Home Tribological behavior of stainless steel in sulfuric acid in the presence of Thymus zygis subsp. gracilis essential oil: experimental and quantum chemical studies
Article Publicly Available

Tribological behavior of stainless steel in sulfuric acid in the presence of Thymus zygis subsp. gracilis essential oil: experimental and quantum chemical studies

  • Mohamed Ouknin , Amal Boumezzourh , Zouhair Lakbaibi , Pierre Ponthiaux , Jean Costa and Lhou Majidi
Published/Copyright: April 12, 2021

Abstract

To reduce the use of synthetic compounds as wear-corrosion inhibitors, and substitute them with new ecological compounds, we are interested in evaluating the effect of the essential oils against the tribocorrosion. The present paper describes the effect of adding Thymus zygis subsp. gracilis (TZ) essential oil (3 g·L−1) on AISI 304L tribocorrosion behavior in 0.5 M H2SO4. As a result, the addition of this essential oil decreases the tribocorrosion rate [E (%) = 84.38], the value of friction coefficient, and the energy dissipated during sliding wear, in comparison to those recorded for dry environment and 0.5 M H2SO4. In addition, the essential oil improves the surface morphology of metal as shown by the scanning electron microscopy and energy dispersive X-ray and the three-dimensional optical profiler analysis. In addition, the modeling approaches of experimental results, involving the density functional theory, the electrostatic potential, and the Monte Carlo methods, show that thymol (42.5% of TZ oil) is the best corrosion and tribocorrosion inhibitor that adhered to the stainless steel surface and offers a greater coverage area to block the metal wear-corrosion. According to the obtained results, the TZ essential oil can be used in the food industry to prevent the wear of materials.

1 Introduction

One of the major problems encountered by the food industry is the presence of metallic debris (3rd body) in large quantities contained in the final product. In fact, food products contaminated by metallic debris, particularly hard and/or sharp or pointed metallic debris, represent a real danger to consumers in the event of accidental ingestion. These foreign bodies become the primary source of consumer complaints in the food industry. This debris is the result of the metals degradation during the production process, this phenomenon is known as tribocorrosion (Ponthiaux et al. 2012; Sun and Bailey 2014). The tribocorrosion is a synergy between the surface mechanical stress and the environment effect that can affect the nature of the degradation mechanism and its amplitude (Ponthiaux et al. 2012; Sun and Bailey 2014). A separate knowledge of tribological and electrochemical behaviors, respectively, in the absence of aggressive medium or in the absence of mechanical impact is not sufficient to derive the tribocorrosion behavior of material torque system (Benea et al. 2009; Berradja et al. 2006; Chen et al. 2015).

Owing to its passivity, stainless steel (304L) is widely used in the food industry, as it meets all the requirements of this sector. The resistance of this metal against the chemical attack of corrosive products is because of its ability to protect itself by spontaneous formation of chromium oxides and hydroxides complex film on its surface, called the “passive layer”. The last one protects the metal substrate from generalized and localized corrosion (Hsissou et al. 2020a; Ouknin et al. 2019a; Saada et al. 2018). However, in some instances, in acidic environments, the corrosion-resistant stainless steels lose their properties (Saada et al. 2018; Yang et al. 2018). The electrochemical and mechanical behavior of the material to the tribocorrosion phenomenon will be evaluated from a mechanical point, such as the measurement of weight loss, the measurement and the evaluation of the energy dissipated in the contact area, the measurement and evaluation of the friction coefficient. In the same way, the nature of the wear mechanism and its comparison with friction in the standard humidity-controlled environment will be retained to highlight the possible modification of the wear mechanism associated with the properties of the studied environment. For the electrochemical aspect, the use of intensity-potential curves allows the evaluation of the potential domains where the essential oil acts, by measuring and monitoring the evolution of the free potential, representing the operating behavior of the system. Therefore, the research and application of green and new high-efficiency corrosion inhibitors have become a requirement to prevent corrosion and wear of stainless steel in an acidic environment (Ouknin et al. 2020a; Yang et al. 2018). In addition, the use of organic inhibitors and epoxy polymer appear a very effective solution to protect metals and alloys from corrosion and wear-corrosion (Hsissou et al. 2020b,c; Lavanya et al. 2019).

To substitute the synthetic lubricant compounds currently used in the food industry with new natural lubricants as tribocorrosion inhibitors, a limited number of research papers have been published. However, Ding et al. (2020), show that fatty acids with phosphorus and sulfur are good organic additives that can reduce friction and wear. Through their oils and extracts, aromatic and medicinal plants may represent a safe solution, because of their richness in organic chemical substances. The study on organic compounds of Thymus willdenowii and Populus nigra var. italica propolis, shows that these compounds can be used against the wear of materials in aggressive solutions in the presence of a mechanical effect (Ouknin et al. 2019a, 2020a). This work constitutes the first step to identify and evaluate the inhibitory properties of Thymus zygis subsp. gracilis essential oils against corrosion and wear resulting from the phenomenon of tribocorrosion, as part of a wider reflection, to replace the lubricants currently used in the food industry with new compatible organic lubricants whose synthesis will involve the chemical industry in the production process. The genus Thymus is abundant species and subspecies from the Lamiaceae that is used as food preservative offering a long shelf-life to dairy products, especially butter. In addition, it is used for its antiseptic, antitussive, expectorant, and carminative proprieties in folk remedies (Dandlen et al. 2010; Dob et al. 2006).

Besides, the modeling approaches such as DFT, ESP, and MC were used to evaluate the tribocorrosion inhibition and adsorption behavior (TIA) of the major components of TZ essential oil on the iron surface (110) (the main constituent of AISI 304L) in 0.5 M H2SO4 as corrosive solution medium. This computational finding allows providing supportive highlights to choose one of the most major constituents of TZ essential oil that will be considered as a good inhibitor of the tribocorrosion of stainless steel surface in 0.5 M H2SO4. The tribocorrosion inhibition process was carried out in acidic solution and consequently, the protonation of inhibitors with both lone pair of electrons and the more nucleophilic sites of carbon atoms will be probably possible; for this, the calculations of nonprotonated forms will be discussed, and then compared to those of the protonated forms.

The goal of this study is to find a new ecologic tribocorrosion inhibitor. For this, we are interested by the evaluation of the T. zygis subsp. gracilis essential oil effect on AISI 304L surface subjected to tribocorrosion in 0.5 M H2SO4 medium using the electrochemical analysis. The three-dimensional optical profiler and scanning electron microscopy and energy dispersive X-ray (SEM-EDX) are used to evaluate the surface morphology. Moreover, the Density Functional Theory (DFT), the Electrostatic Potential (ESP), and the Monte Carlo (MC) methods are used to understand the TIA behavior of TZ essential oil compounds on the AISI 304L surface.

2 Materials and methods

2.1 Essential oil analysis

The TZ essential oil was extracted by hydrodistillation using Clevenger-type apparatus according to the method recommended in the European Pharmacopoeia (1997). The extracted essential oil was analyzed using a PerkinElmer Turbo Mass quadrupole-detector, coupled to a Perkin-Elmer 88 Auto system X, using Helium as carrier gas (1 mL/min). The ion source temperature was 150 °C; the temperature of the oven was programmed from 60 to 230 °C at 2 °C/min and then held isothermally at 230 °C (35 min). The injector temperature was fixed at 280 °C, the energy ionization at 70 eV, electron ionization mass spectra were acquired over the mass range 35–350 uma, split: 1/8, injection volume: 0.2 μL of pure oil.

2.2 Preparation of materials

The coupons studied (AISI 304L) with a surface of 30 × 30 mm2 and thickness of 1 mm, that its chemical composition is shown in Table 1, was polished with several grades of alumina pastes down to 1 µm to obtain a mirror surface (Ra = 0.01 µm), using Buehler Motopol 2000 grinder/polisher machine. Then, rinsed with bidistilled water, degreased in acetone into an ultrasonic bath for 5 min, rinsed with bidistilled water and dried before its uses. The 0.5 M H2SO4 medium was prepared by diluting the H2SO4 analytical grade 95% with bidistilled water. The concentration range of TZ essential oil was chosen upon the maximum solubility that was 3 g·L−1.

