Next Article in Journal
Special Issue “Laser Powder Bed Fusion, Direct Energy Deposition and Hybrid Manufacturing of Metals and Alloys”
Next Article in Special Issue
Flower-like Superhydrophobic Surfaces Fabricated on Stainless Steel as a Barrier against Corrosion in Simulated Acid Rain
Previous Article in Journal
Physiomechanical and Surface Characteristics of 3D-Printed Zirconia: An In Vitro Study
Previous Article in Special Issue
Computational Design of Anticorrosion Properties of Novel, Low-Molecular Weight Schiff Bases
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Effect of Quaternary Ammonium Salts and 1,2,4-Triazole Derivatives on Hydrogen Absorption by Mild Steel in Hydrochloric Acid Solution

by
Yaroslav G. Avdeev
,
Tatyana A. Nenasheva
,
Andrey Yu. Luchkin
*,
Andrey I. Marshakov
and
Yurii I. Kuznetsov
A.N. Frumkin Institute of Physical Chemistry and Electrochemistry, Russian Academy of Sciences, 31 Leninskii Prospect, Moscow 119071, Russia
*
Author to whom correspondence should be addressed.
Materials 2022, 15(19), 6989; https://doi.org/10.3390/ma15196989
Submission received: 18 August 2022 / Revised: 23 September 2022 / Accepted: 30 September 2022 / Published: 8 October 2022
(This article belongs to the Special Issue Advance in Corrosion and Protection of Metals)

Abstract

:
The treatment of low-carbon steel items with hydrochloric acid solutions is used in many industrial technologies. This process is accompanied not only by metal corrosion losses, but also by hydrogen absorption by the metal. In this study, the kinetics of hydrogen cathodic reduction on low-carbon steel in 2 M HCl containing corrosion inhibitors, namely, quaternary ammonium salts and a 3-substituted 1,2,4-triazole, have been studied. Adsorption isotherms of corrosion inhibitors on cathodically polarized steel surface have been obtained. XPS data provide valuable information on the composition and structure of protective layers formed on steel in HCl solutions containing inhibitors. The main rate constants of the stages of gaseous hydrogen evolution and incorporation of hydrogen atoms into the metal have been determined. The addition of quaternary ammonium salts or 3-substituted 1,2,4-triazole inhibits the cathodic reduction of hydrogen and its penetration into steel in the HCl solution. 3-substituted 1,2,4-triazole is the most efficient inhibitor of hydrogen absorption. The inhibitory effect of this compound is caused by a decrease in the ratio of the hydrogen concentration in the metal phase to the degree of surface coverage with hydrogen. The maximum decrease in hydrogen concentration in the metal bulk in the presence of the 3-substituted 1,2,4-triazole is 8.2-fold, which determines the preservation of the plastic properties of steel as it corrodes in HCl solutions. The high efficiency of the 3-substituted 1,2,4-triazole as an inhibitor of hydrogen cathodic reduction and absorption results from strong (chemical) adsorption of this compound on the steel surface and the formation of a polymolecular protective layer.

1. Introduction

Solutions of hydrochloric acid occupy a special place among acid solutions used for various technological processes. Treatment of carbonate-hydrocarbon-containing formations with hydrochloric acid is used in the oil industry to increase oil recovery; in metallurgical enterprises, thermal scale is removed from steel products with HCl solutions; hydrochloric acid cleaning is also an efficient way to clean the internal and external surfaces of steel products of rust and mineral deposits [1,2,3,4,5,6,7,8,9,10]. Compounds of various nature are widely studied as corrosion inhibitors (CIs) of steels in this medium [11,12,13,14,15,16,17,18,19,20]. Of these, quaternary ammonium salts (QASs) [21,22,23,24,25,26,27,28,29,30] often used as components of industrial mixed CIs are important from a practical point of view. Triazole derivatives are a new promising class of acid corrosion inhibitors for steel that possess unique protective properties [31,32,33,34,35]. These compounds make it possible to protect various steel surfaces in mineral acid solutions at temperatures of up to 200 °C [34,35].
The contact of steel products with corrosive aqueous media is often accompanied not only by corrosion losses of the metal due to its chemical reaction with components of the medium, but also by hydrogen absorption into the metal bulk, which can make the metal brittle, thus significantly impairing its mechanical characteristics [36,37,38,39,40]. Most commonly, the absorption of hydrogen by steels occurs in acid solutions where gaseous hydrogen is a corrosion product released on the metal surface. Molecular hydrogen is formed upon the adsorption of atomic hydrogen that partially penetrates the metal bulk.
The CIs used to protect steels in acid media should not only suppress the general corrosion of the metal, but also hydrogen penetration into the steel. Unfortunately, in many studies of organic corrosion inhibitors for steels, the aspect of their effect on the absorption of hydrogen by a steel is most often ignored, although its mechanical properties are important parameters that ensure the reliable operation of industrial products. In rare studies [41,42,43], it has been noted that nitrogen-containing organic CIs decrease the penetration of hydrogen into steel in acid media.
In view of this, it seems important to identify the regularities of the effect of nitrogen-containing organic CIs on the kinetics of cathodic hydrogen evolution and its penetration into steel. It should be shown how the change in the kinetic parameters of the cathodic reaction caused by inhibitors affects overall corrosion, including estimations of the corrosion rate of steel and the effect on hydrogen absorption, as well as the preservation of mechanical properties by the metal. Detailed discussion of the effect of CI on the kinetics of the cathodic reaction of steel requires features of the mechanism of their action to be identified and reviewed. A mixture of alkylbenzyldimethylammonium chlorides (ABDMA) and an inhibitor (TA) belonging to a promising group of triazole compounds [31,32,33,34,35] were chosen for this study.

2. Materials and Methods

2.1. Materials

Samples of mild steel (MS, wt.%: 0.05 C, 0.03 Si, 0.38 Mn, 0.09 Ni, 0.04 S, 0.035 P, 0.05 Cr, 0.15 Cu, and 0.16 Al) and spring steel (SS, wt.%: 0.7 C, 1.52 Si, 0.52 Mn, and 0.3 Cr) were used as the working electrodes.
In this study, 2 M HCl aqueous solution was used as the background electrolyte. The solutions were prepared from a concentrated HCl solution of “chemically pure” grade and distilled water. Solutions de-aerated with argon were used in the electrochemical studies. Corrosion tests were carried out in solutions with free access of air. ABDMA, which is a mixture of alkylbenzyldimethylammonium chlorides ([CnH2n+1N+(CH3)2CH2C6H5]Cl, where n = 10–18), was used as the QAS. Moreover, the TA inhibitor (a 3-substituted 1,2,4-triazole derivative) was used. Due to the low solubility of TA, it was added to the acid solutions as a concentrated solution in ethanol; thus, the concentration of ethanol in the etching solution was 0.24 mol·L−1.

