Mahmoud A. Al-Qudah*a,
Tareq T. Bataineha,
Faten M. Abu Orabib,
Sultan T. Abu-Orabic,
Ghassab M. Al-Mazaidehd and
Abbas I. Alakhrase
aDepartment of Chemistry, Faculty of Science, Yarmouk University, P.O. Box 566, Irbid 21163, Jordan. E-mail: mahmoud.qudah@yu.edu.jo; Fax: +96227211117; Tel: +96277420029
bFaculty of Arts and Sciences, The World Islamic Sciences and Education University, Amman, Jordan. E-mail: faten.aladwan@wise.edu.jo
cDepartment of Medical Analysis, Faculty of Science, Tishk International University, Erbil, KRG, Iraq
dDepartment of Pharmaceutical Chemistry, College of Pharmacy, University of Hafr Al Batin, P. O. Box: 1803, Hafr Al Batin 31991, Saudi Arabia. E-mail: gmazaideh@uhb.edu.sa
eDepartment of Chemistry, College of Science, Imam Mohammad Ibn Saud Islamic University (IMSIU), Riyadh, 11623, Saudi Arabia
First published on 27th March 2025
Acteoside (ACT) isolated from A. orientalis L. was investigated as a corrosion inhibitor of copper in 1.0 M HNO3 acid solutions using conventional weight loss, electrochemical polarization, and electrochemical impedance spectroscopy (EIS) studies. The molecular structure of acteoside (ACT) is supported by all the experimental results from LC-MS, FT-IR, 1H, and 13C-NMR. The findings indicate that ACT is a potent inhibitor, and that its effectiveness increases with both temperature and inhibitor concentration. The highest inhibitor concentration occurs at 48 °C, and the inhibition efficiency peaks at 98.8%. The results indicated that the presence of the inhibitor slightly lessened copper dissolution with rising temperature. ACT adheres to the Langmuir, Freundlich, El-Awady, Temkim, and Redlich–Peterson (R–P) adsorption isotherms on copper surfaces with a high regression coefficient value. The values of the activation parameters (Ea, ΔH*, ΔS*) and the adsorption thermodynamic functions (ΔGads) suggest both physisorption and chemisorption processes. Electrochemical polarization data were used to identify the mixed mode of inhibition. Active molecules adhere to the metal surface and form a protective layer, which causes changes in impedance characteristics, charge transfer resistance, and double layer capacitance in response to variations in ACT concentration. Weight loss and electrochemical data are supported by quantum chemical computations.
Natural products serve as copper corrosion inhibitors due to their environmental friendliness, affordability, and the presence of a variety of organic compounds, some of which have demonstrated their effectiveness as metal corrosion inhibitors by containing heteroatoms like O, N, S, and P.3–10 The use of extracts and isolated compounds from natural sources as corrosion inhibitors for metals is interesting. Previous studies have used natural extracts and compounds as corrosion inhibitors and demonstrated high effectiveness in preventing metal corrosion. Table 1 shows a comparison of the compound used in the study with other inhibitors, including organic inhibitors, plant extracts, and isolated natural compounds.11–20
Inhibitors | [Inhibitors] | IE (%) | Note | Reference |
---|---|---|---|---|
Acteoside | 100–500 ppm | 84.3–98.8 | The findings indicate that ACT is a potent inhibitor, and that its effectiveness increases with both temperature and inhibitor concentration | |
3-Oxocostusic acid | 4.0 × 10−4–16.0 × 10−4 mol L−1 | Up to 95.60 | The inhibition efficiency of 3-oxocostusic acid is high, it means that it effectively prevents copper from corroding in nitric acid | 21 |
Alkaloids (quinine, nicotine) | 10−3 M | 50–85 | These are nitrogen-containing heterocycles that form strong interactions with metal surfaces, providing inhibition | 22 |
Adhatoda vasica leaf extract | 0.01 to 0.1 g L−1 | 73–78 | Spontaneous, governed by physiochemical processes, and occurred according to the Langmuir's adsorption isotherm | 23 |
Vitex negundo leaf extract | 0.01 to 0.1 g L−1 | 93–98 | Spontaneous, governed by physiochemical processes, and occurred according to the Langmuir's adsorption isotherm | 23 |
Saraca asoca leaf extract | 0.01 to 0.1 g L−1 | 53–91 | Spontaneous, governed by physiochemical processes, and occurred according to the Langmuir's adsorption isotherm | 23 |
