Next Article in Journal
Multifunctional Derivatives of Spiropyrrolidine Tethered Indeno-Quinoxaline Heterocyclic Hybrids as Potent Antimicrobial, Antioxidant and Antidiabetic Agents: Design, Synthesis, In Vitro and In Silico Approaches
Previous Article in Journal
Classical vs. Non-Classical Cyclometalated Pt(II) Complexes
Previous Article in Special Issue
Pumice as a Novel Natural Heterogeneous Catalyst for the Designation of 3,4-Dihydropyrimidine-2-(1H)-ones/thiones under Solvent-Free Conditions
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

An Integrated Experimental and Theoretical Studies on the Corrosion Inhibition of Carbon Steel by Harmal Extracts

by
Hassan H. Hammud
1,*,
Sarah A. Maache
2,*,
Nasreen Al Otaibi
1,* and
Nadeem S. Sheikh
3,*
1
Department of Chemistry, College of Science, King Faisal University, Al-Ahsa 31982, Saudi Arabia
2
Leading National Academy, Khobar Niagara College, Khobar 13244, Saudi Arabia
3
Chemical Sciences, Faculty of Science, Universiti Brunei Darussalam, Jalan Tungku Link, Bandar Seri Begawan BE1410, Brunei
*
Authors to whom correspondence should be addressed.
Molecules 2022, 27(21), 7250; https://doi.org/10.3390/molecules27217250
Submission received: 12 September 2022 / Revised: 8 October 2022 / Accepted: 19 October 2022 / Published: 25 October 2022
(This article belongs to the Special Issue Eco-Friendly Materials and Their Applications)

Abstract

:
The corrosion inhibition effect of the three extracts from Harmal roots (HRE), leaves (HLE), and flowers (HFE) were studied for carbon steel corrosion inhibition in 0.25 M H2SO4 solution. The electrochemical impedance study indicated that the three types of extracts decreased corrosion effectively through a charge transfer mechanism. Harmal roots and leaf extracts showed inhibition values of 94.1% and 94.2%, while it was 88.7% for Harmal flower extract at the inhibitor concentration of 82.6 ppm. Potentiodynamic polarization data revealed that Harmal extracts acted through predominant cathodic type inhibition. Both the corrosion current density and corrosion rate decreased significantly in the presence of Harmal extracts compared to blank solution. The corrosion rate (mpy) value was 63.3, 86.1, and 180.7 for HRE, HLE, and HFE, respectively. The adsorption-free energy change ΔGads (kJ·mol−1) values calculated from the Langmuir adsorption isotherm plots were for HRE (−35.08), HLE (−33.17), and HFE (−33.12). Thus, corrosion inhibition occurred due to the adsorption of Harmal extract on the carbon steel surface via the chemisorption mechanism. Moreover, a computational investigation using B3LYP/6-311G++(d,p) basis set in both gaseous and aqueous phases was performed for the major alkaloids (1–8) present in the Harmal extract.

1. Introduction

Corrosion [1] is a relentless and unavoidable process. However, the advancements in corrosion science have rendered a wide range of economically viable solutions. This ubiquitous natural phenomenon has a detrimental industrial impact and undoubtedly consumes billions of dollars annually. According to the statistics, the estimated cost of corrosion in China back in 2015 was ~USD 310 billion. This represented around 3.34% of the nation’s gross domestic product (GDP) [2]. More or less, similar statistics were reported for the corrosion cost in the United States of America [3]. The damaging exposure of metallic surfaces to air and moisture is considered one of the vast challenges on an industrial scale that can cause major economic losses [4]. Despite laser focus on technological advancements to prepare corrosion-resistant materials and develop efficient corrosion inhibitors, corrosion still poses a serious challenge for the scientific community. One of the most sustainable and efficient approaches to corrosion is an application of eco-friendly inhibitors [5,6,7,8] that produce minimal toxic waste. In this regard, the utility of organic substances as corrosion inhibitors is one of the most efficient protocols to mitigate metallic degradation [9]. This protocol is based on the adsorption of the inhibitor on the metallic surface to form a shielding layer that protects the metal from the adverse effect of corrosion. Quite recently, many classes of heterocyclic compounds have been extensively probed as efficient corrosion inhibitors. The presence of several functional groups such as hydroxyl, methoxy, amine, nitro, nitrile, carboxyl, ester, amide, and azo, among others, and multiple p-bonds acted as an adsorption point through which organic compounds bind with the metallic surfaces [10,11,12]. Additionally, polymeric structures [13] and materials [14] have been utilized as inhibitors, although these substances are relatively expensive with low environmental benign properties and polarity. Keeping all these aspects in consideration, current studies focused on exploring natural products as eco-friendly and cost-effective organic inhibitors, including plant extracts [15,16], oleochemicals [17], natural gums [18], and bio-surfactants [19,20].
Peganum harmal plant is one of the pharmaceutically significant herbs that primarily grows in the Asian regions [21]. It has been traditionally utilized as a natural medicine for treating tumors, infections, and inflammations in Greek, Iranian, Indian, and Chinese nations [22]. The significance of this plant is owed to the presence of active alkaloids, as shown in Figure 1a, which are concentrated in the plant’s seeds and roots [23]. Both harmine (1) and harmaline (2) exhibit, more or less, similar pharmacological effects, yet the latter has relatively less toxicity. Moreover, other indoles, quinazoline, and quinazolinone-based alkaloids (38) were also found to be active components in the harmal extracts with substantial medicinal utilities [24]. Specifically, harmine alkaloid (1) possesses anti-diabetic properties and has been used as a traditional medicine to treat diabetes. While harmaline (2), harman, and harmalol alkaloids act as vasorelaxants and angiogenic inhibitors to treat many cardiovascular diseases, including hypertension and bradycardia [25,26,27]. The extract of the harmal plant has also been used as an analgesic medicine to treat severe pain [28] and Parkinson’s disease [29]. In addition, the extract has anti-hallucination, anti-excitation, and anti-depression activities [30]. These active alkaloids have antitumor activity owing to their anti-proliferative effects that reduce the growth of colon cancer cells [31,32]. Other bioactive properties of the harmal extract include its application as an efficient antimicrobial and antibacterial agent against several types of bacteria, such as Proteus vulgaris and Bacillus subtilis [33,34]. Harmane (6) and vasicine (7) showed the most acetylcholinesterase inhibitor effect and can be used to treat Alzheimer’s disease [35]. In terms of respiratory disorders, vasicine (7) and vasicinone (8) have been reported to inhibit cough due to their bronchodilation activity [36]. In addition, the extract is also used at an industrial scale as a natural colorant for fabrics, especially wool, silk, and cotton [37].
DFT calculations are primarily used in inhibition studies as a fast, cost-effective tool to investigate the electron donor/acceptor regions in organic inhibitors. This protocol relies on providing a comprehensive study of chemical reactivity parameters to gain insight into the inhibition mechanism and, thus, support the experimental studies. Contrary to this, theoretical calculations warrant further investigations to accurately predict a relationship between the quantum chemical parameters and the corrosion inhibition efficiency due to limitations associated with the electron flow procedure [38,39].
Herein, we report the preparation of Harmal extracts by sonication from three portions of the plant roots, leaves, and flowers. The extracts were tested as corrosion inhibitors of carbon steel (C-steel) in dilute sulfuric acid solution using impedance and potentiodynamic polarization techniques. Harmal plant extract is an eco-friendly corrosion inhibitor because it is less toxic, biodegradable, highly efficient, available, renewable, and inexpensive. Moreover, the constituents of Harmal extracts have nitrogen and oxygen functional groups that can act as adsorption centers to the metal surface, Figure 1a. Additionally, a computational investigation for the potent bioactive alkaloids present in the Harmal extract has been delineated.

2. Experimental

2.1. Materials

2.1.1. Preparation of Harmal Parts Extracts

Harmal plants were collected from Alhasa in the east regain of Saudi Arabia. Leave root and flower portions were separated, dried in the shade, and then ground to powder. Then, 100 g of powder from each plant portion was added to 100 mL ethanol in the bottom flask. The solution was then continuously sonicated for 1 h at 80 kHz using a POWER SONIC405 [40]. The green extract was filtered then the solvent was removed using a rotary evaporator. The obtained residue was further dried in an air oven at 50 °C for several hours. About one gram of green powder was obtained. A 0.100 g of the powder was dissolved in 10.00 mL methanol to obtain a 10,000 ppm stock solution for each plant portion. Then they were labeled as Harmal roots extract (HRE), Harmal leaves extract (HLE), and Harmal flowers extract (HFE). The green extract solutions were later used as corrosion inhibitors for C-steel in 0.25M H2SO4 solution.

2.1.2. Working Electrodes

A cylindrical carbon steel (C-steel) electrode with an exposed area of 0.5027 cm2 was used as the working electrode. The following is the chemical composition of carbon steel rod: % weight (wt%) C, 0.164; S, 0.001; Mn, 0.710; P, 0.0005; Si, 0.26, Ni, 0.123; Cr, 0.041; balance Fe. The C-steel working electrode was buffed with a series of silicon carbide papers from 600 to 1200 before the beginning of the electrochemical measurements.

2.1.3. Electrolytes

The working electrode was immersed in 0.25 M H2SO4 solution containing 0, 20.8, 29.1, 41.5, 62.1, and 82.6 ppm of the Harmal extract. The standard solutions were prepared by diluting the stock solution (10,000 ppm) with the corresponding dilute water portions.