2.2. Methods

2.2.1. Membrane Test

The rates of hydrogen penetration through the membrane were measured in a Devanathan-Stachurski cell (Figure 1) [44,45]. Mild steel membranes with a diameter of 5 cm and a thickness of 0.1 mm were used. The surface area of the membrane in contact with the electrolyte was 4.25 cm2. Both membrane sides were ground with SiC up to #600 grit, then chemically etched in 16% HCl for 1 min and thoroughly washed with distilled water. Subsequently, a palladium film was electrochemically deposited on the membrane’s exit side for 100 s at a constant current density of 25 mA cm−2. The palladium plating solution contained 25 g L−1 PdCl2 and 20 g L−1 NH4Cl. The solution pH was adjusted to 8.5 by adding the required amount of NH4OH. Prior to the tests, the Pd-plated membranes were degassed at room temperature for at least 2 days.
The diffusion part of the cell was filled with 0.1 M NaOH, and the membrane was polarized at a potential of 0.45 V vs. SHE. The hydrogen flux at the exit surface was measured as its Faradic equivalent ip = i − ibg, where ibg is the background current density. The background current density was less than 5 × 10−3 A m−2. A constant potential (E = const) was applied to the working side of the membrane and the external current (is) was measured. The stationary current of hydrogen penetration through the membrane (ip) was recorded on the anodically polarized (diffusion) side of the membrane. The potential of the working membrane side was shifted from E = −0.4 V in the positive direction. To fill the hydrogen traps that exist both in the metal bulk and on its surface, before starting the measurements, a potential of −0.4 V was set on the working side of the membrane and maintained until a stationary value of the hydrogen penetration current was established (60 min). The cathodic curves and the ip-E plot were obtained at 20 mV increments and 30 min exposure time using an IPC PRO MF potentiostat (Cronas Ltd., Moscow, Russia).
A silver chloride reference electrode and auxiliary platinum electrode were used. All the experiments were carried out at room temperature. Electrode potentials are reported against the SHE.

2.2.2. Electrochemical Impedance Spectroscopy (EIS)

The adsorption of TA was studied by EIS using an IPC-PRO MF potentiostat (Cronas Ltd., Moscow, Russia) combined with a frequency response analyzer (FRA) on a rotating disk electrode (n = 1000 rpm) made of MS steel with a working area of 0.64 cm2. Before each experiment, the electrode was cleaned on abrasive wheels with various grain sizes, polished with diamond paste (ACM 0.5/0), and degreased with acetone. To prevent the dissolution of the steel electrode, it was always kept under cathodic polarization (E = −0.30 V). The working and auxiliary platinum electrodes were mounted in a three-electrode cell. A silver chloride reference electrode was used. The electrochemical impedance of a steel electrode was measured in the frequency range of 10 mHz to 3 kHz at an alternating voltage amplitude of 0.020 V. The results of impedance measurements were processed using the Dummy Circuits Solver program, version 1.7. A steel electrode was placed in the acid solution and kept until a stationary impedance spectrum was obtained (no more than 2 h); then, the required concentration of the inhibitor (Cinh) was added to the solution, and the electrode was kept until a stationary spectrum was obtained (no more than 3 h). The experiments were carried out in 2.0 M HCl solutions de-aerated with argon at t = 22 °C.
The surface coverage with the inhibitor (θinh) was determined using Equation (1):
θ i n h = C d l 0 C d l C d l 0 C d l
where Cdl and Cdlθ are the capacitance of the double electric layer (DEL) of the steel electrode in the background solution, in the inhibited solution, and under conditions of the limiting inhibitor adsorption on the metal, respectively.
The specific capacitance of the DEL was calculated using Equation (2):
C d l s p = C d l S
where Cdl is the capacitance of the DEL of the steel electrode and S is the area of the steel electrode.

2.2.3. Determination of Steel Surface Coverage with an Inhibitor from Stationary Cathodic Current Data

To make the determination of the coverage of MS steel surface with the studied inhibitors more technically feasible, these values were simultaneously calculated from the decrease in the cathodic current (E = −0.30 V) in the presence of the inhibitors (3):
θinh = (ic,0iinh)/ic,0,
where ic,0 is the current density in the background solution and ic,inh is the current density in the solution with the corresponding additive.
The values of the cathodic currents on MS obtained in the studies described in Section 2.2.1 were used in the calculations. The assumption was made that the additives studied predominantly acted by a blocking mechanism. The results obtained in this way for the adsorption of TA on steel highly correlated with the data obtained by the EIS method (Section 2.2.7). This allowed us to subsequently use the values of θinh for ABDMA obtained as the ratio of cathodic currents in the inhibited solution and in the background.

2.2.4. Gravimetric Method

The corrosion rate of SS in 2 M HCl solution was determined from the mass loss of samples (no less than three samples per point) that had a thickness of 0.5 mm and a working surface area of S = 17.6 cm2 (4).
ρ = Δm·S−1 τ−1,·100%
where Δm is the change in the sample mass, g; S is the sample area, m2; and τ is the duration of corrosion tests, h.
Before the start of each experiment, the surface was activated (1 min) with a mixture of acetone and concentrated HCl (10:1) and then wiped dry with a cotton cloth.
The efficiency of inhibitors was estimated as the inhibition factor (5):
Z = (ρ0ρinh) ρ0−1 100%
where ρ0 and ρinh are the corrosion rates in the background solution and in the solution with an additive being studied, respectively.

2.2.5. Determination of the Amount of Hydrogen Absorbed by a Metal Using Vacuum Extraction

The hydrogen concentration in the bulk of SS steel was determined using the vacuum extraction method. After the corrosion test, the sample was placed in a vessel that was then evacuated to a residual pressure of 1.33 × 10−4 Pa and heated to a temperature of t = 500 °C. The amount of hydrogen released upon heating the sample in the vacuum was estimated from the pressure change in 10 min (Ptotal) measured with a McLeod gauge at a constant volume of the evacuated part of the system. The pressure of evolved hydrogen (PH2) was calculated from the change in the total pressure (Ptotal) using Equation (6):
PH2 = PtotalPcorrect,
where Pcorrect is the correction for the blank test.
The molar concentration of hydrogen atoms in the bulk of steel (mol cm−3) was calculated using Equation (7):
C H v . = F P H 2   V 1
where K is a constant related to the volume of the analytical part of the setup, and V is the volume of the steel sample, cm3.
Data on the volume concentration of hydrogen in the metals are reported with correction for metallurgical hydrogen that amounted to 2.4 × 10−6 mol cm−3 in the case of the SS.
The degree of steel protection from hydrogen absorption was determined using Equation (8):
Z v H * = [ ( C H v .   C H v , i n h ) C H . v 1 ] 100 %
where C H v . and C H v , i n h are the molar concentrations of hydrogen in the steel bulk after exposure in the background solution and in the inhibited solution, respectively.

2.2.6. Determination of the Mechanical Properties (Ductility) of Steel

The ductility of SS was estimated with an NG-1-3M device based on the number of bends before breakage of band-shaped samples in the initial state (β0) and after exposure in 2 M HCl solution (β). The steel ductility was determined using Equation (9):
p = β β0−1 100%
The mean value for the SS studied was β0 = 87.

2.2.7. Voltammetric Studies

Electrochemical measurements were carried out on flat SS samples at t = 25 °C in de-aerated 2 M HCl solutions stirred with a magnetic stirrer. An electrode cleaned and degreased with acetone was kept for 30 min in the test solution; then, anodic and cathodic polarization curves were recorded using an IPC-PRO MF potentiostat (Cronas Ltd., Moscow, Russia) at a dynamic potential scan rate of 0.0005 V s−1. The effect of a CI on the rate of the anodic and cathodic processes on the SS in the acid solution was estimated using the inhibition coefficient (10):
γc= ic,0/ic,inh and γa = ia,0/ia,inh,
where ic,0, ia,0, ic,inh, and ia,inh are the cathodic and anodic current densities in the background solution and in the solution with the additive under study, respectively.