Thymus satureioides essential oil | 1200–1600 ppm | 69.72–89.04 | The efficiency increases with the inhibitor concentration | 24 |
Rosmarinus officinalis extract | 300 ppm | 77.0 | The findings showed that when temperature rises, physisorption increases and inhibitory efficiency (%IE) decreases. The adsorption mechanism and Langmuir's adsorption model agreed | 25 |
Amino acid (proline (Pro), phenylalanine (Phe), tyrosine (Tyr), and tryptophan (Try)) | 10−3 M | 35–90 | These amino acids can act as corrosion inhibitors by forming protective layers on the surface of copper, thereby reducing the corrosion rate when exposed to aggressive environments like nitric acid | 26 |
Amino acids (histidine, glycine) | 10−3–10−2 M | 55–80 | Slightly less efficient (55–80%) but have the advantage of being eco-friendly and biodegradable | 27 |
Benzotriazole (BTA) | 10−3–10−2 M | 80–95 | The highest inhibition efficiency (80–95%) due to its strong adsorption and formation of a protective film on the copper surface | 28 |
2-Mercaptobenzothiazole (MBT) | 10−3 M | 70–90 | Strong adsorption on Cu surface | 29 |
Schiff bases | 10−3–10−2 M | 60–85 | Moderately effective (60–85%) due to their chelation properties, though their efficiency depends on their molecular structure | 30 |
Imidazole derivatives | 10−3–10−2 M | 75–92 | Exhibit a similar range (75–92%) as BTA, benefiting from π-electron interactions that improve adsorption | 31 |
To increase the efficacy of possible inhibitors, researchers have recently employed quantum chemistry techniques in corrosion inhibitor experiments.32–34 In this area, the combination of computer science and density functional theory (DFT) has produced useful instruments for examining the molecules of natural products. When compared to conventional post-Hartree–Fock (HF) approaches, the DFT methodology provides a better treatment of electron interactions. DFT has been widely utilized by researchers to quantify several aspects of substances, including their kinetics, thermochemistry, and chemical structure. These methods advance our knowledge of corrosion inhibition and could lead to the creation of future inhibitors that are more effective and long-lasting.
The isolation of natural compounds from plant extracts and analysis of their anticorrosion activity for metals is a new trend. The current study used weight loss, electrochemical polarization, EIS, and quantum chemical calculations techniques to investigate the inhibition of copper corrosion by ACT in a 1.0 M HNO3 solution.
The aqueous methanol extract (75 g) was adsorbed on 75 g of silica gel S and subjected to column chromatography (40.0 × 6.0 cm) using 500 g of the same absorbent. Hexane was used to fill the column and to elute the sample. Then, ethyl acetate was added, gradually raising the polarity. According to their TLC behavior, a total of 60 fractions obtained using this method (each 500 mL) were divided into 4 major groups (AMPM I–AMPM IV). Each of these groups underwent CC and TLC purification or treatment with an appropriate solvent.
IR (KBr) νmax cm−1: 3335, 3075, 1661, 1629, 1548, 1520, 1436, 1349. HRESIMS m/z 623.19864 [M − H] (calcd for [C29H36O15] 624.2054) 1H-NMR (MeOH-d4) δ 1.11(3H, d, J = 6.0 Hz, H6′′), 2.81(2H, m, H7), 3.32(1H, t, J = 9.2 Hz, H4′′), 3.41(1H, m, H2′′′), 3.53(1H, m, H5′′′), 3.56 (1H, m, H5′′), 3.59 (1H, m, H3′′), 3.62 (2H, m, H6′′′), 3.64 (1H, t, J = 8.8 Hz, H3′′′), 3.84 (1H, m, H2′′), 3.96 (2H, m, H8), 4.40(1H, d, J = 7.9 Hz, H1′′′), 4.91 (1H, bs, H4′′′), 5.21 (1H, d, J = 1.4 Hz, H1′′), 6.30 (1H, d, J = 15.9 Hz, H7′), 6.59 (1H, d, J = 7.8, 2.0 Hz, H6), 6.69 (1H, d, J = 7.8 Hz, H5), 6.71 (1H, d, J = 2.0 Hz, H2), 6.82 (1H, d, J = 8.0 Hz, H5′), 6.98 (1H, d, J = 7.8, 2.0 Hz, H6′), 7.09 (1H, d, J = 2.0 Hz, H2′), 7.61 (1H, d, J = 15.8 Hz, H8′). 13C-NMR, (MeOH-d4), δ 18.6 (C6′′), 36.5 (C7), 62.4 (C6′′′), 70.5 (C5′′), 70.9 (C4′′′), 71.0 (C3′′), 72.3 (C8), 72.4 (C2′′), 74.5 (C4′′), 74.8 (C2′′′), 78.1 (C5′′′), 80.4 (C3′′′), 101.7 (C1′′), 102.7 (C1′′′), 114.7 (C7′), 115.3 (C2′), 116.4 (C2), 116.6 (C5′), 117.2 (C5), 121.4 (C6), 123.4 (C6′), 126.3 (C1′), 130.1 (C1), 143.2 (C4), 144.7 (C3), 145.4 (C3′), 146.7 (C4′), 148.4 (C8′), 167.0 (C9′).