Table 1:

Chemical composition of the stainless steel AISI 304L.

Element Cr Ni Si P Fe
Weight (%) 18.51 7.80 0.55 0.08 73.06

2.3 Tribological tests

Before all tests, the alumina balls Al2O3 with a diameter of 10 mm should be rinsed with ethanol. The tribocorrosion tests were conducted at 23 °C. The wear in a dry environment was carried out in ambient air with a relative humidity of 50% HR. For the environment of 0.5 M H2SO4 with and without the addition of 3 g·L−1 of TZ essential oil, the tribocorrosion tests are conducted at room temperature using electrolyte volume of 20 mL with pH 0.2. The AISI 304L surface of 1 cm2 was used as a working electrode (WE). The Ag/AgCl (3 M KCl) represents the reference electrode (RE) and the Pt electrode is used as a counter electrode.

The experimental system used during tribocorrosion tests is schematized in Figure 1. The current measurements and voltage are measured at a resolution of 1 pA and 1 µV, using potentiostat Solartron with electrochemical interface model 1287. After 10 min of coupon immersion into the test solution, the potential results were obtained for 1000 s at 1 Hz. At the end of the tribocorrosion tests, the counterbody was immediately removed from the WE. The electrochemical parameters obtained are compared to ASTM conventions (1994).

Figure 1: 
						Schematic drawing of electrochemical measurement technique connected to the fretting machine.
Figure 1:

Schematic drawing of electrochemical measurement technique connected to the fretting machine.

The tribocorrosion sliding tests were recorded at a normal force of 2 N at 1 Hz, and displacement amplitude of 1 mm, the equipment test used was reported earlier by Mohrbacher et al. (1995). The tribocorrosion tests were evaluated at 23 °C for 10,000 cycles at equally spaced time increments over the whole test duration, the number of cycles, tangential force, normal force, amplitude displacement, and the friction coefficient were logged.

The weight loss measurements were recorded before and after the tribological tests using an analytical balance (accuracy ± 0.0001 mg).

The measure of weight loss is computed according to Equation (1):

(1)Δm=(m0mf)

with, m0 and mf are the coupon weights before and after tribological tests, respectively. The inhibition efficiency IE(%) is determined according to Equation (2):

(2)IE(%)=(Δm0.5 M H2SO4ΔminhΔm0.5 M H2SO4)×100

With, Δm0.5 M H2SO4 and Δminh represent the weight loss in 0.5 M H2SO4 without and with essential oils, respectively.

The electrochemical measurements were recorded using an EG&G Model 273A potentiostat controlled by the CorrWarre software, at a sweep rate of 1 mV/s and the measurement sweep interval was −800 to −200 mV/ESC.

3 Surface and coating characterization

3.1 Scanning electron microscopy and energy dispersive X-rays (SEM-EDX)

The wear tracks morphology was analyzed after the sliding tests, using a Philips scanning electron microscope (XL 30 FEG) connected to an electron dispersive X-ray (EDX) analysis device. Volumetric wear was measured by noncontact white light interferometry analysis using Wyco NT3300 optics profilometer with Vision software (version 2.210).

3.2 Characterization of surface topography and roughness

The surface topography and roughness of wear track on coupons were analyzed by confocal microscopy (NanoFocus µsurf®), comprising a light source (LED), a rotating multi-pinhole disc, an objective lens with a piezo drive, and a CCD camera.

4 Computational procedures

4.1 DFT and ESP calculations

Gaussian 09 program, Revision-D.01 (Frisch et al. 2009) was used to obtain the optimized geometries of TZ essential oil main components such as thymol, p-cymene, γ-terpinene, and borneol using the DFT method (Gece 2008; Obot et al. 2015; Verma et al. 2016) combined to the basis 6-311G++(d, p) (Hariharan and Pople 1973) and the functional B3LYP (Hehre et al. 1972). For these compounds, the distribution densities of HOMO (occupied molecular orbital) and low unoccupied molecular orbital (LUMO) were produced. In addition, we are interested in determining the quantum parameters (QPs) such as the energy of HOMO (EHOMO), the energy of LUMO (ELUMO), the energy gap ΔEinh between EHOMO and ELUMO energies (ΔEinh = ELUMO – EHOMO) (Geerlings et al. 2003; Hsissou et al. 2019; Pearson 1988), the global chemical potential μinh (μinh = 0.5 [EHOMO – ELUMO]) (Benallou et al. 2019; Singh et al. 2018; Vanasundari et al. 2017) and the hardness ηinh (ηinh = ΔEinh) (Benallou et al. 2019; Singh et al. 2018; Vanasundari et al. 2017). The fraction of electrons transferred ΔN from each compound as an inhibitor to bulk iron surface (110), is calculated according to Pearson relationship as follows:

(3)0.5(φFe(110)+μinh)(ηFe(110)+ηinh)1.

where φFe(110) and ηFe(110) are known as the working function and the absolute hardness of bulk iron surface (110), respectively, and then ηinh and μinh are the global chemical potential and the global hardness of inhibitor, respectively (Lukovits et al. 2001; Martinez 2003; Sastri and Perumareddi 1997).

The values of φFe(110) and ηFe(110) are 4.82 and 0 eV, respectively (Kokalj 2012; Parr and Pearson 1983; Tang et al. 2013). Furthermore, the electron back donating character of inhibitor (ΔEbd-inh) is another descriptor that is used to describe the reactivity between metal surface and adsorbent; this parameter is expressed as −ηinh/4 (Olasunkanmi et al. 2016).

Additionally, the reactive centers of the concerned compounds are estimated from three approaches, the spreading analysis of HOMO and LUMO densities (DFMO), the ESP analysis in which: the red color reflects the strong negative electrostatic potential (EP); yellow for a moderately negative EP; blue reflects the strong positive EP and green for the moderately positive EP (Goulart et al. 2013). And, from the computation of electrophilic attacks P+ and nucleophilic attacks P indexes in terms of natural bond orbitals analysis (Domingo et al. 2013).

4.2 Metropolis Monte Carlo (MC) simulations

The adsorption simulation of each inhibitor on the iron surface (110) into corrosive medium (200 of H2O, 20 H3O+, and 10 SO42−) was explored using the adsorption locator module and the COMPASS II force field integrated into the Materials studio version 8 program package (Sun 1998). The single-crystal Fe (110) surface is built using sketching tools and optimized with the smart minimizer using DMolˆ3 at the DNP + basis and B3LYP functional. The single crystal plane (110) was extended to the multi-surface area of Fe (110), using the supercell containing 10 layers with a vacuum layer of 30 Å along the C-axis. The system inhibitor/solution/Fe (110) was simulated in a simulation box of 17.89 × 17.89 × 38.34 Å with periodic boundary conditions. Concerning the crystallographic surface Fe (110) plane, it was chosen for its densely-packed surface (Vanasundari et al. 2017).

5 Results and discussion

5.1 Essential oil chemical composition

To further our research on the activities of TZ essential oil, we are interested to assess the tribocorrosion behavior of this essential oil. Our previous study showed that thymol (42.5%), p-cymene (23%), γ-terpinene (8.9%), and borneol (4.8%) are the main compounds of TZ essential oil (Figure 2) (Ouknin et al. 2020b).

Figure 2: 
						Molecular structures of the main components of TZ essential oil.
Figure 2:

Molecular structures of the main components of TZ essential oil.

5.2 Weight loss measurement

TZ essential oil’s effect on the tribocorrosion behavior of AISI 304L in 0.5 M H2SO4 was studied by weight loss measurements. The weight loss values (Δm) and the inhibition efficiency E (%) are pooled in Table 2.

Table 2:

TZ essential oil effect against tribocorrosion behavior of AISI 304L in three environments.

Test condition Δm (mg) IE (%)
Dry condition 485.20 ± 3.50
0.5 M H2SO4 715.40 ± 4.10
0.5 M H2SO4+ 3 g·L−1 TZ oil 111.75 ± 2.50 84.38 ± 1.13

It is quite clear from the obtained results that in the presence of TZ essential oil, weight loss decreases and the inhibition efficiency is 84.38%. However, this can be attributed to the adsorption of TZ essential oil on the stainless steel surface, thereby blocking the reaction sites and protecting the surface against wear and corrosion (Boumezzourh et al. 2020).