2.2. Methods

2.2.1. Electrochemical Experiment

Impedance study (EIS) and potentiodynamic polarization study (PDP) measurements were carried out using (Gamry, reference 600 Potentiostat/Galvanostat/ZRA, Warminster, PA, USA) and Gamry software v7.07. The cell used consisted of platinum (Pt) wire auxiliary electrode and a saturated silver reference electrode (Ag/AgCl) electrode. A C-steel rod was used as a working electrode, as described above. Before the start of the impedance experiment, the steady-state open circuit potential EOCP was measured by immersing the electrodes in the corrosion solution for 60 min with the indication 10 mV disturbance capacity [41].
Potentiodynamic polarization measurements were carried out in the direction from the cathode to the anode from 500 mV versus the resting potential at a scan rate of 5 mV/s. The percent inhibition IE% and fractions of surface coverage (θ) were evaluated from the following Equations (1) and (2) [42]:
I E % = ( 1 I   corr ( i ) I   corr ( 0 ) ) × 100
θ = 1 I corr ( i ) I corr ( 0 )
where, Icorr(0) and Icorr(i) are the corrosion current density of the steel specimen (mA.cm−2) for the uninhibited and inhibited solutions, respectively.
Electrochemical impedance studies (EIS) were carried out at a potential amplitude of 10 mV, peak to peak (AC signal) in frequencies ranging from 100 kHz to 0.1 Hz.
The obtained impedance results were represented as Nyquist and Bode plots [41]. The following modified Equation (3) was used to calculate the inhibition efficiency (IE%) from the Nyquist plot [42].
I E % = 1 ( R p ( 0 ) R p ( i ) ) × 100
The polarization resistance Rp = Rct + Rf, where Rct is the charge transfer resistance of C-steel and Rf is the coating resistance. (0) indicate a blank solution, and (i) indicates an inhibitor solution. Moreover, the surface coverage (θ) was calculated using the following Equation (4):
θ = 1 R p ( 0 ) R p ( i )

2.2.2. Computational Details

Density functional theory (DFT) [43] calculations were performed using Gaussian 09 (revision E.01) [44], and the Gaussview [45] was used to generate input geometries and visualize the output structures. Geometry optimizations and frequency calculations for the key constituents 1–8 present in the extract of Harmal leaves were performed using the B3LYP functional [46,47,48] with the 6-311++G(d,p) basis set [49]. The calculations were performed in the gaseous phase. However, for comparative purposes and to model the solvation effect, the calculations were carried out in an aqueous medium as well by applying the most commonly used integral equation formalism (IEF) version of the polarized continuum model (PCM) [50,51]. This is essential as the electrochemical corrosion phenomenon takes place in a liquid phase. All stationary points were characterized as minima based on normal vibrational mode analysis. Thermal corrections were computed from unscaled frequencies, assuming a standard state of 298.15 K and 1 atm. The structures described herein are the lowest energy-optimized conformers.

3. Results and Discussions

3.1. Corrosion Inhibition Study

The open-circuit potential, Eocp, for each of the experiments conducted was first examined for 60 min before applying an electrochemical impedance study (EIS) to obtain a steady current reading on the carbon steel surface [52].
The open circuit potential Eocp plots versus time (s) of the C-steel with and without the addition of different concentrations of (a) Harmal Roots extracts, (b) Harmal Leaves extracts, and (c) Harmal Flowers extracts in 0.25M H2SO4 are shown in Supplementary Figure S1.

3.1.1. Electrochemical Impedance Study (EIS)

This study was performed to investigate the amount of current flow and the resistance values occurring on the carbon steel surface with and without the presence of inhibitors [52]. All the processes involved in the electrical response of the system were fitted against the equivalent electrical circuits (EEC) model 1 (Figure 2a) and model 2 (Figure 2b). Nyquist plots are shown in Figure 3 for HRE, Figure 4 for HLE, and Figure 5 for HFE, with (a) for the fitting by model 1 and (b) for the fitting by model 2.
Thus, the equivalent electrical circuit models 1 and 2 with two-time constants were used. They gave a good fitting of impedance data for the blank and the solutions of all ranges of concentrations of Harmal extract roots (Table 1 and Table 2), leaves (Table 3 and Table 4), and flowers (Table 5 and Table 6), respectively.
In general, each electrochemical process in the circuit should be represented by a separate distinct semicircle in the Nyquist plot [42]. But, the first semicircle is not observed due to the high-frequency limit [53]. The fitted model 1 consists of two circuits connected in series and contains solution resistance (Rs), film resistance (Rf), charge-transfer resistance (Rct), constant phase element of the film (CPEf), and the double layer (CPEdl). The double-layer capacitance (Cdl) can also be calculated from the values of Rct, the impedance of CPEdl (ZCPE), and the exponents of CPEdl (n) [41]. The data fitted by model 1 is shown in Table 1, Table 3, and Table 5; Figure 3a, Figure 4a, and Figure 5a for HRE, HLE, and HFE, respectively.
The solution resistance (Rs) describes the ohmic resistance, while the charge transfer resistance (Rct) represents the inhibitor’s resistance towards oxidation of the metal surface, and it is inversely proportional to the corrosion rate [53]. Rf or coating resistance (Rc) is referred to as an intact film/coat (isolation layer), which provides good protective performance [54].
The EEC model 2 (Figure 2b) was applied, which added a diffusion component to model 1 (Figure 2a) to fit the experimental data. If the diffusion region is close to the inhibitor film/metal interface, the mass transfer of reactive species is decreased, resulting in a Warburg impedance W [41,54]. Moreover, a constant phase element (CPE) of model 1 is replaced by a pure double layer capacitor (Cdl) in model 2 to fit the semicircle shape of the Nyquist plot. The results of fitting the data to model 2 are shown in Table 2, Table 4, and Table 6.
A careful investigation of Nyquist plots shows that model 2 showed a slightly better fitting than model 1 when higher concentrations of 62.1 ppm and 82.6 ppm were used, Figure 3, Figure 4 and Figure 5. This indicates that the mass transfer of reactive species at higher concentrations was slowed down next to the inhibitor film/metal interface resulting in a Warburg impedance W.
The higher the inhibitor concentrations, the higher the diameter of the Nyquist plots Figure 3, Figure 4 and Figure 5 [55]. Thus, the higher the polarization resistance values Rp = Rct + Rf, Table 1 (HRE), Table 3 (HLE) and Table 5 (HFE) using model 1 and Table 2 (HRE), Table 4 (HLE) and Table 6 (HFE) using model 2. This caused an increase in the inhibition efficiency of the inhibitor. This could correspond to the strengthening of the inhibitive film on the carbon steel surface [56].
According to Table 1, Table 2, Table 3, Table 4, Table 5 and Table 6, Harmal roots and leaves extracts showed a maximum inhibition value averaged from both models, 94.1% and 94.2% at the concentration of 82.6 ppm, which are higher compared to 88.7% for Harmal flowers extract at the same concentration. The maintained semicircle shapes of the Nyquist plots over the experiments revealed that the corrosion inhibition process occurred through a charge transfer mechanism [52]. The increase in Rct systematically and the decrease in Cdl values with few exceptions upon adding Harmal extracts of various portions of roots, leaves, and flowers indicated a reduction in the corrosion rate. This is because of the replacement of water molecules on the working electrode surface by the inhibitor’s protective films Table 1–6.

Bode and Bode Phase Angle

Bode and Bode phase plots: log (imaginary part of impedance) (log |Z| (ohm)), and phase angle (°) were plotted against log frequency (logf (Hz)) respectively, Figure 6, Figure 7 and Figure 8, for C-steel in 0.25 M H2SO4 in the absence and presence of Harmal extracts [41]. Figure 6a, 7a, and 8a show the log Bode phase angle (°) plotted against the logf (Hz) for the inhibitor HRE, HLE, and HFE, respectively.
It is obvious that increasing the Harmal extract concentration causes more negative values of phase angle at the intermediate frequency (peak), and thus the inhibitive behavior increases [41,42]. In fact, the phase angle values increased from −27.87° for the blank solution to −62.52° (HRE), −66.28° (HLE), and −57.58° (HFE) for the inhibitor solution (82.6 ppm), Table 7.
The α values represent the surface irregularities for C-steel. They were calculated for the Harmal extracts from the slope of the linear region of log|Z| versus the logf plots [42]; (HRE) Figure 6b, (HLE) Figure 7b, and (HFE) Figure 8b; Z (ohms) being the imaginary part of impedance and f (Hz) is the frequency. Hypothetically, the value of α is equal to −1 for an ideal capacitor. In contrast, the coarseness of the C-steel surface occurs for α values less than −1. It can be seen that the presence of Harmal extracts increased the α value from −0.3635 (0 ppm) to −0.7190 (HRE), −0.7671 (HLE), and −0.6571 (HFE) for the Harmal extract (82.6 ppm), Table 7. Thus, the decrease of heterogeneity of the C-steel surface occurred as a result of the inhibition of Harmal extracts.