2.2.8. XPS Study of Steel Surface

The quantitative and qualitative compositions of the surface layers formed by TA inhibitor on MS steel were analyzed using X-ray photoelectron spectroscopy (XPS) on an HB100 Auger microscope Omicron ESCA+ instrument (Germany, Taunusstein) equipped with an additional camera for recording XPS spectra. The vacuum in the analytical chamber was no worse than 10−9 Torr. An Al anode with 200 W power was used as the excitation source. The pass energy of the analyzer was set to 50 eV. Disk electrodes (MS steel, 10 mm in diameter) served as the samples in the XPS studies. Pretreatment of the samples was performed in the same way as in the electrochemical studies.
The binding energy of electrons (Eb) knocked out from internal shells of atoms was calibrated with respect to the XPS peak of C1s electrons from the vapors of the deposited layers of diffusion oil whose binding energy was assumed to be 285.0 eV. The spectrometer was calibrated with respect to the binding energies Eb of Cu2p3/2 (932.7 eV) and Au4f 7/2 (84.0 eV) from copper and gold samples cleaned with argon ions from surface contaminants. The characteristic peaks of the following elements were measured: C1s, O1s, Fe2p, N1s, and Cl2p. Quantitative estimations were based on the photoionization cross-sections of the corresponding electron shells reported elsewhere [46]. The integral peak intensities were obtained after background subtraction by the Shirley method [47] and by fitting the observed peaks by Gaussian curves with a contribution of the Lorentz component. The integral areas under the C1s, O1s, Fe2p, N1s, and Cl2p peaks were used to calculate the film thicknesses.
An important stage of these studies involves the prolonged (up to 18 min) ultrasonic cleaning of the surface of metal samples from a CI in distilled water or in acid solutions. During this procedure, CI molecules retained on the metal surface by physical forces are removed from the surface of samples pre-exposed to an inhibited acid solution. CI molecules bound to the metal surface by chemical forces are not removed by ultrasonic surface cleaning.

3. Results and Discussion

3.1. Kinetics of Cathodic Evolution and the Penetration of Hydrogen into Iron in the Presence of Corrosion Inhibitors

Polarization curves were recorded and plots of the rate of hydrogen penetration into steel vs. potential in hydrochloric acid solution and with addition of 0.01–10 mmol organic CIs (ABDMA and TA) were obtained (Figure 2 and Figure 3). As demonstrated, as the CI concentration increased, the cathodic current and the rate of hydrogen penetration into the metal decrease significantly. At low ABDMA concentrations (Cinh = 0.01–0.1 mmol), a sharp decrease in the rates of cathodic hydrogen evolution (ic) and hydrogen penetration into the metal (ip) was observed in the entire potential (E) region studied, whereas at higher Cinh, the effect of the CI was less pronounced (Figure 2). With an increase in the TA content in the acid solution, the ic and ip values decreased more smoothly (Figure 3).

3.2. Fundamentals of IPZ Analysis of the Dependence of Hydrogen Ion Discharge Rates and Hydrogen Penetration through the Membrane on the Potential [48]

The cathodic evolution of hydrogen on iron in acids occurs by the “discharge—chemical recombination, mixed rate control” or by the “slow discharge—irreversible chemical recombination” mechanisms [48,49,50,51]. The rate constants of the stages of discharge of H+ ions and chemical recombination of hydrogen atoms could be determined by IPZ analysis, i.e., by comparison of the cathodic polarization curve and the plot of the current of hydrogen penetration into the metal vs. potential [48]. IPZ analysis could also be applied if a fraction of the electrode surface is blocked by some adsorbed compound, such as a CI [42]. It is assumed in this case that the discharge of H+ ions occurs on the metal surface not occupied by adsorbed atomic hydrogen; a corrosion inhibitor added to the solution does not change the mechanism of this reaction. The effect of a corrosion inhibitor on the discharge rate of H+ ions (ic) is described as [52] (11):
ic = F kc,0 aH+ (1 − θInh)r1 exp (−s1θInh) exp(−αFE/RT),
where kc,0 is the rate constant H+ ion discharge in the background electrolyte, aH+ is the activity of hydrogen ions, θInh is the coverage of the electrode surface with inhibitor particles, r1 is the number of adsorption sites occupied by hydrogen ions on the surface, s1 is a parameter that characterizes the change in the macro properties of the surface layer and takes the possibility of a specific interaction between the activated complex and adsorbate molecules into account, α is the transfer coefficient of the hydrogen ion discharge reaction, F is the Faraday constant, R is the gas constant, and T is the absolute temperature.
Obviously, if we take the “blocking” effect of both atomic hydrogen and inhibitor particles on the rate of discharge of H+ ions into account, we obtain (12):
ic = Fkc aH+ [(1 − θInh)r1θH] exp(−αFE/RT),
where θH is the surface coverage of the electrode with hydrogen, and kc = kc,0∙exp(−s1)∙is the rate constant of discharge of hydrogen ions in an inhibitor solution at a constant solution acidity.
If the reaction of chemical recombination of H atoms is irreversible, its rate (ir) is determined by the expression (12):
ir = Fkrθ2H
where kr is the rate constant of chemical recombination of H atoms.
The rate of hydrogen transition through the metal surface (ip) and its steady-state diffusion in the membrane are described by the relationships (14):
ip = F(kabcθHkdesCsH)
and (15):
i p = F D C H s L
where kabs and kdes are the constants of absorption and desorption of hydrogen from the metal phase, C H s is the concentration of diffusion-mobile hydrogen in the metal phase near the cathodically polarized membrane surface, L is the membrane thickness, and D is the diffusion coefficient of hydrogen in the metal.
It follows from Equations (14)– ( 16 ) :
θ H = k d e s + D L k a b s C H s = k C H s
where k is the kinetic-diffusion constant that shows the ratio of the concentrations of hydrogen atoms on the surface and in the metal phase.
Using Equations (12), (13), (15), and (16), we can obtain, for steady-state conditions (ic = ip + ir) [48,50,53], (17) and (18):
i c exp ( α F E R T ) = F k 1 a H + ( 1 θ i n h ) r 1 k 1 a H + k L D · i p
i p = D F L k k r · i c i p = D F L k k r i r
By combining Equations (12), (13), (15), and (16), we obtain an expression for calculating the coverage of the steel surface with hydrogen (19):
θ H = ( k 1 , i + D L k ) + ( k 1 , i + D L k ) 2 + 4 k r k 1 , i ( 1 θ i n h ) r 1 2 k r
where k 1 , i = k 1 a H + · exp ( α F E i R T ) is the formal rate constant of hydrogen ion discharge reactions at the Ei potential. The surface concentration of diffusion-mobile hydrogen in the metal ( C H s ) can be calculated from ip values using Equation (15), or from θH values using Equation (16).

3.3. Calculation of the Rate Constants of the Main Stages of Cathodic Hydrogen Evolution and Penetration into Steel, Coverage of the Steel Surface with Hydrogen, and Surface Concentration of Diffusion-Mobile Hydrogen in the Metal