5.3 Friction coefficient results (COF)

Figure 3 provides the AISI 304L friction coefficient (µ) behaviors, as a function of cycle numbers and the environment nature, at frequency of 1 Hz, a cycle numbers of 10,000 cycles, and linear displacement amplitude of 1 mm. The three environments tested are, the first in dry conditions, the second in 0.5 M H2SO4, and the third in 0.5 M H2SO4 with the addition of TZ essential oil (3 g·L−1).

Figure 3: 
						Evolution of the friction coefficient of the tribological torque AISI 304/Al2O3 in the presence of three environments: dry conditions, 0.5 M H2SO4 and 0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1) at 1 Hz, 1 mm, and 10,000 cycles for the normal force (2 N).
Figure 3:

Evolution of the friction coefficient of the tribological torque AISI 304/Al2O3 in the presence of three environments: dry conditions, 0.5 M H2SO4 and 0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1) at 1 Hz, 1 mm, and 10,000 cycles for the normal force (2 N).

The report of the tangential effort Ft (N) on the normal effort Fn (N) allows us to define the friction coefficient value (µ) according to Equation (4).

(4)μ=FTFN

According to Figure 3, it is clear that the evolution of friction coefficient in the function of cycle numbers present two domains. The first domain represents the “running-in” regime, it concerns about 1000 cycles, and this domain is associated with the surface adaptation that is characterized by a low number of mechanical stresses and the surface roughness. In this domain, the friction coefficient value increases significantly as will the cycle numbers. It is also clear that the friction coefficient evolution is relatively sloping in the third environment with essential oil (3 g·L−1), for those in the first and second environment (dry and 0.5 M H2SO4), they follow the same order as the third environment (3 g·L−1 of TZ oil). The second domain is presented between 1000 and 10,000 cycles, this one provides information about the friction coefficient evolution, is named: ‘usage’. In this domain, it is quite clear that the evolution of the friction coefficient value is less fast than the one of the preceding domains. In addition, the evolution of the friction coefficient value is the lowest in the environment of 0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1) and appreciably of the same nature in the two other environments (0.5 M H2SO4 and in dry condition (HR = 50%)).

The lowest friction coefficient value obtained in the environment (0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1)) can be associated with the lubricant effect related to the addition of TZ essential oil (3 g·L−1) to the sulfuric medium 0.5 M. It is also interesting to note that the values of the identified friction coefficient, for the tribological torque AISI 304/Al2O3 in the dry environment (HR = 50%), are similar to those previously reported (Ouknin et al. 2019a, 2020a).

5.4 Dissipated energy evolution during friction

The evolution of dissipated energy Ed (µJ) in the function of the cycle number and the nature of the environment, during the friction test, are presented in Figure 4. The values of dissipated energy for a cycle under total imposed slip condition are obtained by processing the tangential force Ft (N) as a function of the displacement δ. The curves obtained have generally quadratic form, a parallelogram. The area of the parallelogram obtained presents the dissipated energy in the friction test (Benea et al. 2014; Vincent 1994).

Figure 4: 
						Evolution of the dissipated energy in the contact of the tribological torque (AISI 304/Al2O3) in the course of friction and in the presence of three environments: dry conditions, 0.5 M H2SO4 and 0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1) at 1 Hz, 1 mm, 10,000 cycles for the normal force (2 N).
Figure 4:

Evolution of the dissipated energy in the contact of the tribological torque (AISI 304/Al2O3) in the course of friction and in the presence of three environments: dry conditions, 0.5 M H2SO4 and 0.5 M H2SO4 in the presence of TZ essential oil (3 g·L−1) at 1 Hz, 1 mm, 10,000 cycles for the normal force (2 N).

The pace of the three curves shows that the energy dissipated in the contact is higher in domain 1, concerning 1000 cycles corresponding to the surface adaptation. In the domain 2 range of 1000–10,000 cycles, a slight decay of slope is observed in the environment (0.5 M H2SO4 with 3 g·L−1 of TZ essential oil). However, the slope obtained in dry (HR = 50%) and 0.5 M H2SO4 environments are more accentuated and seem to be approximately in the same order of magnitude.

The pace of the three curves indicates a decreasing development of dissipated energy in the function of the number cycles, that reflects the facilitation of sliding by increasing the cycle number, regardless of the environment concerned.

5.5 Potentiodynamic polarization measurements

Figure 5 presents the potentiodynamic polarization curves of AISI 304L coupons in 0.5 M H2SO4 with and without the addition of 3 g·L−1 of TZ essential oil at 298 K. Table 3 gathered the kinetic parameters including corrosion current density (Icorr), the corrosion potential (Ecorr), cathodic slopes (βc) also the inhibition efficiency (IE %).

Figure 5: 
						Polarization curves for AISI 304L in the absence and presence of TZ essential oil in 0.5 M H2SO4.
Figure 5:

Polarization curves for AISI 304L in the absence and presence of TZ essential oil in 0.5 M H2SO4.

Table 3:

Electrochemical parameters of AISI 304L in 0.5 M H2SO4 in the absence and presence of TZ essential oil (3 g·L−1).

Solution E corr (mV) I corr (mA/cm2) βc (mV) IE (%)
0.5 M H2SO4 463 4.32 109.7
TZ oil (3 g·L−1) 488 0.70 165 83.80

As shown by the polarization curves, the current density value decreases in the cathodic and anodic domains in the presence of 3 g·L−1 TZ essential oil. Moreover, the presence of essential oil influences cathodic reactions and the hydrogen reaction mechanism, traduced by the modification of the cathodic Tafel slopes values (βc).

The displacement of Ecorr gives an idea about the inhibitor type, a displacement higher than 85 mV indicates that the inhibitor is cathodic or anodic type. For the one less than 85, it indicates that the inhibitor acts as a mixed inhibitor type. In our case the displacement of Ecorr is 52 mV, indicating that TZ essential oil acts as a mixed inhibitor type (Hsissou et al. 2020b,d; Ouknin et al. 2018, 2020b).

From the data in Table 3, it can be seen that in the presence of 3 g·L−1 of TZ essential oil, the value of the inhibition efficiency (IE %) is 83.80% and that the corrosion current densities (Icorr) decrease considerably.

5.6 Electrochemical behavior during reciprocating sliding tests

The open-circuit potential behavior (OCP) before, during, and after the tribocorrosion test is presented in Figure 6. The studied test is evaluated in the frequency of 1 Hz with displacement amplitude of 1 mm, 10,000 cycles for the normal force (2 N) in 0.5 M H2SO4 without and with the addition of 3 g·L−1 of TZ essential oil.

Figure 6: 
						Evolution of the open circuit potential recorded before, during, and after fretting tests of TZ oil in sulfuric middle at 1 Hz, 1 mm, and 10,000 cycles for the normal force (2 N).
Figure 6:

Evolution of the open circuit potential recorded before, during, and after fretting tests of TZ oil in sulfuric middle at 1 Hz, 1 mm, and 10,000 cycles for the normal force (2 N).

From Figure 6 it is clear that the OCP move to the more cathodic values and is established during the friction test, regardless of the environment tested. On the other hand, the most cathodic values are measured in a 0.5 M H2SO4 environment without the addition of essential oil, the addition of 3 g·L−1 TZ essential oil shows that the OCP range above the previous with 100 mV.

In the environment of 0.5 M H2SO4, the quasi-stationary values of the potential are caused by the surface layer elimination. At the end of the contact, a part of the surface area is totally or partially destroyed on the contact areas. At the end of the sliding test, it is interesting to note that there is a reformation of the surface layer, which is responsible for the value OCP increasing sharply from the one measured before the friction test.

For the environment 0.5 M H2SO4 with the addition of 3 g·L−1 of TZ essential oil, the quasi-stationary value of OCP shows the same trend as the one observed in 0.5 M H2SO4 only. The gap of 100 mV observed in the OCP compared to the one of the 0.5 M H2SO4 environment indicates that the passage of corundum ball (Al2O3) on the contact area does not destroy the surface layer formed and/or the one of essential oil adsorbed on the surface. Concerning the oscillations of OCP during friction (Curve (2) in Figure 6) in the presence of TZ essential oil, they can be attributed to the adsorption or partial removal (depassivation/repassivation) of the essential oil components adsorbed on the metal surface.