3.1.2. Potentiodynamic Polarization Study (PDP)

Figure 9a–c shows the Tafel polarization curves obtained for carbon steel in the absence and presence of Harmal part extracts roots, leaves, and flowers, respectively, in 0.25 M H2SO4 at room temperature.
The analysis of the Tafel polarization curves is computed using the Gamry software v7.07, and the results are presented in Table 8 (HRE), Table 9 (HLE), and Table 10 (HFE). The equilibrium corrosion potential (Ecorr) values shifted to more negative values for the inhibitor in comparison with the blank solution. But the potential Ecorr does not follow a consistent increase with increasing concentration of inhibitors. This was also observed in other reported works [42,56]. The inhibitors could be classified as mixed-type inhibitors [52]. But there is a consistent Icorr current decrease with increased concentration, which clearly indicates the enhancement of inhibition. Thus, the current values are more reliable than the potential values to evaluate inhibition efficiency.
Icorr (μA/cm2) for the diluted sulfuric acid blank is 3401.63, while for 82.6 ppm solution of extract, it is 138.65 (HRE), 188.38 (HLE), and 395.86 (HFE). The greatest decrease in current occurred for HRE. The organic constituents of Harmal part extracts probably adsorbed on the carbon steel surface and formed an insulating film which caused decreasing in corrosion current values [57]. The decrease in the electrical current (I) flow is due to a lowering in the number of electrons being transferred across the metal surface. This suggests that the metal dissolution process (oxidation of iron) was hindered and that the Harmal extracts act as anodic inhibitors [52].
Accordingly, the corrosion rate in mils penetration per year (mpy) decreased from 1554.0 for the blank solution to 63.34 for HRE, 86.1 for HLE, and 180.7 HFE extracts, at an 82.6 ppm inhibitor concentration. Again, the greatest decrease in corrosion rate also occurred for HRE, and the corrosion rate followed the order HRE < HLE < HFE. This indicated that the extracts’ organic constituents inhibited carbon steel’s corrosion process in a dilute H2SO4 solution. Moreover, it is clear from the Tafel plots in Figure 9a–c that the addition of an inhibitor effectively suppressed cathodic reactions. This suggests that Harmal extracts functioned as a cathodic inhibitor. In addition, the corrosion parameters such as Ecorr, Icorr, slope of anodic (βa) and cathodic (βc), and inhibition efficiency values are listed in Table 8, Table 9 and Table 10. The addition of inhibitor decreased the magnitude of βa and βc, indicating that the extracts acted as an anodic and cathodic inhibitor of mild steel samples in H2SO4 solutions [56].
On the other hand, the calculated % inhibition efficiency %IE from the Tafel curves for the highest inhibitor concentration (82.6 ppm) was 95.9% for HRE, 94.5% for HLE, and 88.4% for HFE. The obtained %IE using polarization studies are in close agreement with those obtained using the impedance studies. In conclusion, the inhibitor strength is HRE > HLE > HFE.
In the present work, we have studied three extracts, Harmal roots, leaves, and flowers, instead of only the Harmal leaves extract previously reported [41]. We have concluded that all three types of extracts can be considered efficient inhibitors. The immersion time of C-steel in 0.25M H2SO4 solution in the present work was 60 min compared to 15 min previously. The inhibition of 94.2% in the present work for Harmal leaves extract at 82.6 ppm is much higher than the previously reported 59.7% at 82.6 ppm. One concludes that the longer the time of exposure to inhibitors solutions, the higher the IE%.

3.2. Adsorption Isotherms

In order to study the adsorption behavior of Harmal extracts in H2SO4 solution, several isotherm models were implemented using the inhibition values obtained from polarization studies. These are Langmuir, Frumkin, and Temkin [41]. However, the best fit was obtained for the Langmuir isotherm model for all three extracts Figure 10a–c, and additionally, Frumkin also fitted the data of the HLE extract, Figure 10d.
The Langmuir isotherm model monitors the variation of adsorption coefficient Kads with a concentration of inhibitor C according to the following relationship Equation (5) [56]:
c/θ = c + 1/Kads
The adsorption parameters, regression coefficient R2, Kads, and slope values are obtained by straight line fitting between C/θ and C, Figure 10a (HRE), b (HLE), and c (HFE). The R2 coefficient values are all ≥0.98. It is obvious that the slope values are close to 1.0, as expected from the Langmuir model Equation. According to the Langmuir adsorption isotherm, the tested inhibitor molecules have a typical adsorption site at the metal/solution interface with no interaction with the other molecules adsorbed.
The fitting of the Temkin model was tested by plotting log(θ/C) versus θ. The R2 values obtained from the straight line for HRE, HLE, and HFE are 0.4934, 0.5492, and 0.8343, respectively. Thus, the Temkin model is not an acceptable model in all cases. In contrast, the Frumkin model plot of log[θ/(1 − θ)·C)] versus θ gave R2 values of 0.9086, 0.9929, and 0.8699 for HRE, HLE, and HFE (Figure 10d). These values are also a little smaller than the corresponding R2 values obtained from the Langmuir model, Figure 10a–c. Thus, the Langmuir model is the best model to represent the adsorption of Harmal molecules of the three extracts onto C-steel.
The adsorption equilibrium constant (Kads) was calculated using the inverse of the intercept of the Langmuir plot. In all the adsorption isotherm plots, the concentration used was ppm. The obtained Kads (ppm−1) values from the Langmuir isotherm plots were just converted to Kads (mol/L)−1. The latter values were used to calculate ΔG°ads. The values of Kads for HRE, HLE, and HFE were 24.61 × 103 M−1, 11.42 × 103 M−1, and 11.21 × 103 M−1, respectively. The standard free energy of adsorption ΔG0ads can be calculated from Kads using Equation 6 [40]:
ΔG0ads = −RT ln(55.5Kads)
where R is the gas constant (8.314 J mol K−1), and T is the absolute temperature (K). The obtained values for ΔG0ads are −35.08, −33.17, and −33.12 kJ/mol for HRE, HLE, and HFE, respectively; indicating that the adsorption mechanism of the inhibitors onto the C-steel in 0.25 M H2SO4 solution involved chemisorption adsorption [42]. Finally, it is noted that HRE had the highest value of Kads and ΔG0ads and thus exhibited the most effective corrosion inhibition.

3.3. Inhibition Mechanism

Corrosion inhibition of C-steel in sulfuric acid solution by different Harmal organic molecules present in the three extracts can be explained on the basis of molecular adsorption. The compounds inhibit corrosion by blocking cathodic sites. In a weak acidic solution, the Harmal constituents can exist as neutral or protonated species (by protonation of the nitrogen atom by H+ of sulfuric acid).
The neutral species are chemically adsorbed on steel, while protonated species are physically adsorbed on the carbon steel forming a protective film. This causes a decrease in the exposed surface area of steel in contact with the electrolyte solution and thus causes a decrease in the evolution of hydrogen gas.
The chemical adsorption of Harmal constituents occurs through the formation of coordination bonds between electron-rich functional groups orbitals of Harmal molecules with empty d- orbital of the Fe atom or Fe2+ ion, Figure 11. π electrons of imines >C=N, and aromatic rings and lone pair of electrons of different hetero nitrogen and oxygen atoms from Harmal molecules can be shared with iron orbitals. Chemical adsorption onto the C-steel surface can also occur through π-electrons back donation from Fe dt2g orbital onto empty π* orbital of aromatic rings of the Hamal molecules, Figure 11.
The physical adsorption of the Harmal molecules can occur via electrostatic interactions between the protonated Harmal molecules and the negatively charged steel. This is because the SO42− anion from H2SO4 is attracted to a positively charged steel surface containing (Fe2+) via electrostatic interactions, rendering the surface of the C-steel negatively charged. Therefore, both the physical and chemical adsorption of Harmal molecules favor stronger adsorption and inhibit corrosion by blocking the active site of the iron surface.
The % inhibitor efficiency calculated from impedance and Tafel studies ranges from about 90% to 95% for the three extracts in the order HRE > HLE > HFE. This indicates that the extracts from three parts of the Harmal plants can be used as corrosion inhibitors. Thus, this study indicates that most of the Harmal plants can be collected, grounded, extracted, and used as an inhibitor. Therefore, the production yield to be used from Harmal plant is much higher than focusing on only one extract from leaves; thus, it is of greater economic value for industrial application. Moreover, from FTIR studies (Figure 12), there is an indication that the three extracts contain mainly the same electron-rich functional groups that promote inhibition, such as hydroxyl, amine, imine, alkoxy, and aromatic rings. The stretching vibrations (cm−1) occur for: N-H and O-H at 3278 and 3200 cm−1, C-H aromatic at 2925 and 2840, O-H carboxylic at 2632, >C=O stretch at 1732 cm−1, >C=N and >C=C< alkene at 1640 and aromatic at 1440–1586, >C-O at 1024 and 1250. The bending vibration (cm−1) occurs for C-H aromatic at 748.

3.4. Comparative Studies of Inhibition Efficiency IE%

The corrosion inhibition efficiency of steel by Harmal roots, leaves, and flower extracts in sulfuric acid media was compared with other ones reported in the literature, Table 11. The results showed that the three Harmal extracts studied at a low concentration of 83 ppm with high IE% (93%), averaged from the three extracts, are more effective towards the inhibition of C-steel corrosion compared to other reported ones: Date palm seed (2000 ppm, 91%) [58], tea tree (2250 ppm, 78.6%) [59], Ginkgo leaf extract (200 ppm, 90.0%) [60], and Brassica oleracea L (300 ppm, 92.7%) [61].