The quantum chemical parameters calculated encompassed various descriptors, including the energy gap (ΔEgap) between the highest occupied molecular orbital energy (EHOMO) and the lowest unoccupied molecular orbital energy (ELUMO), the fraction of electron transferred (ΔN), global softness (σ), absolute hardness (η), electrophilicity index (ω), and chemical potential (χ), among others.33,34 These computed parameters provided significant insights into the electronic configuration and chemical stability of the compounds under investigation.
The comprehensive computational methods utilized in this study offer a deeper understanding of the chemical characteristics of the compounds under scrutiny and their potential applications. Furthermore, these computational details help elucidate the properties and possible reactivity of the substances under investigation.
![]() | (1) |
![]() | ||
Fig. 2 Weight loss–time curves for copper corrosion in 1.0 M HNO3 at 28 °C with and without varying ACT concentrations. |
The calculated inhibition efficiency (%IE) values are shown in Table 2. From these tables, the %IE rises steadily as the inhibitor concentration rises and as the temperature rises from 28 to 48 °C. A maximum inhibition efficiency of 98.8% for 500 mg L−1 of ACT in acidic medium was noted at 48 °C. The compound's increased coverage of the metal surface area is most likely the reason for this behavior. It is widely known from the literature that a decrease in inhibitory efficacy with rising temperature frequently indicates the establishment of an electrostatic or physical adsorption layer. When an inhibitor is present, the opposing action (in our case) enhances inhibition efficiency as temperature rises, indicating a chemisorption process.4–9
[ACT] (ppm) | Weight loss (mg) | Corrosion rate (mg h−1 cm−2) | Inhibition efficiency (%IE) | ||||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|---|
301 K | 306 K | 311 K | 316 K | 321 K | 301 K | 306 K | 311 K | 316 K | 321 K | 301 K | 306 K | 311 K | 316 K | 321 K | |
0 | 11.86 | 41.3 | 60.76 | 115 | 158.86 | 0.494 | 1.721 | 2.532 | 4.792 | 6.619 | |||||
100 | 1.86 | 2.4 | 2.7 | 2.8 | 3.03 | 0.078 | 0.083 | 0.113 | 0.117 | 0.126 | 84.3 | 94.2 | 95.6 | 97.6 | 98.1 |
200 | 0.95 | 1.9 | 2.1 | 2.36 | 2.43 | 0.04 | 0.079 | 0.088 | 0.098 | 0.101 | 92.0 | 95.4 | 96.5 | 97.9 | 98.5 |
300 | 0.8 | 1.4 | 1.7 | 1.76 | 2.13 | 0.033 | 0.058 | 0.071 | 0.073 | 0.089 | 93.3 | 96.6 | 97.2 | 98.5 | 98.7 |
400 | 0.65 | 1.26 | 1.6 | 1.66 | 1.9 | 0.027 | 0.053 | 0.067 | 0.069 | 0.079 | 94.5 | 96.9 | 97.4 | 98.6 | 98.8 |
500 | 0.6 | 0.93 | 1.4 | 1.56 | 1.83 | 0.025 | 0.039 | 0.058 | 0.065 | 0.076 | 94.9 | 97.7 | 97.7 | 98.6 | 98.8 |
The weight loss method was carried out at different temperatures (28 °C–48 °C) in the presence of different concentrations of ACT. Table 2 provides a summary of the weight loss, corrosion rates and inhibition efficiency of Cu in 1.0 M HNO3 acid as a function of temperature in the presence and absence of various inhibitor concentrations. From 28 to 48 °C, the corrosion rate of copper increased sharply in the absence of an inhibitor (ACT), but it increased more slowly when an inhibitor was included. Increasing the temperature typically speeds up the corrosion process, which in turn speeds up the metal's rate of dissolution.