In all the solutions studied, both in the background solution and in those containing the CIs, IPZ analysis of experimental data can be applied (Figure 2 and Figure 3) because, in accordance with Equation (18), the plots of ip vs. ir0.5 are almost linear and pass through the origin. As an example, the plots of ip vs. ir0.5 obtained in the background solution are shown (Figure 4). Comparing the values of d E d L o g i c (and d E d L o g i p ) and applying the reported method [48], we calculated the transfer coefficients α used to build the plots of f = i c exp ( α F E R T ) vs. ip. An example of experimental data processing is shown in Figure 5. Similar plots were obtained for all the solutions and CI concentrations studied.
To calculate the constants k1, kr, and k according to Equations (17) and (18), the coverage of the metal surface with an inhibitor (θinh) must be determined. The EIS method was used to determine θinh in solutions containing TA. The impedance spectra of the steel electrode in the background and TA-inhibited 2.0 M HCl solutions presented as Nyquist plots are perfect semicircles described by a simple equivalent circuit that contains the electrical double-layer capacitance (Cdl), reaction resistance (Rct), and solution resistance (Rs) (Figure 6). An increase in the time of steel electrode exposure in the acid solution containing TA results in an increase in the hodograph radius (Figure 6), which indicates that the inhibitor adsorption occurs slowly over time. The EIS polynomials were calculated using the EIS Methods Applying program [54,55]. In view of this, the electrode was exposed to the inhibited acid solution until a stationary value of its capacitance was reached. The stationary values of θinh calculated according to Equation (1) are given in Table 1, while the dependence of steel surface coverage with the TA inhibitor on its concentration in the corrosive medium (adsorption isotherm) is shown in Figure 7.
The TA adsorption isotherm was obtained in the same solutions by comparing the stationary cathodic currents at E = −0.30 V. The θinh values for TA, calculated according to Equations (1) and (3), were nearly the same. For this reason, the θinh values for ABDMA were calculated using Equation (3) only. The data are presented in Table 1, and the ABDMA adsorption isotherm is shown in Figure 7.
Using the values of α and θinh (Table 1) and assuming r1 = 0.3 [56], from the slope of f, ip straight lines (such as those in Figure 5) and the segment cut off on the y-axis (at ip = 0), the kc,i and k values were calculated in accordance with Equation (18). Table 1 shows the values of k and kc,i at Ei = −0.3 V.
Using the k values obtained and assuming that the stationary diffusion coefficient of hydrogen in the membrane is D = 7.3 × 10−5 cm2 s−1 [57], the values of kr were determined in accordance with Equation (17) from the slope of the plots of ip versus ir0.5 (Figure 4) (Table 1).
Thus, the values of kinetic constants of the main stages of hydrogen evolution and penetration into the metal were obtained, both in the background solution and at various Cinh values.
Using the values of the constants (Table 1) and Equation (19), the values of θH on cathodically polarized steel surface at E= −0.3 V in 2M HCl background solution and in the presence of CIs were calculated (Table 1). The subsurface hydrogen concentrations in steel C H s , calculated according to Equation (15) and Equation (16), satisfactorily agreed and differed no more than by 20% (Table 1 shows the average values of C H s ).
As follows from Table 1, the addition of corrosion inhibitors to the acid solution significantly decreased the concentration of hydrogen in the metal. The effect of a CI on the surface coverage with hydrogen is ambiguous: the values of θH can increase with an increase in Cinh. This is due to a decrease in the hydrogen mobilization constant, which, in turn, can be explained by the inhibition of surface diffusion of hydrogen atoms upon the adsorption of inhibitor particles on the metal. At sufficiently high CI concentrations, the values of θH decrease because the discharge rate of H+ ions decreases to a greater extent than the mobilization rate of H atoms.
The compounds studied not only inhibit corrosion, but also hydrogen absorption, because they decrease the rate constant of discharge of H+ ions and increase the k constant, i.e., they change the ratio between the θH and C H s values (Table 1). The latter effect is explained by the fact that a CI blocks hydrogen absorption centers and hinders the transfer of H atoms from the surface into the metal phase.
TA is a more efficient inhibitor of hydrogen absorption. In fact, the C H s value decreases almost 20-fold at its concentration of 5 mM in the solution (Table 1). A significant decrease in the concentration of diffusion-mobile hydrogen in the metal should benefit the resistance of steel to cracking under mechanical stress.

3.4. Effect of Inhibitors on the Corrosion and Mechanical Properties of Steel

The decrease in the hydrogen penetration into the metal by the inhibitors being studied is most noticeable in the case of spring steels that are prone to hydrogen absorption. Indeed, a study of the corrosion of SS steel in 2 M HCl solution showed that these inhibitors reduced the metal mass loss due to corrosion ( ρ ) and the hydrogen concentration in the metal bulk ( C H s ) (Table 2). It should be emphasized that the plastic properties of the metal (p) exhibited almost no change in the presence of TA (Table 2). Therefore, TA is most efficient as an inhibitor of corrosion and hydrogen absorption. It is important that the inhibitory effect of TA is observed not only at 25 °C, but also at 60 °C, which is the optimal temperature for the industrial acid etching of metals. The results obtained agree with the data presented in Section 3.3, i.e., each of the inhibitors decreases both the concentration of diffusion-mobile hydrogen ( C H s ) and the total hydrogen content ( C H s ) in the metal. As a more efficient inhibitor, TA makes it possible to preserve the ductility of SS steel.

3.5. Effect of CIs on the Rate of Electrode Reactions on Steel

In the HCl solution, the addition of the CIs being studied are efficient in slowing down the anodic and cathodic stages of the corrosion of SS steel (Figure 8). The anodic polarization slopes (ba) observed in the presence of ABDMA and TA are 0.15 and 0.16 V, respectively, which are higher than the ba value of 0.12 V observed in the background medium. This effect is more significant for the cathodic reaction because a limiting current is observed in the presence of both CIs, although a value of bc = 0.16 V is observed in the background medium. The addition of these CIs reduces the rate of anodic ionization of steel, for example, by a factor of 8.8 and 15 in the presence of ABDMA and TA, respectively, at E = −0.10 V. The rate of the cathodic reaction at E = −0.30 V decreases 9.4- and 13-fold, respectively, in the presence of these CIs. Other things being equal, the effect of the TA inhibitor on the electrode reactions of SS steel is more significant than that of ABDMA. As a result, the smallest mass loss of steel samples is observed in HCl solutions containing TA (Table 2).

3.6. Nature of the Adsorption Interaction of the TA Inhibitor with Steel Surface

To understand the reasons for the efficient inhibition of electrode reactions on steel by the TA inhibitor, we have to determine the nature of its adsorption interaction with the steel surface. As shown in Figure 7 (curve 1), the adsorption of TA on the steel surface at medium coverages of the metal surface by the CI obeys the Temkin isotherm (20):
θinh = f −1ln [BCinh],
where θinh is the surface coverage with the inhibitor, f is the surface inhomogeneity factor, B is the constant of adsorption equilibrium, and Cinh is the inhibitor concentration in the solution. The calculated values of the parameters are: f = 4.25, B = 5.31 × 105 L mol−1. The free adsorption energy (−ΔGads) was determined using the relationship (21):
B = 1 55.5 e x p [ Δ G a d s R T ]
and amounted to (−ΔGads) = 42 kJ mol−1.
The calculated value of the free energy of TA adsorption on steel surface enabled us to assume the chemisorption nature of the interaction between the metal surface and inhibitor molecules because (−ΔGads) > 40 kJ mol−1 [58]. This mode of interaction between the inhibitor and the surface of steels makes it possible to obtain the highest observed protective effect.

3.7. Composition and Structure of the Protective Layers Formed by the TA Inhibitor on Steel Surface

XPS data provide valuable information on the composition and structure of protective layers formed on steel in HCl solutions containing TA. Based on the positions of the complex Fe2p3/2 and Fe2p1/2 peaks in the XPS spectra of iron and their satellite peaks observed at high binding energies (Figure 9), it may be assumed that a layer consisting of Fe3O4 (Eb = 710.8 eV) exists on the steel surface. The presence of various types of oxygen is demonstrated by the O1s spectrum that can be decomposed into three peaks from adsorbed water molecules (Eb = 533.5 eV), hydroxyl groups (531.8 eV), and oxygen in the iron oxide lattice (530.3 eV) (Figure 10).
Despite the ultrasonic cleaning of samples in distilled water that removed the physically bound inhibitor layers from the metal surface, the complex XPS spectrum of N1s electrons (Figure 11) indicated that an inhibitor film was present on the steel surface exposed for 24 h in 2 M HCl + 5 mmol TA. The observed N1s spectrum can be decomposed into two peaks (401.4 and 399.5 eV) with a ratio of ~1:3, where the second peak should be attributed to the nitrogen atoms of the triazole group.
Based on the quantitative ratios of surface atoms in the XPS spectra of steel pre-exposed in the inhibited HCl solution with and without subsequent ultrasonic (US) cleaning, it can be concluded that a polymolecular layer of the inhibitor with a thickness above 4 nm was formed on the steel. After the ultrasonic cleaning of samples, only a monolayer of the inhibitor no thicker than 2 nm remained on the steel surface. Such a layer was strongly retained on the metal due to the chemisorptive interaction of the surface iron atoms and nitrogen atoms of the triazole cycle in the inhibitor. The inhibitor layers arranged above the chemisorbed layer were weakly bound to it and to each other by physical interactions and were removed upon such washing. The chemisorbed layer was not removed from the metal surface upon ultrasonic washing and in XPS studies under high-vacuum conditions. The XPS spectrum of the steel surface contained no peak of Cl2p electrons, which indicated the absence of chloride anions in the film. The metal surface under that layer was oxidized to iron oxide when steel samples were washed in the air.