5.7 Periodic depassivation and repassivation

The periodic fluctuation of the passive layer thickness during the intermittent sliding is reflected by the OCP cyclic evolution (Figure 7). The OCP drops during sliding and rises during nonsliding periods (toff = 200 s) because it is the result of coupling the passive and depassivated areas.

Figure 7: 
						Evolution of the OCP of AISI 304L with time intermittent (toff = 200 s) sliding tests.
Figure 7:

Evolution of the OCP of AISI 304L with time intermittent (toff = 200 s) sliding tests.

It is clear from Figure 7 that the OCP move to the cathodic value during the sliding. At the end of the sliding test (5 s), it is interesting to indicate that there is the reformation of the surface layer, which is responsible for the raise of the OCP value compared to the one measured before the friction test. Probably, the load of 2 N did not completely destroy the passive film or there was still enough time for the repassivation of the surface, they can be attributed to the adsorption or partial removal of TZ essential oil components adsorbed on the metal surface (Bratu et al. 2007; Ouknin et al. 2020a).

5.8 Wear mechanisms

Figure 8 represents the different features of wear obtained after 10,000 cycles in the three environments tested.

Figure 8: 
						SEM-EDX after reciprocating sliding tests performed of stainless steel (AISI 304): (a) in ambient air, (b) 0.5 M H2SO4 and (c) in 0.5 M H2SO4 with TZ essential oil (3 g·L−1 TZ).
Figure 8:

SEM-EDX after reciprocating sliding tests performed of stainless steel (AISI 304): (a) in ambient air, (b) 0.5 M H2SO4 and (c) in 0.5 M H2SO4 with TZ essential oil (3 g·L−1 TZ).

As shown in Figure 8a the friction in dry condition with relative humidity (HR = 50%), there is the presence of the debris (3rd body), reflecting the wear adhesive mechanism, the debris was retained in the contact area or it was ejected outside the contact zone. However, in a sulfuric environment without essential oil (Figure 8b), a physicochemical attack on the AISI 304L surface as shown in Figure 8b, reflecting the tribochemical wear mechanism, that indicates the difference between the two environments in term of the destroyed surface.

Furthermore, the wear mechanism in the presence of TZ essential oil (3 g·L−1) is different than those observed in the dry and in sulfuric acid without the addition of essential oil. The last one is purely wear abrasive without 3rd body. On the other hand, Figure 8c indicates the presence of the parallel striations to the friction direction.

In the 0.5 M H2SO4 and dry (50% HR) environments, the observation of the features of wear shows that we are in the presence of a wear mechanism, this is confirmed by the high value of the friction coefficient and the decrease of dissipated energy (Figure 4). The high value of the friction coefficient is because of the adhesive type opposing to the movement after the connections rupture, the mechanism involves the 3rd body, facilitate the sliding is one of its consequence identified. In the presence of TZ essential oil, the lubricant aspect of the environment explains the low values of the friction coefficient and those of the dissipated energy. The EDX spectrum of the metal surface in three environments shows that in the presence of TZ essential oil, there is an oxygen bond corresponding to the presence of phenolic compounds adsorbed on the metal surface, as shown at a concentration of 3 g·L−1 the surface is improved compared to that in the absence of the inhibitor.

5.9 Wear volume loss

From the equation of half-ellipse area, the worn area has been calculated using the following Equation (5):

(5)VW=(π×R×r2)δ

R and r represent the large and small radius of the ellipse, respectively. The volume worn is obtained by multiplying the area of the ½ ellipse by the length of the trace rubbed either, δ = 1 mm.

The different depths of wear obtained after 10,000 cycles in the various environments studied are gathered in Figure 9.

Figure 9: 
						(a, b, and c): Wear volume obtained after 10,000 cycles in the three environments studied.
Figure 9: 
						(a, b, and c): Wear volume obtained after 10,000 cycles in the three environments studied.
Figure 9:

(a, b, and c): Wear volume obtained after 10,000 cycles in the three environments studied.

From Figure 9, we note that the maximum wear volume of 3754.14 ± 65.20 µm3 corresponds to the dry environment Air-HR = 50%, followed by that of 0.5 M H2SO4 with a volume worn of 2817.94 ± 47.80 µm3. However, in the presence of TZ essential oil in a sulfuric medium, the volume worn was approximately 1484 ± 21.13 µm3. This reduction of AISI 304L surface damage in the presence of essential oil can be because of the presence of oxygen atom in the functional groups of essential oil compounds, that are generally known by their anticorrosion activity. Those compounds could inhibit the chemical and mechanical effect on the AISI 304L surface, according to the two possible mechanisms. The first mechanism is named chemisorption; in this mechanism, the neutral molecules are adsorbed on the AISI 304L surface and sharing electrons between oxygen atom and iron. Concerning the second mechanism, the organic molecules of TZ essential oil are adhered to the metal/solution interface and decrease the surface area in which the anodic and cathodic reaction takes place, this mechanism can be because of the difficulty of the protonated molecules to approach the positive charge of stainless steel surface (Boumezzourh et al. 2019). The lubricating aspect of the essential oils allows the protection of the surface layer formed and/or the essential oil adsorbed on the surface during friction, which explains the reduction of wear volume in the presence of TZ essential oil (Bratu et al. 2007).

6 Computational study of TIA behavior for the major components of TZ essential oil

The wear-corrosion scientists have turned their interest to describe the TIA behavior of organic compounds on a metal surface using several theoretical approaches, such as the DFT, the ESP, and the MC methods (Verma et al. 2018). The importance of these methods is focused on finding the good correlation between the molecular structure and the inhibition efficiency. In this respect, the modeling approaches such as DFT, ESP, and MC are used to evaluate the TIA behavior of the TZ essential oil major components toward the iron surface (110) in aggressive solution (0.5 M H2SO4). In our previous study, we found that there is no interaction between molecules of TZ essential oil, and the adsorption behavior is governed by a charge transfer process (Ouknin et al. 2020b). Hence, the computational study aimed to highlight the physicochemical properties of TZ essential oil main components and to understand the molecular characteristics that distinguish one compound from others.

The major components of TZ essential oil such as p-cymene (E1), borneol (E2), γ-terpinene (E3), and thymol (E4) and their protonated forms E1H+, E2H+, E3H+, and E4H + are optimized and exposed in Figure 10. The optimization calculations are confirmed by the presence of zero negative frequency in the outputted z-matrix. Further, we are interested in this study by evaluating the aromaticity role of E1 and E4, on their reactivity and adsorption behavior toward Fe (110) surface. The E1H+ and E4H + geometries without aromaticity were designed as E1H + WA and E4H + WA, respectively (Figure 10).

Figure 10: 
					Optimized structures of the major components of TZ essential oil.
Figure 10:

Optimized structures of the major components of TZ essential oil.

6.1 Density functional theory calculations

6.1.1 QP parameters of structures

The HOMO and LUMO orbitals are very important tools to describe the global reactivity of systems with unsaturated bonds and/or heteroatoms such as oxygen, nitrogen, and sulfur, and, the reactive functional groups, for example, –NH, –N=N–, –C=C–, C=O, –CN, –OH, =S, aromatic moiety, etc. In this respect, a molecule with a high value of EHOMO, low value of ELUMO, and low value of ΔEinh is more polarizable and it is generally associated to a high chemical reactivity and low kinetic stability. Interestingly, if ΔN is positive and less than 3.6, the compound has a tendency to donate electrons (π and/or lone pair electrons) to the unoccupied orbital (3d) of the iron surface to form coordinate covalent bonds. Besides, the negative value of ΔEbd-inh indicates that the compound is ready to accept free electrons (3d) from the metal, to form retro-donating bonds.