Table 3 summarizes the corrosion parameter in the 28–48 °C temperature range with and without inhibitor (ACT). Using the Arrhenius equation, the activation energy (Ea) for copper dissolution in 1 M HNO3 was determined from the slope of plots:
![]() | (2) |
[ACT] (ppm) | Ea (kJ mol−1) | ΔS* (kJ mol−1 K−1) | ΔH* (kJ mol−1) |
---|---|---|---|
0 | 100.24 | 75.73 | 97.66 |
100 | 18.27 | −213.81 | 15.68 |
200 | 34.00 | −165.80 | 31.42 |
300 | 35.38 | −163.15 | 32.79 |
400 | 39.22 | −151.69 | 36.64 |
500 | 44.36 | −136.14 | 41.78 |
A plot of log(Rc) versus 1/T gives a straight line according to this equation. The values of Ea were calculated and summarized in (Table 3). In contrast to inhibited solutions, uninhibited solutions have higher energy activation. The value of the activation energy, however, rises as inhibitor concentration does. A decline in inhibition efficiency with rising temperatures and a corresponding increase in corrosion activation energy when an inhibitor is present are frequently interpreted as indicators of the development of a physical (electrostatic) adsorption coating.8 The result, which is consistent with a decrease in activation energy in the presence of an inhibitor and an increase in inhibition efficiency with a rise in temperature, points to a chemisorption mechanism.35 From the trend for the inhibitor suggests that the main effect of inhibiting species' physical adsorption (electrostatic interaction) in 1.0 M HNO3 is predominant.
The spontaneity of metal conversion into corrosion products is largely determined by the thermodynamic properties of the corrosion process. The transition state equation, a different formulation of the Arrhenius equation, was used to calculate the thermodynamic parameters of activation for the corrosion process (enthalpy (ΔH*)), free energy (ΔG), and entropy (ΔS*):
![]() | (3) |
According to this equation, a plot of log(Rc/T) against (1/T) gives a straight line, from which the values of Ea, ΔH* and ΔS* are calculated and listed in Table 2.
The activation energy (Ea) and enthalpy of activation (ΔH*) for the corrosion of copper in 1 M HNO3 is equal to 100.24 kJ mol−1 and 97.66 kJ mol−1, respectively, which is in good agreement with the work of Zarrouk et al. in which he found that the activation energy and enthalpy of activation of copper in 2 M HNO3 is equal to 100.21 kJ mol−1 and 97.53 kJ mol−1, respectively.35
The literature distinguishes three types of inhibitors based on how temperature affects inhibition effectiveness:36,37 inhibitors whose effectiveness decreases with increasing temperature: inhibitors exhibiting a rise in inhibition efficiency as the temperature rises: Ea (uninhibited) < Ea (inhibited) Ea (uninhibited) > Ea (inhibited); inhibitors are not affected by temperature fluctuations in terms of inhibition efficiency: the presence or absence of the inhibitor has no effect on Ea. Nevertheless, the inhibited solutions' ΔH* values in this investigation are lower than those of the uninhibited solutions, and they then steadily rise as the inhibitor's concentration rises, suggesting that both chemical and physical adsorption are responsible for the inhibitory action in solutions.36,37 This more likely indicates comprehensive adsorption, which involves both chemical and physical adsorption. Physical adsorption (electrostatic interaction) may accompany chemical adsorption since the adsorption heat was close to the overall chemical reaction heat.
The endothermic character of the metal dissolving process is reflected in the positive values of ΔH* (15.68–41.78 kJ mol−1). The decrease in Cu corrosion rate is mostly governed by the kinetic parameters of activation, as seen by the increase in ΔH* with increasing inhibitor concentration.37
In both the absence and presence of an inhibitor, the activation entropy (ΔS*) is negative. This suggests that an association rather than a dissociation occurs between reactants and the activated complex in the rate-determining step, which results in a reduction in disordering.38 It is clear that the ΔS* swings to less negative values (less ordered behavior) as inhibition efficiency rises.
![]() | (4) |
For acidic media, the computed Qad values are found to range between 49.06 and 99.26 kJ mol−1. An increase in efficiency at high temperatures is indicated by the positive sign of Qads.39
The corrosion inhibition in the current study was better understood thanks to the isotherms used to illustrate the process of crude adsorption on the copper surface. The type of copper metal, the compound's interaction with the copper surface, the size and structure of the organic compound, and the adsorption mechanism are some of the variables that affect how effective ACT is as a corrosion inhibitor. The most widely used adsorption isotherm systems, including the El-Awady, Freundlich, Flory–Huggins, Langmuir, Redlich–Peterson (R–P), Dubinin–Radushkevich, Temkim, and Frumkin isotherm models, were used to test adsorption modes in order to better understand the mechanism of corrosion inhibitor and the adsorption behavior of ACT on copper surface. Based on correlation (R2) values, which should be near unity, the adsorption process tends to follow the Langmuir, Freundlich, El-Awady, Temkim, and Redlich–Peterson (R–P) adsorption isotherms (Fig. 3). Table 3 lists the adsorption parameters that were thus determined. The Redlich–Peterson (R–P) isotherm provided the greatest match, followed by the Langmuir, El-Awady, Freundlich, and Temkin isotherms, according to the regression coefficient values shown in Table 4. Data from weight loss measures best fit the Langmuir adsorption isotherm, according to attempts to fit the data into several adsorption isotherms. According to Langmuir's assumptions, the degree of surface covering (θ) and the concentration of the adsorbate in the electrolyte's bulk (Cinh) are related in Eq.