4. Conclusions

  • ABDMA and TA inhibitors hinder the cathodic evolution of hydrogen and its penetration into metal under the cathodic polarization of steel in HCl solution. The adsorption isotherms of both corrosion inhibitors on cathodically polarized steel surface have been obtained. Treatment of experimental data through IPZ analysis [48] made it possible to determine the kinetic constants of the processes, both in the background medium and in the presence of the inhibitors. In the presence of a CI, the reaction rate of the discharge of H+ ions decreases, whereas the ratio between the surface coverage with hydrogen and its concentration in the metal phase (kinetic-diffusion constant) increases. As a result, the amount of hydrogen absorbed by steel decreases. TA is the most efficient inhibitor of hydrogen absorption. The addition of 5 mM TA reduces the concentration of diffusion-mobile hydrogen in mild steel almost 20-fold.
  • At the corrosion potential, the TA inhibitor decreases the total hydrogen concentration in spring steel more than 8.2-fold at 25 °C (the degree of steel protection from hydrogen absorption is 87.8%). As a result, the plastic properties of steel exhibit almost no change, and its cracking resistance increases considerably. On raising the solution temperature to 60 °C, the degree of steel protection from hydrogen absorption decreases to 50.8%, but even at this inhibitor efficiency, it increases the resistance of steel to cracking.
  • The TA inhibitor significantly decreases the rate of steel anodic dissolution in the HCl solution. This effect and the hindrance of the rate of cathodic hydrogen evolution determine the efficiency of TA as an inhibitor of the acid corrosion of steels. At a TA concentration of 5 mM, the minimum degree of protection is 94.8%.
  • The high efficiency of TA in slowing down the corrosion of steels and in preserving the ductility of the metal is determined by the mechanism of the inhibitor’s protective action. This compound forms a polymolecular protective layer of triazole molecules up to 4 nm thick on the metal in HCl solutions. The triazole monolayer directly adjacent to the metal is chemically bound to it, whereas the overlying layers are bound to it and to each other due to physical interactions.

Author Contributions

Conceptualization, Y.G.A. and Y.I.K.; methodology, A.I.M. and T.A.N.; validation, Y.G.A. and A.I.M.; investigation, A.Y.L., T.A.N., and Y.G.A.; writing—original draft Y.G.A. and A.I.M. preparation, A.Y.L., T.A.N., and Y.G.A.; writing—review and editing, Y.G.A. and A.I.M.; visualization, A.Y.L.; supervision, Y.I.K. and Y.G.A.; project administration, Y.G.A. All authors have read and agreed to the published version of the manuscript.