Therefore, the high value of ΔN and μinh and the lower value of ΔEinh reflect that the compound exhibits a strong capacity to share their electrons with the iron surface (110), and consequently the compound is more adsorbed on the iron surface (110). Furthermore, the negative value of ΔEbd-inh means that the compound exhibits a stronger ability to receive electrons from the metal. In this study, the calculated QPs aimed to evaluate global reactivity of nonprotonated forms (Ei) i.e., E1, E2, E3, and E4, and also their protonated forms (EiH+) i.e., E1H+, E2H+, E3H+, and E4H+; the QP parameters are exposed in Table 4 and in Figure 11. Accordingly, the protonation of the studied compounds leads to an increase of ΔN and μinh and a decrease of ΔEinh. This result suggests that the transfer of electrons from the studied compounds to the iron surface is very easy when these compounds are in their protonated form. This indicates the prominent role of the protonation to ensure the formation of adsorbent bonds. It is also worth noting that the retro-donating character of these compounds becomes very important in their protonated forms. All these results point to the fact that the thymol is considered to be a more effective adsorbent onto the iron surface, because of its high value of ΔN (0.82 e), lower value of ΔEinh (2.26 eV), and lower negative value of ΔEbd-inh (−0.57 eV). From all QPs results of the studied protonated compounds, we conclude that the strength of adsorption can be ranged as follow: E4H+ > E1H+ > E2H+ > E3H+. Figure 11 shows that E4H + structure is different from that of the E1H+, by the presence of hydroxyl group (–OH). Consequently, the high adsorption of E4H + on the iron surface, with respect to the E1H + could be attributed to the number of lone pair electrons present on the oxygen atom of the hydroxyl group. Similarly, to investigate the role of aromaticity attached to E4H+ and E1H+, the QPs of E4H+ and E1H + without aromaticity (E4H + WA and E1H + WA) are calculated (Table 4 and Figure 11). The inspection of Table 4 and Figure 11 shows that the destruction of the aromaticity of E4H+ and E1H + causes a strong decrease of ΔN and a great increase of ΔEinh, which indicate clearly the effect of aromaticity on the adsorption of E4H + toward Fe (110) surface.

Table 4:

Calculated QPs for the components as demonstrated in Figure 10.

E HOMO (eV) E LUMO (eV) ΔEinh (eV) μ inh (eV) η inh (eV) ΔN (e) ΔEbd-inh (eV)
E1 −6.20 −0.30 5.90 −2.95 5.90 0.16 −1.48
E2 −6.46 −0.37 6.09 −3.05 6.09 0.15 −1.52
E3 −7.36 −0.38 6.98 −3.49 6.98 0.10 −1.75
E4 −6.14 −0.45 5.69 −2.85 5.69 0.17 −1.42
E1H+ −4.61 −1.59 3.01 −1.51 3.01 0.55 −0.75
E2H+ −4.74 −1.64 3.10 −1.55 3.10 0.53 −0.78
E3H+ −4.12 −0.88 3.24 −1.62 3.24 0.49 −0.81
E4H+ −3.26 −1.00 2.26 −1.13 2.26 0.82 −0.57
E1H + WA −7.85 −0.21 7.64 −3.82 7.64 0.07 −1.91
E4H + WA −7.44 −0.33 7.11 −3.56 7.11 0.09 −1.78
Figure 11: 
							Energy gap ΔEinh, electron back donating energy ΔEbd-inh and fraction of electrons transferred ΔN for all investigated structures.
Figure 11:

Energy gap ΔEinh, electron back donating energy ΔEbd-inh and fraction of electrons transferred ΔN for all investigated structures.

6.1.2 Local reactivity study

6.1.2.1 FMO and ESP analysis

To determine the active regions of the studied compounds, the distribution of HOMO and LUMO densities (FMO) and the depiction of the ESP surfaces are calculated. The ESP analysis is widely used to provide more information about the overall molecular charge distribution and predicting reactive positions of nucleophilic and/or electrophilic attacks. The red color region represents the high negative charges of electrostatic potential (EP), the blue color region represents strongly positive EP, the green region in the ESP surfaces corresponds to a moderate positive charge EP, and the yellow region is moderate negative charge EP. Figure 12 shows the representations of FMO and ESP surfaces related to the compounds Ei and EiH+. From Figure 12, we can see that the protonation of Ei compounds influences the LUMO densities. In contrast, the HOMO densities of these compounds are almost unchanged. For example, for E4 both the HOMO of the benzene ring and HOMO of the hydroxyl group (–OH) are not influenced by the protonation, but it must be noted that the protonation effect generates a high density of LUMO on the HO–benzene ring bond, which is absent in the nonprotonated form of E4. This observation is confirmed by the ESP analysis of E4, in which the density of positive charge is mainly sited around the hydrogen atom of the hydroxyl group (–OH), the last becomes relatively important in its protonated form E4H+. Furthermore, we note that E4H + presents both a strong density of the negative charge and a moderate density of positive charge throughout carbon atoms of the benzene ring. Moreover, for E4H + we can observe a moderate density of negative charge on the oxygen atom of the hydroxyl group (–OH), this result indicates clearly that the protonated form of thymol (E4H+) have a strong active site when this compound is interacting with the iron surface by donor-acceptor interaction to form chemical coordination high bonds. Similarly, the ESP pictures of E2 and E2H+, given in Figure 11, show for both forms that, the region of high negative charges is observed around carbon atoms of the benzene ring, that suggesting the high ability of E1 to donate its electrons to the iron surface. The ESP picture of E3 shows a high negative charge around the two double bonds (C=C) in the cycle ring, which is approximately dispreaded in its protonation form of E3H+; suggesting the low reactivity of E3 toward the iron surface. In other words, the E2 and its protonated form E2H + even if they are not plane, present active sites for the nucleophilic and electrophilic attacks. This reflects evidently the more reactivity of E2 and E2H + compared to E3 and E3H+. Combining with the above results, we conclude that the strong interaction EiH+…Fe surface in this study is related to the protonated form of thymol (E4H+), resulting from its high ability to donate electrons (π and/or lone pair electrons) to the unoccupied orbital (3d) of iron surface, to form coordinate covalent bonds, and accept free electrons (3d) from the iron surface to form retro-donating bonds. Consequently, the thymol will be the appropriate tribocorrosion inhibitor among the compounds under study.

Figure 12: 
								Depictions of HOMO and LUMO densities and ESP surfaces for E1H+, E2H+, E3H+, and E4H + compounds. Note to ESP plots: (red) strong negative electrostatic potential (EP); (yellow) moderately negative (EP); (blue) strong positive (EP); (green) moderately positive (EP).
Figure 12:

Depictions of HOMO and LUMO densities and ESP surfaces for E1H+, E2H+, E3H+, and E4H + compounds. Note to ESP plots: (red) strong negative electrostatic potential (EP); (yellow) moderately negative (EP); (blue) strong positive (EP); (green) moderately positive (EP).

6.1.2.2 Parr functions analysis of E4 and E4H+

To identify the most nucleophilic and electrophilic sites of organic compounds, Parr functions P− and P+ are extensively used to give clear information about atomic sites responsible for the nucleophilic and electrophilic interactions, respectively. However, the high values of P− and P+ correspond to the most nucleophilic and electrophilic centers, respectively. Although, the atoms with negative values (or the values close to zero) of P− and P+ are considered as nonreactive atoms. The values of electrophilic P+ and the nucleophilic P− Parr functions are calculated for the significant atoms of the nonprotonated and protonated form of thymol and regrouped in Table 5. Then, the good nucleophilic and electrophilic centers are characterized by maximum values of P− and P+, respectively, and vice versa. As shown in Table 5, the protonation process modifies the nucleophilic P− and electrophilic P+ Parr functions. Surprisingly, we note an increase in the number of the nucleophilic centers around the benzene ring (Table 5). In other words, we particularly note that the protonation of thymol causes an increase of the nucleophilic character of oxygen atom O7 of hydroxyl group, and also causes a significant increase of the electrophilic character for such carbon atoms of the benzene ring (C2, C3, C4, C5, and C6). Which indicates that the nucleophilic and electrophilic behavior of thymol will be more favored in its protonated form than in the nonprotonated one. In the light of these results, we note that the protonation process increases the nucleophilic and electrophilic attacks throughout the benzene ring and hydroxyl group of thymol toward the iron surface, and ensuring high adsorption between the thymol and Fe (110) surface.

