations, pitfalls, and conceptualization of molecule–surface bonding. Corros. Sci. 2021, 193, 109650. [Google Scholar] [CrossRef]
  234. Ebenso, E.E.; Verma, C.; Olasunkanmi, L.O.; Akpan, E.D.; Verma, D.K.; Lgaz, H.; Guo, L.; Kaya, S.; Quraishi, M.A. Molecular modelling of compounds used for corrosion inhibition studies: A review. Phys. Chem. Chem. Phys. 2021, 23, 19987–20027. [Google Scholar] [CrossRef]
  235. Rodriguez, J.A.; Cruz-Borbolla, J.; Arizpe-Carreon, P.A.; Gutierrez, E. Mathematical Models Generated for the Prediction of Corrosion Inhibition Using Different Theoretical Chemistry Simulations. Materials 2020, 13, 5656. [Google Scholar] [CrossRef]
  236. Ikeuba, A.I.; Ita, B.I.; Okonkwo, C.P.; Ekuri, P.E.; Edet, H.O.; Amajama, J.; Iwuji, P.C. Review of computational methods used in the evaluation corrosion inhibition of metallic materials. Discov. Chem. Eng. 2024, 4, 28. [Google Scholar] [CrossRef]
  237. Al-Sabagh, A.M.; El Basiony, N.M.; Sadeek, S.A.; Migahed, M.A. Scale and corrosion inhibition performance of the newly synthesized anionic surfactant in desalination plants: Experimental, and theoretical investigations. Desalination 2018, 437, 45–58. [Google Scholar] [CrossRef]
  238. Abdulazeez, M.; Oyebamiji, A.; Semire, B. DFT-QSAR studies on corrosion inhibition efficiency of derivatives of thiadiazole, oxadiazole and triazole. Int. J. Corros. Scale Inhib. 2016, 5, 248–262. [Google Scholar] [CrossRef]
  239. Radovanović, M.B.; Tasić, Ž.Z.; Mihajlović, M.B.P.; Simonović, A.T.; Antonijević, M.M. Electrochemical and DFT studies of brass corrosion inhibition in 3% NaCl in the presence of environmentally friendly compounds. Sci. Rep. 2019, 9, 16081. [Google Scholar] [CrossRef] [PubMed]
  240. Chiter, F.; Bulteau, Y.; Bonin, P.; Pébère, N.; Lacaze-Dufaure, C. On the identification of favourable factors for corrosion inhibition of aluminium by 8-hydroxyquinoline and its derivatives: DFT and electrochemical studies. Corros. Sci. 2024, 233, 112104. [Google Scholar] [CrossRef]
  241. Cen, H.; Zhu, Z.; Chen, M.; Guo, X.; Chen, Z. Effect of Magnetic Field on the Inhibition Performance of Corrosion Inhibitors with Different Dipole Moment Gradients. Met. Mater. Int. 2021, 27, 5046–5058. [Google Scholar] [CrossRef]
  242. Kuruvilla, M.; Prasad, A.R.; John, S.; Joseph, A. Enhanced Inhibition of the Corrosion of Metallic Copper Exposed in Sulphuric Acid Through the Synergistic Interaction of Cysteine and Alanine: Electrochemical and Computational Studies. J. Bio- Tribo-Corros. 2016, 3, 5. [Google Scholar] [CrossRef]
  243. Challouf, H.; Souissi, N.; Ben Messaouda, M.; Rym, A.; Madani, A. Origanum majorana Extracts as Mild Steel Corrosion Green Inhibitors in Aqueous Chloride Medium. J. Environ. Prot. 2016, 07, 532–544. [Google Scholar] [CrossRef]
  244. Khaled, K.; Abdel-Rehim, S.S.; Saker, G. On the corrosion inhibition of iron in hydrochloric acid solutions, Part I: Electrochemical DC and AC studies. Arab. J. Chem. 2010, 5, 213–218. [Google Scholar] [CrossRef]
  245. Zheng, H.; Zhang, B.; Wang, X.; Lu, Y.; Li, F.; Li, C. Improved corrosion resistance of carbon steel in soft water with dendritic-polymer corrosion inhibitors. Chem. Eng. J. 2023, 452, 139043. [Google Scholar] [CrossRef]
  246. Lukovits, I.; Kálmán, E.; Zucchi, F. Corrosion Inhibitors—Correlation between Electronic Structure and Efficiency. Corrosion 2001, 57, 3–8. [Google Scholar] [CrossRef]
  247. Pérez, P.; Domingo, L.; Aizman, A.; Contreras, R. The Electrophilicity Index in Organic Chemistry. Theor. Comput. Chem. 2007, 19, 139–201. [Google Scholar] [CrossRef]
  248. Lakbaibi, Z.; Damej, M.; Molhi, A.; Benmessaoud, M.; Tighadouini, S.; Jaafar, A.; Tariq, B.; Ansari, A.; Driouich, A.; Tabyaoui, M. Evaluation of inhibitive corrosion potential of symmetrical hydrazine derivatives containing nitrophenyl moiety in 1M HCl for C38 steel: Experimental and Theoretical studies. Heliyon 2022, 8, e09087. [Google Scholar] [CrossRef]
  249. Prakashaiah, B.G.; Jayaprakash, G.K.; Rani, B.E.A. A Study of Corrosion Behavior of (E)-2-(3,4-dihydroxybenzylidene)hydrazinecarbothioamide and Bis [[3,4-dihydroxyphenylmethylene] Carbonothioicdihydrazide]-Sealed Anodized AA2024-T3. J. Bio- Tribo-Corros. 2022, 8, 75. [Google Scholar] [CrossRef]
  250. Mahdi, B.S.; Abbass, M.K.; Mohsin, M.K.; Al-azzawi, W.K.; Hanoon, M.M.; Al-kaabi, M.H.; Shaker, L.M.; Al-amiery, A.A.; Isahak, W.N.; Kadhum, A.A.H.; et al. Corrosion Inhibition of Mild Steel in Hydrochloric Acid Environment Using Terephthaldehyde Based on Schiff Base: Gravimetric, Thermodynamic, and Computational Studies. Molecules 2022, 27, 4857. [Google Scholar] [CrossRef]
  251. Razali, N.Z.K.; Wan Hassan, W.N.S.; Sheikh Mohd Ghazali, S.A.I.; Mohd Shotor, S.N.; Dzulkifli, N.N. DFT, Fukui indices, and molecular dynamic simulation studies on corrosion inhibition characteristics: A review. Chem. Pap. 2024, 78, 715–731. [Google Scholar] [CrossRef]
  252. Huerta-Servin, R.; Ahuatzi-Chacón, D.; Salmerón-Alcocer, A.; García-Medina, S.; Mena-Cervantes, V.Y.; Hernández-Altamirano, R. Effect of solubility and partition coefficient on the toxicity of corrosion inhibitors used in the oil industry. Chem. Pap. 2024, 78, 3063–3081. [Google Scholar] [CrossRef]
Figure 1. Distribution of corrosion costs in the United States in 2013 [1].
Figure 1. Distribution of corrosion costs in the United States in 2013 [1].
Technologies 13 00103 g001
Figure 2. Complex setup of factors that enhance corrosion processes.
Figure 2. Complex setup of factors that enhance corrosion processes.
Technologies 13 00103 g002
Figure 3. Classification of corrosion forms.
Figure 3. Classification of corrosion forms.
Technologies 13 00103 g003
Figure 4. Reactions that occur at metal surfaces: (a) general form, (b) adapted for iron corrosion.
Figure 4. Reactions that occur at metal surfaces: (a) general form, (b) adapted for iron corrosion.
Technologies 13 00103 g004
Figure 5. Pitting corrosion of iron-based materials.
Figure 5. Pitting corrosion of iron-based materials.
Technologies 13 00103 g005
Figure 6. General process of biological processes.
Figure 6. General process of biological processes.
Technologies 13 00103 g006
Figure 7. Schematic representation of corrosion forms.
Figure 7. Schematic representation of corrosion forms.
Technologies 13 00103 g007
Figure 8. Corrosion protection techniques.
Figure 8. Corrosion protection techniques.
Technologies 13 00103 g008
Figure 10. Mechanism of action of inhibitors on the surface affected by corrosion.
Figure 10. Mechanism of action of inhibitors on the surface affected by corrosion.
Technologies 13 00103 g010
Figure 11. Interactions between corrosion inhibitors and metal surface.
Figure 11. Interactions between corrosion inhibitors and metal surface.
Technologies 13 00103 g011
Table 1. Recent plant extract studies as inhibitors of corrosion.
Table 1. Recent plant extract studies as inhibitors of corrosion.
Plant and Part of the Plant
Used as GIC
Extraction MediumMetalActive Phyto-ComponentsCorrosive MediumMIE
%
Temp
°C
Max ConcAnalysis TechniquesRef.
Camellia chrysantha- (Matricaria recutita L.) flowerEthanolic.Al.Bisabolol, camphen, cineole.Artificial seawater
(3.5 wt.% NaCl solution).
75.6619. 8520 mL L−1OCP, Tafel, kinetic thermodynamic adsorption isotherm (Langmuir, Freundlich, Temkin), FTIR, AFM, SEM.[167]
Feverfew rootDeionized water.Q235 carbon steel.Parthenolide.0.5 mol/L H2SO4.97.225400 mg L−1Weight loss, EIS, PDP, OCP, FTIR, SEM, XPS.[168]
Tea leaves5 mL of 10% H2SO4
and
100 mL of water.
Al alloy 1100.-10% H2SO4.71.43
85.7
--Weight loss, OCP, EIS, SEM, EDS, XRD, FTIR.[169]
Tulsi plantsH2SO4
and
100 mL of water.
Al alloy 1100.-10% H2SO4.85.7--Weight loss, OCP, EIS, SEM, EDS, XRD, FTIR.[169]
Castor beansDouble-distilled water.Mid steel.Ricinoleic acid.1 M HCl.94.245200 mg L−1Weight loss, OCP, EIS, LP, PPC, adsorption isotherm (Langmuir, Temkin, Flory–Huggins, El-Awady), EDS, SDS-PAGE and FTIR.[164]
Justicia brandegeean
Aerial parts (leaves, flowers, and stems) autumn and spring
Ethanol.Carbon steel AISI 1020.Alcohols (68.42%), aldehydes and ketones (9.92%), esters (6.59%), and alkanes (1.10%) palmitic acid; 86.49% unsaturated fatty acids, the main component being linoleic acid with 37.4%. H2SO4
1 mol L−1
Solution.
93.54 301500 ppmFTIR, weight loss, NMR, ATR, OCP, EIS, LPR, PP, SEM, Adsorption isotherm (Frumkin, Langmuir, Temkin, Flory–Huggins, El-Awady).[170]
Solanum macrocarpon
leaves
Methyl alcohol extract.Mild steel.Alkaloids, flavonoids, tannins, saponins, steroids.2 M H2SO4.91.429.850.5% w/vFTIR, gasometrical techniques, hydrogen evolution data, kinetic and thermodynamic data, CR.[171]
Erigeron bonariensis leaves, flowers, and stemsEthanol.Weathering steel–mild steel.Quercetin, rutin, naringenin, luteolin, caffeic acid, apigenin.1 M H2SO4.99.5 (leaf)
94.35 (flower)
85.22 (steam)
27 ± 12000
mg L−1
Weight loss, absorption studies, UV spectroscopy, phytochemical analysis, PDP, OCP, EIS, SEM.[172]
Dactylocte-nium aegyptium
entire plant
Ethanolic extract.Steel.Tricin, Vanillic acid, p-hydroxybenzaldehyde, p-hydroxybenzoic acid. 0.5 M HCl solution.95.7 -800 ppmUV–Vis, IR, SEM, electrochemical measurements, adsorption isotherm models (Langmuir, Temkin and Freundlich).[173]
Okra
leaves
Distilled water.N80 steel.Tannin, lectin, saponin, many phenolic compounds.1 M H2SO4.9630100
mL L−1
Gravimetrical and electrochemical techniques, FTIR, UV, optical microscopy, AFM, SEM, TGA.[174]
Lemon verbena
leaves
Water.Mild steel (st37).-0.5 M H2SO4.
1 M HCl.
90 (H2SO4)
94 (HCl)
25 2000 ppm
(H2SO4)
2500 
ppm (HCl)
ECN, EIS, PP, SEM. Quantum chemical computation. MD simulation.[175]
Dillenia suffruticosa
leaves
Ethanolic extract.Mild steel.Flavonoids and glycosides, anthraquinone glycosides, phenolic derivatives and tannins, saponins, steroids, and triterpenoids.HCl.81.422.351000
mg L−1
Gravimetric and electrochemical methods, surface analysis of the corroded steel samples: SEM, FTIR.[176]
Pyracantha fortuneana fruitAbsolute ethanol.Copper.HTP, DTT, AGA, APA, ACA, DTP.H2SO4.95-600
mg L−1
Electrochemical test. Morphology analysis, FTIR, SEM, AFM.[177]
Verbena officinalis leavesEthanol.Carbon steel.Luteolin, diosmosing -7-neohesperidoside.0.5 M H2SO4.91.1251000 ppmWeight loss, PDP, EIS, AFM, XPS. [178]
Trifolium
repens
entire plant
Dichloro-methane.Carbon steel API5LX60.Flavones: acacetin, luteolin, and several others with hydroxy and methoxy groups.3.5% NaCl.9820–2520 ppmFTIR, PP, EIS, SEM, AFM.[179]
Millettia aboensis
leaves
Methanolic extract.Mild steel. Alkaloids, tannins, glycosides, phenolic compounds and flavonoids.0.3 M HCl.88.6-3 g L−1 Electrochemical
measurements, XPS, SEM, GC–MS.
[180]
Zea mays bractsEthanol aqueous solution
4:1 v/v%.
Mild steel.Flavonoids, carbohydrates phenolic compounds phenyl-propanoids.1 M HCl. 96.2455.0 g L−1 FT-IR, UV–Vis, XPS, PDP, EIS.[181]
AFM—atomic force microscopy; ATR—attenuated total reflection; CR—corrosion rate; ECN—electrochemical current noise; EDS—energy-dispersive spectroscopy; EIS—electrochemical impedance spectroscopy; EIS—electrochemical impedance spectroscopy; FTIR—Fourier transform infrared spectroscopy; GC–MS—gas chromatography–mass spectrometry; GIC—green inhibitors corrosion; IR—infrared spectroscopy; LPR—linear polarization resistance; Max. conc—maximum corrosion inhibition was achieved; MD—molecular dynamics; MIE—maximum inhibition efficacy; NMR—nuclear magnetic resonance; OCP—open circuit potential; PDP—linear polarization; PP—potentiodynamic polarization; PPC—potentiodynamic polarization curves; SDS-PAGE—electrophoresis method that allows protein separation by mass; SEM—scanning electron microscopy; TGA—thermogravimetric analysis; UV–Vis—ultraviolet–visible spectroscopy; XPS—X-ray photoelectron spectroscopy; XRD—X-ray diffraction.
Table 2. Parameters involved in corrosion/corrosion inhibition processes.
Table 2. Parameters involved in corrosion/corrosion inhibition processes.
ParameterSymbolSignificationReference
Energy of the highest occupied molecular orbitalEHOMOElectron-donating ability of a molecule.[237]
Energy of the lowest unoccupied molecular orbitalELUMOAbility of the molecule to accept electrons.[238]
Ionization potentialIPCapacity of a chemical compound to eliminate an electron.[239]
Electron affinityACapacity of a molecule to interact with nucleophile.[240]
Dipole momentµDRelated to asymmetry of charge in a molecule and a good indicator of the stability of complex on a metal surface.[241]
Energy gapΔE
ΔE = EHOMO − ELUMO
Related to a good adsorption on surface of inhibitor molecule. Lower ΔE results in higher stability of the metal–inhibitor interaction.[242]
Fraction of electron transferredΔNIf ΔN > 3.6, the inhibition efficiency increases.[243]
Global hardnessηResistance of an atom to transfer its charge.[244]
SoftnessσHow easily an inhibitor performs a charge transfer.[245]
ElectronegativityχAttraction of electrons by inhibitor molecules.[246]
Electrophilicity indexωAssessment of electrophilic properties of a molecule.[247]
Electro-donating powerΩ−Assessment of the ability of a species to donate electrons.[248]
Electro-accepting powerω+Assessment of the ability of a species to accept electrons.[249]
Dipole polarizabilityαα is a measure of polarizability. Large values of α results in a strong adsorption process.[250]
Fukui functionsfk+, fk−This is an indicator for the zones of a molecule with nucleophilic, electrophilic, or potential radical properties.[251]
Partition coefficientLog PLog P, or octanol–water partition coefficient, is a measure of how hydrophilic or hydrophobic a molecule is.[252]
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

Răuță, D.-I.; Matei, E.; Avramescu, S.-M. Recent Development of Corrosion Inhibitors: Types, Mechanisms, Electrochemical Behavior, Efficiency, and Environmental Impact. Technologies 2025, 13, 103. https://doi.org/10.3390/technologies13030103

AMA Style

Răuță D-I, Matei E, Avramescu S-M. Recent Development of Corrosion Inhibitors: Types, Mechanisms, Electrochemical Behavior, Efficiency, and Environmental Impact. Technologies. 2025; 13(3):103. https://doi.org/10.3390/technologies13030103

Chicago/Turabian Style

Răuță, Denisa-Ioana (Gheorghe), Ecaterina Matei, and Sorin-Marius Avramescu. 2025. "Recent Development of Corrosion Inhibitors: Types, Mechanisms, Electrochemical Behavior, Efficiency, and Environmental Impact" Technologies 13, no. 3: 103. https://doi.org/10.3390/technologies13030103

APA Style

Răuță, D.-I., Matei, E., & Avramescu, S.-M. (2025). Recent Development of Corrosion Inhibitors: Types, Mechanisms, Electrochemical Behavior, Efficiency, and Environmental Impact. Technologies, 13(3), 103. https://doi.org/10.3390/technologies13030103

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