![]() | (5) |
![]() | ||
Fig. 3 Adsorption isotherms for copper specimen with ACT in 1.0 M HNO3 for 4 h at different temperatures. |
Adsorption isotherm | 306 K | 311 K | 316 K | 321 K | 326 K | |
---|---|---|---|---|---|---|
Langmuir adsorption isotherm | R2 | 0.9999 | 0.9999 | 1.0000 | 1.0000 | 1.0000 |
Slope | 1.0219 | 1.0147 | 1.0171 | 1.0103 | 1.0093 | |
Intercept | 14.9510 | 5.5628 | 3.4008 | 1.7237 | 1.1480 | |
Kads | 0.0669 | 0.1798 | 0.2940 | 0.5802 | 0.8710 | |
ΔGads (kJ mol−1) | −27.71 | −31.29 | −33.09 | −35.42 | −37.08 | |
RL | 0.1357 | 0.0263 | 0.0111 | 0.0043 | 0.0023 | |
Freundlich adsorption isotherm | R2 | 0.9778 | 0.9575 | 0.9909 | 0.9663 | 0.9945 |
Slope | 0.0807 | 0.0197 | 0.0141 | 0.0074 | 0.0050 | |
Intercept | −0.2757 | −0.0642 | −0.0477 | −0.0255 | −0.0182 | |
Kf | 0.5301 | 0.8627 | 0.8959 | 0.9430 | 0.9589 | |
1/n | 0.0807 | 0.0197 | 0.0141 | 0.0074 | 0.0050 | |
ΔGads (kJ mol−1) | −33.81 | −35.35 | −36.01 | −36.72 | −37.34 | |
El-Awady adsorption isotherm | R2 | 0.9671 | 0.9108 | 0.9863 | 0.9620 | 0.9984 |
Slope | 0.7722 | 0.5306 | 0.4320 | 0.4000 | 0.3337 | |
Intercept | −0.7771 | 0.1503 | 0.4626 | 0.7921 | 1.0428 | |
1/y | 1.30 | 1.88 | 2.32 | 2.50 | 3.00 | |
Kads | 0.10 | 1.92 | 11.78 | 95.59 | 1.33 × 103 | |
ΔGads (kJ mol−1) | −30.60 | −36.62 | −39.10 | −41.75 | −43.96 | |
Redlich–Peterson (R–P) isotherm | R2 | 0.9994 | 1.0000 | 1.0000 | 1.0000 | 1.0000 |
Slope | 1.0731 | 1.0152 | 1.0094 | 1.0025 | 1.0001 | |
Intercept | −0.4948 | −0.1188 | −0.0797 | −0.0271 | −0.0101 | |
αR | 0.6097 | 0.8880 | 0.9234 | 0.9732 | 0.9899 | |
β | 1.0731 | 1.0152 | 1.0094 | 1.0025 | 1.0001 | |
KR | 1 | 1 | 1 | 1 | 1 |
Ideal simulating 1 is indicated by the slope of the Cinh/θ vs. Cinh plots being unity, as predicted by the Langmuir adsorption isotherm. One possible explanation is the interactions between the adsorbed species on the copper surface.1,36 The free energy of adsorption (ΔGads) and adsorption equilibrium constant (Kads) were calculated and given in Table 4 using the following relationship.
ΔG0ads = −R × T × Ln(55.5 × Kads) | (6) |
In general, electrostatic interactions are consistent with ΔGads values of −20 kJ mol−1 or less, while chemical interactions are defined as those with values of −40 kJ mol−1 or greater.1,40 In this investigation, the computed ΔGads values in an acidic medium ranged from −3.34 to −10.51 kJ mol−1 as the temperature rose from 28 to 48 °C. This suggests that the inhibitor's adsorption on the copper surface occurs spontaneously and validates the physical adsorption mechanism.
Additionally, the value of Kads has been observed to rise with temperature, suggesting that the adsorption of inhibitor molecules on the copper surface was more advantageous at higher temperatures. These findings suggest that, in this investigation, the interactions between the adsorbed molecules and the metal surface intensify as the temperature rises to 48 °C. This finding explains why the inhibition efficiency rises as the temperature rises. Nonetheless, the modest reported values for the free energy shift ΔGads validate the tested ACT inhibitor's physisorption action on the copper surface.