Table 5:

Significant nucleophilic P− and electrophilic P+ Parr functions of the thymol in its protonated (E4H+) and nonprotonated form (E4).

E4 E4H+
Atoms P− P+ Atoms P− P+
C1 0.351 −2.576 C1 0.266 −1.261
C2 −0.163 0.291 C2 0.162 0.992
C3 0.326 −0.050 C3 0.213 0.322
C4 0.159 0.624 C4 0.253 6.927
C5 −0.083 2.028 C5 0.391 3.097
C6 0.210 0.298 C6 0.049 0.992
O7 0.158 0.099 O24 0.206 −0.004
  1. For atom numbering see Figure 10.

6.2 Adsorption behavior analysis

For the studied compounds the planarity is an important factor that should be considered to understand the adsorption mode of these compounds on the Fe (110) surface. In this context, before starting this study, some significant dihedral angles of the protonated compounds are calculated and illustrated in Figure 13. According to this Figure, we note that the dihedral angles calculations around the full atoms of E1H+, E3H+, and E4H + are close to ±0 or 180° (except two methyls of isopropyl group), that signify high planarity of these compounds. However, we evidenced that despite these compounds having a planar area, the E4H+ is one among compounds adsorbing almost on flat orientation onto the iron surface (110) to maximize surface coverage and contact (see later in Figure 14). Thus, is probably because of the presence of hydroxyl group attached to the plane of the benzyl ring moiety of E4H+ (protonated thymol), in this case, the high inhibition efficiency of the protonated thymol could be attributed to the number of lone pair electrons present on the oxygen atom of the hydroxyl group. Otherwise, we note that the p-cymene molecule is relatively neighboring to the Fe (110) surface with respect to the γ-terpinene, which indicates the role of the aromatic ring (delocalization of π electrons) for stronger adsorption. In addition, it appears that the hydroxyl group of the borneol molecule is not adhered to the metal surface, outstanding to the presence enormously of the steric effect caused by the presence of the methyl group as well to the nonplanarity of this molecule, which provides certainly a weak interaction between the borneol and the iron surface.

Figure 13: 
						Dihedral angles of inhibitors given in degree.
Figure 13:

Dihedral angles of inhibitors given in degree.

Figure 14: 
						Side views of stable adsorption configurations of inhibitors in their protonated form on the Fe (110) surface in solution (100 molecules of H2O, 10H3O+, and 5SO42−) at 297.15 K.
Figure 14:

Side views of stable adsorption configurations of inhibitors in their protonated form on the Fe (110) surface in solution (100 molecules of H2O, 10H3O+, and 5SO42−) at 297.15 K.

The nature of inhibitor interactions and their orientation onto Fe surface (110) was studied by the MC method. This technique was performed to predicting the strength and behavior adsorptions of each inhibitor. Side views of stable adsorption configurations of the studied interfaces (inhibitor/100water/20H3O+/10SO42−/Fe (110)) using adsorp-tion Locator modules are given in Figure 14. From this figure, we suggest that the adsorption regions of all inhibitors on Fe (110) are unsaturated double bonds C=C and/or oxygen atom (–O–). To estimate the most adsorption between the inhibitor and iron surface such energies are calculated: the total energy (ET) reports the global energy for the systems under study (inhibitor/solution/Fe(110)), this parameter is calculated by optimizing the whole system. Therefore, the adsorption energy (Eads), which displays the energy, required when the relaxed inhibitors are adsorbed on the Fe surface (110), this parameter is commonly recognized to evaluate the strength and mechanism of adsorption. Then, the rigid adsorption energy (ERads) reflects the energy released, when the unrelaxed inhibitor molecule is adsorbed on the Fe surface (110) before its optimization step. The deformation energy (EDef) is the energy released when the adsorbed inhibitor molecule is relaxed on the Fe (110) surface; another term of energy namely the desorption energy dEads/dNi, represents the required energy when one of the molecule inhibitors was removed from the system of inhibitor/solution/Fe (110) surface.

Table 6 shows that the system of (E4H+/solution/Fe(110)) is the most stable (low value of ET) among the four systems under study. Further, the negative values of Eads mean that the studied adsorption can be spontaneously (Gece 2008). Moreover, we observed that the system of E4H+/solution/Fe(110) shows the high negative value of Eads (−825.45 kcal/mol) comparing to the other systems under study, which indicates a stronger interaction between the thymol inhibitor and the iron surface (110) in the acidic solution (100 H2O, 20 H3O+, 10 SO42−), this is mainly because of the presence of electron-donating group (–OH) attached to the benzene ring characterized by delocalization π-electrons, that serves as supplementary reactive sites of the larger thymol adsorption on Fe (110) surface (Emregul and Atakol 2004). As consequence, we conclude that the thymol is the best tribocorrosion inhibitor among the four compounds under study. The obtained results are in accordance with the experimental results revealing that the adsorption nature of extract TZ essential oil on the stainless mild in 0.5 H2SO4 is a typical chemisorption process.

Table 6:

Total energy ET, adsorption energy Eads, rigid adsorption ERads, deformation energy EDef and necessary energy to exclude inhibitor from the iron surface dEads/dNi, in kcal/mol, are given in solution (100 of H2O, 10 H3O+, and 5 SO42−).

E1H+ E2H+ E3H+ E4H+
E T −636.35 −626.16 −622.19 −641.41
E ads −767.82 −765.70 −684.20 −825.45
E Rads −685.13 −693.93 −678.58 −699.67
E Def −82.69 −71.77 −5.62 −125.78
dEads/dNi −157.60 −148.34 −61.24 −214.20

7 Conclusions

This study reports for the first time the effect of TZ essential oil against the tribocorrosion effect on the stainless steel AISI 304L in 0.5 M H2SO4 medium. The obtained results show that the addition of 3 g·L−1 of TZ essential oil in 0.5 M H2SO4 medium, gives a lubricant aspect to the studied environment that is confirmed by the low values of friction coefficient and the dissipated energy, compared to those observed in dry and 0.5 M H2SO4 environment. However, in the presence of TZ essential oil, there is an absence of metal debris (3rd body) and the presence of striations parallel to the friction direction. The potentiodynamic study revealed that in the presence of this oil the cathodic reaction appears to be much influenced than the anodic reaction, suggesting that the phenolic components act as predominantly cathodic inhibitors, with an inhibition efficiency of 83.80%. The SEM-EDX and microtopography analysis are in compliance with the weight loss measurements and electrochemical results. The theoretical studies show that the thymol acts as the best wear-corrosion inhibitor among the major constituents of TZ essential oil. From the obtained results, we can suggest that TZ essential oil is a new natural substance that can be used against material wear in aggressive solutions in the presence of a mechanical effect.


Corresponding author: Lhou Majidi, Laboratory of Natural Substances & Synthesis and Molecular Dynamics, Sciences and Technics Faculty, Moulay Ismail University, Errachidia, BP 509, Boutalamine, Errachidia, Meknes50000, Morocco, E-mail:

Funding source: EU-FP Grant Oil & Sugar

Award Identifier / Grant number: 295202

Acknowledgments

The authors extend their appreciation to the Moroccan Association of theoretical chemists (AMCT) for access to the computational facility.

  1. Author contributions: All the authors have accepted responsibility for the entire content of this submitted manuscript and approved submission.

  2. Research funding: This work was partially supported by EU-FP Grant Oil & Sugar (295202).

  3. Conflicts of interest: The authors declare no conflicts of interest regarding this article.

References

ASTM International Designation, G. (1994). Guide for determining synergism between wear and corrosion. West Conshohocken, PA, USA: ASTM, pp. 119–04.Search in Google Scholar

Benallou, A., Lakbaibi, Z., Garmes, H., and EL Abdallaoui, H.E.A. (2019). The role of the polarity on the mechanism and selectivity in the [3+2] cycloaddition reaction between CF3-ynone ylide and azide group: a quantum chemical investigation. J. Fluor. Chem. 219: 79–91, https://doi.org/10.1016/j.jfluchem.2018.12.008.Search in Google Scholar

Benea, L., Wenger, F., Ponthiaux, P., and Celis, J.P. (2009). Tribocorrosion behaviour of Ni–SiC nanostructured composite coatings obtained by electrode position. Wear 266: 398–405, https://doi.org/10.1016/j.wear.2008.04.018.Search in Google Scholar

Benea, L., Mardare-Danaila, E., Mardare, M., and Celis, J.P. (2014). Preparation of titanium oxide and hydroxyapatite on Ti–6Al–4V alloy surface and electrochemical behaviour in bio-simulated fluid solution. Corrosion Sci. 80: 331–338, https://doi.org/10.1016/j.corsci.2013.11.059.Search in Google Scholar

Berradja, A., Déforge, D., Nogueira, R.P., Ponthiaux, P., Wenger, F., and Celis, J.P. (2006). An electrochemical noise study of tribocorrosion processes of AISI 304 L in Cl- and media. J. Phys. D Appl. Phys. 39: 3184–3192, https://doi.org/10.1088/0022-3727/39/15/s08.Search in Google Scholar

Boumezzourh, A., Ouknin, M., Bouyanzer, A., Costa, J., and Majidi, L. (2019). Ammodaucus Leucotrichus Cosson & Durieu fruits essential oil as corrosion inhibitor of tinplate in 0.5 M oxalic acid medium and its thermodynamic properties. Moroc. J. Chem. 7: 7–1.Search in Google Scholar

Boumezzourh, A., Ouknin, M., Chibane, E., Costa, J., Bouyanzer, A., Hammouti, B., and Majidi, L. (2020). Inhibition of tinplate corrosion in 0.5M H2C2O4 medium by Mentha pulegium essential oil. Int. J. Corros. Scale Inhib. 9: 152–170.10.17675/2305-6894-2020-9-1-9Search in Google Scholar

Bratu, F., Benea, L., and Celis, J.P. (2007). Tribocorrosion behaviour of Ni–SiC composite coatings under lubricated conditions. Surf. Coating. Technol. 201: 6940–6946, https://doi.org/10.1016/j.surfcoat.2006.12.027.Search in Google Scholar

Chen, J., Wang, J., Yan, F., Zhang, Q., and Li, Q. (2015). Effect of applied potential on the tribocorrosion behaviors of Monel K500 alloy in artificial seawater. Tribol. Int. l81: 1–8, https://doi.org/10.1016/j.triboint.2014.07.014.Search in Google Scholar

Dandlen, A.S., Lima, A.S., Mendes, M.D., Miguel, M.G., Faleiro, M.L., Sousa, M.J., Pedro, L.G., Barroso, J.G., and Figueiredo, A.C. (2010). Antioxidant activity of six Portuguese thyme species essential oils. Flavour Fragrance J. 25: 150–155, https://doi.org/10.1002/ffj.1972.Search in Google Scholar

Ding, H., Yang, X., Xu, L., Li, M., Li, S., Zhang, S., and Xia, J. (2020). Analysis and comparison of tribological performance of fatty acid-based lubricant additives with phosphorus and sulfur. J Bioresour Bioprod 5: 134–142, https://doi.org/10.1016/j.jobab.2020.04.007.Search in Google Scholar

Dob, T., Dahmane, D., Benabdelkader, T., and Chelghoum, C. (2006). Studies on the essential oil composition and antimicrobial activity of Thymus algeriensis Boiss. Int. J. Aromather. 16: 95–100, https://doi.org/10.1016/j.ijat.2006.04.003.Search in Google Scholar

Domingo, L.R., Pérez, P., and Sáez, J.A. (2013). Understanding the local reactivity in polar organic reactions through electrophilic and nucleophilic Parr functions. RSC Adv. 3: 1486–1494, doi:https://doi.org/10.1039/c2ra22886f.Search in Google Scholar

Emregul, K.C. and Atakol, O. (2004). Corrosion inhibition of iron in 1M HCl solution with Schiff base compounds and derivatives. Mater. Chem. Phys. 83: 373–379, https://doi.org/10.1016/j.matchemphys.2003.11.008.Search in Google Scholar

European Pharmacopoeia. (1997). 3rd ed. Council of Europe, Strasbourg.Search in Google Scholar

Frisch, M.J., Trucks, G.W., Schlegel, H.B., Scuseria, G.E., Robb, M.A., Cheeseman, J.R., Scalmani, G., Barone, V., Mennucci, B., Petersson, G.A., et al.. (2009). Gaussian 09, Vol. 32. Gaussian, Inc., Wallingford, CT, pp. 5648–5652.Search in Google Scholar

Gece, G. (2008). The use of quantum chemical methods in corrosion inhibitor studies. Corrosion Sci. 50: 2981–2992, https://doi.org/10.1016/j.corsci.2008.08.043.Search in Google Scholar

Geerlings, P., De Pro, F., and Langenaeker, W. (2003). Conceptual density functional theory. Chem. Rev. 103: 1793–1874, https://doi.org/10.1021/cr990029p.Search in Google Scholar PubMed

Goulart, C.M., Esteves-Souza, A., Martinez-Huitle, C.A., Rodrigues, C.J.F., Maciel, M.A.M., and Echevarria, A. (2013). Experimental and theoretical evaluation of semicarbazones and thiosemicarbazones as organic corrosion inhibitors. Corrosion Sci. 67: 281–291, https://doi.org/10.1016/j.corsci.2012.10.029.Search in Google Scholar

Hariharan, P.C. and Pople, J.A. (1973). The influence of polarization functions on molecular orbital hydrogenation energies. Theor. Chim. Acta 28: 213–222, https://doi.org/10.1007/bf00533485.Search in Google Scholar

Hehre, W.J., Ditchfield, R., and Pople, J.A. (1972). Self-consistent molecular orbital methods. XII. Further extensions of Gaussian-type basis sets for use in molecular orbital studies of organic molecules. J. Chem. Phys. 56: 2257–2261, https://doi.org/10.1063/1.1677527.Search in Google Scholar

Hsissou, R., Benzidia, B., Hajjaji, N., and Elharfi, A. (2019). Elaboration, electrochemical investigation and morphological study of the coating behavior of a new polymeric polyepoxide architecture: crosslinked and hybrid decaglycidyl of phosphorus penta methylene dianiline on E24 carbon steel in 3.5% NaCl. Port. Electrochim. Acta 37: 179–191.10.4152/pea.201903179Search in Google Scholar

Hsissou, R., Benhiba, F., Dagdag, O., El Bouchti, M., Nouneh, K., Assouag, M., Briche, S., Zarrouk, A., and Elharfi, A. (2020a). Development and potential performance of prepolymer in corrosion inhibition for carbon steel in 1.0 M HCl: outlooks from experimental and computational investigations. J. Colloid Interface Sci. 574: 43–60, https://doi.org/10.1016/j.jcis.2020.04.022.Search in Google Scholar PubMed

Hsissou, R., Benhiba, F., Abbout, S., Dagdag, O., Benkhaya, S., Berisha, A., Erramli, H., and Elharfi, A. (2020b). Trifunctional epoxy polymer as corrosion inhibition material for carbon steel in 1.0 M HCl: MD simulations, DFT and complexation computations. Inorg. Chem. Commun. 115: 107858, https://doi.org/10.1016/j.inoche.2020.107858.Search in Google Scholar

Hsissou, R., Benhiba, F., Echihi, S., Benkhaya, S., Hilali, M., Berisha, A., Briche, S., Zarrouk, A., Nouneh, K., and Elharfi, A. (2020c). New epoxy composite polymers as a potential anticorrosive coatings for carbon steel in 3.5% NaCl solution: experimental and computational approaches. Chem. Data Collect. 31: 100619.10.1016/j.cdc.2020.100619Search in Google Scholar

Hsissou, R., Abbout, S., Seghiri, R., Rehioui, M., Berisha, A., Erramli, H., Assouag, M., and Elharfi, A. (2020d). Evaluation of corrosion inhibition performance of phosphorus polymer for carbon steel in [1M] HCl: computational studies (DFT, MC and MD simulations). J. Mater. Res. Technol. 9: 2691–2703, https://doi.org/10.1016/j.jmrt.2020.01.002.Search in Google Scholar

Kokalj, A. (2012). On the HSAB based estimate of charge transfer between adsorbates and metal surfaces. Chem. Phys. 393: 1–12, https://doi.org/10.1016/j.chemphys.2011.10.021.Search in Google Scholar

Lavanya, M., Murthy, V.R., and Rao, P. (2019). Performance evaluation of a potent green inhibitor on 6061 aluminum alloy under liquid/solid jet impingement. J. Bio- Tribo-Corrosion 5: 93, https://doi.org/10.1007/s40735-019-0288-7.Search in Google Scholar

Lukovits, I., Kálmán, E., and Zucchi, F. (2001). Corrosion inhibitors - correlation between electronic structure and efficiency. Corrosion Sci. 57: 3–8, https://doi.org/10.5006/1.3290328.Search in Google Scholar

Martinez, S. (2003). Inhibitory mechanism of mimosa tannin using molecular modeling and substitutional adsorption isotherms. Mater. Chem. Phys. 77: 97–102, https://doi.org/10.1016/s0254-0584(01)00569-7.Search in Google Scholar

Mohrbacher, H., Blanpain, B., Celis, J.P., and Roos, J.R. (1995). The influence of humidity on the fretting behavior of PVD TiN coatings. Wear 180: 43–52, https://doi.org/10.1016/0043-1648(94)06538-1.Search in Google Scholar

Obot, I., Macdonald, D., and Gasem, Z. (2015). Density functional theory (DFT) as a powerful tool for designing new organic corrosion inhibitors. Part 1: an overview. Corrosion Sci. 99: 1–30, https://doi.org/10.1016/j.corsci.2015.01.037.Search in Google Scholar

Olasunkanmi, L.O., Kabanda, M.M., and Ebenso, E.E. (2016). Quinoxaline derivatives as corrosion inhibitors for mild steel in hydrochloric acid medium: electrochemical and quantum chemical studies. Phys. E Low-dimens. Syst. Nanostruct. 76: 109–126, https://doi.org/10.1016/j.physe.2015.10.005.Search in Google Scholar

Ouknin, M., Romane, A., Costa, J., Ponthiaux, P., and Majidi, L. (2018). Anticorrosion properties of Thymus munbyanus Boiss & eut. Moroc. J. Chem. 6: 6–3.Search in Google Scholar

Ouknin, M., Yang, Y., Paolini, J., Costa, J., Ponthiaux, P., and Majidi, L. (2019a). The effect of Corsican poplar leaf buds (Populus nigra var. italica) essential oil on the tribocorrosion behavior of 304L stainless steel in the sulfuric medium. J. Bio- Tribo-Corrosion 5: 83, https://doi.org/10.1007/s40735-019-0275-z.Search in Google Scholar

Ouknin, M., Costa, J., and Majidi, L. (2020a). Tribocorrosion and electrochemical behavior of AISI 304L stainless steel in acid medium and Thymus willdenowii Boiss & Reut essential oil effect. Chem. Data Collect. 20: 100389, https://doi.org/10.1016/j.cdc.2020.100389.Search in Google Scholar

Ouknin, M., Romane, A., Ponthiaux, J.P., Costa, J., and Majidi, L. (2020b). Evaluation of corrosion inhibition and adsorption behavior of Thymus zygis subsp. gracilis volatile compounds on mild steel surface in 1M HCl. Corrosion Rev. 38: 137–149, https://doi.org/10.1515/corrrev-2019-0055.Search in Google Scholar

Parr, R.G. and Pearson, R.G. (1983). Absolute hardness: companion parameter to absolute electronegativity. J. Am. Chem. Soc. 105: 7512–7516, https://doi.org/10.1021/ja00364a005.Search in Google Scholar

Pearson, R.G. (1988). Absolute electronegativity and hardness: application to inorganic chemistry. Inorg. Chem. 27: 734–740, https://doi.org/10.1021/ic00277a030.Search in Google Scholar

Ponthiaux, P., Wenger, F., and Celis, J.P. (2012). Tribocorrosion: material behaviour under combined conditions of corrosion and mechanical loading. In: Shih, Dr. (Ed.). Corrosion resistance. In Tech, pp. 81–106.10.5772/35634Search in Google Scholar

Saada, F.B., Antar, Z., Elleuch, K., Ponthiaux, P., and Gey, N. (2018). The effect of nanocrystallized surface on the tribocorrosion behavior of 304L stainless steel. Wear 394: 71–79, https://doi.org/10.1016/j.wear.2017.10.007.Search in Google Scholar

Sastri, V.S. and Perumareddi, J.R. (1997). Molecular orbital theoretical studies of some organic corrosion inhibitors. Corrosion 53: 617–622, https://doi.org/10.5006/1.3290294.Search in Google Scholar

Singh, A., Ansari, K.R., Haque, J., Dohare, P., Lgaz, H., Salghi, R., and Quraishi, M.A. (2018). Effect of electron donating functional groups on corrosion inhibition of mild steel in hydrochloric acid: experimental and quantum chemical study. J. Taiwan Inst. Chem. Eng. 82: 233–251, https://doi.org/10.1016/j.jtice.2017.09.021.Search in Google Scholar

Sun, H. (1998). Compass: an ab initio force-field optimized for condensed-phase applications overview with details on alkane and benzene compounds. J. Phys. Chem. B 102: 7338–7364, https://doi.org/10.1021/jp980939v.Search in Google Scholar

Sun, Y. and Bailey, R. (2014). Improvement in tribocorrosion behavior of 304 stainless steel by surface mechanical attrition treatment. Surf. Coating. Technol. 253: 284–291, https://doi.org/10.1016/j.surfcoat.2014.05.057.Search in Google Scholar

Tang, Y., Zhang, F., Hu, S., Cao, Z., Wu, Z., and Jing, W. (2013). Novel benzimidazole derivatives as corrosion inhibitors of mild steel in the acidic media. Part I: gravimetric, electrochemical, SEM and XPS studies. Corrosion Sci. 74: 271–282, https://doi.org/10.1016/j.corsci.2013.04.053.Search in Google Scholar

Vanasundari, K., Balachandran, V., Kavimani, M., and Narayana, B. (2017). Spectroscopic investigation, vibrational assignments, Fukui functions, HOMO-LUMO, MEP and molecular docking evaluation of 4–[(3, 4 dichlorophenyl) amino] 2–methylidene 4 – oxo butanoic acid by DFT method. J. Mol. Struct. 1147: 136–147, https://doi.org/10.1016/j.molstruc.2017.06.096.Search in Google Scholar

Verma, C., Olasunkanmi, L.O., Obot, I., Ebenso, E.E., and Quraishi, M. (2016). 2,4-Diamino-5-(phenylthio)-5 H-chromeno [2,3-b] pyridine-3-carbonitriles as green and effective corrosion inhibitors: gravimetric, electrochemical, surface morphology and theoretical studies. RSC Adv. 6: 53933–53948, https://doi.org/10.1039/c6ra04900a.Search in Google Scholar

Verma, C., Lgaz, H., Verma, D.K., Ebenso, E.E., Bahadur, I., and Quraishi, M.A. (2018). Molecular dynamics and Monte Carlo simulations as powerful tools for study of interfacial adsorption behavior of corrosion inhibitors in aqueous phase: a review. J. Mol. Liq. 260: 99–120, https://doi.org/10.1016/j.molliq.2018.03.045.Search in Google Scholar

Vincent, L. (1994). Material and fretting, ESIS 18. Mechanical Engineering Publication, London, pp. 323–337.Search in Google Scholar

Yang, F., Su, B., Zhang, A., Meng, J., Han, J., and Wu, Y. (2018). Tribological properties and wear mechanisms of Mo2FeB2 based cermets at high temperatures. Tribol. Int. 120: 391–397, https://doi.org/10.1016/j.triboint.2017.12.038.Search in Google Scholar

Received: 2020-05-29
Accepted: 2021-02-16
Published Online: 2021-04-12
Published in Print: 2021-06-25

© 2021 Walter de Gruyter GmbH, Berlin/Boston

Downloaded on 1.10.2025 from https://www.degruyterbrill.com/document/doi/10.1515/corrrev-2020-0053/html
Scroll to top button