Funding

The study was carried out within the framework of R&D (2022–2024): “Chemical resistance of materials, protection of metals and other materials from corrosion and oxidation” (registration number in EGISU 122011300078-1, inventory number FFZS-2022-0013).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Obot, I.B.; Meroufel, A.; Onyeachu, I.B.; Alenazi, A.; Sorour, A.A. Corrosion Inhibitors for Acid Cleaning of Desalination Heat Exchangers: Progress, Challenges and Future Perspectives. J. Mol. Liq. 2019, 296, 111760. [Google Scholar] [CrossRef]
  2. Verma, C.; Quraishi, M.A.; Ebenso, E.E. Corrosive electrolytes. Int. J. Corros. Scale Inhib. 2020, 9, 1261–1276. [Google Scholar] [CrossRef]
  3. Leong, V.H.; Ben Mahmud, H. A Preliminary Screening and Characterization of Suitable Acids for Sandstone Matrix Acidizing Technique: A Comprehensive Review. J. Pet. Explor. Prod. Technol. 2018, 9, 753–778. [Google Scholar] [CrossRef] [Green Version]
  4. Van Hong, L.; Ben Mahmud, H. A Comparative Study of Different Acids used for Sandstone Acid Stimulation: A Literature Review. IOP Conf. Series Mater. Sci. Eng. 2017, 217, 12018. [Google Scholar] [CrossRef] [Green Version]
  5. Shafiq, M.U.; Ben Mahmud, H. Sandstone Matrix Acidizing Knowledge and Future Development. J. Pet. Explor. Prod. Technol. 2017, 7, 1205–1216. [Google Scholar] [CrossRef] [Green Version]
  6. Guo, B.; Liu, X.; Tan, X. Chapter 13. Acidizing. In Petroleum Production Engineering, 2nd ed.; Gulf Professional Publishing: Houston, TX, USA, 2017; pp. 367–387. [Google Scholar] [CrossRef]
  7. Agrawal, A.; Sahu, K.K. An Overview of the Recovery of Acid from Spent Acidic Solutions from Steel and Electroplating Industries. J. Hazard. Mater. 2009, 171, 61–75. [Google Scholar] [CrossRef] [PubMed]
  8. Avdeev, Y.G.; Kuznetsov, Y.I. Inhibitory Protection of Steels from High-Temperature Corrosion in Acid Solutions. A Review. Part 1. Int. J. Corros. Scale Inhib. 2020, 9, 394–426. [Google Scholar] [CrossRef]
  9. Finšgar, M.; Jackson, J. Application of Corrosion Inhibitors for Steels in Acidic Media for the Oil and Gas Industry: A Review. Corros. Sci. 2014, 86, 17–41. [Google Scholar] [CrossRef] [Green Version]
  10. Richardson, J.; Bhuiyan, S.H. Corrosion in Hydrogen Halides and Hydrohalic Acids. In Reference Module in Materials Science and Materials Engineering; Elsevier Inc.: Amsterdam, The Netherlands, 2017; pp. 1–22. [Google Scholar] [CrossRef]
  11. Chen, Y.; Chen, Z.; Zhuo, Y. Newly Synthesized Morpholinyl Mannich Bases as Corrosion Inhibitors for N80 Steel in Acid Environment. Materials 2022, 15, 4218. [Google Scholar] [CrossRef] [PubMed]
  12. Al-Senani, G.M.; Al-Saeedi, S.I. The Use of Synthesized CoO/Co3O4 Nanoparticles as A Corrosion Inhibitor of Low-Carbon Steel in 1 M HCl. Materials 2022, 15, 3129. [Google Scholar] [CrossRef] [PubMed]
  13. Khowdiary, M.M.; Taha, N.A.; Saleh, N.M.; Elhenawy, A.A. Synthesis of Novel Nano-Sulfonamide Metal-Based Corrosion Inhibitor Surfactants. Materials 2022, 15, 1146. [Google Scholar] [CrossRef]
  14. Malinowski, S.; Wróbel, M.; Woszuk, A. Quantum Chemical Analysis of the Corrosion Inhibition Potential by Aliphatic Amines. Materials 2021, 14, 6197. [Google Scholar] [CrossRef]
  15. Al-Senani, G.M. Synthesis of ZnO-NPs Using a Convolvulus arvensis Leaf Extract and Proving Its Efficiency as an Inhibitor of Carbon Steel Corrosion. Materials 2020, 13, 890. [Google Scholar] [CrossRef] [Green Version]
  16. Varvara, S.; Berghian-Grosan, C.; Damian, G.; Popa, M.; Popa, F. Combined Electrochemical, Raman Analysis and Machine Learning Assessments of the Inhibitive Properties of an 1,3,4-Oxadiazole-2-Thiol Derivative against Carbon Steel Corrosion in HCl Solution. Materials 2022, 15, 2224. [Google Scholar] [CrossRef] [PubMed]
  17. Hsissou, R.; Benhiba, F.; Abbout, S.; Dagdag, O.; Benkhaya, S.; Berisha, A.; Erramli, H.; Elharfi, A. Trifunctional Epoxy Polymer as Corrosion Inhibition Material for Carbon Steel in 1.0 M Hcl: MD Simulations, DFT and Complexation Computations. Inorg. Chem. Commun. 2020, 115, 107858. [Google Scholar] [CrossRef]
  18. Hsissou, R.; Abbout, S.; Benhiba, F.; Seghiri, R.; Safi, Z.; Kaya, S.; Briche, S.; Serdaroğlu, G.; Erramli, H.; Elbachiri, A.; et al. Insight into the Corrosion Inhibition of Novel Macromolecular Epoxy Resin as Highly Efficient Inhibitor for Carbon Steel in Acidic Mediums: Synthesis, Characterization, Electrochemical Techniques, AFM/UV–Visible and Computational Investigations. J. Mol. Liq. 2021, 337, 116492. [Google Scholar] [CrossRef]
  19. Rbaa, M.; Benhiba, F.; Hssisou, R.; Lakhrissi, Y.; Touhami, M.E.; Warad, I.; Zarrouk, A. Green Synthesis of Novel Carbohydrate Polymer Chitosan Oligosaccharide Grafted On D-Glucose Derivative as Bio-Based Corrosion Inhibitor. J. Mol. Liq. 2020, 322, 114549. [Google Scholar] [CrossRef]
  20. Hsissou, R.; Abbout, S.; Safi, Z.; Benhiba, F.; Wazzan, N.; Guo, L.; Nouneh, K.; Briche, S.; Erramli, H.; Touhami, M.E.; et al. Synthesis and Anticorrosive Properties of Epoxy Polymer for CS in [1 M] HCl Solution: Electrochemical, AFM, DFT and MD Simulations. Constr. Build. Mater. 2020, 270, 121454. [Google Scholar] [CrossRef]
  21. Bin-Hudayb, N.S.; Badr, E.E.; Hegazy, M. Adsorption and Corrosion Performance of New Cationic Gemini Surfactants Derivatives of Fatty Amido Ethyl Aminium Chloride with Ester Spacer for Mild Steel in Acidic Solutions. Materials 2020, 13, 2790. [Google Scholar] [CrossRef]
  22. Chen, L.; Lu, D.; Zhang, Y. Organic Compounds as Corrosion Inhibitors for Carbon Steel in HCl Solution: A Comprehensive Review. Materials 2022, 15, 2023. [Google Scholar] [CrossRef]
  23. Gao, C.; Wang, S.; Dong, X.; Liu, K.; Zhao, X.; Kong, F. Construction of a Novel Lignin-Based Quaternary Ammonium Material with Excellent Corrosion Resistant Behavior and Its Application for Corrosion Protection. Materials 2019, 12, 1776. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  24. Abd-Elaal, A.; Shaban, S.; Tawfik, S.M. Three Gemini Cationic Surfactants Based on Polyethylene Glycol as Effective Corrosion Inhibitor for Mild Steel in Acidic Environment. J. Assoc. Arab Univ. Basic Appl. Sci. 2017, 24, 54–65. [Google Scholar] [CrossRef] [Green Version]
  25. Chauhan, D.S.; Quraishi, M.A.; Mazumder, M.A.J.; Ali, S.A.; Aljeaban, N.A.; Alharbi, B.G. Design and Synthesis of a Novel Corrosion Inhibitor Embedded with Quaternary Ammonium, Amide and Amine Motifs for Protection of Carbon Steel in 1 M HCl. J. Mol. Liq. 2020, 317, 113917. [Google Scholar] [CrossRef]
  26. Hegazy, M.; Abdallah, M.; Awad, M.; Rezk, M. Three Novel Di-Quaternary Ammonium Salts as Corrosion Inhibitors for API X65 Steel Pipeline in Acidic Solution. Part I: Experimental Results. Corros. Sci. 2014, 81, 54–64. [Google Scholar] [CrossRef]
  27. Kaczerewska, O.; Leiva-Garcia, R.; Akid, R.; Brycki, B. Efficiency of Cationic Gemini Surfactants with 3-Azamethylpentamethylene Spacer as Corrosion Inhibitors for Stainless Steel in Hydrochloric Acid. J. Mol. Liq. 2017, 247, 6–13. [Google Scholar] [CrossRef] [Green Version]