The equilibrium parameter RL, also known as the separation factor or equilibrium parameter, is a dimensionless constant that can be used to express the essential characteristics of the Langmuir isotherm.41–43
![]() | (7) |
![]() | ||
Fig. 4 Schematic illustration of inhibition mechanism of ACT adsorption on copper in 1.0 M HNO3 yielding the formation of protective layer. |
El-Awady isotherm also best fits the experimental data.33 This equation describes the El-Awady isotherm.
![]() | (8) |
![]() | (9) |
If 1/y is less than one, the inhibitor is likely to form numerous layers on the metal surface; if 1/y is more than one, the inhibitor molecule is likely to occupy many active sites.42–44 Each inhibitory molecule was attached to several active sites on the copper surface, as evidenced by the current values of 1/y being greater than one. It might be because the compound hydrolyzed in the presence of HNO3 solution, resulting in the presence of more than one anodic site with the compound.
The Freundlich isotherm describes the degree of adsorbent surface heterogeneity in multilayer adsorption.42,43 The linear form of the Freundlich isotherm equation is provided by the following formula.
![]() | (10) |
The intensity and capacity of adsorption are related by the Freundlich isotherm constants Kf (L g−1) and 1/n (dimensionless), respectively. The constants 1/n and Kf are determined using the slope and intercept of the linear plot of lnqe against lnCe shown in Fig. 3. Table 4 displays the Freundlich isotherm constants. According to the R2 score, this isotherm fit the experimental data quite well. The ease of adsorption is described by the value of 1/n. Adsorption is typically thought to be easy when 0 < 1/n < 1 and moderate or challenging when 1/n = 1 or 1/n > 1, respectively.44 An effective physical adsorption process between copper ions and ACT is indicated by the obtained 1/n value, which is again smaller than unity.
The Redlich–Peterson (R–P) isotherm is quite versatile and can be applied to both homogeneous and heterogeneous systems. When compared to the Langmuir, Freundlich, and D–R isotherms, the R–P isotherm incorporates three parameters into a single equation.43 This is the linear form of the R–P isotherm.
![]() | (11) |
Another well-known adsorption isotherm that is frequently used to explain how corrosion inhibitors work is the Temkin model. Unlike those covered thus far, it provides some information about the kind of interactions taking place in the adsorbed layer. The model is expressed as follows:
e2aθ = K × Ce | (12) |
The sign of the molecular interaction parameter (α) is utilized to establish whether repulsion or attraction takes place in the adsorbed layer.43,45 The linearized form of the equation can be used to create linear plots of θ against lnC (Fig. 4).
![]() | (13) |
The strength of the inhibitor molecules' adsorption on the metal surface is indicated by the value of K.
![]() | ||
Fig. 5 Copper polarization anodic and cathodic curves in 1.0 M HNO3 with and without different ACT concentrations. |
The shape of the polarization curve (Fig. 5) makes it evident that anodic and cathodic reactions are inhibited. They are evident from the data in Table 5 that as the concentration of the inhibitor (ACT) increases, the values of the anodic (βa) and cathodic (βc) Tafel constants also alter. This bolsters the mixed mode (cathodic and anodic) inhibitory activity of the acteoside molecule.
[Inhibitor] (ppm) | Ecorr (mV s−1) (Ag/AgCl) | Icorr (mA cm−2) | Rp (Ω cm2) | βa (mV dec−1) | βc (mV dec−1) | %IE |
---|---|---|---|---|---|---|
Blank | −4.23 | 0.697 | 29.80 | 188.89 | 64.00 | |
100 | 1.75 | 0.266 | 54.33 | 60.92 | 73.37 | 61.8 |
200 | 12.70 | 0.210 | 111.43 | 71.26 | 221.26 | 69.8 |
300 | 19.60 | 0.126 | 230.46 | 72.48 | 824.33 | 82.0 |
400 | 29.10 | 0.103 | 296.20 | 75.34 | 1033.16 | 85.2 |
500 | 38.60 | 0.091 | 356.88 | 80.11 | 1095.29 | 87.0 |
Copper's corrosion potential (Ecorr), corrosion current density (Icorr), anodic Tafel constant (βa), cathodic Tafel constant (βc), and inhibition efficiency (%IE) were all determined using the values of its polarization curves in 1.0 M of HNO3 with and without different inhibitor concentrations. Table 4 then provides a summary of this data.
In specimens containing 500 ppm of ACT at 25 °C, the value of βc > βa. The anode ion breaks down into the HNO3 solution when the value of βc > βa. In contrast, the values of βc < βa for the remaining instances. The absence of metal ion breakdown at the anode is indicated by a value of βc < βa.8
Table 5 shows that the corrosion current density (Icorr) values decrease from 0.697 mA cm−2 to 0.091 mA cm−2 when different concentrations of ACT are added, demonstrating that ACT has a significant inhibitory effect on the corrosion of copper in these media. The %IE increases from 61.8% to 87.0% as the Icorr decreases. At a concentration of 500 ppm inhibitor, the maximum %IE was reached at 87.0%. Generally speaking, if a compound's displacement in Ecorr is greater than 85 mV relative to the Ecorr of the blank, it can be classified as either an anodic or cathodic kind of inhibitor; if it is less than 85, it can be considered a mixed type.8 The maximum displacement in our study was 38.60 mV, indicating that the ACT is a mixed-type inhibitor. Additionally, it is discovered that Rp value rises as inhibitor concentration is increased, suggesting that copper corrosion is slowed down in inhibited solution versus uninhibited. When the inhibitor concentration grew from 100 ppm to 500 ppm, the maximum %IE determined from the Tafel polarization ranged between 61.8% and 87.0%. However, the %IE attained with weight loss (79.5–94.6%) is a little bit greater than that attained through electrochemical investigations. While the stabilization period for electrochemical measurements is only 30 min, the 2 h immersion time in weight loss measures may be the cause of this difference.
The acteoside (ACT) compound offers very good protection to copper against corrosion in acidic media, according to the results of both weight loss and polarization procedures. This is explained by the acteoside film's relative stability, which developed on copper's surface.
![]() | ||
Fig. 6 Nyquist plots for Cu corrosion at 303 K using the investigated inhibitor ACT and in a 1.0 M HNO3 solution. |
![]() | ||
Fig. 7 The EIS data for Cu corrosion in 1.0 M HNO3 solution and with the investigated inhibitor ACT were fitted using an electrochemical equivalent circuit. |
Conc. (ppm) | RS, Ω cm2 | Rct, Ω cm2 | Cdl, μF cm−2 | CPEP | %IE |
---|---|---|---|---|---|
Blank | 1.75 | 240.00 | 286.81 | 0.77 | |
100 | 1.42 | 1719.40 | 235.04 | 0.77 | 86.04 |
200 | 1.42 | 1973.70 | 221.57 | 0.78 | 87.84 |
300 | 1.31 | 2286.50 | 214.40 | 0.75 | 89.50 |
400 | 1.14 | 2612.20 | 203.87 | 0.78 | 90.81 |
500 | 1.06 | 3789.50 | 196.46 | 0.75 | 93.67 |
The inhibition efficiency from the impedance data was computed by comparing the charge transfer resistance values with and without the inhibitor.
![]() | (14) |
![]() | ||
Fig. 8 SEM image and EDS mapping of copper surface (A) without immersion (B) in 1.0 M HNO3 (C) presence of 300 ppm of ACT. |
This study uses a novel method to examine a molecule's reactivity to adsorption on a copper-metallic surface. The emphasis is on assessing the ΔEgap, a critical metric that represents the inhibitor's effectiveness and level of reactivity. Higher reactivity and better inhibitory performance are indicated by a decrease in the ΔEgap value. The improved inhibitory potential that results from the complex formed between the primary organic component and the surface is the cause of this effect.33
The σ and η parameters, which are obtained from ΔEgap and offer more information, are also included in the analysis. The σ parameter, which is associated with the polarizability of the molecule, emphasizes how responsive a softer molecule with a smaller ΔEgap is, leading to increased reactivity.50,51 To fully explore these intriguing phenomena, a number of quantum chemistry and molecular dynamics parameters, such as ΔEgap, σ, ΔN, η, X and ω, were carefully computed for the ACT molecule. Table 7 presents the results, which provide fresh insight into the compound's adsorption behavior.