3.5. Surface Morphology

Figure 13 shows SEM micrographs of C-steel immersed for 3 h in (a) free 0.25 M H2SO4 and (b) 0.25 M H2SO4 containing 300 ppm of Harmal leaf extracts HLE. The morphology in Figure 13a shows a rough surface containing numerous white particles of iron oxides pointing out from the surface, which characterizes the uniform corrosion of C-steel in sulfuric acid solution. While in the presence of HLE extracts (Figure 13b), a smoother surface can be observed because of the formation of the eco-friendly inhibitor’s protective film on the metal surface. This finding may confirm that the presence of Harmal extracts molecules improves the inhibition efficiency of the C-steel.

3.6. Computational Studies

The optimized structure, depicted in Figure 14, represents the stable conformers of eight key organic compounds present in the Harmal extract. All these alkaloids present in the Harmal extract have structurally related units; however, they vary due to the presence of methoxy and hydroxyl groups. Another noticeable feature is the presence of the carbonyl group in alkaloid 8, which may impact its behavior and display dissimilar values of electronic properties than structurally related natural alkaloid 7. Moreover, the aromatic systems in compounds 1 and 4 offer an extended conjugation and, thus, are presumed to experience similar electronic properties and behavior despite having methoxy and hydroxy substituents, respectively.

3.6.1. Frontier Molecular Orbital Analysis and Electrostatic Potential Map

The analysis of the frontier molecular orbitals (FMOs) comprising the highest occupied molecular orbitals (HOMOs) and lowest unoccupied molecular orbitals (LUMOs) is fundamental to understanding the chemical reactivity, stability, identification of active sites, and hardness and softness-related features of any molecule [62,63,64,65,66]. The energy difference between these orbitals is referred to as the energy gap or band gap, and it is strongly affected by the substitution pattern present in a molecule. The energy of HOMO (EHOMO) indicates the electron donation capability of a molecule. Hence, high values of EHOMO show the molecule’s tendency to donate electrons to electrophilic species, in addition to its correlation with the ionization potential of a molecule. The energy of LUMO (ELUMO) represents the ability to gain electrons, which helps to determine the electrophilic regions in a molecule and corresponds with its electron affinity. The calculated values of EHOMO and ELUMO for the compounds 18 present in the extract of the Harmal plant in the aqueous phase are shown in Figure 15. As expected, the shapes of the frontier molecular orbitals are similar for compounds 1 and 4 due to the presence of identical structural features. A similar trend can be observed for the pair of compounds 2 and 5. However, it appears to be a little different for compounds 7 and 8, and it is believed to be due to the presence of the carbonyl group in the indole system in compound 8. The latter compounds appear to be structurally different compared to the rest of the alkaloids present in Harmal extract, as they contain benzylic rings adjacent to the indole system, which may greatly impact their electronic properties. Moreover, harmane (6) can be compared with tetrahydroharmine (3) in terms of the effect of methoxy substituent on the inhibition efficiency of both molecules. In addition to FMOs, we have also probed the molecular electrostatic potential (ESP) for all the investigated compounds [67,68]. This helps to visualize a three-dimensional distribution of the electronic charge within a molecule. The blue-colored region has the highest electrostatic potential energy and represents the electron-deficient site, while the red-colored zone indicates the lowest potential and is an indication of the electron-affluent part of the molecule. As shown in Figure 15, the nitrogen atoms in the six-membered ring attached to the indole skeleton contain the highest electronic density due to the lone pair. At the same time, the resonance phenomenon renders the nitrogen atoms of the indole skeleton to be the most electron-deficient regions in the molecules.

3.6.2. EHOMO and ELUMO and ΔE

The values of FMOs are calculated and depicted in Figure 16. The order of inhibition efficiency varies according to the changing values of FMOs. Compound 3 has the largest EHOMO in both states, while compound 8 has the lowest ELUMO in both states. Subsequently, the energy gap values (ΔE) followed a descending order commencing with compound 8 with a value of 5.019 eV in a gaseous phase, which also remains in the same range throughout all the phases in both neutral and protonated forms. Surprisingly, ΔE starts decreasing in the following order 8 > 3 > 1 > 7 > 4 > 6 > 2 > 5. As illustrated in Figure 16, compound 5 shows the lowest energy gap in both the gaseous and solvent states. Accordingly, it is expected to show high inhibition potency in comparison with other compounds. Similarly, compound 2 shows a similar behavior having ΔE of 3.475 eV, which is relatively close to compound 5. Compounds 1, 4, 6, and 7 have greater ΔE, and this is assumed to be related to the extended conjugation in compounds 1 and 4 and the presence of two electron-rich atoms in compounds that can give relative stability to the compounds and lower their reactivity accordingly. Compound 3 has a high energy gap, although its only difference in the conformation is the absence of the double bond between C11 and C10, which implies its impact on the reactivity. The largest ΔE value was noted for compound 8, which can be referred to as the presence of the carbonyl and the methoxy groups in the compound that can enhance the energy difference between both FMOs, leading to low inhibition efficiency. In the protonated form (gaseous phase), compound 3 still has the largest EHOMO, and it starts to show a significant decrease for other compounds. As previously mentioned, compounds 1 and 4 have structural similarities (extended conjugation), which is why they both show relatively close values. On the other hand, the high EHOMO shown by compound 2 in comparison to 5 can be influenced by the presence of methoxy/hydroxyl groups. Compound 8 has the lowest ELUMO in this gaseous phase, and the values start increasing in the following order 2 < 4 < 1 < 3. High ELUMO implies a high tendency to accept electrons. EHOMO, ELUMO, and ΔE of the solvent state showed a slight elevation than the gaseous state. Compared to the neutral form, the same results are given by the same compounds (8 and 5) in the protonated form in terms of EHOMO and ELUMO. However, the lowest ΔE here is calculated for compound 2 in both gaseous and aqueous phases. Further study is underway in our group to calculate the quantum chemical parameters for the major alkaloids present in the harmal extract by applying various available methods, including electron propagator theory (EPT) [69], and will be disseminated in due course.

4. Conclusions

This study revealed that the three types of Harmal extracts roots, leaves, and flowers (HRE, HLE, HFE) acted as efficient green corrosion inhibitors on carbon steel in 0.25 M H2SO4 media, and HRE extract showed the highest inhibitory activity amongst the three extracts studied. The % corrosion inhibition efficiency %E calculated from impedance and potentiodynamic polarization is in the order HRE = HLE > HFE. In contrast, the corrosion rate follows the order HRE < HLE < HFE. Moreover, the high negative values of ΔGads calculated from the Langmuir isotherm revealed that corrosion inhibition occurred due to the adsorption of the extract on C-steel through the chemisorption mechanism. In addition, DFT calculations were performed to visualize the frontier molecular orbitals and the energy gap between the HOMO and LUMO for the major alkaloids present in the harmal extract.

Supplementary Materials

The following are available online at https://www.mdpi.com/article/10.3390/molecules27217250/s1, Figure S1: The open circuit potential EOCP plot versus time (s) of C-steel with and without the addition of different concentrations of Harmal extracts (a) Roots, (b) Leaves and (c) Flowers in 0.25M H2SO4, Table S1: Computed energies for the compounds 1-8 present in the Harmal extract calculated at B3LYP/6-311G++(d,p) [values are in Hartree], Table S2: Cartesian coordinates for the compounds 1-8 present in the Harmal extract computed at B3LYP/6-311G++(d,p).

Author Contributions

H.H.H. and S.A.M. contributed equally to this work. H.H.H. and N.S.S. conceptualized, designed, supervised, and acquired the funding for the project. N.A.O. carried out the extraction and corrosion inhibition studies, and S.A.M. performed the computational investigation. All authors analyzed the results and got involved in writing the manuscript. Both H.H.H. and N.S.S. edited to generate the final version of the manuscript, agreed upon by all the authors. All authors have read and agreed to the published version of the manuscript.

Funding

This work was supported through the Annual Funding track by the Deanship of Scientific Research, Vice Presidency for Graduate Studies and Scientific Research, King Faisal University, Saudi Arabia [Project No. AN000551].

Data Availability Statement

All data are available in the main text or the Electronic Supplementary Information (ESI).

Acknowledgments

The authors acknowledge the Deanship of Scientific Research at King Faisal University: Saudi Arabia: for the financial support under the Annual Funding track (Grant No. AN000551).

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

The extracted samples of the Harmal plant are available from the authors.

References

  1. Zehra, S.; Mobin, M.; Aslam, J. An overview of the corrosion chemistry. In Environmentally Sustainainable Corrosion Inhibitors; Elsevier: Amsterdam, The Netherlands, 2022; pp. 3–23. [Google Scholar] [CrossRef]
  2. Hou, B.; Li, X.; Ma, X.; Du, C.; Zhang, D.; Zheng, M.; Xu, W.; Lu, D.; Ma, F. The cost of corrosion in China, Npj. Mater. Degrad. 2017, 1, 4. [Google Scholar] [CrossRef]
  3. Koch, G.H.; Brongers, M.P.; Thompson, N.G.; Virmani, Y.P.; Payer, J.H. Cost of corrosion in the United States. In Handbook of Environmental Degradation of Materials; Elsevier: Amsterdam, The Netherlands, 2005; pp. 3–24. [Google Scholar] [CrossRef]
  4. Alamri, A.H. Localized corrosion and mitigation approach of steel materials used in oil and gas pipelines—An overview. Eng. Fail. Anal. 2020, 116, 104735. [Google Scholar] [CrossRef]
  5. Panchal, J.; Shah, D.; Patel, R.; Shah, S.; Prajapati, M.; Shah, M. Comprehensive Review and Critical Data Analysis on Corrosion and Emphasizing on Green Eco-friendly Corrosion Inhibitors for Oil and Gas Industries. J. Bio-Tribo-Corros. 2021, 7, 107. [Google Scholar] [CrossRef]
  6. Kaur, J.; Daksh, N.; Saxena, A. Corrosion Inhibition Applications of Natural and Eco-Friendly Corrosion Inhibitors on Steel in the Acidic Environment: An Overview. Arab. J. Sci. Eng. 2022, 47, 57–74. [Google Scholar] [CrossRef]
  7. Radha, K.V.; Patel, D.; Kumar, N.; Devasena, T. Investigation of Eco-friendly Corrosion Inhibitor for Low Carbon Steel Using Extract of Physalis Minima Leaves. J. Bio-Tribo-Corros. 2022, 8, 47. [Google Scholar] [CrossRef]
  8. Zomorodian, A.; Bagonyi, R.; Al-Tabbaa, A. The efficiency of eco-friendly corrosion inhibitors in protecting steel reinforcement. J. Build. Eng. 2021, 38, 102171. [Google Scholar] [CrossRef]
  9. Verma, C.; Verma, D.K.; Ebenso, E.E.; Quraishi, M.A. Quraishi, Sulfur and phosphorus heteroatom-containing compounds as corrosion inhibitors: An overview. Heteroat. Chem. 2018, 29, e21437. [Google Scholar] [CrossRef] [Green Version]
  10. Assad, H.; Kumar, A. Understanding functional group effect on corrosion inhibition efficiency of selected organic compounds. J. Mol. Liq. 2021, 344, 117755. [Google Scholar] [CrossRef]
  11. Chaubey, N.; Qurashi, A.; Chauhan, D.S.; Quraishi, M. Frontiers and advances in green and sustainable inhibitors for corrosion applications: A critical review. J. Mol. Liq. 2021, 321, 114385. [Google Scholar] [CrossRef]
  12. Serdaroğlu, G.; Kaya, S. Organic and Inorganic Corrosion Inhibitors. In Organic Corrosion Inhibitors; Wiley: Hoboken, NJ, USA, 2021; pp. 59–73. [Google Scholar] [CrossRef]
  13. Al Hashmi, S.; Al Zadjali, S.; Rajput, N.S.; Mohammedture, M.; Gutierrez, M.; Esawi, A.M.K. Polymers and Graphene-Based Materials as Barrier Coatings. In Handbook of Research on Green Synthesis and Applications of Nanomaterials; IGI Global: Hershey, PA, USA, 2022; pp. 129–151. [Google Scholar] [CrossRef]
  14. Aslam, R.; Mobin, M.; Aslam, J. Nanomaterials as corrosion inhibitors. In Inorganic Anticorrosive Materials; Elsevier: Amsterdam, The Netherlands, 2022; pp. 3–20. [Google Scholar] [CrossRef]
  15. Murungi, P.I.; Sulaimon, A.A. Ideal corrosion inhibitors: A review of plant extracts as corrosion inhibitors for metal surfaces. Corros. Rev. 2022, 40, 127–136. [Google Scholar] [CrossRef]
  16. Fazal, B.R.; Becker, T.; Kinsella, B.; Lepkova, K. A review of plant extracts as green corrosion inhibitors for CO2 corrosion of carbon steel. Npj Mater. Degrad. 2022, 6, 5. [Google Scholar] [CrossRef]
  17. Ansari, F.A.; Chauhan, D.S.; Quraishi, M.A. Oleochemicals as Corrosion Inhibitors. In Organic Corrosion Inhibitors: Synthesis, Characterization, Mechanism, and Applications; Wiley: Hoboken, NJ, USA, 2021; pp. 343–369. [Google Scholar] [CrossRef]
  18. Vaidya, N.R.; Aklujkar, P.; Rao, A.R. Modification of natural gums for application as corrosion inhibitor: A review. J. Coatings Technol. Res. 2022, 19, 223–239. [Google Scholar] [CrossRef]
  19. Fawzy, A.; Abdallah, M.; Alfakeer, M.; Altass, H.M.; Althagafi, I.I.; El-Ossaily, Y.A. Performance of unprecedented synthesized biosurfactants as green inhibitors for the corrosion of mild steel-37-2 in neutral solutions: A mechanistic approach. Green Chem. Lett. Rev. 2021, 14, 488–499. [Google Scholar] [CrossRef]
  20. Lou, Y.; Chang, W.; Cui, T.; Wang, J.; Qian, H.; Ma, L.; Hao, X.; Zhang, D. Microbiologically influenced corrosion inhibition mechanisms in corrosion protection: A review. Bioelectrochemistry 2021, 141, 107883. [Google Scholar] [CrossRef]
  21. Moloudizargari, M.; Mikaili, P.; Aghajanshakeri, S.; Asghari, M.H.; Shayegh, J. Pharmacological and therapeutic effects of Peganum harmala and its main alkaloids. Pharmacogn. Rev. 2013, 7, 199. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  22. Niroumand, M.C.; Farzaei, M.H.; Amin, G. Medicinal properties of Peganum harmala L. in traditional Iranian medicine and modern phytotherapy: A review. J. Tradit. Chinese Med. 2015, 35, 104–109. [Google Scholar] [CrossRef]
  23. Filban, F.; Ravanbakhsh, M.; Poormohammadi, A.; Khaghani, S.; Sadeghi-Nejad, B.; Neisi, A.; Goudarzi, G. Antimicrobial properties of Peganum harmala L. seeds’ smoke in indoors: Applications and prospects. Environ. Monit. Assess. 2022, 194, 17. [Google Scholar] [CrossRef]
  24. Karasawa, M.M.G.; Mohan, C. Fruits as Prospective Reserves of bioactive Compounds: A Review. Nat. Products Bioprospect. 2018, 8, 335–346. [Google Scholar] [CrossRef] [Green Version]
  25. Tahraoui, A.; El-Hilaly, J.; Israili, Z.; Lyoussi, B. Ethnopharmacological survey of plants used in the traditional treatment of hypertension and diabetes in south-eastern Morocco (Errachidia province). J. Ethnopharmacol. 2007, 110, 105–117. [Google Scholar] [CrossRef]
  26. Shi, C.-C.; Liao, J.-F.; Chen, C.-F. Comparative Study on the Vasorelaxant Effects of Three Harmala Alkaloids In Vitro. Jpn. J. Pharmacol. 2001, 85, 299–305. [Google Scholar] [CrossRef] [Green Version]
  27. Berrougui, H.; Cordero, C.M.; Khalil, A.; Hmamouchi, M.; Ettaib, A.; Marhuenda, E.; Herrera, M. Vasorelaxant effects of harmine and harmaline extracted from Peganum harmala L. seed’s in isolated rat aorta. Pharmacol. Res. 2006, 54, 150–157. [Google Scholar] [CrossRef] [PubMed]
  28. Shahrajabian, M.H.; Sun, W.; Cheng, Q. Improving health benefits with considering traditional and modern health benefits of Peganum harmala. Clin. Phytosci. 2021, 7, 18. [Google Scholar] [CrossRef]
  29. Herraiz, T.; González, D.; Ancín-Azpilicueta, C.; Arán, V.; Guillén, H. β-Carboline alkaloids in Peganum harmala and inhibition of human monoamine oxidase (MAO). Food Chem. Toxicol. 2010, 48, 839–845. [Google Scholar] [CrossRef] [PubMed]
  30. Farouk, L.; Laroubi, A.; Aboufatima, R.; Benharref, A.; Chait, A. Evaluation of the analgesic effect of alkaloid extract of Peganum harmala L.: Possible mechanisms involved. J. Ethnopharmacol. 2008, 115, 449–454. [Google Scholar] [CrossRef] [PubMed]
  31. Zhang, Q.; Zan, Y.-H.; Yang, H.-G.; Yang, M.-Y.; Liu, F.-S.; Li, S.-G.; Peng, X.-H.; Lin, B.; Li, Z.-L.; Li, D.-H.; et al. Anti-tumor alkaloids from Peganum harmala. Phytochemistry 2022, 197, 113107. [Google Scholar] [CrossRef]
  32. Elkady, A.I.; A Hussein, R.; El-Assouli, S.M. Harmal Extract Induces Apoptosis of HCT116 Human Colon Cancer Cells, Mediated by Inhibition of Nuclear Factor-κB and Activator Protein-1 Signaling Pathways and Induction of Cytoprotective Genes. Asian Pacific J. Cancer Prev. 2016, 17, 1947–1959. [Google Scholar] [CrossRef] [Green Version]
  33. Khadraoui, N.; Essid, R.; Jallouli, S.; Damergi, B.; Ben Takfa, I.; Abid, G.; Jedidi, I.; Bachali, A.; Ayed, A.; Limam, F.; et al. Antibacterial and antibiofilm activity of Peganum harmala seed extract against multidrug-resistant Pseudomonas aeruginosa pathogenic isolates and molecular mechanism of action. Arch. Microbiol. 2022, 204, 133. [Google Scholar] [CrossRef]
  34. Shaheen, H.A.; Issa, M.Y. In vitro and in vivo activity of Peganum harmala L. alkaloids against phytopathogenic bacteria. Sci. Hortic. 2020, 264, 108940. [Google Scholar] [CrossRef]
  35. Zheng, X.-Y.; Zhang, Z.-J.; Chou, G.-X.; Wu, T.; Cheng, X.-M.; Wang, C.-H.; Wang, Z.-T. Acetylcholinesterase inhibitive activity-guided isolation of two new alkaloids from seeds of Peganum nigellastrum Bunge by an in vitro TLC- bioautographic assay. Arch. Pharm. Res. 2009, 32, 1245–1251. [Google Scholar] [CrossRef]
  36. Liu, W.; Wang, Y.; He, D.-D.; Li, S.-P.; Zhu, Y.-D.; Jiang, B.; Cheng, X.-M.; Wang, Z.-T.; Wang, C.-H. Antitussive, expectorant, and bronchodilating effects of quinazoline alkaloids (±)-vasicine, deoxyvasicine, and (±)-vasicinone from aerial parts of Peganum harmala L. Phytomedicine 2015, 22, 1088–1095. [Google Scholar] [CrossRef]
  37. Adeel, S.; Rehman, F.-U.; Zuber, M.; Batool, F.; Habib, N.; Hussaan, M.; Amin, N. Environmental friendly application of ultrasonic rays for extraction of natural colorant from Harmal (P. harmala) for dyeing of bio-mordanted silk. J. Eng. Fiber. Fabr. 2021, 16, 155892502110638. [Google Scholar] [CrossRef]
  38. Kokalj, A. Molecular modeling of organic corrosion inhibitors: Calculations, pitfalls, and conceptualization of molecule–surface bonding. Corros. Sci. 2021, 193, 109650. [Google Scholar] [CrossRef]
  39. 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] [PubMed]
  40. El-Faham, A.; Dahlous, K.A.; Al Othman, Z.A.; Al-Lohedan, H.A.; El-Mahdy, G.A. sym-Trisubstituted 1,3,5-Triazine Derivatives as Promising Organic Corrosion Inhibitors for Steel in Acidic Solution. Molecules 2016, 21, 436. [Google Scholar] [CrossRef] [PubMed]
  41. Al Otaibi, N.; Hammud, H.H. Corrosion Inhibition Using Harmal Leaf Extract as an Eco-Friendly Corrosion Inhibitor. Molecules 2021, 26, 7024. [Google Scholar] [CrossRef]
  42. Ali, A.I.; Mahrous, Y.S. Corrosion inhibition of C-steel in acidic media from fruiting bodies of Melia azedarach L. extract and a synergistic Ni 2+ additive. RSC Adv. 2017, 7, 23687–23698. [Google Scholar] [CrossRef] [Green Version]
  43. Parr, R.G.; Weitao, Y. Density-Functional Theory of Atoms and Molecules; Oxford University Press: Oxford, UK, 1989. [Google Scholar] [CrossRef]
  44. Frisch, M.J.; Trucks, G.W.; Schlegel, H.B.; Scuseria, G.E.; Robb, M.A.; Cheeseman, J.R.; Scalmani, G.; Barone, V.; Petersson, G.A.; Nakatsuji, H. Gaussian; Gaussian, Inc.: Wallingford, CT, USA, 2016; Volume 16. [Google Scholar]
  45. Dennington, R.; Keith, T.; Millam, J. GaussView, Version 5; Semichem Inc.: Shawnee Mission, KS, USA, 2009. [Google Scholar]
  46. Becke, A.D. Density-functional thermochemistry. III. The role of exact exchange. J. Chem. Phys. 1993, 98, 5648–5652. [Google Scholar] [CrossRef] [Green Version]
  47. Becke, A.D. A new mixing of Hartree–Fock and local density-functional theories. J. Chem. Phys. 1993, 98, 1372–1377. [Google Scholar] [CrossRef]
  48. Lee, C.; Yang, W.; Parr, R.G. Development of the Colle-Salvetti correlation-energy formula into a functional of the electron density. Phys. Rev. B 1988, 37, 785–789. [Google Scholar] [CrossRef] [Green Version]
  49. Obot, I.; Obi-Egbedi, N. Theoretical study of benzimidazole and its derivatives and their potential activity as corrosion inhibitors. Corros. Sci. 2010, 52, 657–660. [Google Scholar] [CrossRef]
  50. Mennucci, B.; Cancès, E.; Tomasi, J. Evaluation of Solvent Effects in Isotropic and Anisotropic Dielectrics and in Ionic Solutions with a Unified Integral Equation Method: Theoretical Bases, Computational Implementation, and Numerical Applications. J. Phys. Chem. B 1997, 101, 10506–10517. [Google Scholar] [CrossRef]
  51. Tomasi, J.; Mennucci, B.; Cammi, R. Quantum Mechanical Continuum Solvation Models. Chem. Rev. 2005, 105, 2999–3094. [Google Scholar] [CrossRef] [PubMed]
  52. Faiz, M.; Zahari, A.; Awang, K.; Hussin, H. Corrosion inhibition on mild steel in 1 M HCl solution by Cryptocarya nigra extracts and three of its constituents (alkaloids). RSC Adv. 2020, 10, 6547–6562. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, J.T.; Hu, J.M.; Zhang, J.Q.; Cao, C.N. Studies of impedance models and water transport behaviors of polypropylene coated metals in NaCl solution. Prog. Org. Coatings 2004, 49, 293–301. [Google Scholar] [CrossRef]
  54. Meng, F.; Liu, L.; Tian, W.; Wu, H.; Li, Y.; Zhang, T.; Wang, F. The influence of the chemically bonded interface between fillers and binder on the failure behaviour of an epoxy coating under marine alternating hydrostatic pressure. Corros. Sci. 2015, 101, 139–154. [Google Scholar] [CrossRef]
  55. Jüttner, K. Electrochemical impedance spectroscopy (EIS) of corrosion processes on inhomogeneous surfaces. Electrochim. Acta. 1990, 35, 1501–1508. [Google Scholar] [CrossRef]
  56. Ji, G.; Anjum, S.; Sundaram, S.; Prakash, R. Musa paradisica peel extract as green corrosion inhibitor for mild steel in HCl solution. Corros. Sci. 2015, 90, 107–117. [Google Scholar] [CrossRef]
  57. Oguzie, E.E.; Adindu, C.B.; Enenebeaku, C.K.; Ogukwe, C.E.; Chidiebere, M.A.; Oguzie, K.L. Natural Products for Materials Protection: Mechanism of Corrosion Inhibition of Mild Steel by Acid Extracts of Piper guineense. J. Phys. Chem. C. 2012, 116, 13603–13615. [Google Scholar] [CrossRef]
  58. Mohammed, N.; Othman, N.; Taib, M.; Samat, M.; Yahya, S. Experimental and Theoretical Studies on Extract of Date Palm Seed as a Green Anti-Corrosion Agent in Hydrochloric Acid Solution. Molecules 2021, 26, 3535. [Google Scholar] [CrossRef]
  59. Kim, J.-Y.; Shin, I.; Byeon, J.-W. Corrosion Inhibition of Mild Steel and 304 Stainless Steel in 1 M Hydrochloric Acid Solution by Tea Tree Extract and Its Main Constituents. Materials 2021, 14, 5016. [Google Scholar] [CrossRef]
  60. Qiang, Y.; Zhang, S.; Tan, B.; Chen, S. Evaluation of Ginkgo leaf extract as an eco-friendly corrosion inhibitor of X70 steel in HCl solution. Corros. Sci. 2018, 133, 6–16. [Google Scholar] [CrossRef]
  61. Li, H.; Qiang, Y.; Zhao, W.; Zhang, S. A green Brassica oleracea L. extract as a novel corrosion inhibitor for Q235 steel in two typical acid media Colloids and Surfaces A. Physicochem. Eng. Asp. 2021, 616, 126077. [Google Scholar] [CrossRef]
  62. Marahatta, A.; Marahatta, A.B. A DFT Analysis for the Electronic Structure, Mulliken Charges Distribution and Frontier Molecular Orbitals of Monolayer Graphene Sheet. Int. J. Progress. Sci. Technol. 2019, 16, 51–65. [Google Scholar]
  63. Efil, K.; Obot, I.B. Quantum Chemical Investigation of the Relationship Between Molecular Structure and Corrosion Inhibition Efficiency of Benzotriazole and its Alkyl-Derivatives on Iron. Prot. Met. Phys. Chem. Surfaces 2017, 53, 1139–1149. [Google Scholar] [CrossRef]
  64. Harismah, K.; Dhumad, A.M.; Ibraheem, H.S.; Zandi, H.; Majeed, H.J. A DFT approach on tioguanine: Exploring tio-tiol tautomers, frontier molecular orbitals, IR and UV spectra, and quadrupole coupling constants. J. Mol. Liq. 2021, 334, 116018. [Google Scholar] [CrossRef]
  65. Benhiba, F.; Hsissou, R.; Abderrahim, K.; Serrar, H.; Rouifi, Z.; Boukhris, S.; Kaichouh, G.; Bellaouchou, A.; Guenbour, A.; Oudda, H.; et al. Development of New Pyrimidine Derivative Inhibitor for Mild Steel Corrosion in Acid Medium. J. Bio- Tribo-Corros. 2022, 8, 36. [Google Scholar] [CrossRef]
  66. Nazir, U.; Akhter, Z.; Ali, N.Z.; Hussain, R.; Liaqat, F.; Tahir, A.; Qamar, S. Corrosion inhibition studies of ferrocenyl Schiff bases in a mild acidic medium through experimental methods and DFT calculations. New J. Chem. 2022, 46, 3925–3938. [Google Scholar] [CrossRef]
  67. Oyeneyin, O.E.; Ojo, N.D.; Ipinloju, N.; James, A.C.; Agbaffa, E.B. Investigation of Corrosion Inhibition Potentials of Some Aminopyridine Schiff Bases Using Density Functional Theory and Monte Carlo Simulation. Chem. Africa 2022, 5, 319–332. [Google Scholar] [CrossRef]
  68. Suresh, C.H.; Remya, G.S.; Anjalikrishna, P.K. Molecular electrostatic potential analysis: A powerful tool to interpret and predict chemical reactivity. WIREs Comput. Mol. Sci. 2022, 12, e1601. [Google Scholar] [CrossRef]
  69. El-Hendawy, M.M.; Kamel, A.M.; Mohamed, M.M.A. The anti-corrosive behavior of benzo-fused N-heterocycles: An in silico study towards developing organic corrosion inhibitors. Phy. Chem. Chem. Phy. 2022, 24, 743. [Google Scholar] [CrossRef]
Figure 1. (a) Key bioactive indole-, quinazoline- and quinazolinone-based alkaloids present in the extract of the harmal plant and (b) generalized selected uses of the harmal extract [21].
Figure 1. (a) Key bioactive indole-, quinazoline- and quinazolinone-based alkaloids present in the extract of the harmal plant and (b) generalized selected uses of the harmal extract [21].
Molecules 27 07250 g001
Figure 2. Equivalent circuits (a) Model 1 and (b) Model 2 used to fit the impedance data of the Harmal extracts.
Figure 2. Equivalent circuits (a) Model 1 and (b) Model 2 used to fit the impedance data of the Harmal extracts.
Molecules 27 07250 g002
Figure 3. Nyquist plots for all concentrations of Harmal roots extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Figure 3. Nyquist plots for all concentrations of Harmal roots extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Molecules 27 07250 g003
Figure 4. Nyquist plots for all concentrations of Harmal leaves extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Figure 4. Nyquist plots for all concentrations of Harmal leaves extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Molecules 27 07250 g004
Figure 5. Nyquist plots for all concentrations of Harmal flower extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Figure 5. Nyquist plots for all concentrations of Harmal flower extract inhibitors in 0.25 M H2SO4 at 298 K fitted by (a) Model 1 and (b) Model 2.
Molecules 27 07250 g005
Figure 6. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of Harmal roots extract.
Figure 6. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of Harmal roots extract.
Molecules 27 07250 g006
Figure 7. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of leaf extract.
Figure 7. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of leaf extract.
Molecules 27 07250 g007
Figure 8. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of Harmal flower extract.
Figure 8. Bode phase (a) and Bode (b) plots of a C-steel electrode in 0.25 M H2SO4 solution with and without different concentrations of Harmal flower extract.
Molecules 27 07250 g008
Figure 9. Tafel plots for all concentrations of (a) Harmal roots, (b) Harmal leaves, and (c) Harmal flowers extract inhibitors in 0.25M H2SO4 at 298 K.
Figure 9. Tafel plots for all concentrations of (a) Harmal roots, (b) Harmal leaves, and (c) Harmal flowers extract inhibitors in 0.25M H2SO4 at 298 K.
Molecules 27 07250 g009aMolecules 27 07250 g009b
Figure 10. Langmuir adsorption isotherm of Harmal (a) roots, (b) leaves, and (c) flower extract, and (d) Frumkin adsorption isotherm of Harmal leaves extract; for all concentrations on carbon steel electrode in 0.25M H2SO4 at 298 K.
Figure 10. Langmuir adsorption isotherm of Harmal (a) roots, (b) leaves, and (c) flower extract, and (d) Frumkin adsorption isotherm of Harmal leaves extract; for all concentrations on carbon steel electrode in 0.25M H2SO4 at 298 K.
Molecules 27 07250 g010aMolecules 27 07250 g010bMolecules 27 07250 g010c
Figure 11. Inhibition mechanism based on chemical adsorption of the Harmal molecules from Harmal plant extracts HRE, HLE, and HFE. (a) σ bonds and (b) π back bonds are shown for the coordinate bonds between the Harmal molecules and Fe or Fe2+ on the C-steel surface.
Figure 11. Inhibition mechanism based on chemical adsorption of the Harmal molecules from Harmal plant extracts HRE, HLE, and HFE. (a) σ bonds and (b) π back bonds are shown for the coordinate bonds between the Harmal molecules and Fe or Fe2+ on the C-steel surface.
Molecules 27 07250 g011
Figure 12. FTIR spectra of the extracts of Harmal roots (HRE), Harmal leaves (HLE), and Harmal flowers (HFE).
Figure 12. FTIR spectra of the extracts of Harmal roots (HRE), Harmal leaves (HLE), and Harmal flowers (HFE).
Molecules 27 07250 g012
Figure 13. SEM images of C-steel immersed for 3 h in (a) free 0.25 M H2SO4 and (b) 0.25M H2SO4 containing 300 ppm of Harmal leaf extract (HLE).
Figure 13. SEM images of C-steel immersed for 3 h in (a) free 0.25 M H2SO4 and (b) 0.25M H2SO4 containing 300 ppm of Harmal leaf extract (HLE).
Molecules 27 07250 g013aMolecules 27 07250 g013b
Figure 14. Optimized structures of the stable conformers for the neutral form of organic molecules 18 present in the extract of Harmal leaves computed at the B3LYP/6-311G++(d,p) in water using the PCM solvation model [49].
Figure 14. Optimized structures of the stable conformers for the neutral form of organic molecules 18 present in the extract of Harmal leaves computed at the B3LYP/6-311G++(d,p) in water using the PCM solvation model [49].
Molecules 27 07250 g014aMolecules 27 07250 g014b
Figure 15. Frontier molecular orbitals (HOMO and LUMO) and electrostatic potential (ESP) maps of the stable conformers for the neutral form of compounds 18 by B3LYP method with 6-311G++(d,p) basis set using water as the solvent [49].
Figure 15. Frontier molecular orbitals (HOMO and LUMO) and electrostatic potential (ESP) maps of the stable conformers for the neutral form of compounds 18 by B3LYP method with 6-311G++(d,p) basis set using water as the solvent [49].
Molecules 27 07250 g015aMolecules 27 07250 g015b
Figure 16. Comparative representation of EHOMO and ELUMO for the neutral and protonated forms of studied compounds 18 calculated in the gaseous and solvent phases using water and by applying PCM model at the B3LYP/6-311G++(d,p).
Figure 16. Comparative representation of EHOMO and ELUMO for the neutral and protonated forms of studied compounds 18 calculated in the gaseous and solvent phases using water and by applying PCM model at the B3LYP/6-311G++(d,p).
Molecules 27 07250 g016
Table 1. Impedance parameters for the Harmal roots extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Table 1. Impedance parameters for the Harmal roots extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Harmal RootsRsRfnfZCPEfRctnZCPECdlRf + RctθIE%
Model 1Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2 µΩ−1s n cm−2μF/cm2Ω cm2
0 ppm6.639.230.6592936.149.940.9764.30 × 102375.6619.170.0000.00
20.8 ppm6.6864.250.924131.3715.530.6148.76 × 10258.6279.780.76075.97
29.1 ppm5.2755.850.923171.5371.030.7381.87 × 10240.25126.880.84984.89
41.5 ppm4.9485.110.937131.41123.010.7611.37 × 10237.78208.120.90890.79
62.1 ppm5.87222.140.87867.8317.370.5891.65 × 103138.10239.520.92092.00
82.6 ppm7.3722.480.80062.44319.370.8404.42 × 10119.29341.840.94494.39
Table 2. Impedance parameters for the Harmal roots extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Table 2. Impedance parameters for the Harmal roots extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Harmal RootsRsRfnfZCPEfRctCdlRf + RctθIE%W
Model 2Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2μF/cm2Ω cm2 Ss1/2
0 ppm6.5815.110.902430.084.1230475.4319.230.0000.000.00873
20.8 ppm6.4228.591.000305.3550.9291.2579.440.75875.790.00127
29.1 ppm5.17118.790.824105.798.4682.38127.250.84984.890.00296
41.5 ppm4.80119.490.81595.8687.22128.09206.710.90790.700.00325
62.1 ppm5.73210.780.90070.0627.42599.36238.200.91991.930.00178
82.6 ppm4.91131.960.8844.86176.1557.11308.100.93893.760.00254
Table 3. Impedance parameters for the Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Table 3. Impedance parameters for the Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Harmal LeavesRsRfnfZCPEfRctnZCPECdlRf + RctθIE%
Model 1Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2 µΩ−1s n cm−2μF/cm2Ω cm2
0 ppm6.69.230.65882936.149.940.9764.30 × 102375.6619.170.0000.00
20.8 ppm6.222.640.8894265.762.450.6133.18 × 103147.4825.100.23623.61
29.1 ppm6.066.260.8968123.877.480.6321.55 × 103115.5473.730.74074.00
41.5 ppm5.965.650.660699.52147.840.6981.87 × 10239.54213.500.91091.02
62.1 ppm5.5164.280.951553.0117.230.6668.83 × 102107.89181.510.89489.44
82.6 ppm4.5333.090.887547.050.900.2655.49 × 10412.80333.990.94394.26
Table 4. Impedance parameters for the Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Table 4. Impedance parameters for the Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Harmal LeavesRsRfnfZCPEfRctCdlRf + RctθIE%W
Model 2Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2μF/cm2Ω cm2 Ss1/2
0 ppm6.615.110.9024430.084.1230,475.4319.230.0000.008.73 × 10−3
20.8 ppm6.12.101.0000559.1822.92185.2025.010.23123.113.95 × 10−3
29.1 ppm5.841.160.9701196.0032.1384.7673.290.73873.762.27 × 10−3
41.5 ppm6.345.080.9751151.3839.6443.8684.720.77377.301.10 × 10−3
62.1 ppm5.3132.161.000063.8848.5162.88180.670.89489.362.06 × 10−3
82.6 ppm4.7143.770.892468.09186.1571.61329.920.94294.175.10 × 10−3
Table 5. Impedance parameters for the Harmal flower extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Table 5. Impedance parameters for the Harmal flower extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 1.
Harmal FlowersRsRfnfZCPEfRctnZCPECdlRf + RctθIE%
Model 1Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2 µΩ−1s n cm−2μF/cm2Ω cm2
0 ppm6.639.230.6592936.149.940.9764.30 × 102375.6619.170.000−0.01
20.8 ppm6.3832.730.937238.7120.290.7345.72 × 102114.1553.020.63863.85
29.1 ppm6.2955.650.895154.822.990.5384.73 × 103121.9058.640.67367.31
41.5 ppm4.59102.300.876140.744.040.5424.56 × 103155.72106.340.82081.97
62.1 ppm6.09126.380.83286.062.500.5862.05 × 10349.32128.880.85185.13
82.6 ppm4.903.890.5404887.61166.590.876.94 × 1011.56170.490.88888.76
Table 6. Impedance parameters for the Harmal flower extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Table 6. Impedance parameters for the Harmal flower extract on the C-steel in 0.25 M H2SO4 medium by fitting the equivalent circuit model 2.
Harmal FlowersRsRfnfZCPEfRctCdlRf + RctθIE%W
Model 2Ω cm2Ω cm2 µΩ−1s n cm−2Ω cm2μF/cm2Ω cm2 Ss1/2
0 ppm6.5815.110.902430.084.1230,475.4319.230.0000.008.73 × 10−3
20.8 ppm6.2339.070.924229.1613.83127.9752.890.63663.643.61 × 10−3
29.1 ppm6.2515.200.954277.1043.04163.8258.230.67066.983.01 × 10−3
41.5 ppm4.5443.970.884239.3161.53190.31105.500.81881.772.98 × 10−3
62.1 ppm5.9184.910.918115.5642.7749.33127.680.84984.941.49 × 10−3
82.6 ppm4.8275.360.8799.8293.5597.23168.910.88688.622.77 × 10−3
Table 7. Phase angles and alpha values (slopes) from the Bode phase and Bode plots for different types of Harmal extract at various concentrations.
Table 7. Phase angles and alpha values (slopes) from the Bode phase and Bode plots for different types of Harmal extract at various concentrations.
Harmal Extract
0 ppm20.8 ppm29.1 ppm41.5 ppm62.1 ppm82.6 ppm
Harmal RootsPhase angle °−27.87−41.30−49.06−56.00−57.16−62.52
Frequency Hz79.00125.60252.40252.40158.40315.50
Slope α−0.3635−0.4804−0.5491−0.6417−0.6619−0.7190
R20.96460.99550.99730.99820.99990.9968
Harmal FlowersPhase angle °−27.87−40.23−44.08−53.23−50.23−57.58
Frequency Hz75.00125.60125.60125.60252.40198.60
Slope α−0.3635−0.4945−0.5492−0.6324−0.5995−0.6571
R20.94920.98750.97760.98860.96520.9964
Harmal LeavesPhase angle °−27.87−31.78−45.74−41.88−59.68−66.28
Frequency Hz79.00125.60125.60252.40158.40198.60
Slope α−0.3635−0.3995−0.5557−0.4851−0.7075−0.7671
R20.98920.99580.99690.99300.99820.9984
Table 8. Polarization parameters for various concentrations of Harmal roots extract on the C-steel in 0.25 M H2SO4 medium.
Table 8. Polarization parameters for various concentrations of Harmal roots extract on the C-steel in 0.25 M H2SO4 medium.
Harmal RootsIcorr
μA
Ecorr
mV
βa
V/Decade
βc
V/Decade
Corrosion Rate mpyχ2Icorr μA/cm2θIE%
0 ppm1710.0−397.00.41280.49701554.006.5323401.630.0000.0
20.8 ppm585.0−444.00.26810.3553532.0021.511163.720.65865.8
29.1 ppm206.6−417.00.20660.3011187.4033.64410.980.87987.9
41.5 ppm102.2−432.00.18080.236992.4740.54203.300.94094.0
62.1 ppm102.0−443.00.18860.233392.7338.78202.900.94094.0
82.6 ppm69.7−446.00.16870.224463.3443.41138.650.95995.9
Table 9. Polarization parameters for various concentrations of Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium.
Table 9. Polarization parameters for various concentrations of Harmal leaf extract on the C-steel in 0.25 M H2SO4 medium.
Harmal LeavesIcorr
μA
Ecorr
mV
βa
V/Decade
βc
V/Decade
Corrosion Rate mpyχ2Icorr μA/cm2θIE%
0 ppm1710.0−397.00.41280.49701554.06.5323401.630.0000.0
20.8 ppm1550.0−457.00.35390.44471412.011.013083.350.0949.4
29.1 ppm567.0−449.00.26690.3593515.722.631127.910.66866.8
41.5 ppm338.0−462.00.24230.3060307.230.61672.370.80280.2
62.1 ppm180.0−436.00.21860.2565163.738.45358.070.89589.5
82.6 ppm94.7−452.00.18230.213686.151.29188.380.94594.5
Table 10. Polarization parameters for various concentrations of Harmal flower extract on the C-steel in 0.25 M H2SO4 medium.
Table 10. Polarization parameters for various concentrations of Harmal flower extract on the C-steel in 0.25 M H2SO4 medium.
Harmal FlowersIcorr
μA
Ecorr
mV
βa
V/Decade
βc
V/Decade
Corrosion Rate mpyχ2Icorr μA/cm2θIE%
0 ppm1710.0−397.00.41280.49701554.006.5323401.630.0000.0
20.8 ppm767.0−451.00.28970.3984696.9016.571525.760.55155.1
29.1 ppm653.0−444.00.26840.3540593.8023.171298.990.61861.8
41.5 ppm344.0−425.00.22390.2850313.0033.80684.300.79979.9
62.1 ppm261.0−464.00.22810.2783237.1028.18519.200.84784.7
82.6 ppm199.0−460.00.20070.2584180.7037.51395.860.88488.4
Table 11. Percent inhibition efficiency (IE%) toward steel corrosion in sulfuric acid by Harmal roots, leaves, and flower extracts compared with other reported systems at various concentrations (ppm) [58,59,60,61].
Table 11. Percent inhibition efficiency (IE%) toward steel corrosion in sulfuric acid by Harmal roots, leaves, and flower extracts compared with other reported systems at various concentrations (ppm) [58,59,60,61].
Harmal
(ppm)
This Work
HRE
IE%
EIS
0.25 M
H2SO4
C-Steel
HLE
IE%
EIS
0.25 M
H2SO4
C-Steel
HFE
IE%
EIS
0.25 M
H2SO4
C-Steel
Date
Palm
Seed
(ppm)
IE%
WLS
0.5 M HCl
C-Steel
Tea
Tree
(ppm)
IE%
PDP
1 M HCl
Mild Steel
Ginkgo Leaf Extract
(ppm)
IE%
PDP
1 M HCl X-70 Steel
Brassica Oleracea L
(ppm)
IE%
PDP
0.50 M
H2SO4
Q235
Steel
20.865.89.455.18007115056.52562.05081.2
29.187.966.861.89008230064.05070.810086.0
41.594.080.279.912008875075.610086.320090.0
62.194.089.584.7140095225078.620090.030092.7
82.695.994.588.4200091
This work[58][59][60][61]
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Hammud, H.H.; Maache, S.A.; Al Otaibi, N.; Sheikh, N.S. An Integrated Experimental and Theoretical Studies on the Corrosion Inhibition of Carbon Steel by Harmal Extracts. Molecules 2022, 27, 7250. https://doi.org/10.3390/molecules27217250

AMA Style

Hammud HH, Maache SA, Al Otaibi N, Sheikh NS. An Integrated Experimental and Theoretical Studies on the Corrosion Inhibition of Carbon Steel by Harmal Extracts. Molecules. 2022; 27(21):7250. https://doi.org/10.3390/molecules27217250

Chicago/Turabian Style

Hammud, Hassan H., Sarah A. Maache, Nasreen Al Otaibi, and Nadeem S. Sheikh. 2022. "An Integrated Experimental and Theoretical Studies on the Corrosion Inhibition of Carbon Steel by Harmal Extracts" Molecules 27, no. 21: 7250. https://doi.org/10.3390/molecules27217250

Article Metrics

Back to TopTop