  28. Mahdavian, M.; Tehrani-Bagha, A.R.; Alibakhshi, E.; Ashhari, S.; Palimi, M.J.; Farashi, S.; Javadian, S.; Ektefa, F. Corrosion of Mild Steel in Hydrochloric Acid Solution in the Presence of Two Cationic Gemini Surfactants with and Without Hydroxyl Substituted Spacers. Corros. Sci. 2018, 137, 62–75. [Google Scholar] [CrossRef]
  29. Prathibha, B.; Nagaswarupa, H.; Kotteeswaran, P.; BheemaRaju, V. Inhibiting effect of Quaternary Ammonium Compound on the Corrosion of Mild Steel in 1 M Hydrochloric Acid Solution, Its Adsorption and Kinetic Characteristics. Mater. Today: Proc. 2017, 4, 12245–12254. [Google Scholar] [CrossRef]
  30. Tawfik, S.M. Ionic Liquids Based Gemini Cationic Surfactants as Corrosion Inhibitors for Carbon Steel in Hydrochloric Acid Solution. J. Mol. Liq. 2016, 216, 624–635. [Google Scholar] [CrossRef]
  31. Ansari, K.R.; Chauhan, D.S.; Singh, A.; Saji, V.S.; Quraishi, M.A. Corrosion Inhibitors for Acidizing Process in Oil and Gas Sectors. In Corrosion Inhibitors in the Oil and Gas Industry; Saji, V.S., Umoren, S.A., Eds.; Wiley-VCH Verlag GmbH & Co. KGaA: Weinheim, Germany, 2020; pp. 153–176. [Google Scholar] [CrossRef]
  32. Quraishi, M.A.; Chauhan, D.S.; Saji, V.S. Heterocyclic Corrosion Inhibitors for Acid Environments. In Heterocyclic Organic Corrosion Inhibitors; Quraishi, M.A., Chauhan, D.S., Saji, V.S., Eds.; Elsevier Inc.: Amsterdam, The Netherlands, 2020; pp. 87–131. [Google Scholar] [CrossRef]
  33. Swathi, N.P.; Alva, V.D.P.; Samshuddin, S. A Review on 1,2,4-Triazole Derivatives as Corrosion Inhibitors. J. Bio Tribo-Corrosion 2017, 3, 42. [Google Scholar] [CrossRef]
  34. Nitrogen-Containing Five-Membered Heterocyclic Compounds as Corrosion Inhibitors for Metals in Solutions of Mineral Acids—An Overview. Int. J. Corros. Scale Inhib. 2021, 10, 480–540. [CrossRef]
  35. Avdeev, Y.G.; Kuznetsov, Y.I. Inhibitory Protection of Steels in Acid Solutions Under High-Temperature Corrosion Conditions. A Review. Part 3. Int. J. Corros. Scale Inhib. 2020, 9, 1194–1236. [Google Scholar] [CrossRef]
  36. Pradhan, A.; Vishwakarma, M.; Dwivedi, S.K. A Review: The Impact of Hydrogen Embrittlement on the Fatigue Strength of High Strength Steel. Mater. Today: Proc. 2020, 26, 3015–3019. [Google Scholar] [CrossRef]
  37. Ohaeri, E.; Eduok, U.; Szpunar, J. Hydrogen Related Degradation in Pipeline Steel: A Review. Int. J. Hydrog. Energy 2018, 43, 14584–14617. [Google Scholar] [CrossRef]
  38. Liu, Q.; Zhou, Q.; Venezuela, J.; Zhang, M.-X.; Wang, J.; Atrens, A. A Review of the Influence of Hydrogen on the Mechanical Properties Of DP, TRIP, and TWIP Advanced High-Strength Steels for Auto Construction. Corros. Rev. 2016, 34, 127–152. [Google Scholar] [CrossRef]
  39. Lunarska, E.; Nikiforov, K. Hydrogen Degradation of the Refinery and Electric Power Installations. Corros. Rev. 2008, 26, 173–213. [Google Scholar] [CrossRef]
  40. Ramamurthy, S.; Atrens, A. Stress Corrosion Cracking of High-Strength Steels. Corros. Rev. 2013, 31, 1–31. [Google Scholar] [CrossRef]
  41. Muralidharan, S.; Quraishi, M.; Iyer, S. The Effect of Molecular Structure on Hydrogen Permeation and the Corrosion Inhibition of Mild Steel in Acidic Solutions. Corros. Sci. 1995, 37, 1739–1750. [Google Scholar] [CrossRef]
  42. Marshakov, A.I.; Nenasheva, T.A.; Rybkina, A.A.; Maleeva, M. On Inhibiting the Iron Anodic Dissolution in Acid Sulfate Electrolyte by Tetrabutylammonium Cations. Prot. Met. 2007, 43, 77–83. [Google Scholar] [CrossRef]
  43. Kumar, H.; Karthikeyan, S.; Vivekanand, P.A.; Kamaraj, P. The Inhibitive Effect of Cloxacillin on Mild Steel Corrosion In 2 N Sulphuric Acid Medium. Mater. Today Proc. 2020, 36, 898–902. [Google Scholar] [CrossRef]
  44. Devanathan, M.A.V.; Stachurski, Z. The Adsorption and Diffusion of Electrolytic Hydrogen in Palladium. Proc. R. Soc. Lond. Ser. A Math. Phys. Sci. 1962, 270, 90–102. [Google Scholar] [CrossRef]
  45. Devanathan, M.A.V.; Stachurski, Z. The Mechanism of Hydrogen Evolution on Iron in Acid Solutions by Determination of Permeation Rates. J. Electrochem. Soc. 1964, 111, 619–623. [Google Scholar] [CrossRef]
  46. Wagner, C.D.; Davis, L.E.; Zeller, M.V.; Taylor, J.A.; Raymond, R.H.; Gale, L.H. Empirical Atomic Sensitivity Factors for Quantitative Analysis by Electron Spectroscopy for Chemical Analysis. Surf. Interface Anal. 1981, 3, 211–225. [Google Scholar] [CrossRef]
  47. Shirley, D.A. High-Resolution X-Ray Photoemission Spectrum of the Valence Bands of Gold. Phys. Rev. B 1972, 5, 4709. [Google Scholar] [CrossRef] [Green Version]
  48. Iyer, R.N.; Pickering, H.W.; Zamanzadeh, M. Analysis of Hydrogen Evolution and Entry into Metals for the Discharge-Recombination Process. J. Electrochem. Soc. 1989, 136, 2463–2470. [Google Scholar] [CrossRef]
  49. Popov, B.N.; Lee, J.-W.; Djukic, M.B. Hydrogen Permeation and Hydrogen-Induced Cracking. Cracking. Chapter 7. In Handbook of Environmental Degradation of Materials, 3rd. ed.; Elsevier Inc.: Amsterdam, The Netherlands, 2018; pp. 133–162. [Google Scholar] [CrossRef]
  50. Marshakov, A.I.; Rybkina, A.A.; Skuratnik, Y.B. Effect of Absorbed Hydrogen on the Iron Dissolution. Russ. J. Electrochem. 2000, 36, 1101–1107. [Google Scholar] [CrossRef]
  51. Marshakov, A.; Nenasheva, T.A. The Effect of Sorbed Hydrogen on the Dissolution of Iron in a Thiocyanate-Containing Sulfuric Acid. Prot. Met. Phys. Chem. Surf. 2001, 37, 543–552. [Google Scholar] [CrossRef]
  52. Damaskin, B.B.; Afanas’ev, B.N. Current State of the Theory of the Effect of the Adsorption of Organic Substances on the Kinetics of Electrochemical Reactions. Sov. Elektrokhimiya 1977, 13, 1099–1117. [Google Scholar]
  53. Marshakov, A.; Nenasheva, T.A. Effect of Sorbed Hydrogen on Iron Dissolution in the Presence of Tetraethylammonium Cations. Prot. Met. Phys. Chem. Surf 2002, 38, 556–562. [Google Scholar] [CrossRef]
  54. Kasatkin, V.E.; Kornienko, L.P.; Shcherbakov, A.I.; Korosteleva, I.G.; Kasatkina, I.V.; Dorofeeva, V.N. Effect of Sodium Hydroxide Concentration on Nickel Surface State: Experience in EIS Methods Applying. Electrocatalysis 2022, 13, 539–550. [Google Scholar] [CrossRef]
  55. Marshakov, A.I.; Nenasheva, T.A. Study of the Surface State of Nickel Electrode During Polarization in an Alkaline Medium by Electrochemical Impedance Spectroscopy Method. Int. J. Corros. Scale Inhib. 2020, 9, 1516–1529. [Google Scholar] [CrossRef]
  56. Afanas’ev, B.N.; Skobochkina, Y.P. Physicochemical Bases of the Action of Cor-Rosion Inhibitors. In Izhevsk; Publishing House of UdGu: Izhevsk, Russia, 1990; 20p. (In Russian) [Google Scholar]
  57. Kiuchi, K.; McLellan, R. The Solubility and Diffusivity of Hydrogen in Well-Annealed and Deformed Iron. Acta Met. 1983, 31, 961–984. [Google Scholar] [CrossRef]
  58. Harvey, T.; Walsh, F.; Nahlé, A. A Review of Inhibitors for the Corrosion of Transition Metals in Aqueous Acids. J. Mol. Liq. 2018, 266, 160–175. [Google Scholar] [CrossRef]