Physical state | EHOMO (eV) | ELUMO (eV) | ΔEgap (eV) | σ | η | X | ω | ΔN | Dipole moment, μ (Debye) |
---|---|---|---|---|---|---|---|---|---|
Gas | −8.117 | −5.428 | 2.689 | 0.743 | 1.344 | 6.772 | 7.460 | 0.249 | 3.325 |
Acidic | −5.508 | −1.492 | 4.016 | 0.498 | 2.008 | 3.500 | 4.997 | 0.093 | 4.643 |
Both the gas and acidic states of the ACT compound have shown promising potential in protecting copper surfaces and serving as environmentally friendly inhibitors, as shown in Table 7 and Fig. 10. DFT calculations have revealed no significant differences in the calculated quantum parameters between the two states. The ΔEgap values of the ACT compound exhibit an increasing trend from ACT (gas) to ACT (acidic). Notably, ACT (gas) has the highest softness (σ) and the lowest global hardness (η) values. In addition, compared to the acidic media, ACT (gas) displays a higher electrophilicity index (ω), which is supported by its elevated ELUMO value. The inhibitor's capacity to take up electrons from the metal surface is shown by the electrophilicity index (ω).51 The ELUMO value (−5.428 eV) of ACT (gas) indicates its strong capacity to accept electrons from copper. Furthermore, the fraction of transferred electrons (ΔN) reveals that ACT (gas) contributes a higher proportion of transferred electrons to the copper surface (ΔN = 0.249) compared to ACT (acidic). It is worth noting that inhibitors with greater electron-donating ability at the metal surface tend to exhibit increased efficiency when ΔN is less than 3.6.51
![]() | ||
Fig. 10 The quantum parameters and calculated properties for the ACT compound in the gas and acidic states. |
Based on their ΔN values, which are less than 3.6, the compounds examined in both stages of the current investigation can be classified as green inhibitors. This chemical is the most effective inhibitor among those studied in both gaseous and acidic settings because it acts as the primary contributor to electron transfer onto the copper surface.
The presence of oxygen hetero-atoms and π-electrons in the aromatic rings of the ACT compound contributes to its adsorption in both phases.33,52 The results show that ACT may donate electrons from electron-rich centers to the copper's unoccupied d orbitals and accept electrons from the copper surface to establish a back-donating connection, allowing it to absorb onto the surface in all phases. The inhibitor's spatial orientation in the optimized structure determines whether this bond forms. The findings imply that the ACT molecule donates the unshared pair of electrons from oxygen atoms to the unoccupied d-orbitals of copper, hence demonstrating the strongest capacity to adsorb onto the copper surface. Consequently, it is anticipated that the ACT's adsorption on the copper surface will be the main cause of its inhibitory efficiency. In summary, physisorption and chemisorption are the two ways that ACT adsorbs on the Cu surface in acidic media. The difference between the inhibitors' and copper metal's chemical potential determines how much adsorption occurs. ACT (gas) and ACT (acidic) have computed ΔGads values of −52.854 and −22.599, respectively. Since the ΔGads values fall between zero and −40 kJ mol−1, they indicate that the inhibitors' adsorption mechanism is spontaneous physisorption.
(1) Acteoside compound (ACT) was proposed as a green corrosion inhibitor for copper in an acidic medium (1.0 M HNO3) based on the results of weight loss experiments and polarization techniques.
(2) The inhibition effectiveness rose as ACT concentration was increased, reaching a maximum value at 500 ppm at 48 °C. The inhibition efficiency also increased as temperature was raised.
(3) In the absence of an inhibitor, pure copper was found to dissolve in solution with a higher activation energy (Ea) than when an inhibitor was present. This implies that ACT-induced copper corrosion was successfully avoided by both the passivation procedure and barrier protection.
(4) The equilibrium isotherms for the adsorption of ACT on copper in 1.0 M HNO3 at different temperatures were investigated using four well-known isotherm models: the Langmuir, Freundlich, El-Awady, and Redlich–Peterson (R–P) isotherms.
(5) The sign of the free energy of adsorption indicates that the process is spontaneous, and the free Gibbs energy value showed that the physical adsorption in acidice midia could be ascribed to the inhibitory mechanism.
(6) The inhibitor (ACT) is a mixed type inhibitor (anodic and cathodic), according to polarization measurements.
(7) The ACT compound exhibited the most notable adsorption capacity on the copper surface, in both gas and acidic environments, thereby positioning it as a highly promising copper corrosion inhibitor. The adsorption mechanism was found to encompass a combination of physisorption and chemisorption. The favorable and spontaneous nature of the adsorption process is corroborated by the calculated ΔGads values.
(8) Potentiodynamic polarization studies reveal that ACT acts as mixed type inhibitors.
(9) The EIS analysis indicates that corrosion resistance or impedance increase with ACT by the formation of a protective layer on metal.
(10) SEM and EDS analysis show that ACT forms a protective film on the copper surface.
(11) The experimental and DFT results showcased in this study greatly augment our comprehension of the molecular interactions existing between inhibitors and metal surfaces. These findings offer invaluable insights that can be harnessed for the advancement of innovative and efficacious strategies for corrosion inhibition. Further investigation is imperative to explore the pragmatic implementation of ACT compound and other similar inhibitor in real-life scenarios of corrosion protection. This inquiry will serve to bridge the disparity between theoretical predictions and practical implementation, thereby propelling the progress of corrosion prevention techniques.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ra01657f |
This journal is © The Royal Society of Chemistry 2025 |