Hind Boughazi*ab,
Yamina Boudinar
ab,
Samira Tlili
c,
Amel Djedouanid and
Noura Naili
ae
aDepartment of Chemistry, Faculty of Sciences, University of 20 August 1955, Skikda 21000, Algeria. E-mail: h.boughazi@univ-skikda.dz; hind.boughazi21@gmail.com
bLaboratory of Physico-Chemistry Research on Surfaces and Interfaces, University of Skikda, 21000, ALgeria
cResearch Center in Industrial Technologies CRTI, Echahid Mohammed Abassi, BP. Box 64, Cheraga, 16014, Algiers, Algeria
dNormal Higher School Assia Djebar Constantine 3, 25000, Algeria
eResearch Unit of Environmental and Structural Molecular Chemistry, University of Constantine 1, Constantine 25000, Algeria
First published on 16th October 2025
In this work, (E)-2-((2-hydroxybenzylidene))hydrazine-1-carboxamide (HBHC) was investigated as a new organic and eco-friendly corrosion inhibitor for mild steel in acidic medium through electrochemical measurements including potentiodynamic polarization (PP), electrochemical impedance spectroscopy (EIS), long-term immersion tests, surface characterization, theoretical calculations, and ADMET studies. HBHC demonstrated excellent inhibition performance, achieving 94.50% efficiency by PP and 93.33% by EIS at 200 ppm, and retained remarkable stability over 30 days of immersion with 97.64% efficiency. Adsorption behavior was consistent with the Langmuir isotherm, with negative values indicating a mixed physisorption–chemisorption mechanism. SEM micrographs, EDX analysis, and elemental mapping analysis confirmed the formation of a uniform protective film enriched with heteroatoms on the steel surface. DFT calculations, including analysis of HOMO–LUMO frontier orbitals, revealed a low HOMO–LUMO energy gap (ΔE), supporting the high reactivity of HBHC and its strong donor–acceptor interactions with the Fe(110) surface, while MD simulations further confirmed its adsorption stability. Furthermore, ADMET predictions indicated low toxicity and good bioavailability, supporting the environmentally benign character of HBHC.
Hydrazine derivatives such as hydrazides, hydrazones, and hydrazinones are widely recognized as effective organic inhibitors for steel in acidic media.21,22 Their efficiency is mainly attributed to electron-donating heteroatoms and conjugated systems that facilitate adsorption and the formation of stable protective layers. Literature reports indicate that these compounds generally act as mixed-type inhibitors, with inhibition mechanisms governed by physicochemical adsorption described by Langmuir or Temkin isotherms.23–27 Their inhibition efficiencies are often high, with several studies reporting values above 95% at low concentrations.28,29 The adsorption of these inhibitors generally depends on the molecular structure and the presence of functional groups such as methoxy, nitro, or thioamide moieties.23,24,27,30,31 Quantum chemical calculations and molecular dynamics simulations have provided insight into the electronic properties and adsorption mechanisms, confirming that electron-donating groups and planarity enhance inhibitor performance.27,29,30,32 Surface analyses (SEM, AFM) further confirm the formation of uniform, protective layers that mitigate corrosion damage.19,21,26,29
Cherrak et al. reported that 1,5-dimethyl-1H-pyrazole-3-carbohydrazide (PyHz) exhibited up to 96% inhibition efficiency in 1 M HCl, with Langmuir-type adsorption confirmed by electrochemical techniques, SEM/EDX, and theoretical approaches (DFT and Monte Carlo).33 More recently, Kumari et al. investigated N′-[(4-methyl-1H-imidazole-5-yl)methylidene]-2-(naphthalen-2-yloxy)acetohydrazide (IMNH), which also displayed high anticorrosion activity in 1 M HCl, with efficiency increasing proportionally to concentration. EIS and PDP analyses confirmed its behavior as a Langmuir-type mixed inhibitor, while SEM and AFM observations revealed a smoother and less damaged surface for the inhibited steel.34 Collectively, these studies underline the strong potential of hydrazone and hydrazide derivatives as promising corrosion inhibitors in acidic environments. In general, hydrazide derivatives have emerged as highly efficient organic corrosion inhibitors. Their protective action is attributed to adsorption involving electrostatic interactions, coordination bonds, and the presence of aromatic or polar functional groups, leading to dense protective films. These features make them attractive alternatives to conventional inhibitors for safeguarding metallic infrastructures in aggressive acidic environments.
In this context, the present study investigates the efficiency of a novel organic inhibitor, HBHC [(E)-2-(2-hydroxybenzylidene) hydrazine-1-carboxamide], for the protection of A179 mild steel in an acidic medium. HBHC was selected as a corrosion inhibitor because of its multifunctional structure, which offers several adsorption centers. The hydrazine core (–NH–NH–) contains nitrogen atoms with lone pairs of electrons that serve as active sites for coordination with vacant d-orbitals of the metal surface. The aromatic benzylidene ring enhances π–electron interactions, while the hydroxyl substituent increases polarity and facilitates hydrogen bonding and additional coordination. Finally, the carboxamide group (–CONH2) contributes to molecular stability and promotes the cohesion of the protective film. These synergistic features make HBHC a strong candidate for efficient corrosion inhibition. Previous studies have reported that hydrazide derivatives, structurally related to HBHC, act as efficient corrosion inhibitors in acidic environments. However, many of these compounds still suffer from limitations such as toxicity, environmental concerns, moderate efficiency, or poor long-term stability. To the best of our knowledge, the corrosion inhibitive performance of HBHC on mild steel in acidic medium has not yet been explored. This study addresses that gap by demonstrating that HBHC combines strong adsorption capacity with long-term stability while maintaining an environmentally benign and non-toxic profile. Moreover, through the integration of electrochemical measurements, surface characterization, theoretical simulations, and ADMET studies as a new dimension in corrosion inhibitor research, the work delivers comprehensive mechanistic and safety insights that distinctly differentiate HBHC from previously reported hydrazide-based inhibitors.
![]() | ||
Fig. 1 Molecular structure of (E)-2-(2-hydroxybenzylidene)hydrazine-1-carboxamide HBHC; formula = C8H9N3O2, molar mass = 179.07 g mol−1. |
Element | C | Mn | Si | S | P | Fe |
---|---|---|---|---|---|---|
Value (%) | 0.06–0.18 | 0.27–0.63 | 0.25 max | 0.035 max | 0.035 max | Rest |
The surface was prepared by polishing with SiC papers (400–2000 grit), then washed with distilled water, rinsed with acetone, rinsed again with distilled water, and finally dried under an air stream.
The polarization experiments and electrochemical impedance spectroscopic investigations (EIS) were performed using an AUTOLAB model PGSTAT 302 N. Data were analyzed using Nova 2.1.4 software.
The electrochemical measurement and polarization corrosion experiments were conducted in a 1 M HCl solution, with two essential parameters being modified:
• The influence of the HBHC inhibitor concentration was examined by immersing the electrode in solutions with concentrations ranging from 0 to 200 ppm at room temperature.
• The influence of immersion duration on the corrosion inhibition performance of HBHC at optimal concentration (200 ppm).
Initially, a steady-state open circuit potential (Eocp) was established by immersing the working electrode in the test solutions for 30 minutes. Subsequently, potentiodynamic polarization curves were acquired at a scan rate of 0.5 mV s−1 within a potential range of ±250 mV relative to the Eocp. Corrosion current density values were determined by extrapolating the anodic and the cathodic Tafel lines to their intersection at the corrosion potential (Ecorr).
The inhibition efficiency (IE %) was calculated using eqn (1) based on data obtained from the Tafel plots:35
![]() | (1) |
The EIS measurements were carried out in the range from 105 Hz to 10−1 Hz, with a 10 mV sinusoidal signal perturbation. The frequency points per decade were set at 10 points per decade. The resulting data were utilized to estimate the inhibition effectiveness (EEIS) using the following expression:35
![]() | (2) |
Key quantum chemical descriptors were calculated, including ionization potential (I), electron affinity (A), chemical hardness (η), electronegativity (χ), softness (σ), electrophilicity index (ω), and chemical potential (ε), computed using standard equations39,40 (eqn (3)–(9)).
I = −EHOMO | (3) |
A = −ELUMO | (4) |
η = (I − A)/2 | (5) |
χ = (I + A)/2 | (6) |
σ = 1/η | (7) |
ω = χ2/(2η) | (8) |
ε = −χ | (9) |
ΔN110 = (χFe(110) − χHBHC)/[2 × (ηFe(110) + ηHBHC)] | (10) |
The Fukui indices are defined as follows:39
For a nucleophilic attack (addition of an electron) on atom k:
fk+ = qk(N + 1) − qk(N) | (11) |
• For an electrophilic attack (removal of an electron) on atom k:
fk− = qk(N) − qk(N − 1) | (12) |
• For a radical attack (average variation) on atom k:
f0k = [qk(N + 1) − qk(N − 1)]/2 | (13) |
V(r) = ΣZa/|Ra − r| − ∫[ρ(r′)/|r′ − r|]dr′ | (14) |
The first sum represents the positive contribution of the nuclei. The integral represents the negative contribution of the electrons. Analysis of the ESP map allows for the visual identification of active sites for inhibitor metal surface interaction, as these sites are generally correlated with regions of high negative potential, favoring coordination to metal atoms.43
In this research, the interactions between the studied molecules and the Fe(110) surface were examined using molecular dynamics (MD) simulations with the Forcite module in Materials Studio 2023. The simulation cell measured 24.82 × 24.82 × 30 Å3 and included 621 water molecules, 9 chloride ions (Cl−), and 9 hydronium ions (H3O+). The simulation ran for 500 picoseconds (ps) with a time step of 1.0 femtosecond (fs), while the temperature was maintained at 300 K using the NVT ensemble and the Andersen thermostat. The COMPASS III force field was applied to accurately model the interactions within the system. This approach allows for a detailed and reliable analysis of molecular behavior at the metal–solution interface, supporting a better understanding of corrosion inhibition mechanisms.43
Absorptions in the range of 1349–1132 cm−1 are indicative of C–N, N–N, and C–O stretching modes, reflecting the presence of amine and phenolic functionalities. Finally, the bands at 943, 755, 629, and 543 cm−1 correspond to aromatic C–H out-of-plane bending vibrations, confirming the aromatic nature of the compound. Collectively, the FTIR data support the proposed molecular structure and confirm the presence of hydroxyl, carbonyl, imine, amine, and aromatic functional groups.
![]() | ||
Fig. 3 Open circuit potential curves of A179 carbon steel potential without and with different concentrations of HBHC inhibitor at 25 °C. |
Overall, the OCP curves highlight the increasing effectiveness of the inhibitor, with optimal performance observed at 200 ppm.
![]() | ||
Fig. 4 Potentiodynamic polarization curves of A179 in a 1 M HCl solution in the absence and presence of HBHC at different concentrations at 298 K. |
Medium | Ecorr (mV/Ag/AgCl) | icorr (μA cm−2) | βc (mV dec−1) | βa (mV dec−1) | Rp (Ω) | Surface coverage (θ) | IE (%) |
---|---|---|---|---|---|---|---|
Blank | −451.01 ± 10.42 | 7.001 ± 0.08 | 338.6 | 200.64 | 8420 ± 78 | — | — |
50 ppm | −460.53 ± 8.47 | 3.567 ± 0.11 | 344.2 | 183.48 | 15![]() |
0.4905 | 49.05 |
100 ppm | −425.53 ± 12.44 | 2.0516 ± 0.07 | 332.96 | 175.25 | 38![]() |
0.7069 | 70.69 |
150 ppm | −385.28 ± 7.85 | 1.2644 ± 0.04 | 356.34 | 156.34 | 45![]() |
0.8194 | 81.94 |
200 ppm | −390.93 ± 8.77 | 0.3848 ± 0.02 | 302.51 | 81.22 | 51![]() |
0.9450 | 94.50 |
An initial analysis of these curves shows that the addition of the inhibitor affects both the anodic and cathodic reactions.
In the 1 M HCl solution, the inhibitor reduces the anodic partial current associated with metal dissolution and the reduction in the cathodic current corresponds to a decrease in proton concentration.
The cathodic curves present a linear Tafel region, confirming that the hydrogen evolution reaction on the steel is controlled entirely by activation.48,49 The Tafel polarization curves show the influence of different concentrations of the corrosion inhibitor on the electrochemical behavior of the metal surface. As the inhibitor concentration increases, a noticeable decrease in the corrosion current density (icorr) is observed, showing a notable decrease in the rate of corrosion. Among the tested concentrations, 200 ppm is the optimal inhibitor concentration, demonstrating the lowest (icorr) value and a notable shift in the corrosion potential (Ecorr) toward more noble values, indicating enhanced protection of the surface. The calculated inhibition efficiency at this concentration is 94.50%, providing excellent corrosion inhibition under the studied conditions.
Table 2 demonstrates that even a small amount of the inhibitor significantly reduces the corrosion current density (icorr). The icorr of the blank solution decreases from 7.001 μA cm−2 to 0.3848 μA cm−2 with the addition of 200 ppm of HBHC. Furthermore, polarization resistance (Rp) increases considerably, rising from 8420 Ω cm2 to 51679 Ω cm2 without and with the inhibitor at 200 ppm, respectively. Table 2 also indicates that HBHC acts as a mixed-type inhibitor. This is evidenced by the fact that the variation in corrosion potential (Ecorr) in the presence of the inhibitor is less than 85 mV compared to the blank value.31,50,51 In the present case, the maximum displacement is less than 70 mV vs. Ag/AgCl. Additionally, both anodic and cathodic partial current densities are significantly decreased, indicating that the inhibitor effectively suppresses both anodic metal dissolution and cathodic reduction reactions.
The reduction in the corrosion rate due to the adsorption of inhibitors on the metal surface is a clear indicator that the inhibitor molecules are successfully protecting the metal. The type of adsorption, whether physisorption, chemisorption, or a combination of both, determines the strength, reversibility, and longevity of the corrosion protection.52 The functional groups in HBHC inhibitor, such as oxygen and nitrogen, facilitate the adsorption of the inhibitor molecules onto the mild steel surface.53 The interactions formed between the inhibitor and the metal surface lead to a robust and stable protective layer that decreases the corrosion rate, positioning HBHC as an efficient corrosion inhibitor. This protective mechanism guarantees extended corrosion protection by preventing corrosive agents from reaching the metal surface.
![]() | ||
Fig. 5 Nyquist diagrams (a) and Bode diagrams (b) of A179 carbon steel in 1 M HCl solution in the absence and presence of HBHC at different concentrations after 1 h of immersion at 25 °C. |
The impedance spectra in the absence of HBHC present one capacitive loop, which depicts a depressed semicircle adjusted with electrical circuit Fig. 6a. The depressed semicircle is attributed to the frequency dispersion effect and surface irregularities and heterogeneities.29,54 This loop can be attributed to a charge transfer process.
In the presence of the HBHC inhibitor, examination of the impedance spectra reveals the presence of two distinct loops. The first loop, observed at high frequencies, is associated with the adsorption of the inhibitor film, whereas the second loop, at low frequencies, corresponds to the charge-transfer processes occurring at the metal/electrolyte interface. An increase in the size of the loops is noted as the inhibitor concentration increases. This result suggests that the inhibitor completely changes the corrosion process at the A179 carbon steel/1 M HCl solution interface.
Fig. 5b presents the Bode plots for A179 steel in the absence and presence of various concentrations of HBHC. An increase in log(Z) values, accompanied by a shift toward lower frequency values, is observed with rising inhibitor concentrations. This phenomenon can be associated with the development of a protective film on the A179 steel surface.55 On the other hand, the phase angle values increase up to a more negative value of −80° for an optimal concentration of 200 ppm compared to the blank (around −50°), which implies that the tested inhibitor is effective.49
Fig. 6a presents the equivalent circuit (EC) employed to analyze the impedance spectra in the absence of the inhibitor. In this model, the constant phase element (CPE) is introduced in place of the ideal double-layer capacitance, owing to the depressed capacitive semicircles observed in the Nyquist plots. The use of CPE in such circuits is a well-established practice in corrosion studies, as the measured macroscopic impedance is strongly influenced by various microscopic factors, including surface roughness, charge inhomogeneities, and complex electrochemical reactions. The circuit elements consist of the solution resistance (Rs), the constant phase element (CPE) and the charge-transfer resistance (Rct).
The equivalent electrical circuit (shown in Fig. 6b) was proposed to fit the EIS data in presence of HCBC inhibitor, the high-frequency loop is modeled by a Constant Phase Element of the film (CPEf), which accounts for the non-ideal capacitive behavior of the protective layer, connected in parallel with the film resistance (Rf). The double-layer Constant Phase Element (CPEdl) in parallel with the charge-transfer resistance (Rct) represents the low-frequency loop.
From the electrochemical parameters obtained by fitting the experimental data with the equivalent circuit (Table 3), it can be observed that the capacitance of the constant phase element of the double layer (CPEdl) decreases with increasing inhibitor concentration, while the charge transfer resistance (Rct) increases. This behavior is attributed to the gradual replacement of water molecules by HBHC inhibitor molecules at the electrode surface, which reduces the number of active sites and thereby retards the corrosion process.56 Finally, the inhibition efficiency (IE %) for metal corrosion, calculated using eqn (2), in the presence of HBHC increases with rising inhibitor concentration. This indicates that a greater number of inhibitor molecules are able to adsorb onto the metal surface, thereby forming a thicker surface film or enhancing surface coverage.57 The value of the inhibition efficiency reached a maximum of 93.33% of effectivity at 200 ppm. Furthermore, Table 3 shows that the electrolyte resistance Rs remains nearly constant across different HBHC concentrations, indicating that the inhibitor does not significantly affect the solution conductivity.58,59
C (ppm) | Rs (Ω cm2) | CPEf | Rf (Ω cm2) | CPEdl | Rct (Ω cm2) | IE (%) | ||
---|---|---|---|---|---|---|---|---|
Qf (F sn−1) × 10−7 | n | Qdl (F sn−1) × 10−4 | n | |||||
Blank | 2.78 ± 0.07 | — | — | — | 2.291 ± 1.658 | 0.898 ± 0.012 | 666.1 ± 29.3 | — |
50 ppm | 3.45 ± 1.05 | 2.87 ± 0.45 | 0.964 ± 0.005 | 337,4 ± 28.27 | 0.838 ± 0.781 | 0.864 ± 0.023 | 5251.9 ± 55.3 | 87.32 |
100 ppm | 2.47 ± 0.14 | 0.755 ± 0.385 | 0.974 ± 0.025 | 477.8 ± 71.8 | 0.571 ± 0.568 | 0.875 ± 0.011 | 8952.7 ± 75.2 | 92.55 |
150 ppm | 3.54 ± 0.05 | 0.581 ± 0.104 | 0.988 ± 0.001 | 510.8 ± 124.9 | 0.327 ± 0.711 | 0.858 ± 0.014 | 9551.4 ± 122.4 | 93.03 |
200 ppm | 3.61 ± 0.12 | 0.302 ± 0.021 | 0.989 ± 0.002 | 582.9 ± 135.2 | 0.192 ± 1.025 | 0.839 ± 0.035 | 9998.6 ± 107.9 | 93.33 |
The electrochemical performance of HBHC was compared with literature data for structurally related hydrazide derivatives. Abdallah et al. reported inhibition efficiencies of 92–93% for H2HEH and H2MEH in 1 M HCl, with Rct values up to 7285 Ω cm2 at 10−4 M. In our study, HBHC achieved similar efficiencies (93%) but with higher Rct values (9998.6 Ω cm2 at 200 ppm). This close agreement indicates that HBHC follows a comparable inhibition mechanism to other nitrogen-rich hydrazide derivatives, supporting the reliability of our findings.60
![]() | ||
Fig. 7 Adsorption isotherms (a) Langmuir, (b) Freundlish, (c) Temkin, and (d) Frumkin for A179 MS in 1 M HCl inhibited by 200 ppm of HBHC at 298 K. |
Mathematically, Langmuir's adsorption isotherm is expressed as in eqn (15):63
![]() | (15) |
On the other hand, the Freundlich isotherm describes multilayer adsorption of the tested inhibitor onto the steel surface, and it can be expressed in linear form by eqn (16) as follows:64
![]() | (16) |
The Temkin isotherm considers the enthalpy change due to inhibitor interactions. Where all molecules in the layer decrease linearly with coverage due to inhibitor–inhibitor interactions, and can be explained in linear form by eqn (17) as follows:64
![]() | (17) |
The Frumkin isotherm includes an interaction parameter to account for attractions or repulsions between adsorbed inhibitormolecules. This can be expressed in linear form by eqn (18) as follows:64
![]() | (18) |
In Fig. 7, the Langmuir plot consists of a straight line with a 0.9999 regression coefficient. The slope values that were obtained were close to 1, which validated the plot's well-fitting value.64 The ratio 1/Kads corresponds to the intercept of the Langmuir isotherm (Fig. 7a). The adsorption equilibrium constant Kads obtained from this intercept was subsequently used to calculate the standard free energy of adsorption using eqn (19).65
![]() | (19) |
According to the Langmuir isotherm, adsorption occurs as a monolayer of inhibitor molecules uniformly covering distinct adsorption sites on the metal surface, where each adsorption site holds only one inhibitor molecule, with no interaction between adsorbed molecules.66,67
The value of Kads is 44752.74 L mol−1. The high value of adsorption equilibrium constant Kads shows good adsorption of the HBHC inhibitor.68 The nature of the adsorption process was further evaluated through thermodynamic analysis; the standard Gibbs free energy of adsorption
provides insight into the spontaneity of the adsorption process and the nature of the adsorption mechanism. The resulting
was found to be −36.483 kJ mol−1. A negative
value confirms the spontaneous adsorption of the HBHC inhibitor onto the steel surface and implies a strong interaction between the inhibitor molecules and the metal substrate.69 Generally,
values less than or equal to −20 kJ mol−1 are indicative of physisorption, which involves weak van der Waals forces, whereas values equal to or more negative than −40 kJ mol−1 are characteristic of chemisorption, where stronger chemical bonds are formed. In this study, the calculated
of −36.483 kJ mol−1 indicates that the adsorption of HBHC is thought to involve two different types of interaction: chemisorption and physisorption, since this value is close to the chemisorption threshold, it can be concluded that adsorption is mainly dominated by chemisorption.67,70 This is common for organic inhibitors that interact with metal surfaces through mechanisms like electron donation, π–π interaction, or coordination with metal sites.71,72
![]() | ||
Fig. 8 Measured EIS response of the A179 steel samples immersed for 1, 6, 12, 18, 24, and 30 days in (a) 1 M HCl solution and (b) 1 M HCl solution containing 200 ppm of HBHC inhibitor at 298 K. |
The Nyquist plots presented in Fig. 8a, illustrate the evolution of the metal's corrosion behavior over time in the absence of the HBHC corrosion inhibitor. All the curves display a single semicircle, which means the corrosion process is mainly controlled by charge transfer at the metal–solution interface. As the period of immersion increases from 1 day to 30 days, the sizes of the semicircles consistently decrease. This decrease indicates a gradual lowering of charge transfer resistance (Rct), which implies that the metal is becoming more susceptible to corrosion. In other words, with longer exposure times, the metal surface becomes increasingly reactive and more likely to corrode. The equivalent circuit used to fit the data, shown in the inset, includes a solution resistance (Rs) and a constant phase element (CPEdl) parallel to Rct. The use of a CPE instead of an ideal capacitor suggests some irregularity at the surface, likely due to surface roughness corrosion product accumulation, or uneven degradation.
When 200 ppm of HBHC inhibitor is present, the impedance values increase steadily up to 18 days, indicating the progressive formation and stabilization of the protective film. The slight variation observed between days 24 and 30 falls within the experimental margin of error and does not reflect a real decrease in protection. Overall, the impedance values remain stable at longer immersion times and are significantly higher than those of the uninhibited sample, confirming the durability and long-term protective effect of HBHC.
As shown in Fig. 8b, the impedance diagrams exhibit a similar shape at different immersion times, characterized by the appearance of two distinct loops. These correspond to two time constants: the high-frequency loop is related to the development of the inhibitor film, while the low-frequency loop is associated with charge transfer resistance. Over time, the capacitive loops expand, which is attributed to the adsorption of HBHC molecules on the steel surface, forming a protective barrier against corrosion.
The equivalent circuit model used to fit the impedance data is presented in the inset of Fig. 8b. The high-frequency loop is represented by a Constant Phase Element (CPEf), simulating the non-ideal capacitance of the film, in parallel with the film resistance (Rf). The low-frequency loop corresponds to the double-layer Constant Phase Element (CPEdl) in parallel with the charge transfer resistance (Rct).
A comparison of the electrochemical impedance parameters is provided in Table 4 for steel samples immersed in the corrosive medium with and without 200 ppm HBHC. In the absence of the inhibitor, the charge transfer resistance (Rct) decreases markedly from 542.1 Ω cm2 on day 1 to 212.2 Ω cm2 on day 30, indicating a continuous increase in corrosion activity. In contrast, the presence of HBHC maintains much higher Rct values, confirming its protective performance over prolonged exposure.
Time (days) | Rs (Ω cm2) | CPEf | Rf (Ω cm2) | CPEdl | Rct (Ω cm2) | IE (%) | ||
---|---|---|---|---|---|---|---|---|
Qf (F sn−1) × 10−5 | n | Qdl (F sn−1) × 10−6 | n | |||||
Blank | ||||||||
1 day | 2.22 ± 0.04 | — | — | — | 1.29 ± 0.05 | 0.878 ± 0.089 | 542.1 ± 98.2 | — |
6 days | 1.58 ± 0.12 | — | — | — | 1.76 ± 0.11 | 0.854 ± 0.157 | 499.5 ± 87.5 | — |
12 days | 1.14 ± 0.03 | — | — | — | 3.82 ± 0.09 | 0.888 ± 0.042 | 385.4 ± 112.3 | — |
18 days | 2.85 ± 0.22 | — | — | — | 6.41 ± 0.06 | 0.865 ± 0.047 | 302.7 ± 103.3 | — |
24 days | 3.05 ± 0.17 | — | — | — | 22.5 ± 0.08 | 0.834 ± 0.149 | 284.9 ± 85.7 | — |
30 days | 2.87 ± 0.28 | — | — | — | 32.7 ± 0.07 | 0.824 ± 0.258 | 212.2 ± 45.2 | — |
![]() |
||||||||
200 ppm HBHC | ||||||||
1 day | 8.25 ± 0.15 | 2.32 ± 0.17 | 0.902 ± 0.05 | 9240.2 ± 125.8 | 2.54 ± 0.148 | 0.911 ± 0.047 | 2572.6 ± 225.1 | 78.92 |
6 days | 7.55 ± 0.09 | 1.51 ± 0.24 | 0.944 ± 0.04 | 9895.1 ± 144.2 | 1.52 ± 0.097 | 0.865 ± 0.044 | 3645.4 ± 205.3 | 86.29 |
12 days | 8.04 ± 1.05 | 0.762 ± 0.18 | 0.915 ± 0.02 | 10![]() |
0.855 ± 0.284 | 0.922 ± 0.055 | 5542.1 ± 388.2 | 93.04 |
18 days | 7.92 ± 0.08 | 0.514 ± 0.09 | 0.878 ± 0.05 | 11![]() |
0.654 ± 0.254 | 0.887 ± 0.037 | 7445.9 ± 378.8 | 95.93 |
24 days | 8.33 ± 0.16 | 0.454 ± 0.16 | 0.889 ± 0.01 | 11![]() |
0.551 ± 0.015 | 0.898 ± 0.087 | 9247.8 ± 258.5 | 96.91 |
30 days | 7.56 ± 1.12 | 0.478 ± 0.08 | 0.887 ± 0.02 | 10![]() |
0.674 ± 0.087 | 0.881 ± 0.048 | 9025.7 ± 178.7 | 97.64 |
At the same time, the double layer capacitance (Qdl) increases, suggesting a rise in the buildup of corrosion products and deterioration of the surface.
In contrast, when 200 ppm of HBHC is added, a substantial increase in both film resistance (Rf) and charge transfer Rct is observed, with Rct peaking at 9247.8 Ω cm2 from day 1 to day 30. The lower Qdl values and higher “n” factors (around 0.9) also indicate a more uniform and capacitive behavior, consistent with the presence of an effective barrier layer. This trend reflects the gradual development of a stable and adherent film on the metal surface that limits charge transfer and thus reduces corrosion activity. The corresponding rise in inhibition efficiency (IE %), as illustrated in Table 5, from 87.92% to 97.64%, supports this interpretation and demonstrates that the HBHC compound offers an intense and persistent corrosion inhibition performance, suggesting its reliability for use in extended corrosive exposure.
The shiny coupons showed no significant flaws except for polishing scratches (Fig. 9a). In the case of HCl without inhibitor, the surface becomes noticeably damaged even after just one day (Fig. 9b) with clear signs of corrosion characterized by the existence of numerous corrosive pits distributed across the surface. After 30 days, the damage becomes much more severe. The surface of the A179 MS was strongly corroded with notable localized pits (Fig. 9c), which could be attributed to the corrosive effect of the 1 M HCl solution. When comparing Fig. 9b and c of MS in acidic medium without inhibitor to Fig. 9a of the polished MS, it is evident that the surface exhibits uniform corrosion accompanied by localized corrosion. On the other hand, the samples treated with 200 ppm of HBHC inhibitor show a stark contrast. Fig. 9d and e showed the changes in the surface appearance caused by the existence of a protective layer on the surface. After 1 day, the micrograph shows the absence of corrosion, and the surface remains relatively smooth and uniform, indicating that the protective film formed by the inhibitor is already active. Remarkably, even after 30 days of exposure, the inhibited surface still appears relatively intact and smooth, with only minimal changes.
The results demonstrate that the HBHC inhibitor efficiently adsorbs onto the A179 carbon steel surface, offering protective coverage. Therefore, the HBHC inhibitor not only reduces corrosion from the beginning but also offers continued protection to the material over extended exposure periods.
In the spectrum (Fig. 10d), the sample immersed for 30 days in the HCl with 200 ppm HBHC solution still shows a high iron content (77.73 wt%), indicating excellent long-term corrosion resistance. Nitrogen increases to 5.16 wt%, and carbon remains stable at 8.30 wt%, indicating that the inhibitor layer remains present and may even become more established over time. The oxygen content slightly drops to 8.53 wt%, and chlorine decreases further to 0.27 wt%, demonstrating continued suppression of both oxidation and chloride attack. Overall, the consistent carbon levels, rising nitrogen levels, and declining chlorine levels suggest that the HBHC inhibitor offers enduring protection by creating a stable and firmly adhering film on the metal surface, which substantially lowers corrosion during extended exposure.
![]() | ||
Fig. 11 Percentage distribution of different forms of HBHC as a function of pH, obtained by Marvin Sketch software. |
At low pH values (below 3), the protonated species at the hydrazine nitrogen and the neutral form coexist in comparable proportions rather than a single dominant species. This coexistence is evident in Fig. 11, where each species constitutes approximately 50% of the total distribution at around pH 1.08. Hence, the molecular population at acidic pH reflects a dynamic equilibrium between these two forms, affecting the adsorption characteristics and inhibitory performance.77
In the intermediate pH range (approximately 3 to 10), the neutral form becomes stabilized, partly due to intramolecular hydrogen bonding, and emerges as the predominant species. At alkaline pH values (above 11), both the phenolic hydroxyl (–OH) and the hydrazine nitrogen moieties undergo deprotonation in a narrow pH window, leading to the formation and coexistence of several anionic species. The overlapping protonation and deprotonation equilibria, due to closely spaced pKa values of multiple functional groups in HBHC, result in the non-linear and complex distribution curves observed. This behavior is characteristic of polyfunctional organic molecules with multiple ionizable sites and reflects genuine chemical equilibria rather than random fluctuations. Additionally, the electronic properties and reactivity of the neutral and protonated forms were analyzed through optimized molecular structures and spatial distributions of HOMO and LUMO orbitals, providing insights into the protonation-dependent functionality of HBHC as a corrosion inhibitor.
![]() | ||
Fig. 12 Optimized geometry, HOMO, and LUMO orbitals of HBHC and HBHC+ in their neutral and protonated forms in aqueous solution. |
In the neutral form, the HOMO is primarily localized on nitrogen and oxygen atoms, indicating potential electron-donating sites, while the LUMO is distributed over the aromatic ring, suggesting electron-accepting regions. Upon protonation, significant changes in the localization of the frontier orbitals are observed, reflecting the altered electronic structure and reactivity.78 The calculated quantum chemical descriptors are summarized in Table 5.
Table 5 presents the quantum chemical descriptors and reactivity indices for the HBHC molecule in both its neutral and protonated forms, calculated using DFT at the M06-2X/6-31G(d) level. The optimized energy of the protonated form (−624.973 hartree) is lower than that of the neutral form (−624.541 hartree), indicating that protonation leads to a more stable molecular structure. This increased stability may contribute to the enhanced persistence of the inhibitor under acidic conditions.79
The EHOMO value, which reflects the ability of the molecule to donate electrons,80 decreases from −7.327 eV in the neutral form to −8.457 eV in the protonated form. This decrease suggests that the protonated HBHC is less likely to act as an electron donor compared to the neutral form. Conversely, the ELUMO value, which indicates the ability to accept electrons,80 also decreases significantly from −0.384 eV (neutral) to −2.063 eV (protonated). This substantial reduction demonstrates that the protonated form is a much better electron acceptor.80
The HOMO–LUMO gap narrows from 6.943 eV in the neutral form to 6.394 eV in the protonated form, smaller HOMO–LUMO gap generally means higher chemical reactivity and greater ease of electron excitation, making the molecule more likely to participate in chemical interactions with the metal surface.81 Regarding the ionization potential (I) and electron affinity (A), both parameters increase upon protonation (I: 7.327 → 8.457 eV; A: 0.384 → 2.063 eV). The higher ionization potential indicates that the protonated molecule is more resistant to losing electrons, while the increased electron affinity confirms its enhanced capacity to accept electrons.82
The chemical hardness (η) decreases slightly from 3.472 eV in the neutral form to 3.197 eV in the protonated form. Lower hardness is associated with higher reactivity, suggesting that the protonated HBHC is more chemically active and may interact more readily with the Fe(110) surface.83 The electronegativity (χ) rises from 3.855 eV to 5.260 eV with protonation, reflecting a stronger tendency of the protonated form to attract electrons. This is further supported by the chemical potential (ε), which becomes more negative upon protonation (−3.855 → −5.260 eV), indicating a greater drive to acquire electrons from the environment.84 The softness (σ), which is the inverse of hardness, increases from 0.288 eV (neutral) to 0.313 eV (protonated), reinforcing the observation that the protonated form is more reactive.85 A notable increase is observed in the electrophilicity index (ω), which nearly doubles from 2.140 eV in the neutral form to 4.327 eV in the protonated form. This substantial rise demonstrates that the protonated HBHC is a much stronger electrophile, making it more effective at interacting with the electron-rich sites on the iron surface.86 The electron transfer fraction (ΔN110) changes from a positive value (0.139) in the neutral form to a negative value (−0.069) in the protonated form. A positive ΔN110 suggests electron transfer from the inhibitor to the metal, while a negative value indicates the reverse.87 This shift implies that the direction and mechanism of charge transfer are affected by protonation, which could influence the adsorption mode and inhibition efficiency.87
These changes are expected to enhance the adsorption of the protonated molecule onto the Fe(110) surface, consequently improving its performance as a corrosion inhibitor, especially in acidic environments where protonated species predominate.
Importantly, these theoretical descriptors correlate well with the experimental inhibition trends. The relatively high EHOMO of the neutral form supports moderate inhibition efficiencies under near-neutral conditions, where electron donation from HBHC to Fe is favored. In contrast, the protonated form exhibits a much lower ELUMO, narrower energy gap, and higher electrophilicity index, all of which enhance its ability to accept electron density from the Fe surface via back-donation. This dual charge-transfer pathway, as reflected in the change of ΔN from positive to negative values, explains the stronger and more stable adsorption of protonated HBHC observed experimentally, and hence its superior inhibition performance in acidic environments.
Molecule | Atom | f− (Nucleophilic) | f+ (Electrophilic) | CDD |
---|---|---|---|---|
HBHC | N(10) | 0.1171 | 0.0272 | −0.0899 |
O(12) | 0.1065 | 0.0687 | −0.0378 | |
N(9) | 0.0577 | 0.1192 | 0.0614 | |
HBHC+ | N(9) | −0.1951 | 0.1105 | 0.3056 |
C(2) | −0.1971 | 0.0875 | 0.2847 | |
O(12) | −0.1386 | 0.0694 | 0.2079 |
Structures | Total energy | Adsorption energy | Rigid adsorption energy | Deformation energy | dEad/dNi |
---|---|---|---|---|---|
HBHC | −333.225 | −1698.608 | −169.209 | −1529.399 | −1698.608 |
HBHC+ | −349.531 | −1688.522 | −109.009 | −1579.513 | −1688.522 |
Both forms exhibit highly negative adsorption energies (HBHC: −1698.6080 kcal mol−1; HBHC+: −1688.5220 kcal mol−1), confirming strong surface binding. The neutral form exhibits a more negative rigid adsorption energy (−169.2090 kcal mol−1), indicating a stronger initial interaction with the surface. In contrast, the protonated form displays a significantly higher deformation energy (−1579.5130 kcal mol−1), suggesting greater structural adjustment upon adsorption.89 These findings indicate that both forms possess high affinity and adaptability toward the surface, with the protonated species demonstrating a particularly pronounced ability to accommodate structural changes during adsorption.
![]() | ||
Fig. 14 Adsorption geometries of neutral and protonated HBHC on Fe(110) surface: top and side perspectives from MD simulations. |
The distances between the main reactive atoms of HBHC and the Fe(110) surface, summarized in Table 8, further clarify the adsorption behavior. For the protonated form, the Fe–N(9) distance is reduced to 2.932 Å compared to 3.144 Å for Fe–N(10) in the neutral form, indicating stronger and more direct interaction. Similarly, Fe–C(2) and Fe–O(12) distances are slightly shorter or comparable in the protonated form, supporting the observation of enhanced adsorption. These results are consistent with previous studies, which have shown that protonation enhances the adsorption and inhibitory performance of organic inhibitors on metal surfaces by increasing their electrophilicity and interaction strength.89–91
Molecule | Fe/atom pair | Distance (Å) |
---|---|---|
HBHC | Fe/N(10) | 3.144 |
Fe/O(12) | 3.564 | |
Fe/N(9) | 3.328 | |
Fe/C(2) | 3.208 | |
HBHC+ | Fe/N(9) | 2.932 |
Fe/C(2) | 3.195 | |
Fe/O(12) | 3.625 |
![]() | ||
Fig. 15 Comparative NCI (RDG) isosurfaces and scatter plots for (a) protonated and (b) neutral HBHC inhibitors. |
In the neutral HBHC structure, the RDG scatter plot reveals prominent blue regions at sign(λ2)ρ values around −0.020 a.u., corresponding to strong hydrogen bonding interactions. These are predominantly localized near the nitrogen and oxygen atoms, with RDG values peaking at approximately 1.5, indicating significant stabilization through hydrogen bonding.
Green regions, centered near sign(λ2)ρ = 0 a.u. and RDG values between 0.5 and 1.2, reflect the presence of van der Waals forces, particularly across the aromatic rings and methoxy groups.
Red regions, observed at positive sign(λ2)ρ values up to 0.020 a.u., signify steric repulsion, but their relatively limited intensity (RDG < 0.5) suggests that the molecular geometry of HBHC efficiently mitigates unfavorable crowding at the interface.
Upon protonation, the HBHC molecule exhibits an intensification of blue regions in the RDG plot, with sign(λ2)ρ values extending to −0.025 a.u. and RDG maxima approaching 1.8. This enhancement indicates stronger and more numerous hydrogen bonds, especially around the protonated nitrogen centers, which act as effective electrophilic sites for interaction with the electron-rich steel surface.91 The green van der Waals regions remain well distributed, supporting stable adsorption, while the red steric repulsion zones do not increase significantly, confirming that protonation does not introduce substantial spatial hindrance. Quantitatively, the RDG isovalue range from −0.035 to 0.020 a.u. across both forms of HBHC demonstrates a balanced interplay between attractive and repulsive interactions. The predominance of blue and green regions, coupled with the limited presence of red, underscores the favorable adsorption profile of HBHC, ensuring robust surface coverage and effective inhibition of corrosive species. The analysis confirms that both neutral and protonated HBHC molecules form stable, protective layers on carbon steel by maximizing hydrogen bonding and van der Waals interactions while minimizing steric repulsion. The protonated form, in particular, enhances adsorption strength, which is advantageous for corrosion inhibition in acidic environments. These findings are consistent with the observed trends in the RDG plots and support the molecular design principles for efficient organic corrosion inhibitors.
Finally, there is a retro-donation process where electrons from the Fe d-orbitals are back donated into the π* antibonding orbitals of HBHC (d → π*), shown as dashed green lines, which further stabilizes the metal–inhibitor complex by strengthening the interaction with a partial covalent character. On the other hand, physical adsorption involves electrostatic attractions between the positively charged Fe2+/Fe3+ ions on the metal surface and the negatively charged or Polar regions of the inhibitor, such as lone pair-bearing atoms or deprotonated groups. Red dashed lines depict these interactions. Although weaker than chemisorption, physisorption can facilitate initial binding, especially in the early stages of inhibition. The diagram clearly distinguishes the adsorption behavior in the neutral versus protonated forms of HBHC. Protonation enhances the electron-withdrawing character of the molecule, often increasing the molecule's dipole moment and solubility in acidic environments (such as 1 M HCl), thereby improving its adsorption efficiency.
In the protonated form, new sites become positively charged, promoting stronger electrostatic interactions with the metal surface and possibly improving hydrogen bonding with adsorbed water molecules. Overall, this dual adsorption mechanism, involving both donor–acceptor interactions (chemisorption) and electrostatic forces (physisorption), enables HBHC to form a stable protective layer on the iron surface. This layer effectively isolates the metal from the corrosive environment, reducing both anodic and cathodic reactions and confirming HBHC's potential as a high-performance corrosion inhibitor in acidic media.
Property | HBHC | HBHC+ |
---|---|---|
a DILI: Drug-Induced Liver Injury.b Ames: Ames Mutagenicity Test.c LC50FM: Lethal Concentration 50% – Fathead Minnow.d LC50DM: Lethal Concentration 50% – Daphnia magna.e log![]() |
||
Human toxicity | Low | Low |
Hepatotoxicity: 0.029 | Hepatotoxicity: 0.033 | |
DILI: 0.77 | DILI: 0.913 | |
Ames: 0.417 | Ames: 0.407 | |
Aquatic toxicity | Moderate | Low |
Bioaccumulation: 0.586, LC50FM: 4.58, LC50DM: 4.318 | Bioaccumulation: 0.593 LC50FM: 4.474, LC50DM: 4.316 | |
Water solubility (log![]() |
Good (−2.748) | Good (−2.945) |
Chemical stability | High | High |
Stable under operational conditions, no instability alerts | Stable under operational conditions, no instability alerts | |
Bioaccumulation | Low | Low |
Bioaccumulation factor: 0.586 | Bioaccumulation factor: 0.593 |
Regarding systemic distribution, both molecules display high plasma protein binding (>90%) and a strong likelihood of blood–brain barrier penetration. Predicted clearance rates (7.5–8.5 mL min−1 kg−1) indicate moderate elimination. In terms of toxicity, HBHC and HBHC+ show potential risks of mutagenicity (Ames positive probability ∼0.4), drug-induced liver injury (DILI), and skin sensitization, while hepatotoxicity and carcinogenicity probabilities are relatively low. Both compounds also show low predicted human toxicity overall, supporting worker safety and suitability for industrial use.
From an environmental and operational perspective, HBHC and HBHC+ demonstrate good water solubility (logS between −2 and −3), chemical stability under operational conditions, and low bioaccumulation, ensuring effective dispersion, uniform metal surface coverage, long-term corrosion protection, and minimal long-term environmental impact.93,94 HBHC+ exhibits slightly lower aquatic toxicity and bioaccumulation, making it more environmentally favorable in contexts where discharge into water systems is possible.
Overall, the ADMET predictions indicate that HBHC and HBHC+ possess drug-like physicochemical profiles, moderate bioavailability, and acceptable safety margins, with certain toxicity alerts (DILI, skin sensitization) that require further experimental validation. Both compounds meet key criteria for operational performance, worker safety, and environmental protection, making them effective, stable, and safe corrosion inhibitors for industrial applications, with HBHC+ preferred in environmentally sensitive contexts.
The promising anticorrosion performance of HBHC underscores its potential for practical applications in acidic environments. Its high efficiency at low concentrations and long-term stability make it suitable for protecting steel pipelines and equipment in oil and gas industries, as well as in chemical processing systems. Furthermore, its favorable ADMET profile suggests that HBHC can serve as a safer and more sustainable alternative to conventional toxic inhibitors.
Supplementary information (SI): additional results on predicted quantitative ADMET properties of HBHC and its protonated form (HBHC+). See DOI: https://doi.org/10.1039/d5ra05876g.
This journal is © The Royal Society of Chemistry 2025 |