Figure 1. Devanatkhan-Stakhursky electrochemical cell: 1—working electrode (membrane); 2—Teflon holder; 3—working part of the cell; 3′—diffusion part of the cell; 4, 4′—auxiliary electrode cell; 5, 5′—auxiliary electrode; 6, 6’—electrolytic switch; 7, 7′—reference electrode; 8, 8’—tap for draining the solution; 9, 9′—solution input into the cell; 10—argon input into the cell; 11—water seal.
Figure 1. Devanatkhan-Stakhursky electrochemical cell: 1—working electrode (membrane); 2—Teflon holder; 3—working part of the cell; 3′—diffusion part of the cell; 4, 4′—auxiliary electrode cell; 5, 5′—auxiliary electrode; 6, 6’—electrolytic switch; 7, 7′—reference electrode; 8, 8’—tap for draining the solution; 9, 9′—solution input into the cell; 10—argon input into the cell; 11—water seal.
Materials 15 06989 g001
Figure 2. Cathodic polarization curves on steel (A) and plots of the rate of hydrogen penetration into steel vs. potential (B) in 2 M HCl containing ABDMA.
Figure 2. Cathodic polarization curves on steel (A) and plots of the rate of hydrogen penetration into steel vs. potential (B) in 2 M HCl containing ABDMA.
Materials 15 06989 g002
Figure 3. Cathodic polarization curves on steel (A) and plots of the rate of hydrogen penetration into steel vs. potential (B) in 2 M HCl containing TA.
Figure 3. Cathodic polarization curves on steel (A) and plots of the rate of hydrogen penetration into steel vs. potential (B) in 2 M HCl containing TA.
Materials 15 06989 g003
Figure 4. Plot of the current of hydrogen penetration through the membrane vs. the rate of its chemical recombination in 2 M HCl.
Figure 4. Plot of the current of hydrogen penetration through the membrane vs. the rate of its chemical recombination in 2 M HCl.
Materials 15 06989 g004
Figure 5. Plot of the function f = i c exp ( α F E R T ) vs. the current of hydrogen penetration through the membrane in 2 M HCl.
Figure 5. Plot of the function f = i c exp ( α F E R T ) vs. the current of hydrogen penetration through the membrane in 2 M HCl.
Materials 15 06989 g005
Figure 6. Equivalent electrical circuit and Nyquist plots of a steel electrode (0.64 cm2) in 2.0 M HCl recorded after adding 0.01 mmol TA to the solution with various exposure times.
Figure 6. Equivalent electrical circuit and Nyquist plots of a steel electrode (0.64 cm2) in 2.0 M HCl recorded after adding 0.01 mmol TA to the solution with various exposure times.
Materials 15 06989 g006
Figure 7. Adsorption isotherm of TA and ABDMA on MS steel (E = 0.30 V) from 2.0 M HCl.
Figure 7. Adsorption isotherm of TA and ABDMA on MS steel (E = 0.30 V) from 2.0 M HCl.
Materials 15 06989 g007
Figure 8. Polarization curves of SS steel at t = 25 °C in 2M HCl in the presence of the following inhibitors ABDMA and TA.
Figure 8. Polarization curves of SS steel at t = 25 °C in 2M HCl in the presence of the following inhibitors ABDMA and TA.
Materials 15 06989 g008
Figure 9. Standard XPS spectrum of Fe2p electrons (spin orbital splitting—doublet) of steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h).
Figure 9. Standard XPS spectrum of Fe2p electrons (spin orbital splitting—doublet) of steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h).
Materials 15 06989 g009
Figure 10. XPS spectra of O1s electrons on the steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h).
Figure 10. XPS spectra of O1s electrons on the steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h).
Materials 15 06989 g010
Figure 11. XPS spectra of N1s electrons on the steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h) followed by washing in an ultrasonic bath.
Figure 11. XPS spectra of N1s electrons on the steel surface after preliminary inhibitor adsorption (2.0 M HCl + 5.0 mmol TA, 25 °C, 24 h) followed by washing in an ultrasonic bath.
Materials 15 06989 g011
Table 1. Kinetic constants, coverage of the metal surface with the inhibitors (θinh) and hydrogen atoms (θH), and surface concentration of diffusion-mobile hydrogen ( C H s ) under the cathodic polarization (E = −0.3 V) of steel in 2 M HCl solution containing ABDMA and TA.
Table 1. Kinetic constants, coverage of the metal surface with the inhibitors (θinh) and hydrogen atoms (θH), and surface concentration of diffusion-mobile hydrogen ( C H s ) under the cathodic polarization (E = −0.3 V) of steel in 2 M HCl solution containing ABDMA and TA.
Inhibitor θinhk1,i,
mol cm−2 s−1
k,
cm3 mol−1
kr,
mol cm−2 s−1
θH ×100 C H s . ,   × 10 8
mol cm−3
Background 9.73 × 10−93.5 × 1057.5 × 10−63.4 ± 0.110.0 ± 0.2
ABDMA
+0.01 mM0.754.7 × 10−93.8 × 1056.8 × 10−62.0 ± 0.14.4 ± 0.4
+0.05 mM0.922.0 × 10−94.9 × 1055.4 × 10−61.2 ± 0.12.0 ± 0.2
+0.1 mM0.941.5 × 10−91.3 × 1061.0 × 10−62.3 ± 0.11.6 ± 0.07
+0.5 mM0.941.3 × 10−91.8 × 1065.5 × 10−72.8 ± 0.11.4 ± 0.07
+1 mM0.941.5 × 10−92.2 × 1062.9 × 10−74.2 ± 0.11.8 ± 0.06
+5 mM0.951.5 × 10−92.6 × 1063.0 × 10−73.9 ± 0.11.5 ± 0.04
+10 mM0.961.4 × 10−93.0 × 1062.9 × 10−73.4 ± 0.11.3 ± 0.04
TA
+0.01 mM0.427.2 × 10−94.7 × 1052.5 × 10−64.6 ± 0.18.9 ± 0.5
+0.05 mM0.725.0 × 10−98.9 × 1059.1 × 10−75.6 ± 0.25.6 ± 0.3
+0.1 mM0.863.2 × 10−99.2 × 1051.1 × 10−63.7 ± 0.13.4 ± 0.2
+0.5 mM0.961.2 × 10−92.4 × 1064.0 × 10−73.0 ± 0.11.4 ± 0.07
+1 mM0.986.3 × 10−102.4 × 1063.9 × 10−71.9 ± 0.10.75 ± 0.1
+5 mM0.991.1 × 10−102.5 × 1066.0 × 10−71.1 ± 0.10.55 ± 0.5
+10 mM0.993.4 × 10−103.3 × 1064.4 × 10−71.1 ± 0.10.41 ± 0.3
Table 2. Effect of 5 mmol of nitrogen-containing inhibitors on the corrosion, hydrogen absorption, and ductility of SS steel in 2 M HCl.
Table 2. Effect of 5 mmol of nitrogen-containing inhibitors on the corrosion, hydrogen absorption, and ductility of SS steel in 2 M HCl.
Inhibitort, °Cρ, g m−2·h−1Zcor, % ( C H v . ) ,   mol   cm 3 ZvH*, %p, %
Background2511.4-3.2 × 10−5-- *
ABDMA256.3644.21.2 × 10−562.7-
TA250.5994.83.9 × 10−687.8100
Background6048.7-1.8 × 10−5--
ABDMA6016.566.16.9 × 10−661.3-
TA602.495.18.8 × 10−650.891
* Total loss of ductility.
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Avdeev, Y.G.; Nenasheva, T.A.; Luchkin, A.Y.; Marshakov, A.I.; Kuznetsov, Y.I. Effect of Quaternary Ammonium Salts and 1,2,4-Triazole Derivatives on Hydrogen Absorption by Mild Steel in Hydrochloric Acid Solution. Materials 2022, 15, 6989. https://doi.org/10.3390/ma15196989

AMA Style

Avdeev YG, Nenasheva TA, Luchkin AY, Marshakov AI, Kuznetsov YI. Effect of Quaternary Ammonium Salts and 1,2,4-Triazole Derivatives on Hydrogen Absorption by Mild Steel in Hydrochloric Acid Solution. Materials. 2022; 15(19):6989. https://doi.org/10.3390/ma15196989

Chicago/Turabian Style

Avdeev, Yaroslav G., Tatyana A. Nenasheva, Andrey Yu. Luchkin, Andrey I. Marshakov, and Yurii I. Kuznetsov. 2022. "Effect of Quaternary Ammonium Salts and 1,2,4-Triazole Derivatives on Hydrogen Absorption by Mild Steel in Hydrochloric Acid Solution" Materials 15, no. 19: 6989. https://doi.org/10.3390/ma15196989

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop