Next Article in Journal / Special Issue
Study on the Galvanic Corrosion between 13Cr Alloy Tubing and Downhole Tools of 9Cr and P110: Experimental Investigation and Numerical Simulation
Previous Article in Journal
State of the Art and Perspectives on Surface-Strengthening Process and Associated Mechanisms by Shot Peening
Previous Article in Special Issue
Corrosion Behaviors of N80 and 1Cr Tubing Steels in CO2 Containing Downhole Environment—A Case Study of Underground Gas Storage in LiaoHe Oil Field
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Convolvulus microphyllus Extract as a Green, Effective, and Affordable Corrosion Inhibitor: Theoretical Calculations and Experimental Studies

1
School of Chemical Engineering, Yeungnam University, Gyeongsan 38541, Republic of Korea
2
Centre for Materials Science, College of Science, Engineering and Technology, University of South Africa, Johannesburg 1710, South Africa
3
Department of Chemistry, Faculty of Natural and Mathematics Science, University of Prishtina, 10020 Prishtina, Kosovo
*
Authors to whom correspondence should be addressed.
Coatings 2023, 13(5), 860; https://doi.org/10.3390/coatings13050860
Submission received: 13 April 2023 / Revised: 28 April 2023 / Accepted: 29 April 2023 / Published: 30 April 2023
(This article belongs to the Special Issue Investigation on Structure and Corrosion Resistance of Steels/Alloys)

Abstract

:
This study demonstrates the ability of Convolvulus microphyllus extract to prevent low-carbon steel corrosion (LCS) by varying inhibitor concentration. The effectiveness of the corrosion reaction was examined using gravimetric techniques and electrochemical procedures in a corrosive 0.5 M sulfuric acid medium. The results of polarization show a mixed adsorption nature on the LCS surface. C. microphyllus-derived film (extract) had an inhibition efficiency (IE) of 92.47% at an inhibitor concentration of 600 mg/L and a temperature 298 K. To examine the morphology, a scanning electron microscope (SEM) and atomic force microscope (AFM) were used to analyze the external films that protect LCS from sulfuric acid. A thin protective coat of inhibitor outside the LCS substrate follows the Langmuir adsorption isotherm. Additionally, computational exploration provided vital insights. The results of these experimental inhibitory outcomes are consistent with those of molecular dynamic simulations.

1. Introduction

Low carbon steel (LCS) is a basic metal that contains qualities from several different materials [1,2]. In addition to its use in acidic media and the consequent disintegration of low-carbon metals, it is now involved in numerous procedures such as cleaning, descaling, and pickling [3]. To cure or safeguard metals from environmental corrosion, inhibitors with phosphorus, oxygen, nitrogen, sulfur, and π-electronic frameworks are adsorbed on metal surfaces [4,5,6,7]. There are two methods for carrying out this adsorption: physisorption and chemisorption [8,9]. Physical adsorption happens at low temperatures and worsens as the temperature rises [10], and it has a lower activation energy. At normal temperatures, chemisorption occurs and worsens as the temperature rises [11,12]. In a natural process known as corrosion, metals start to corrode when they come into contact with moisture [13]. It is a combined development of hydrogen evolution and anodic corrosion of metals [14]. Another significant financial issue is this that metals must be shielded from corrosion. There are several techniques to prevent corrosion in metals, but the majority of them are both expensive and unfriendly to the environment [15]. Various agricultural materials are non-toxic, non-hazardous, and environmentally beneficial, such as organic goods, seeds, dried leaves, bark, and the peels of some fruits. These materials are experts in corrosion resistance and are more advantageous due to their simple accessibility and affordable price [16]. The corrosion inhibition action depends on the key components in the phytochemicals. The greater the number of heteroatoms and aromatic rings present in the major phytochemicals, the higher the corrosion inhibition efficiency. The main advantage of using C. microphyllus aerial parts extract as a corrosion inhibitor over other reported inhibitors is the presence of key phytochemicals present in this plant extract. C. microphyllus-derived film (extract) contains two key phytochemicals, kaempferol and p-hydroxycinnamic acid, which contain many heteroatoms and aromatic structures and can efficiently inhibit the corrosion process. Moreover, the geometry of these key phytochemicals plays a vital role in inhibiting action, where the presence of a linear structure is more desirable to cover the entire surface of the LCS surface, leading to maximum adsorption and thus enhancing the anti-corrosion properties.
The aerial parts of C. microphyllus-derived film (extract) (eco-waste) are generally present in our surroundings. To dispose of this waste, burning is often the common practice, which may cause air pollution. In this study, we aimed to use this plant waste as a natural corrosion inhibitor for steel in acidic media. This makes the waste a significant component in a financially sustainable environment. Another aspect of our research is to use this plant waste as a valuable product with a low manufacturing cost. In this study, we utilized a biomass extract as a natural corrosion inhibitor with high inhibitory performance at a low inhibitor concentration. Although there are numerous corrosion-resistant materials available, a large portion of them utilize synthetic engineered items that are poisonous and harmful to the climate [17,18,19,20]. Therefore, it is important to develop harmless, eco-friendly corrosion inhibitors. C. microphyllus is a remarkable plant that belongs to the Convolvuaceae family. The leaves are small, round, and nearly chipped, bearing 1–33 blossoms. Many these species have been featured in business shows known as shankhpushpi and bindweed. These plants are found across the world, from India to the coast of West Africa. They may also be found in Rajasthan, which is in the northern region of India. The two main phytochemicals found in C. microphyllus-derived film (extract) are kaempferol and p-hydroxycinnamic acid [21].
Our research is innovative in that it seeks to replace synthetic corrosion inhibitors with natural ones. It is also novel and efficient in that it uses natural waste as a corrosion inhibitor. The advantages of using natural extract as a green inhibitor include the following: (1) natural goods are safe for both humans and the environment, and (2) natural goods create a usable substance out of waste materials at a low cost. Additionally, compared to similar studies, the film (extract) made from C. microphyllus offers greater surface adsorption and corrosion inhibition effectiveness at lower inhibitor concentrations.
The corrosion resistance capabilities of C. microphyllus-derived film (extract) have been detected in sulfuric acid corrosive conditions through weight loss estimations and electrochemical evaluations. Additionally, using SEM and AFM technologies, the adsorption of the green inhibitor on the LCS surface was investigated. FTIR spectroscopy revealed various functional groups in the plant extract, and the adsorption phenomena were confirmed by UV-visible spectroscopy. Additionally, computational analyses using methmolecular dynamics simulations and density functional theory were conducted to validate the results of the experiments.

2. Experimental

2.1. Materials and Methods

The aerial parts of the C. microphyllus plant were collected from Dehradun, Uttrakhand, India. All synthetic compounds were purchased from Sigma Aldrich Incorporation (St. Louis, MO, USA). The low-carbon steel was purchased from a local hardware store. The steel sheet was cut as per the requirements of the studies. The chemical composition of the low carbon specimen used was 0.08% (C), 0.12% (P), 0.039% (Si), 0.43% (Mn), 0.45% (Cr), 0.43% (Cu), 0.27% (Ni), and 97.83% (Fe) (weight %) [22]. Low carbon steel specimens measuring 1.00 cm × 1.00 cm × 0.03 cm were used for the weight loss studies [23,24,25]. For electrochemical studies, low carbon steel specimens measuring 1.00 cm x 1.00 cm x 8.00 cm were used, with only 1.00 cm2 of the surface exposed to the electrolyte (the remainder was covered with epoxy) [26]. The low-carbon steel specimens were sequentially abraded with 220, 400, 600, 800, 1000, and 1200 grades of sandpaper, then dried, cleaned with acetone, and washed with distilled water before use. In order to create the electrolyte (0.5 M H2SO4), 99.999% sulfuric acid (mw 98.08) was employed. The pH of the corrosive solution was approximately 2.

2.2. Preparation of Plant Extract

The extraction process involved collecting C. microphyllus and verifying it with a botanist, Professor Bhatt from LPU, India. The raw plant materials were washed under tap water, cleaned, and then dried under natural conditions. After the drying process, it was crushed and converted into a powder. The extraction process was carried out using the Soxhlet apparatus by adding 400 g of C. microphyllus powder to 1000 mL of MeOH. The obtained materials were separated and concentrated [27], resulting in an extraction yield of approximately 20%. The extract was then stored in a vacuum desiccation under a controlled temperature for further corrosion inhibition studies. Figure 1 presents a schematic illustration of the plant extract preparation process. The pH of the pure plant extract was approximately 9.

3. Methods

3.1. Extract Characterization

3.1.1. FT-IR Analysis

In order to detect distinct active and aromatic groups, FT-IR data was employed. For the creation of a standard pallet, the plant extract and KBr were combined. Shimadzu’s FTIR 8400S frequency spectrometer (Kyoto, Japan) was used to analyze this material over a range of 400 to 4000 cm−1 [28].

3.1.2. UV-Visible Analysis

The test solution was examined using a Shimadzu UV-1800 UV-visible spectrophotometer before and after the corrosion test. The resultant spectrum was used to verify the procedure [28].

3.2. Weight Loss Study

The weight loss test was conducted in accordance with the American Society for Testing and Materials (ASTM) G 31-72. All the weight loss measurements took place between 298 K and 318 K over 24 h [29,30,31]. A water thermostat (PPI Fini X 48) was used to maintain the temperature throughout the process. For each experiment, 500 mL of newly made corrosive solution was utilized [32]. The metal sample was weighed for corrosion testing. The metallic samples were taken out of the test solution after the corrosion test (which lasted 24 h) and cleaned with distilled water as well as acetone, then dried with nitrogen flow. They were then weighed again to determine any weight loss [33]. All the weight measurements were made using Shimadzu electronic scales (BL-220H/D455006313).

3.3. Electrochemical Study

Potentiodynamic Polarization (PDP) measurements were performed after conducting electrochemical tests using electrochemical impedance spectroscopy (EIS) [34,35]. All electrochemical experiments were conducted using a CHI760E electrochemical workstation at 25 °C in 0.5 M sulfuric acid with varying inhibitor concentrations. A three-electrode model was used, which included a saturated calomel electrode (SCE) as the reference electrode, a platinum plate as the counter electrode, and a square of low carbon steel (1 cm2) as the working anode [36,37,38,39]. A freshly prepared electrolyte solution of 250 mL was used for all electrochemical tests [40]. The EIS curve was analyzed in the frequency range of 100,000 Hz to 0.01 Hz with an amplitude perturbation of 5.00 mV. When compared to the saturated calomel electrode, the Tafel curves were evaluated at a potential of ±250 mV and a scan rate of 1.00 mV s−1 [41]. In the past, to acquire the Open Circuit Potential (OCP) constant, the working electrodes were submerged in the solution for 3600 s [42,43].

3.4. Surface Study

In addition to SEM and AFM analyses, this study employed a low-carbon steel model for an experiment on weight loss [44,45,46]. The metal samples were subjected to surface testing in an acidic environment with an inhibiting concentration of 600 mg/L. Studies on surface morphology provided evidence of the adsorption of inhibitor compounds on metal surfaces. The SEM and AFM models demonstrated the advantage of enhancing the surface adsorption of inhibitor molecules while producing a protective coating on the metal surface that shields it from an acidic environment [47]. The SEM experimental parameters, such as acceleration voltage, were set at 15 kV with a pressure of 20–10 Pa. The NT-MDT-INTEGRA set was designed for AFM, and the LEO-435 VP set was used for SEM microphotographs.

3.5. Computational Studies

3.5.1. DFT

Computational research has enabled a deeper comprehension of adsorption methods and confirmed the occurrence of this experiment. Here, quantum computation was employed in combination with theoretical tests. Plants are known to have multiple phytochemical components, and in this analysis, we selected the most important component in the plant C. microphyllus [21]. The selection of molecules or phytochemicals was achieved using quick handling Density Functional Theory (DFT) with a HyperChem software package using a DFT/B3LyP functional and 6-31G (d,p) basis set [48,49,50]. Crucial values were obtained from the optimized structures of examined inhibitor molecules.

3.5.2. MD Simulations

In order to ascertain the strength and nature of the molecular interactions on the surface of metallic Fe(110), a molecular dynamics simulation (MDS) was carried out [51,52]. The modeling was carried out using Material Studio, with a simulated grid size of 32.270 × 32.270 × 34.134Å3. Periodic boundary conditions with composites (inhibitors, 20H3O+ + 10SO42− + 500H2O) and Fe(110) were also used. The design was evaluated at 298 K with an NVT simulation duration of 1000 ps and a time step of 1.0 fs under the direction of an Andersen thermostat [53].

3.6. Statistical Analysis

All experiments were performed in triplicate, and the results were presented according to the mean of different measurements ± standard deviation. The statistical analysis was performed using the ANOVA method according to the Dunnett experiment with SPSS version 23 (SPSS Inc., Chicago, IL, USA), or the Kruskal-Wallis and Mann-Whitney test.

4. Results and Discussions

4.1. Characterization of C. microphyllus Extract

4.1.1. FT-IR Study

The FT-IR strategy was utilized to identify the presence of different functional groups and aromatic gatherings in the plant material. The FTIR spectra of the pure C. microphyllus plant extract are shown in Figure 2.
The specific peaks found in C. microphyllus plant extract were related to the various functional groups that are listed in Table 1. In the pure C. microphyllus plant extract, the obtained spectra confirmed the presence of CH2 (aromatic), C–O, –OH, and C = O groups, suggesting the availability of aromatic and heteroatoms in the sample [54]. These heteroatoms can easily transfer their electrons, which do not bind the vacant d-orbit to the LCS metal to form a compound. Aromatic molecules or parts can also use their electron segregation to aid in the surface adsorption of inhibitory molecules on the metal surface. As the plant extract is a complex organic mixture, it may contain many C–C and C=C bonds [55]. The small peaks near 1620 cm−1 to 1680 cm−1 were due to the presence of C=C bonds, while a few small peaks around 1200 cm−1 to 1400 cm−1 were due to the presence of C–C bonds [56].

4.1.2. UV-Visible Study

Further evaluation of the adsorption accuracy was done using UV-visible spectroscopy. The test was conducted both before and after the weight loss measurement. Figure 3 shows the UV-vis spectra of the inhibitory solution for both species.
The corrosion inhibitor solution exhibited two peaks at 218.00 nm as well as 287.00 nm prior to the weight loss examination, indicating n-π* and π-π* transitions, respectively [57]. The signal at 218.00 nm indicated a π-π* transition associated with the C=C conjugation of aromatic rings. A visible peak suggesting an n-π* transition of several functional groups, including heteroatoms such as oxygen, was detected in the frequency range of 250.00 to 350.00 nm. These results indicate that the adsorption peaks underwent a hypsochromic shift following the weight loss assessment [58,59]. The absorption maxima ( λ m a x ) before immersion (218.00 nm 287.00 nm) shifted to a shorter wavelength (202.00 nm and 270.00 nm). The peak related to before immersion showed an absorbance maximum intensity of approximately 2.9 a.u. and 0.75 a.u., which decreased to approximately 2.65 a.u. and 0.65 a.u. after immersion. Additionally, the strength of the adsorption peaks was dramatically decreased. These modifications demonstrate that complex formation among inhibitor molecules and iron particles occurs at a certain point. When comparing the two spectra, it is generally agreed that the inhibitor molecules were able to detect an essential shift in the inhibitory adsorption particles’ outer adsorption band.

4.2. Weight-Loss Study

A simple and manual testing procedure, the weight-loss approach provides insight into the link between inhibitor concentration and inhibition efficiency. This test was performed using the ASTM technique at 298 ± 0.5 K to 318 ± 0.5 K for 24 h with various inhibitor doses [14,53,60]. The following equations were used to obtain important parameters such as corrosion rate, inhibitory effectiveness, and surface coverage value [45]:
ω = ω o ω i
C R = 87,600 × ω A t D
where:
CR = corrosion rate (mmy−1);
ωo = initial weight of the LCS (g);
ωi = final weight of the LCS (g);
∆ω = weight loss of the LCS (g);
D = 7.86 g cm−3 (density of iron);
t = immersion time (h);
A = area of the LCS (cm2).
The following formulas were used to calculate the surface coverage values and the inhibitor’s effectiveness [47]:
I E % = C o r r o s i o n R a t e C o r r o s i o n R a t e i n h C o r r o s i o n R a t e × 100
θ = C o r r o s i o n R a t e C o r r o s i o n R a t e i n h C o r r o s i o n R a t e
where:
θ = surface coverage values;
CorrosionRateinh = corrosion rate of the LCS with inhibitor;
CorrosionRate = corrosion rate of the LCS without inhibitor.
Table 2 displays the results of the weight loss tests, which indicate that as the inhibitor concentration increases, inhibition efficiency also increases, and the corrosion rate decreases [61]. The corrosion rate reaches up to 11.33 mmy−1 without the inhibitor, but at a concentration of 600 mg/L, the inhibitor exhibits the highest corrosion inhibition efficiency of 86.95% and reduces the corrosion rate to 1.47 mmy−1. After 600 mg/L, no significant change was observed in inhibition efficacy, suggesting that 600 mg/L is the optimum or highest inhibitor concentration for achieving the best anti-corrosion performance. Based on these results, the Langmuir adsorption isotherm was also plotted [62]. In addition, we can observe that as the temperature increased, the corrosion inhibition performance of C. microphyllus-derived film (extract) decreased, which is attributed to the degradation of phytochemicals because phytochemicals are organic molecules that are known to easily degrade at higher temperatures.

4.3. Electrochemical Studies

4.3.1. OCP

Electrochemical experiments were conducted to assess the inhibitory power of the investigated plant extract and to determine the corrosion resistance of metals. The initial electrochemical test carried out was open circuit potential (OCP) measurements. Low-carbon steel samples were submerged in 0.5 M H2SO4 in both the absence and presence (600 mg/L) of the investigated inhibitor, and the corrosion potential was determined. Figure 4a shows a stable OCP both with and without an inhibitory concentration of 600 mg/L. When the potential rapidly drops after the first immersion, it usually indicates that the native oxide layer is dissolving. The creation of a protective coating on the surface is typically linked to the subsequent movement of the potential towards more positive values. In the EIS and PDP tests, low-carbon steel samples were immersed in the corrosive environment for one hour to ensure a stable open circuit potential. To describe the EIS diagram, we used a simple electronic equation model [63]. Measurement of the OCP took 200 s [14]. The low-carbon steel OCP was first reduced without inhibitors and then stabilized at 465 mVvsSCE. The stabilizer at 600 mg/L was stable at 463 mVvs SCE, which was due to the formation of an insoluble metal oxide layer on the metal surface. This indicates that there was a barrier molecule that changed the state of interaction between the carbon materials and electrolytes.

4.3.2. PDP

Polarization measurements in 0.5 M sulfuric acid made it possible to determine the corrosion resistance of low-carbon steel during adsorption. The polarization curve of low-carbon steel in the aggressive 0.5 M sulfuric acid environment at 298 K was observed with and without varying concentrations of C. microphyllus-derived film (extract), as shown in Figure 4b.
PDP measurements provide key parameters such as corrosion potential (Ecorr), anodic tafel slope (βa), cathodic tafel slope (βc), and corrosion current density (icorr). The corrosion resistance performance or corrosion inhibition efficiency was calculated using the following equation [64]:
I E % = i c o r r i c o r r i n h i c o r r × 100
where:
icorr = the corrosion current density without inhibitor;
icorrinh = the corrosion current density with inhibitor.
The results of PDP for LCS both with and without C. microphyllus-derived film (extract) are shown in Table 3. The inhibitory-active chemicals in the C. microphyllus extract were adsorbed on the surface of the LCS and prevented corrosion of the metal in an acidic environment. The anodic as well as cathodic Tafel slopes became less steep as a result of this process [65]. The anodic tafel slope (βa) and cathodic tafel slope (βc) were altered with every increment of C. microphyllus-derived film (extract) concentration, demonstrating that the inhibitor molecules hinder the metal-dissolving process and the hydrogen evolution reaction. The observation that the cathodic and anodic Tafel slopes remained essentially unchanged before and after the addition of the C. microphyllus extract proves that charge transfer regulates the cathodic and anodic reaction processes. As a result, without altering the reaction mechanism, the main components of the C. microphyllus extract were adsorbed into the steel surface, preventing LCS corrosion. Several natural substances found in the C. microphyllus extract may electrostatically adsorb to the surface of steel and form strong coordination interactions with iron through their heteroatoms. The functional particles of the C. microphyllus extract are strongly bound to the steel surface due to the combination of physisorption and chemisorption. This protects the steel from corrosion, as evidenced by the decline in corrosion current density values. These results show that the corrosion current density decreases as the concentration of the C. microphyllus extract increases.
Additionally, it has been observed that as the concentration of C. microphyllus-derived film (extract) increases, the values of icorr (corrosion current density) gradually decrease.
This might be due to the strong coordination link that exists between the free electrons of heteroatoms and the unoccupied d-orbitals of iron in the C. microphillus extract [66]. The lowest icorr value (67.06 µA cm−2) was observed with the highest concentration of C. microphyllus-derived film (extract) at 600 mg/L concentration, resulting in an IE of 92.47%. Beyond 600 mg/L inhibitor concentration, there was no significant change observed in inhibition efficacy, indicating that 600 mg/L is the optimum or highest inhibitor concentration where it shows the best corrosion inhibition performance.

4.3.3. EIS

The EIS spectra were drawn using a Randle electric equivalent circuit. The Nyquist plots are shown in Figure 5a, and the Bode plots are shown in Figure 5b. The results are shown in Table 4. The following equations were used to determine the corrosion inhibition effectiveness and other important factors [67]:
I E % = R c t i n h R c t R c t i n h × 100
The following equation was used to estimate the double-layer capacitance (Cdl) [68]:
C d l = Υ 0 ( 2 π f m a x ) n 1
where:
Rctinh = the charge transfer resistance with inhibitor;
Rct = the charge transfer resistance without inhibitor;
Υ0 = the magnitude of the CPE;
fmax = the frequency at maximum imaginary impedance;
n = 2α/π;
α = the phase angle.
The corrosion of LCS is controlled by the charge transfer mechanism of corrosion, as represented by the capacitive loops. Figure 5a illustrates how the diameter of the semicircles grows as C. microphyllus extract is added to the corrosive media. The greater the corrosion inhibition, the larger the loop’s diameter. In comparison to other concentrations, the C. microphyllus-derived film (extract) concentration at 600 mg/L exhibited the widest semicircle, indicating its superior corrosion resistance ability [69].
With increasing inhibitor concentration, Cdl decreased and Rct increased, as shown in Table 4. The inductive loops associated with the EIS curves for the initial solution were at low frequencies due to the relaxation of the adsorbed product intermediates from the steel surface. Typically, the demise of the inductive loop is considered a “degradation” phenomenon for EIS for the remaining concentrations. The Nyquist plot’s semicircle diameter variation trend was consistent with the impedance value variation trend in the Bode plot using C. microphyllus extract concentration in the Bode modulus curves. On all phase angle-frequency curves, a single wave could be observed, validating the Nyquist plot’s one-time constant. In accordance with the literature, if the phase angle is equal to 90° or 0°, the electrochemical reaction at the steel solution interface is capacitive or resistive, respectively.
These findings demonstrate the remarkable surface adsorption of the film (extract) generated from C. microphyllus. The results of the EIS approach were consistent with the PDP as well as measures of weight reduction. C. microphyllus-derived film (extract) exhibited 89.87% inhibition at 600 mg/L and Rct (155.13 Ω cm2). Beyond 600 mg/L inhibitor concentration, there was no significant change observed in inhibition efficacy, indicating that 600 mg/L is the optimum or highest inhibitor concentration for achieving the best corrosion inhibition performance.

4.4. Adsorption Isotherm

The Langmuir adsorption isotherm graph was created using data from trials on weight loss [70]. Figure 6 shows a graph of the Langmuir adsorption isotherm, which represents the relationship between the inhibitor concentration/surface coverage value and the inhibitor concentration (C/θ and C). Adsorption isotherms provide a clearer illustration of the interaction between the inhibitor molecules and the LCS surface. The ability of an inhibitor to inhibit corrosion depends on how efficiently its particle adsorbs to a metal surface. Reintroducing water particles to the metal solution interface is a step in the adsorption process. It was previously understood that the adsorption of natural inhibitor particulates between the metal and the solution is the first stage in reducing metal corrosion. Surface coverage is determined by the concentration of the inhibitor and the inhibition efficacy, expressed as (θ = IE (%)/100). With an increase in inhibitor concentration, more molecules are adsorbed onto the surface.
It was previously stated that the inhibitor would follow Langmuir adsorption and form a monolayer on the metal surface if the regression coefficient and slope values were close to one. Our results indicate that the inhibitor follows the Langmuir isotherm and forms a protective monolayer over the metal surface to protect it from corrosive environments. The obtained values were extremely close to one. This occurrence can be described as [71]:
C θ = 1 K a d s + C
where:
θ = the surface coverage;
C = inhibition concentration;
Kads = the equilibrium adsorption constant.

4.5. Surface Studies

SEM and AFM Analysis

Using SEM and AFM methods, surface morphological investigations were conducted to obtain surface micrographs of LCS after the weight loss study with and without the inhibitor in the corrosive medium [72]. Figure 7 displays photomicrographs (SEM and AFM) of polished LCS surfaces following weight reduction with and without an inhibitor (600 mg/L). SEM and AFM photomicrographs of the used LCS surface before the test are shown in Figure 7a, which are referred to as polished LCS. The surface roughness was 2.99 nm, and the maximum height was 60 nm. SEM and AFM photomicrographs of the LCS surface during a weight loss test without an inhibitor are shown in Figure 7b [68]. The uneven SEM micrographs were apparent because of the high corrosion rate. The LCS surface was corroded, resulting in a severely corroded surface with a very high surface roughness value (138.81 nm) and a maximum height of 2100 nm. Figure 7c shows the LCS surface following the testing for weight loss with a 600 mg/L inhibitor concentration. This demonstrates that the the inhibitor forms a shield on the LCS surface, shielding it from the corrosive medium and giving it a surface that is relatively smooth with a reduced surface roughness value (15.25 nm) and a maximum height of 160 nm [73]. The surface roughness values and maximum height values decreased from uninhibited to inhibited LCS. These lower values of surface roughness and maximum height for the inhibited LCS compared to the corroded LCS suggest surface adsorption and the creation of a protective thin film on the metal surface to protect it from the corrosive media. This shows that C. microphyllus-derived film (extract) absorbs on the metal surface, creating a protective layer to prevent corrosion on LCS surfaces.

4.6. Computational Studies

4.6.1. DFT

Density functional theory has been used in computational investigations [74]. As referenced before, C. microphyllus-derived film (extract) is rich in phytochemicals, making it very hard to predict the corrosion inhibition activity of any specific molecule present in the C. microphyllus extract. Therefore, we chose two key/major phytochemical constituents for the DFT studies: kaempferol (M1) and p-hydroxycinnamic acid (M2), which are present in C. microphyllus-derived film (extract) [21]. Figure 8 displays the HOMO, LUMO, and MEP (Molecular Electrostatic Potential) directions of kaempferol (M1) and p-hydroxycinnamic acid (M2). The key parameters were calculated using the following equations [60]:
E = E L U M O E H O M O
χ = 1 2 ( E L U M O + E H O M O )
η = 1 2 ( E L U M O E H O M O )
σ = 1 η
E B a c k D o n a t i o n = η 4
Δ N = χ F e χ i n h 2 ( η F e η i n h )
π = χ
where:
χinh = the electronegativity of the inhibitor molecule;
ηinh = hardness of inhibitor molecule;
χFe = the electronegativity of the iron;
ηFe = hardness of the iron;
EHOMO = energy of highest occupied molecular orbital;
ELUMO = energy of lowest un-occupied molecular orbital;
ΔE = energy gap;
σ = global softness;
ΔN = energy from inhibitor to metals;
π = chemical potential.
The vital components of kaempferol (M1) and p-hydroxycinnamic acid (M2) are shown in Table 5, which are significant due to their influence on the electron bonding of the inhibitor molecule with the metal ions. A low ELUMO suggests that the molecule under evaluation has a strong potential for absorbing electrons [75]. A high EHOMO score implies that the molecule under examination has a high potential for free electron donation. A stable inhibitor-metal combination is indicated by a low ΔE value. The capacity of the tested inhibitor molecule to donate electrons is indicated by a low value of χ [76,77].

4.6.2. MDS

The aim of this study was to theoretically or computationally investigate the influence of the test substances’ adsorption on low carbon steel. Molecular dynamics simulations (MDS) were performed on the Fe(110) surface in a medium containing H2SO4 [78]. The optimal adsorption system for the inhibitor molecules (M1 and also M2) at the attack site is shown in Figure 9a. The investigated molecules (M1 and M2) were found to effectively adsorb at the molecular limit to reduce the density of the metal and the surface area of the corrosion cycle. Furthermore, surface and horizontal images showed how the C. microphyllus extract adhereed to the surface of Fe(110), improving the performance of the film (extract) made from C. microphyllus on the primary iron layer.
The energies of adsorption and binding were used to compute the functions of chemical molecule adsorption and dissociation on the surface of the LCS. The binding energy (Ebinding) and interaction energy (Einteraction) were calculated using the following equation [79]:
E i n t e r a c t i o n = E t o t a l ( E s u r f a c e + s o l u t i o n + E i n h i b i t o r )
E b i n d i n g = E i n t e r a c t i o n
All the results are shown in Table 6. An inhibitor molecule’s adsorption on the metal surface is indicated by a high binding value and a low interaction value, implying that C. microphyllus-derived film (extract) molecules are highly reactive with the LCS surface area [80].
The findings demonstrate that p-hydroxycinnamic acid (M2) has a higher adsorption efficiency than kaempferol (M1) for subsequent uses. The outcomes of this experiment were validated using this system. The Monte Carlo simulation of the (Fe(110)-inhibitor) adsorption is depicted in Figure 9b. These findings imply that substances adhere to metal surfaces via coordination bonds [81].

4.7. Inhibition Mechanism

To ensure the impact of C. microphyllus-derived film (extract) on the surface of low carbon steel in a corrosive environment containing 0.5 M sulfuric acid, the outcomes of this investigation need to be analyzed in terms of the electronic, chemical, and structural properties of inhibitory molecules. C. microphyllus-derived film (extract) contains different phytochemicals, which have various aromatic rings and heteroatoms [60]. As these molecules behave like a Lewis base, they establish coordination bonds with the free d-orbital of iron by donating their non-bonding electrons. They are then adsorbed onto the metal surface. This procedure adds a protective coating to the metal substrate’s surface to protect it from damage. Therefore, it is possible to anticipate that chemisorption and physisorption will affect the relationship between inhibitory particles and iron. The aromatic rings’ π-electron content may be used for retro-donation [70]. Theoretical studies and polarization findings both support this conclusion. The structure of the kaempferol molecule for adsorption on a metal surface is shown in Figure 10. Figure 11 compares the relative corrosion inhibition efficiencies attained using weight loss, PDP, and EIS for various factors. Table 7 compares the effectiveness of previously published work with the current study in terms of inhibition. Li et al. [82] studied the suppression of copper in alkaline electrolytes using M@0.33P. The M@0.33P nanohybrids protection mechanism is explained by the corrosion inhibition mechanism, which points to a superior barrier effect and the interfacial passivation-maintenance process. Dong et al. [83] hypothesized that the corrosion attack sequence on the TiO2(110) surface was S2− > CO32− > Cl > HS > HCO3− when corrosive ions (Cl, HS, S2−, HCO3−, and CO32−) adsorbed on the surface. The environment containing H2S will cause substantial corrosion in TC4 titanium alloy, followed by the environment containing CO2, which has comparatively strong corrosion resistance to Cl. The corrosion of the TC4 titanium alloy will be more severe when the corrosion products H2S, CO2, and Cl coexist in the corrosion medium. Liu et al. [84] investigated supermolecular assemblies based on benzothiazole derivatives as effective copper corrosion inhibitors in artificial seawater. Following a cathodic-dominated mixed-type inhibition mechanism, the authors hypothesized that MBT and ABT were stably integrated into β-CD with differing thermodynamic characteristics, yielding optimum inhibition efficiencies of 95.6% and 89.2% at a 0.5 g/L dose. The corrosion inhibition of metal involves two reactions (anodic and cathodic reactions). The anodic reaction mechanism of iron in sulfuric acid solution takes place in two continuous steps:
Fe ⇌ Fe+ + e
Fe+ ⇌ Fe2+ + e
FeSO42− → (FeSO42−)abs
(FeSO42−)abs → (FeSO4)abs + 2e
(FeSO4)abs → Fe2+ + SO42−
The cathodic reaction is as follows:
2H+ + 2e → H2
O2 + 2e + 4H+ → 2H2O
O2 + 2e + 2H+ → H2O2
H2O2 + 2e +2H+ → 2H2O

5. Conclusions

The anti-corrosive properties of C. microphyllus-derived film (extract) in 0.5 M sulfuric acid were found to be effective for low carbon steel. Based on the results obtained, the following observations can be made:
Electrochemical analysis revealed that the inhibitory effect of C. microphyllus-derived film (extract) at a concentration of 600 mg/L provides approximately 92% protection against corrosion for low carbon steel at 298 K.
Potentiodynamic polarization estimates indicated that this natural product is a mixed-type inhibitor and also functions as an adsorptive inhibitor. Therefore, the efficiency of the inhibition may be improved by increasing the inhibitor concentration.
When the concentration of the inhibitor was increased, electrochemical impedance spectroscopy showed a decrease in the electrical double layer’s CPE and an increase in the charge transfer resistance.
These findings further demonstrate that the C. microphyllus-derived film (extract) extract only affects the metal/solution interface by adsorption.
The FT-IR technique demonstrated the presence of heteroatoms and unsaturated mixtures that were used as inhibitors.
The coordination connections between the inhibitor molecules and Fe2+ were confirmed by the UV-visible spectroscopic technique.
The SEM the AFM approaches explored the adsorption of the C. microphyllus-derived film (extract) inhibitor on a metal substrate.
The experimental results indicate that these inhibitor compounds are strongly physisorbed and chemisorbed onto the LCS surface.
Studies in DFT, MC, and MDS demonstrated that inhibitors frequently cured and safeguarded metals due to their distinct ability to attract and withdraw electrons from metals.

Author Contributions

Conceptualization, R.H. and R.V.; methodology, O.D.; software, A.B.; validation, R.H. and O.D.; formal analysis, R.H.; investigation, R.H.; resources, S.-C.K.; data curation, R.V.; writing—original draft preparation, R.H.; writing—review and editing, O.D.; visualization, R.H.; supervision, O.D. and S.-C.K.; project administration, S.-C.K.; funding acquisition, S.-C.K. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by the NRF (National Research Foundation) of Korea funded by the Basic Science Research Program through the Ministry of Education (2020R1I1A3052258). In addition, this work was also supported by the Technology Development Program (S3060516) funded by the Ministry of SMEs and Startups (MSS, Korea) in 2021.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data supporting the findings of this study are available within the article.

Conflicts of Interest

The authors declare no conflict of interest.

Abbreviations

AFMAtomic force microscopyM1Kaempferol
ASTMAmerican society for testing and materialsM2p-hydroxycinnamic acid
DFTDensity functional theoryMSDMolecular dynamic simulation
EISElectrochemical impedance spectroscopyOCPOpen circuit potential
FTIRFourier transform infrared spectroscopyPDPPotentiodynamic polarization
IEInhibition efficiencySEMScanning electron microscopy
LCSLow carbon steelSCESaturated calomel electrode

References

  1. Verma, D.K.; Kazi, M.; Alqahtani, M.S.; Syed, R.; Berdimurodov, E.; Kaya, S.; Salim, R.; Asatkar, A.; Haldhar, R. N–hydroxybenzothioamide derivatives as green and efficient corrosion inhibitors for mild steel: Experimental, DFT and MC simulation approach. J. Mol. Struct. 2021, 1241, 130648. [Google Scholar] [CrossRef]
  2. Dehghani, A.; Bahlakeh, G.; Ramezanzadeh, B.; Ramezanzadeh, M. A combined experimental and theoretical study of green corrosion inhibition of mild steel in HCl solution by aqueous Citrullus lanatus fruit (CLF) extract. J. Mol. Liq. 2019, 279, 603–624. [Google Scholar] [CrossRef]
  3. Majd, M.T.; Asaldoust, S.; Bahlakeh, G.; Ramezanzadeh, B.; Ramezanzadeh, M. Green method of carbon steel effective corrosion mitigation in 1 M HCl medium protected by Primula vulgaris flower aqueous extract via experimental, atomic-level MC/MD simulation and electronic-level DFT theoretical elucidation. J. Mol. Liq. 2019, 284, 658–674. [Google Scholar] [CrossRef]
  4. Sanaei, Z.; Bahlakeh, G.; Ramezanzadeh, B.; Ramezanzadeh, M. Application of green molecules from Chicory aqueous extract for steel corrosion mitigation against chloride ions attack; the experimental examinations and electronic/atomic level computational studies. J. Mol. Liq. 2019, 290, 111176. [Google Scholar] [CrossRef]
  5. Tabatabaei majd, M.; Bahlakeh, G.; Dehghani, A.; Ramezanzadeh, B.; Ramezanzadeh, M. A green complex film based on the extract of Persian Echium amoenum and zinc nitrate for mild steel protection in saline solution; Electrochemical and surface explorations besides dynamic simulation. J. Mol. Liq. 2019, 291, 111281. [Google Scholar] [CrossRef]
  6. Dehghani, A.; Bahlakeh, G.; Ramezanzadeh, B. Green Eucalyptus leaf extract: A potent source of bio-active corrosion inhibitors for mild steel. Bioelectrochemistry 2019, 130, 107339. [Google Scholar] [CrossRef]
  7. Dehghani, A.; Bahlakeh, G.; Ramezanzadeh, B.; Ramezanzadeh, M. Aloysia citrodora leaves extract corrosion retardation effect on mild-steel in acidic solution: Molecular/atomic scales and electrochemical explorations. J. Mol. Liq. 2020, 310, 113221. [Google Scholar] [CrossRef]
  8. Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Green synthesis of reduced graphene oxide nanosheets decorated with zinc-centered metal-organic film for epoxy-ester composite coating reinforcement: DFT-D modeling and experimental explorations. J. Taiwan Inst. Chem. Eng. 2020, 114, 311–330. [Google Scholar] [CrossRef]
  9. Majd, M.T.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Steel corrosion lowering in front of the saline solution by a nitrogen-rich source of green inhibitors: Detailed surface, electrochemical and computational studies. Constr. Build. Mater. 2020, 254, 119266. [Google Scholar] [CrossRef]
  10. Asfia, M.P.; Rezaei, M.; Bahlakeh, G. Corrosion prevention of AISI 304 stainless steel in hydrochloric acid medium using garlic extract as a green corrosion inhibitor: Electrochemical and theoretical studies. J. Mol. Liq. 2020, 315, 113679. [Google Scholar] [CrossRef]
  11. Keramatinia, M.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Synthesis of a multi-functional zinc-centered nitrogen-rich graphene-like thin film from natural sources on the steel surface for achieving superior anti-corrosion properties. Corros. Sci. 2021, 178, 109077. [Google Scholar] [CrossRef]
  12. Mofidabadi, A.H.J.; Bahlakeh, G.; Ramezanzadeh, B. Anti-corrosion performance of the mild steel substrate treated by a novel nanostructure europium oxide-based conversion coating: Electrochemical and surface studies. Colloids Surf. A Physicochem. Eng. Asp. 2021, 609, 125689. [Google Scholar] [CrossRef]
  13. Lashgari, S.M.; Bahlakeh, G.; Ramezanzadeh, B. Detailed theoretical DFT computation/molecular simulation and electrochemical explorations of Thymus vulgaris leave extract for effective mild-steel corrosion retardation in HCl solution. J. Mol. Liq. 2021, 335, 115897. [Google Scholar] [CrossRef]
  14. Haldhar, R.; Prasad, D.; Saxena, A.; Singh, P. Valeriana wallichii root extract as a green & sustainable corrosion inhibitor for mild steel in acidic environments: Experimental and theoretical study. Mater. Chem. Front. 2018, 2, 1225–1237. [Google Scholar] [CrossRef]
  15. Shahini, M.H.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Superior inhibition action of the Mish Gush (MG) leaves extract toward mild steel corrosion in HCl solution: Theoretical and electrochemical studies. J. Mol. Liq. 2021, 332, 115876. [Google Scholar] [CrossRef]
  16. Salmasifar, A.; Edraki, M.; Alibakhshi, E.; Ramezanzadeh, B.; Bahlakeh, G. Theoretical design coupled with experimental study of the effectiveness of the inhibitive molecules based on Cynara scolymus L extract toward chloride-induced corrosion of steel. J. Mol. Liq. 2021, 332, 115742. [Google Scholar] [CrossRef]
  17. Mostafatabar, A.H.; Bahlakeh, G.; Ramezanzadeh, B.; Dehghani, A.; Ramezanzadeh, M. A comprehensive electronic-scale DFT modeling, atomic-level MC/MD simulation, and electrochemical/surface exploration of active nature-inspired phytochemicals based on Heracleum persicum seeds phytoextract for effective retardation of the acidic-induced c. J. Mol. Liq. 2021, 331, 115764. [Google Scholar] [CrossRef]
  18. Tehrani, M.E.H.N.; Ghahremani, P.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B. Theoretical and experimental assessment of a green corrosion inhibitor extracted from Malva sylvestris. J. Environ. Chem. Eng. 2021, 9, 105256. [Google Scholar] [CrossRef]
  19. Salmasifar, A.; Edraki, M.; Alibakhshi, E.; Ramezanzadeh, B.; Bahlakeh, G. Combined electrochemical/surface investigations and computer modeling of the aquatic Artichoke extract molecules corrosion inhibition properties on the mild steel surface immersed in the acidic medium. J. Mol. Liq. 2021, 327, 114856. [Google Scholar] [CrossRef]
  20. Lashgari, S.M.; Yari, H.; Mahdavian, M.; Ramezanzadeh, B.; Bahlakeh, G.; Ramezanzadeh, M. Synthesis of graphene oxide nanosheets decorated by nanoporous zeolite-imidazole (ZIF-67) based metal-organic framework with controlled-release corrosion inhibitor performance: Experimental and detailed DFT-D theoretical explorations. J. Hazard. Mater. 2021, 404, 124068. [Google Scholar] [CrossRef]
  21. Wagner, H.; Schwarting, G. Struktur der microphyllinsäure aus dem harz von Convolvulus microphyllus. Phytochemistry 1977, 16, 715–717. [Google Scholar] [CrossRef]
  22. Koch, G.; Varney, J.; Thompson, N.; Moghissi, O.; Gould, M.; Payer, J. International Measures of Prevention, Application, and Economics of Corrosion Technologies Study, 1st ed.; NACE International: Houston, TX, USA, 2016. [Google Scholar]
  23. Dehghani, A.; Bahlakeh, G.; Ramezanzadeh, B.; Mofidabadi, A.H.J. Construction of a high-potency anti-corrosive metal-organic film based on europium (III)-benzimidazole: Theoretical and electrochemical investigations. Constr. Build. Mater. 2021, 269, 121271. [Google Scholar] [CrossRef]
  24. Keshmiri, N.; Najmi, P.; Ramezanzadeh, B.; Ramezanzadeh, M.; Bahlakeh, G. Nano-scale P, Zn-codoped reduced-graphene oxide incorporated epoxy composite; synthesis, electronic-level DFT-D modeling, and anti-corrosion properties. Prog. Org. Coat. 2021, 159, 106416. [Google Scholar] [CrossRef]
  25. Kumar, S.; Vashisht, H.; Olasunkanmi, L.O.; Bahadur, I.; Verma, H.; Goyal, M.; Singh, G.; Ebenso, E.E. Polyurethane Based Triblock Copolymers as Corrosion Inhibitors for Mild Steel in 0.5 M H2SO4. Ind. Eng. Chem. Res. 2017, 56, 441–456. [Google Scholar] [CrossRef]
  26. Belghiti, M.; Echihi, S.; Mahsoune, A.; Karzazi, Y.; Aboulmouhajir, A.; Dafali, A.; Bahadur, I. Piperine derivatives as green corrosion inhibitors on iron surface; DFT, Monte Carlo dynamics study and complexation modes. J. Mol. Liq. 2018, 261, 62–75. [Google Scholar] [CrossRef]
  27. Verma, D.; Khan, F.; Bahadur, I.; Salman, M.; Quraishi, M.; Verma, C.; Ebenso, E.E. Inhibition performance of Glycine max, Cuscuta reflexa and Spirogyra extracts for mild steel dissolution in acidic medium: Density functional theory and experimental studies. Results Phys. 2018, 10, 665–674. [Google Scholar] [CrossRef]
  28. Goyal, M.; Vashisht, H.; Kumar, S.; Bahadur, I. Anti-corrosion performance of eco-friendly inhibitor (2-aminobenzyl) triphenylphosphonium bromide ionic liquid on mild steel in 0.5M sulfuric acid. J. Mol. Liq. 2018, 261, 162–173. [Google Scholar] [CrossRef]
  29. Verma, C.; Ebenso, E.E.; Bahadur, I.; Quraishi, M.A. An overview on plant extracts as environmental sustainable and green corrosion inhibitors for metals and alloys in aggressive corrosive media. J. Mol. Liq. 2018, 266, 577–590. [Google Scholar] [CrossRef]
  30. Goyal, M.; Kumar, S.; Verma, C.; Bahadur, I.; Ebenso, E.E.; Lgaz, H.; Chung, I.-M. Interfacial adsorption behavior of quaternary phosphonium based ionic liquids on metal-electrolyte interface: Electrochemical, surface characterization and computational approaches. J. Mol. Liq. 2020, 298, 111995. [Google Scholar] [CrossRef]
  31. Goyal, M.; Vashisht, H.; Kumar, A.; Kumar, S.; Bahadur, I.; Benhiba, F.; Zarrouk, A. Isopentyltriphenylphosphonium bromideionic liquid as a newly effective corrosion inhibitor on metal-electrolyte interface in acidic medium: Experimental, surface morphological (SEM-EDX & AFM) and computational analysis. J. Mol. Liq. 2020, 316, 113838. [Google Scholar] [CrossRef]
  32. Sadik, K.; El Hamdani, N.; Hachim, M.E.; Byadi, S.; Bahadur, I.; Aboulmouhajir, A. Towards a theoretical understanding of alkaloid-extract Cytisine derivatives of Retama monosperma (L.) Boiss. Seeds, as eco-friendly inhibitor for carbon steel corrosion in acidic 1M HCl solution. J. Theor. Comput. Chem. 2020, 19, 2050013. [Google Scholar] [CrossRef]
  33. Goyal, M.; Vashist, H.; Kumar, S.; Bahadur, I.; Benhiba, F.; Zarrouk, A. Acid corrosion inhibition of ferrous and non-ferrous metal by nature friendly Ethoxycarbonylmethyltriphenylphosphonium Bromide (ECMTPB): Experimental and MD simulation evaluation. J. Mol. Liq. 2020, 315, 113705. [Google Scholar] [CrossRef]
  34. Kumar, B.; Vashisht, H.; Goyal, M.; Kumar, A.; Benhiba, F.; Prasad, A.K.; Kumar, S.; Bahadur, I.; Zarrouk, A. Study of adsorption mechanism of chalcone derivatives on mild steel-sulfuric acid interface. J. Mol. Liq. 2020, 318, 113890. [Google Scholar] [CrossRef]
  35. Daoudi, W.; El Aatiaoui, A.; Dagdag, O.; Zaidi, K.; Haldhar, R.; Kim, S.C.; Oussaid, A.; Aouinti, A.; Berisha, A.; Benhiba, F.; et al. Anti-corrosion coating formation by a biopolymeric extract of Artemisia herba-alba plant: Experimental and theoretical investigations. Coatings 2023, 13, 611. [Google Scholar] [CrossRef]
  36. Kashani, F.R.; Rezaei, M. Electrochemical studies and molecular simulations on the use of molybdic acid for stabilization of AISI 304 stainless steel passive film in sulfuric acid medium. J. Mol. Liq. 2021, 344, 117733. [Google Scholar] [CrossRef]
  37. Verma, D.K.; Al Fantazi, A.; Verma, C.; Khan, F.; Asatkar, A.; Hussain, C.M.; Ebenso, E.E. Experimental and computational studies on hydroxamic acids as environmental friendly chelating corrosion inhibitors for mild steel in aqueous acidic medium. J. Mol. Liq. 2020, 314, 113651. [Google Scholar] [CrossRef]
  38. Kashani, F.R.; Rezaei, M. Improving the localized corrosion resistance of 304 stainless steel in HCl solution by adsorption of molybdate ions: Interaction mechanisms at the interface using molecular dynamics simulation and electrochemical noise analysis. Colloids Surf. A Physicochem. Eng. Asp. 2022, 647, 129085. [Google Scholar] [CrossRef]
  39. Verma, D.K.; Aslam, R.; Aslam, J.; Quraishi, M.A.; Ebenso, E.E.; Verma, C. Computational Modeling: Theoretical Predictive Tools for Designing of Potential Organic Corrosion Inhibitors. J. Mol. Struct. 2021, 1236, 130294. [Google Scholar] [CrossRef]
  40. Alibakhshi, E.; Ramezanzadeh, M.; Bahlakeh, G.; Ramezanzadeh, B.; Mahdavian, M.; Motamedi, M. Glycyrrhiza glabra leaves extract as a green corrosion inhibitor for mild steel in 1 M hydrochloric acid solution: Experimental, molecular dynamics, Monte Carlo and quantum mechanics study. J. Mol. Liq. 2018, 255, 185–198. [Google Scholar] [CrossRef]
  41. Jmiai, A.; Tara, A.; El Issami, S.; Hilali, M.; Jbara, O.; Bazzi, L. A new trend in corrosion protection of copper in acidic medium by using Jujube shell extract as an effective green and environmentally safe corrosion inhibitor: Experimental, quantum chemistry approach and Monte Carlo simulation study. J. Mol. Liq. 2021, 322, 114509. [Google Scholar] [CrossRef]
  42. Naseri, E.; Hajisafari, M.; Kosari, A.; Talari, M.; Hosseinpour, S.; Davoodi, A. Inhibitive effect of Clopidogrel as a green corrosion inhibitor for mild steel; statistical modeling and quantum Monte Carlo simulation studies. J. Mol. Liq. 2018, 269, 193–202. [Google Scholar] [CrossRef]
  43. Berrissoul, A.; Ouarhach, A.; Benhiba, F.; Romane, A.; Zarrouk, A.; Guenbour, A.; Dikici, B.; Dafali, A. Evaluation of Lavandula mairei extract as green inhibitor for mild steel corrosion in 1 M HCl solution. Experimental and theoretical approach. J. Mol. Liq. 2020, 313, 113493. [Google Scholar] [CrossRef]
  44. Zuo, X.; Li, W.; Luo, W.; Zhang, X.; Qiang, Y.; Zhang, J.; Li, H.; Tan, B. Research of Lilium brownii leaves extract as a commendable and green inhibitor for X70 steel corrosion in hydrochloric acid. J. Mol. Liq. 2021, 321, 114914. [Google Scholar] [CrossRef]
  45. Abdallah, M.; Altass, H.M.; Al-Gorair, A.S.; Al-Fahemi, J.H.; Jahdaly, B.A.A.L.; Soliman, K.A. Natural nutmeg oil as a green corrosion inhibitor for carbon steel in 1.0 M HCl solution: Chemical, electrochemical, and computational methods. J. Mol. Liq. 2021, 323, 115036. [Google Scholar] [CrossRef]
  46. Khadom, A.A.; Kadhim, M.M.; Anaee, R.A.; Mahood, H.B.; Mahdi, M.S.; Salman, A.W. Theoritical evaluation of Citrus Aurantium Leaf Extract as green inhibitor for chemical and biological corrosion of mild steel in acidic solution: Statistical, molecular dynamics, Docking, and quantum mechanics study. J. Mol. Liq. 2021, 343, 116978. [Google Scholar] [CrossRef]
  47. Dagdag, O.; Safi, Z.; Hamed, O.; Jodeh, S.; Wazzan, N.; Haldhar, R.; Safi, S.K.; Berisha, A.; El Gouri, M. Comparative study of some epoxy polymers based on bisphenolic and aromatic diamines: Synthesis, viscosity, thermal properties computational and statistical approaches. J. Polym. Res. 2021, 28, 165. [Google Scholar] [CrossRef]
  48. Dagdag, O.; El Bachiri, A.; Hamed, O.; Haldhar, R.; Verma, C.; Ebenso, E.; El Gouri, M. Dendrimeric Epoxy Resins Based on Hexachlorocyclotriphosphazene as a Reactive Flame Retardant Polymeric Materials: A Review. J. Inorg. Organomet. Polym. Mater. 2021, 31, 3240–3261. [Google Scholar] [CrossRef]
  49. El-Aouni, N.; Hsissou, R.; Safi, Z.; Abbout, S.; Benhiba, F.; El Azzaoui, J.; Haldhar, R.; Wazzan, N.; Guo, L.; Erramli, H.; et al. Performance of two new epoxy resins as potential corrosion inhibitors for carbon steel in 1MHCl medium: Combining experimental and computational approaches. Colloids Surf. A Physicochem. Eng. Asp. 2021, 626, 127066. [Google Scholar] [CrossRef]
  50. Dagdag, O.; Guo, L.; Safi, Z.; Verma, C.; Ebenso, E.E.; Wazzan, N.; Masroor, S.; Haldhar, R.; Jodeh, S.; El Gouri, M. Epoxy resin and TiO2 composite as anticorrosive material for carbon steel in 3% NaCl medium: Experimental and computational studies. J. Mol. Liq. 2020, 317, 114249. [Google Scholar] [CrossRef]
  51. Erramli, H.; Dagdag, O.; Safi, Z.; Wazzan, N.; Guo, L.; Abbout, S.; Ebenso, E.; Verma, C.; Haldhar, R.; El Gouri, M. Trifunctional epoxy resin as anticorrosive material for carbon steel in 1 M HCl: Experimental and computational studies. Surf. Interfaces 2020, 21, 100707. [Google Scholar] [CrossRef]
  52. Saxena, A.; Prasad, D.; Haldhar, R. Investigation of Corrosion Inhibition Effect and Adsorption Activities of Achyranthes aspera Extract for Mild Steel in 0.5 M H2SO4. J. Fail. Anal. Prev. 2018, 18, 957–968. [Google Scholar] [CrossRef]
  53. Haldhar, R.; Prasad, D.; Bahadur, I.; Dagdag, O.; Kaya, S.; Verma, D.K.; Kim, S.-C. Investigation of plant waste as a renewable biomass source to develop efficient, economical and eco-friendly corrosion inhibitor. J. Mol. Liq. 2021, 335, 116184. [Google Scholar] [CrossRef]
  54. Saxena, A.; Prasad, D.; Haldhar, R. Use of Syzygium aromaticum extract as green corrosion inhibitor for mild steel in 0.5M H2SO4. Surf. Rev. Lett. 2018, 26, 1850200. [Google Scholar] [CrossRef]
  55. Zavidovskiy, I.A.; Streletskiy, O.A.; Nuriahmetov, I.F.; Nishchak, O.Y.; Savchenko, N.F.; Tatarintsev, A.A.; Pavlikov, A.V. Highly selective polyene-polyyne resistive gas sensors: Response tuning by low-energy ion irradiation. J. Compos. Sci. 2023, 7, 156. [Google Scholar] [CrossRef]
  56. Tucureanu, V.; Matei, A.; Avram, A.M. FTIR spectroscopy for carbon family study. Crit. Rev. Anal. Chem. 2016, 46, 502–520. [Google Scholar] [CrossRef]
  57. Saxena, A.; Prasad, D.; Haldhar, R.; Singh, G.; Kumar, A. Use of Sida cordifolia extract as green corrosion inhibitor for mild steel in 0.5 M H2SO4. J. Environ. Chem. Eng. 2018, 6, 694–700. [Google Scholar] [CrossRef]
  58. Saxena, A.; Prasad, D.; Haldhar, R. Investigation of corrosion inhibition effect and adsorption activities of Cuscuta reflexa extract for mild steel in 0.5 M H2SO4. Bioelectrochemistry 2018, 124, 156–164. [Google Scholar] [CrossRef] [PubMed]
  59. Saxena, A.; Prasad, D.; Haldhar, R.; Singh, G.; Kumar, A. Use of Saraca ashoka extract as green corrosion inhibitor for mild steel in 0.5 M H2SO4. J. Mol. Liq. 2018, 258, 89–97. [Google Scholar] [CrossRef]
  60. Haldhar, R.; Kim, S.-C.; Prasad, D.; Bedair, M.; Bahadur, I.; Kaya, S.; Dagdag, O.; Guo, L. Papaver somniferum as an efficient corrosion inhibitor for iron alloy in acidic condition: DFT, MC simulation, LCMS and electrochemical studies. J. Mol. Struct. 2021, 1242, 130822. [Google Scholar] [CrossRef]
  61. Haldhar, R.; Prasad, D.; Saxena, A. Myristica fragrans extract as an eco-friendly corrosion inhibitor for mild steel in 0.5 M H2SO4 solution. J. Environ. Chem. Eng. 2018, 6, 2290–2301. [Google Scholar] [CrossRef]
  62. Haldhar, R.; Prasad, D.; Saxena, A.; Kumar, R. Experimental and theoretical studies of Ficus religiosa as green corrosion inhibitor for mild steel in 0.5 M H2SO4 solution. Sustain. Chem. Pharm. 2018, 9, 95–105. [Google Scholar] [CrossRef]
  63. Haldhar, R.; Prasad, D.; Bhardwaj, N. Extraction and experimental studies of Citrus aurantifolia as an economical and green corrosion inhibitor for mild steel in acidic media. J. Adhes. Sci. Technol. 2019, 33, 1169–1183. [Google Scholar] [CrossRef]
  64. Haldhar, R.; Prasad, D.; Saxena, A. Armoracia rusticana as sustainable and eco-friendly corrosion inhibitor for mild steel in 0.5M sulphuric acid: Experimental and theoretical investigations. J. Environ. Chem. Eng. 2018, 6, 5230–5238. [Google Scholar] [CrossRef]
  65. Haldhar, R.; Prasad, D.; Bhardwaj, N. Surface Adsorption and Corrosion Resistance Performance of Acacia concinna Pod Extract: An Efficient Inhibitor for Mild Steel in Acidic Environment. Arab. J. Sci. Eng. 2020, 45, 131–141. [Google Scholar] [CrossRef]
  66. Haldhar, R.; Prasad, D.; Saxena, A.; Kaur, A. Corrosion resistance of mild steel in 0.5 M H2SO4 solution by plant extract of Alkana tinctoria: Experimental and theoretical studies. Eur. Phys. J. Plus 2018, 133, 356. [Google Scholar] [CrossRef]
  67. Haldhar, R.; Prasad, D.; Bahadur, I.; Dagdag, O.; Berisha, A. Evaluation of Gloriosa superba seeds extract as corrosion inhibition for low carbon steel in sulfuric acidic medium: A combined experimental and computational studies. J. Mol. Liq. 2021, 323, 114958. [Google Scholar] [CrossRef]
  68. Haldhar, R.; Prasad, D.; Mandal, N.; Benhiba, F.; Bahadur, I.; Dagdag, O. Anticorrosive properties of a green and sustainable inhibitor from leaves extract of Cannabis sativa plant: Experimental and theoretical approach. Colloids Surf. A Physicochem. Eng. Asp. 2021, 614, 126211. [Google Scholar] [CrossRef]
  69. Haldhar, R.; Prasad, D.; Bhardwaj, N. Experimental and Theoretical Evaluation of Acacia catechu Extract as a Natural, Economical and Effective Corrosion Inhibitor for Mild Steel in an Acidic Environment. J. Bio- Tribo-Corros. 2020, 6, 76. [Google Scholar] [CrossRef]
  70. Haldhar, R.; Prasad, D.; Kamboj, D.; Kaya, S.; Dagdag, O.; Guo, L. Corrosion inhibition, surface adsorption and computational studies of Momordica charantia extract: A sustainable and green approach. SN Appl. Sci. 2021, 3, 25. [Google Scholar] [CrossRef]
  71. Haldhar, R.; Prasad, D. Corrosion Resistance and Surface Protective Performance of Waste Material of Eucalyptus globulus for Low Carbon Steel. J. Bio- Tribo-Corros. 2020, 6, 48. [Google Scholar] [CrossRef]
  72. Saxena, A.; Prasad, D.; Haldhar, R. Withania somnifera extract as green inhibitor for mild steel in 8 % H2SO4. Asian J. Chem. 2016, 28, 2471–2474. [Google Scholar] [CrossRef]
  73. Haldhar, R.; Prasad, D.; Saharan, H. Performance of Pfaffia paniculata extract towards corrosion mitigation of low-carbon steel in an acidic environment. Int. J. Ind. Chem. 2020, 11, 1–12. [Google Scholar] [CrossRef]
  74. Haldhar, R.; Prasad, D.; Nguyen, L.T.; Kaya, S.; Bahadur, I.; Dagdag, O.; Kim, S.-C. Corrosion inhibition, surface adsorption and computational studies of Swertia chirata extract: A sustainable and green approach. Mater. Chem. Phys. 2021, 267, 124613. [Google Scholar] [CrossRef]
  75. Benhiba, F.; Hsissou, R.; Benzekri, Z.; Echihi, S.; El-Blilak, J.; Boukhris, S.; Bellaouchou, A.; Guenbour, A.; Oudda, H.; Warad, I.; et al. DFT/electronic scale, MD simulation and evaluation of 6-methyl-2-(p-tolyl)-1,4-dihydroquinoxaline as a potential corrosion inhibition. J. Mol. Liq. 2021, 335, 116539. [Google Scholar] [CrossRef]
  76. Dagdag, O.; Safi, Z.; Wazzan, N.; Erramli, H.; Guo, L.; Mkadmh, A.M.; Verma, C.; Ebenso, E.; El Gana, L.; El Harfi, A. Highly functionalized epoxy macromolecule as an anti-corrosive material for carbon steel: Computational (DFT, MDS), surface (SEM-EDS) and electrochemical (OCP, PDP, EIS) studies. J. Mol. Liq. 2020, 302, 112535. [Google Scholar] [CrossRef]
  77. Salman, M.; Ansari, K.R.; Srivastava, V.; Chauhan, D.S.; Haque, J.; Quraishi, M.A. Chromeno naphthyridines based heterocyclic compounds as novel acidizing corrosion inhibitors: Experimental, surface and computational study. J. Mol. Liq. 2021, 322, 114825. [Google Scholar] [CrossRef]
  78. Kumar, H.; Yadav, V. Highly efficient and eco-friendly acid corrosion inhibitor for mild steel: Experimental and theoretical study. J. Mol. Liq. 2021, 335, 116220. [Google Scholar] [CrossRef]
  79. Kumar, H.; Yadav, V.; Anu; Saha, S.K.; Kang, N. Adsorption and inhibition mechanism of efficient and environment friendly corrosion inhibitor for mild steel: Experimental and theoretical study. J. Mol. Liq. 2021, 338, 116634. [Google Scholar] [CrossRef]
  80. Kumar, H.; Dhanda, T. Cyclohexylamine an effective corrosion inhibitor for mild steel in 0.1 N H2SO4: Experimental and theoretical (molecular dynamics simulation and FMO) study. J. Mol. Liq. 2021, 327, 114847. [Google Scholar] [CrossRef]
  81. Mechbal, N.; Belghiti, M.; Benzbiria, N.; Lai, C.-H.; Kaddouri, Y.; Karzazi, Y.; Touzani, R.; Zertoubi, M. Correlation between corrosion inhibition efficiency in sulfuric acid medium and the molecular structures of two newly eco-friendly pyrazole derivatives on iron oxide surface. J. Mol. Liq. 2021, 331, 115656. [Google Scholar] [CrossRef]
  82. Liu, Y.; Fan, B.; Xu, B.; Yang, B. Ambient-stable polyethyleneimine functionalized Ti3C2Tx nanohybrid corrosion inhibitor for copper in alkaline electrolyte. Mater. Lett. 2023, 337, 133979. [Google Scholar] [CrossRef]
  83. Dong, P.; Zhang, Y.; Zhu, S.; Nie, Z.; Ma, H.; Liu, Q.; Li, J. First-Principles Study on the Adsorption Characteristics of Corrosive Species on Passive Film TiO2 in a NaCl Solution Containing H2S and CO2. Metals 2022, 12, 1160. [Google Scholar] [CrossRef]
  84. Liu, Z.; Fan, B.; Zhao, J.; Yang, B.; Zheng, X. Benzothiazole derivatives-based supramolecular assemblies as efficient corrosion inhibitors for copper in artificial seawater: Formation, interfacial release and protective mechanisms. Corros. Dcience 2023, 212, 110957. [Google Scholar] [CrossRef]
  85. Bhardwaj, N.; Prasad, D.; Haldhar, R. Study of the Aegle marmelos as a Green Corrosion Inhibitor for Mild Steel in Acidic Medium: Experimental and Theoretical Approach. J. Bio- Tribo-Corros. 2018, 4, 61. [Google Scholar] [CrossRef]
  86. Saxena, A.; Prasad, D.; Haldhar, R. Use of Asparagus racemosus extract as green corrosion inhibitor for mild steel in 0.5 M H2SO4. J. Mater. Sci. 2018, 53, 8523–8535. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of the process of plant extract preparation.
Figure 1. Schematic representation of the process of plant extract preparation.
Coatings 13 00860 g001
Figure 2. FTIR spectra of pure C. microphyllus plant extract.
Figure 2. FTIR spectra of pure C. microphyllus plant extract.
Coatings 13 00860 g002
Figure 3. The inhibitor solution’s UV-vis spectra in both modes.
Figure 3. The inhibitor solution’s UV-vis spectra in both modes.
Coatings 13 00860 g003
Figure 4. Stable OCP (a) and PDP curves (b) with and without several concentrations of inhibitor solution.
Figure 4. Stable OCP (a) and PDP curves (b) with and without several concentrations of inhibitor solution.
Coatings 13 00860 g004
Figure 5. Nyquist curves (a) and Bode curves (b) for LCS with and without several concentrations of inhibitor solution (Z′ = real, Z″ = imaginary, Z = vector of length, freq = Frequency).
Figure 5. Nyquist curves (a) and Bode curves (b) for LCS with and without several concentrations of inhibitor solution (Z′ = real, Z″ = imaginary, Z = vector of length, freq = Frequency).
Coatings 13 00860 g005
Figure 6. The Langmuir adsorption isotherm plot for C. microphyllus inhibitor.
Figure 6. The Langmuir adsorption isotherm plot for C. microphyllus inhibitor.
Coatings 13 00860 g006
Figure 7. SEM and AFM micrographs of (a) polished, (b) corroded, and (c) protected surfaces of LCS. Reprinted with permission from [68]. Copyright 2021, Elsevier.
Figure 7. SEM and AFM micrographs of (a) polished, (b) corroded, and (c) protected surfaces of LCS. Reprinted with permission from [68]. Copyright 2021, Elsevier.
Coatings 13 00860 g007
Figure 8. Optimized, HOMO, LUMO, and MEP structures of the M1 and M2 molecules.
Figure 8. Optimized, HOMO, LUMO, and MEP structures of the M1 and M2 molecules.
Coatings 13 00860 g008
Figure 9. MDS (a) and MC (b) simulation of studied molecules on the metal (Fe(110)) surface.
Figure 9. MDS (a) and MC (b) simulation of studied molecules on the metal (Fe(110)) surface.
Coatings 13 00860 g009
Figure 10. Proposed adsorption phenomenon of kaempferol molecules on the LSC surface.
Figure 10. Proposed adsorption phenomenon of kaempferol molecules on the LSC surface.
Coatings 13 00860 g010
Figure 11. Inhibition efficiencies of C. microphyllus-derived film (extract) obtained by different techniques at several inhibitor concentrations.
Figure 11. Inhibition efficiencies of C. microphyllus-derived film (extract) obtained by different techniques at several inhibitor concentrations.
Coatings 13 00860 g011
Table 1. FT-IR spectra analysis of C. microphyllus extract.
Table 1. FT-IR spectra analysis of C. microphyllus extract.
FT-IR (Wavenumber; cm−1)
2931.89C–H stretching
1118.89C–O stretching
1610.61C=O stretching
3389.04O–H stretching
1670.63C=C stretching
1305.25C–C stretching
Table 2. Corrosion rate and inhibition efficiency of C. microphyllus-derived film (extract) with different concentrations, both with and without inhibitor.
Table 2. Corrosion rate and inhibition efficiency of C. microphyllus-derived film (extract) with different concentrations, both with and without inhibitor.
Temperature (K)Inhibitor Concentration
C (mg/L)
Corrosion Rate
C R (mmy−1)
Inhibition Efficiency
IE (%)
Blank11.33-
1003.6767.55
2003.0872.74
2983002.4278.59
4001.8983.27
5001.7284.78
6001.4786.95
Blank15.29-
1005.5967.61
2004.7072.81
3083004.0778.64
4003.5183.31
5002.4084.81
6002.3987.02
Blank19.14-
1007.5763.44
2006.3369.26
3183005.7573.38
4004.9177.04
5004.0778.73
6003.8879.73
Table 3. PDP parameters of C. microphyllus-derived film (extract) for LCS in 0.5 M sulphuric acid.
Table 3. PDP parameters of C. microphyllus-derived film (extract) for LCS in 0.5 M sulphuric acid.
C
(mg/L)
E c o r r
(mV vs. SCE)
i c o r r
(μA cm−2)
β a
(mV/dec)
β c
(mV/dec)
IE
(%)
Blank465890.90141.66164.25-
100239251.7053.00106.6271.74
200452226.2062.22107.0074.61
300625144.4068.28106.7083.79
400460141.8045.21108.4584.08
500469103.0058.96122.8088.44
60046367.0634.79107.4092.47
Table 4. EIS parameters for C. microphyllus-derived film (extract) for LCS in 0.5 M sulphuric acid.
Table 4. EIS parameters for C. microphyllus-derived film (extract) for LCS in 0.5 M sulphuric acid.
C
(mg/L)
R c t
(Ω cm−2)
C d l
(μF cm−2)
IE
(%)
Blank15.71269.00-
10049.98247.4368.56
20072.58232.1978.35
30076.40220.2179.43
400104.60211.3484.98
500117.09205.0586.58
600155.13155.1589.87
Table 5. Theoretical parameters of kaempferol (M1) and p-hydroxycinnamic acid (M2).
Table 5. Theoretical parameters of kaempferol (M1) and p-hydroxycinnamic acid (M2).
MoleculeEHOMO (eV)ELUMO (eV)ΔE (eV)ΔN
(eV)
ΔEBack-Donation
(eV)
η
(eV)
σ
(eV−1)
χ
(eV)
π
(eV)
Kaempferol (M1)−5.36−2.412.940.22−0.361.470.673.88−3.88
p-hydroxycinnamic acid (M2)−5.70−2.513.190.27−0.391.590.624.11−4.11
Table 6. Einteraction and Ebinding energies of Kaempferol (M1) and p-hydroxycinnamic acid (M2).
Table 6. Einteraction and Ebinding energies of Kaempferol (M1) and p-hydroxycinnamic acid (M2).
Systems E i n t e r a c t i o n
(kJ/mol)
E b i n d i n g
(kJ/mol)
Fe(110)/M1−620.00620.00
Fe(110)/M2−545.00545.00
Table 7. Comparison of inhibition efficiency between previously reported and the current study.
Table 7. Comparison of inhibition efficiency between previously reported and the current study.
Sr. No.InhibitorOptimum
Concentration
Corrosive
Media
IE (%)Ref.
1Citrullus lanatus800 ppm1 M HCl91[2]
2Eucalyptus800 ppm1 M HCl88[6]
3Garlic10 cm3/L0.5 M H2SO488[10]
4Cuscuta reflexa2 g/L1 M HCl81[27]
5Glycyrrhiza glabra800 ppm1 M HCl88[40]
6Lavandula mairei0.4 g/L1 M HCl92[43]
7Lilium brownii200 mg/L0.5 M H2SO485[44]
8Achyranthes aspera500 ppm0.5 M H2SO490[52]
9Aegle marmelos500 ppm0.5 M H2SO483[85]
10Asparagus racemosus100 mg/L0.5 M H2SO493[86]
11Myristica fragrans500 mg/L0.5 M H2SO487[61]
12Ficus religiosa500 mg/L0.5 M H2SO492[62]
13Alkana tinctoria500 mg/L0.5 M H2SO491[66]
14Pfaffia paniculate600 mg/L0.5 M H2SO488[73]
15Convolvulus microphyllus600 mg/L0.5 M H2SO492This work
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

Haldhar, R.; Vanaraj, R.; Dagdag, O.; Berisha, A.; Kim, S.-C. Convolvulus microphyllus Extract as a Green, Effective, and Affordable Corrosion Inhibitor: Theoretical Calculations and Experimental Studies. Coatings 2023, 13, 860. https://doi.org/10.3390/coatings13050860

AMA Style

Haldhar R, Vanaraj R, Dagdag O, Berisha A, Kim S-C. Convolvulus microphyllus Extract as a Green, Effective, and Affordable Corrosion Inhibitor: Theoretical Calculations and Experimental Studies. Coatings. 2023; 13(5):860. https://doi.org/10.3390/coatings13050860

Chicago/Turabian Style

Haldhar, Rajesh, Ramkumar Vanaraj, Omar Dagdag, Avni Berisha, and Seong-Cheol Kim. 2023. "Convolvulus microphyllus Extract as a Green, Effective, and Affordable Corrosion Inhibitor: Theoretical Calculations and Experimental Studies" Coatings 13, no. 5: 860. https://doi.org/10.3390/coatings13050860

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