Performance and computational studies of new soluble triazole as corrosion inhibitor for carbon steel in HCl

Performance and computational studies of new soluble triazole as corrosion inhibitor for carbon steel in HCl

Chemical Data Collections 22 (2019) 100242 Contents lists available at ScienceDirect Chemical Data Collections journal homepage: www.elsevier.com/lo...

3MB Sizes 0 Downloads 82 Views

Chemical Data Collections 22 (2019) 100242

Contents lists available at ScienceDirect

Chemical Data Collections journal homepage: www.elsevier.com/locate/cdc

Data Article

Performance and computational studies of new soluble triazole as corrosion inhibitor for carbon steel in HCl Z. Rouifi a, F. Benhiba a, M. El Faydy b, T. Laabaissi a, H. About a, H. Oudda a, I. Warad c, A. Guenbour d, B. Lakhrissi b, A. Zarrouk d,∗ a

Laboratory of Separation Procedure, Faculty of Sciences, IbnTofail University, Kenitra, Morocco Laboratory of Agro-Resources, Polymers and Process Engineering, Faculty of Sciences, Ibn Tofail University, Kenitra, Morocco c Department of Chemistry, AN-Najah National University, P.O. Box 7, Nablus, Palestine d Laboratory of Materials, Nanotechnology and Environment, Faculty of Sciences, Mohammed V University, Av. Ibn Battouta, P.O. Box 1014 Agdal-Rabat, Morocco b

a r t i c l e

i n f o

Article history: Received 13 January 2019 Revised 24 May 2019 Accepted 7 June 2019 Available online 11 June 2019

a b s t r a c t The current investigation reports the corrosion inhibition and adsorption behavior of new 4-amino-1-((8-hydroxyquinolin-5-yl)methyl)-1,2,3-triazole-5-carboxylate (MHTC) for our substrate [carbon steel (CS)] exposed to 1 M HCl. Tafel curves, electrochemical impendence spectroscopy (EIS) measurements and mass loss (ML) estimations were performed to analyze the corrosion performance ability of MHTC. The inhibitory efficacy of MHTC increases with its concentration, it gives a maximum inhibition efficiency of 91% at 1Mm. Polarization studies indicate that MHTC acts as a mixed corrosion inhibitor. The adsorption of MHTC on the surface of CS obeyed the Langmuir isotherm. UV–visible and SEM techniques were carried out for surface investigation of the CS. The reactivity of MHTC was quantum chemically analyzed by the DFT method to clarify the effectiveness of this soluble inhibitor. Adsorption model of MHTC molecule on CS surface was investigated using Monte Carlo (MC) simulations, which revealed that MHTC adsorb on Fe(110) surface in flat orientation. © 2019 Elsevier B.V. All rights reserved.

Keywords: Synthesis Carbon steel Corrosion inhibition UV-visible spectrophotometry DFT

Specifications Table Subject area Compounds Data category Data acquisition format Data type Procedure

Data accessibility



Organic Chemistry, Spectroscopy, Computational Chemistry, Physical Chemistry, Corrosion Science. New triazole derivative. Spectral, synthesized, mass loss, polarisation curves, electrochemical impedance spectroscopy, Scanning electron microscopy, UV-visible, SEM, computational simulations. NMR, elemental analysis, computational simulations, molecular dynamics, polarisation curves,electrochemical impedance spectroscopy, scanning electron microscopy. Electrochemical impedance spectroscopy and polarization curves measurement. Impedance diagrams and Bode plots. Fitted data. The evaluation was carried out using mass loss, electrochemical impedance spectroscopy and polarization curves measurement. Impedance diagrams and Bode plots for uninhibited and inhibited systems was analysed using Zview program. The fitted data observed trails nearly the same pattern as the experimental results data is with this article or in public repository.

Corresponding author. E-mail address: [email protected] (A. Zarrouk).

https://doi.org/10.1016/j.cdc.2019.100242 2405-8300/© 2019 Elsevier B.V. All rights reserved.

2

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

1. Rationale These days the investigation of corrosion of carbon steel in acidic solution has find broad application in different chemical process enterprises including: petrochemical, oil well acidizing, corrosive pickling, and so forth [1,2]. A few examinations on the use of some organic compounds as efficient corrosion inhibitors for carbon steel should be evaluated according to the certqin parameters of the system, since preventive measures, utilized successfully in a given environment, may be harmful under other conditions. The use of inhibitors, in particular, heterocyclic organic compounds, is a very common means of use to fight against corrosion. These compounds contain active centers such as ring π electrons and heteroatoms, for example: (nitrogen, oxygen, and sulfur). This promotes their absorption and leads to the reduction of corrosion of the metal in question [3–6]. Several factors can influence the adsorption of inhibitors. Let’s mention the load metal electronics at the metalmedium interface, the molecular structure and functional group of the inhibitor, the interactions with the solvent, with other inhibitors, the oxidation–reduction reactions that undergoes the inhibitor [7,8]. The inhibitory properties of N-(8hydroxyquinolin-5-yl)-methyl)-N-phenylacetamide [9], 5-azidomethyl-8-quinolinol [10] has been studied for CS exposed to 1 M HCl. As of late, The calculation of DFT quantum chemistry is very effective in studying the link between the molecular properties of inhibitors and its effectiveness in inhibiting corrosion. Quantum chemical calculations which are based on the density functional theory (DFT) have assisted in the explaining of experimental data and allow access to phenomena that are hard to observe experimentally [11]. Thus, the purpose of this study is to investigate MHTC as an inhibitor for CS in 1 M HCl by electrochemical methods, weight loss and surface morphology analysis (SEM and UV-Visible). Further study by means of theoretical calculation and MC simulation was aimed at inhibition mechanism. Rationale behind the creation of the dataset(s), experimental procedure, protocol, computational tool, etc. 2. Procedure 2.1. Experimental details 2.1.1. Materials The elemental composition (wt.%) of utilized CS samples were: C = 0.37%, Mn = 0.68%, Si = 0.23%, Cu = 0.16%, S = 0.016%, Cr = 0.077%, Co = 0.09%, Ti = 0.011% Ni = 0.059% and Fe = 98.307%. The electrolyte is a solution of hydrochloric acid molar (1 M HCl), obtained by diluting concentrated acid with a density of 1.19 and a purity of 37%. The choice concentration of MHTC worn was 10−6 to 10− 3 M. 2.1.2. Synthesis(MHTC) The preparation of MHTC is laid out in Scheme 1, beginning from (AZHQ) which was synthesized according to the method described by Himmi et al. [12]. The compound (MHTC) was synthesized by condensation reaction of 1 mol of (AZHQ), 2 mol (MCA) and 1 mol of Potassium carbonate in chloroform. The reactants were heated at 60 °C during one day. The reaction was monitored by TLC. The mixture was cooled and then washed with 50 mL of distilled water, the extraction was carried out with 3 × 30 mL of ethyl acetate. The organic phases were combined, dried and evaporated. The product was then purified by silica column chromatography (benzene/acetone) in a gradient (9: 1 to 5: 5). To furnish the product MHTC (0.79 g, 75%) as white solid, mp 168–170 °C. The compound was characterised using 1 H NMR and 13 C NMR. 1 H NMR (DMSO – d 6 ): δ ppm = 3.69 (s, 3 H, CH3 ), 7.030–8.49 (m,5 H, aromatic quinoline), 4.50 (s,2H,quinoline-CH2 -triazole), 8.86 (s,2 H,NH 2 ), 3.40 (s, H of trace H2 O present into DMSO–d6) (Fig. 1). 13C NMR(DMSO – d6): δ ppm = 30.63 (CH3 -C = O),53.708 (quinoline-CH2 -),111.056–153.389(CH. C quinoline and triazole),166.980(CH3 -C = O) (Fig. 2). Elemental analysis for C14 H13 N5 O3 C, 56.18; H, 4.38; N, 23.40; O, 16.04. 2.2. Corrosion test 2.2.1. Mass loss(LM) method The details of instrumental techniques (weigt loss) in this study are the same as discussed previously [13]. The value of this quantity (w) expressed in mg cm−2 h−1 was obtained by exploiting the next formula:

w=

m S×t

(1)

where m is the mass loss of CS, S is the total area of CS specimen with a size of (3 × 3 × 0.3 cm3 ), and t is the immersion period (6 h). The inhibition performance, ηLM (%), was calculated as [14–16]:

  w ηLM (% ) = 1 − i × 100 w0

(2)

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 1.

1

Fig. 2.

13

H NMR spectrum of (MHTC).

C NMR spectrum of (MHTC).

3

4

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

where w0 and wi are the values of the corrosion rates of CS specimen in the absence and presence of MHTC. 2.2.2. Electrochemical studies The details of instrumental techniques (electrochemical techniques) and scientific software utilized in this study are the same as discussed previously [13]. The inhibition efficiency for polarization curves using the following relation (3) [17,18].

  i ηTa f el (% ) = 1 − corr (i) × 100 icorr

(3)

Where icor and icor(i) are the currents of corrosion existence and nonexistence, correspondingly. The inhibition efficiency for EIS measurements was calculated in terms of polarization resistance using the following Eq. (4) [17,18].



ηEIS (% ) = 1 −

Rp



R p/inh

× 100

(4)

wherever Rp and Rp/inh are the values of polarization resistance observed in the nonexistence and the existence of inhibitor respectively. 2.3. Morphology measurements(SEM) Surface analyzes employ the SEM was performed for secondary electron imaging observation giving to the topography of the specimen. Our SEM observations were based on CS samples after immersion for 6 h at 298 K in 1 M HCl alone and with the adding of 1 mM MHTC in the electrolyte. 2.4. UV-visible spectroscopy In this study, UV-visible spectrometry is utilized to analyze corrosive solutions containing 1 mM MHTC. The electron absorption spectra (UV-Visible) were wear out on a Jenway spectrophotometer (series 67). A quartz cell with 1 cm path length was utilized. 2.5. Computational chemical details The study of the structure, the distribution of electrons and the adsorption of molecules on the surfaces of metal and oxide, the mechanisms of inhibition in the depths explored. The quantum chemical method utilizing DFT for performance ability of inhibitors with their molecular properties [19,20]. in this study the complete optimization of MHTC was carried out by the method of DFT / (B3LYP) and 6–31 G (d, p) by the Gaussian logciel 09. For these seek, molecular descriptors, such as ELUMO , EHOMO . From these values on the various quantum parametric parameters such as the gap energy (E), the dipole moment (μ), the absolute electronegativity (χ ), the number of electrons transferred (N) and the overall duration (η).

IE = −EHOMO

(5)

EA = −ELUMO

(6)

The absolute electronegativity values χ , the absolute hardness η and the softness σ (The inverse of hardness) can be determined from the values of EI and EA utilizing the following expressions [20]:

χ=

IE + EA 2

(7)

η=

IE − EA 2

(8)

σ=

1

η

=−

2 EHOMO − ELUMO

(9)

Therefore, the fraction of electrons transferred (N) from the inhibiting molecule to the metal atom is calculated according to the following equation [21,22]:

N110 =

 − χinh 2(ηF e + ηinh )

(10)

Where φ is work function, χ inh denote the absolute electronegativity of inhibitor molecule. Also, ηFe and ηinh denote the absolute hardness of iron and the inhibitor molecule respectively. In this study, we use the theoretical value of φ = 4.82 eV and ηFe = 0, for calculating the number of electron transferred.

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

5

Table 1 Calculated values of corrosion rate and performance capacity obtained by weight loss at various concentrations of MHTC in 1 M hydrochloric acid medium. Medium

Conc (M)

W (mg cm−2 h−1 )

ηWL (%)

Blank MHTC

1 10−3 10−4 10−5 10−6

0.429 0.061 0.067 0.103 0.185

— 85 84 76 56

2.6. Monte Carlo (MC) simulations MC simulations of the tested inhibitor were carried out in a simulation box with periodic boundary conditions utilizing Materials Studio 8.0 (from Accelrys Inc.). First the molecular structures of the inhibitor (MHTC), water (H2 O) and Hydronium ion (H3 O+ ) were geometrically fully optimized utilizing the Forcite module. Then the iron crystal was imported and cleaved along (110) plane and a slab of 6 A˚ was employed. So Fe (110) surface was enlarged to a (12 × 12) supercell to provide a large surface for the interaction of the inhibitor with utilizing a vacuum of 29 A˚ along the C-axis. An iron maille with a ˚ contains 500 H2 O, 5Cl− , 5H3 O+ and a molecule of tested inhibitor was created. The values size of a = b = 27.45 A˚ c = 29.14 A, of binding energies were obtained for the most stable configurations of MHTC these values are calculated by the following equations:



Einteraction = Etotal − Esur f ace+solution + Einhibitor



Ebinding = −Einteraction

(11) (12)

Where Etotal is the total energy of the entire system. Esurface+solution referred to the total energy of Fe (110) surface and solution without the inhibitor and Einhibitor represent the total energy of inhibitor. 3. Data, value and validation 3.1. Loss mass(LM) analysis The loss mass experiments were carried out in 1 M HCl with the nonexistence and the existence of diverse concentrations of MHTC (10−3 M, 10− 4 M, 10− 5 M and 10− 6 M) at ambient temperature for 6 h immersion. Corrosion rate (w) and performance ability(ηWL ) values after 6 h of CS immersion in the acidic environment at different concentrations of MHTC are illustrated in Table 1. The results reveal that weight loss and the corrosion rate vary inversely while inhibition effectiveness vary directly with increasing inhibitor dosage up to the optimum concentration. The corrosion rate of 0.429 mg cm−2 h−1 declined to 0.061 mg cm−2 h−1 in corrosive system containing 10−3 M MHTC and the corresponding η is 85%. It is commonly recognized that the organic molecules perform throughout adsorption (physisorptions or /and chemisorption) at the CS/corrosive solution interface [23]. So, the adsorption of MHTC under the neutral form might occur because of the formation of links with the d-orbital of metal atoms, linking the displacement of water(H2 O) molecules from the metal surface, and the lone pairs exist on the five oxygen atoms and three nitrogen atoms and π -orbitals in the traizole and quinoline rings act by blocking the active sites of acid attack, reflecting a diminish of corrosion rate [24]. 3.2. Electrochemical experiments(Tafel curves and EIS) 3.2.1. Open circuit potential The variation in open circuit potential (Eocp) as a function of immersion time for CS in 1 M HCl in the absence and presence of MHTC is represented in Fig 3. In blank solution, the Eocp values are starting at −458 mV/SCE, then shifted anodically and reach the steady state after 270 s; this can be construed as the initial dissolution of the air formed oxide film on CS and the commencement of the attack of the metal by the acid molecules. With the addition of MHTC, Eocp started at a relatively positive potential with respect to in its absence, then shifted cathodically. The steady OCP increases toward the positive direction with increasing concentrations of MHTC. 3.2.2. Tafel curves analysis Polarization measurements have been employed bring together information concerning the kinetics of anodic and cathodic reactions. The potentiodynamic polarization curves for CS in 1 M HCl solution in the nonexistence and the existence of diverse concentrations of MHTC are publicized in Fig. 4. Their extrapolation parameters and performance capacity values are given in Table 2.

6

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 3. The variation in open circuit potential (Eocp) as a function of immersion time for CS in 1 M HCl in the absence and presence of MHTC.

Fig. 4. Potentiodynamic polarization curves for CS in 1 M HCl containing different concentrations of MHTC.

Table 2 Electrochemical quantities and performance ability of the corrosion of CS in 1 M HCl not including and through addition of MHTC at different concentrations. Medium

Cinh (M)

-Ecorr (mV/SCE)

-β c (mV dec−1 )

icor (μA cm−2 )

ηTafel (%)

θ

Blank MHTC

1 10−3 10−4 10−5 10−6

454.4 467.8 452.8 467.8 451.6

101.1 84.7 91.5 84.1 86.1

590.8 54.6 70.1 74.1 246.0

— 91 88 87 58

— 0.91 0.88 0.87 0.58

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

7

Fig. 5. The Nyquist, phase angle and Bode curves at the steel / MHTC / HCl interface at 298 K.

Fig. 6. Equivalent circuit diagram used to fit impedance data, blank and presence of MHTC.

Inspection of Fig. 4 shows that the addition of MHTC has an inhibitive effect on both anodic and cathodic parts of the polarization curves and shifts both the anodic and cathodic curves to lower current densities. This may be ascribed to adsorption of the inhibitor over the metal surface. These results lead to MHTC reducing anodic dissolution as well as delaying the reduction of H+ protons. Therefore, the largest displacement in Ecorr value is less than 85 mV, signifying that MHTC is a typical mixed (anodic and cathodic effect) inhibitor for CS in 1 M HCl. However, the inhibitor addition does not change the hydrogen evolution reaction mechanism such as indicated by the slight changes in the cathodic slopes (β c ) values. The values of polarization parameters are displayed in Table 2. It is evidently seen that the analysis in Table 2 indicates that corrosion currents (icor ) decrease with increasing concentrations of MHTC. We note that the icor is diminished in the presence of 1 mm MHTC and becomes only 54.6 μA cm−2 . The performance ability of MHTC increases with its concentration. This increase indicates performance indicates that the MHTC molecules are adsorbed on the metal surface (CS). The use of the triazole derivatives as corrosion inhibitors have been widely reported by several authors [25–29]. As an example, Table 3 reports the percentage inhibition efficiency for some selected triazole derivatives used as corrosion inhibitors in 1 M HCl medium [30–32]. The values of inhibition efficiency, given in this table, were obtained using polarization curves measurement after 30 min of immersion in 1 M HCl solution containing 10−3 M of trizaole derivative at 298 K. The variation in inhibitive efficiency mainly depends on the type and the nature of the substituents present in the inhibitor molecule. it is remarkable that the alkylation of triazole nitrogen atoms by groups which does not promote the delocalization of the electrons makes the molecule less stable, i.e. less inhibition. So, the decreased the inhibitory efficiency of our molecule comparing it with the other molecules (Table 3) due to no conjugated system between traizole and 8-hydroxyquinoline moieties. 3.2.3. Electrochemical impedance spectroscopic(EIS) studies In order to obtain more dynamic and electrode interface information, the EIS experiments were tested at 298 K for CS within hydrochloric acid(HCl) in the nonexistence and the existence of diverse concentrations of MHTC. Fig. 5. present the Niquist and Bode plots for MHTC. In Fig. 5, it can be found that the impedance spectrum in a blank solution without a MHTC shows a semicircle arc in high frequency area and a straight line in low frequency area. As shown in Fig. 5, it can be seen that the radius of the capacitive arc becomes visibly larger in the high frequency area with the increase of the concentrations of MHTC. The capacitive arc of the high frequency area corresponds to the charge transfer resistance and electric double layer capacitance [33]. Therefore, the increase of capacitive arc radius indicates MHTC molecules adsorption on CS surface increases the polarization resistance and effectively suppresses the corrosion of CS. The equivalent circuit diagram (Fig. 6) is used to fit the impedance spectrum in the blank solution, and the impedance spectrum with MHTC in the test solution. The electrochemical parameters after fitting are shown in Table 3. These important parameters include Cdl , Rs , Rp and CPE which represent double layer capacitance, solution resistance, polarization resistance, and constant phase angle element, respectively. Using Eq. (13), the impedance of the CPE can be obtained as following [34–36]:

ZCPE =

1 Y0 ( jω )n

(13)

8

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Table 3 Percentage inhibition efficiency for different triazole derivatives in 1 M HCl (the concentration used is 10−3 M). Triazole derivative

Highest inhibition efficiency (%)∗

Metal exposed

Reference

99.3

Mild steel

[30]

98.8

Mild steel

[30]

97.5

Mild steel

[31]

94.0

Mild steel

[31]

94.0

Mild steel

[31]

90.0

Ordinary steel

(A)1-[2-(4-Nitro-phenyl)−5-[1,2,4]triazol-1-ylmethyl-[1,2,3]oxadiazol-3yl]-enthanone (NTOE)

(B)1-(4-Methoxy-phenyl)−2-(5-[1,3,4]triazol-1-ylmethyl-4H-[1,2,4]triazol-3ylsulfanyl)-ethanone (MTTE)

4–chloro-acetophenone-O-1 -(1,3,4 -triazolyl)-metheneoxime (CATM)

4-methoxyl-acetophenone-O-1-(1,3,4-triazolyl)-metheneoxime (MATM)

4-fluoro-acetophenone-O-1-(1,3,4-triazolyl)-metheneoxime (FATM)) [32]

1-((4–bromo-2-(2,4-dichlorophenyl)tetrahydrofuran-2-yl)methyl) −1H-1,2,4-triazole (Bromuconazole) (continued on next page)

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

9

Table 3 (continued) Triazole derivative

Highest inhibition efficiency (%)∗

Metal exposed

Reference

92.0

Ordinary steel

[32]

91.0

Carbon steel

This work

2-(2,4-dichlorophenyl)−1-(1H-1,2,4-triazol-1-yl)hexan-2-ol((Hexaconazole)

4-amino-1-((8-hydroxyquinolin-5-yl)methyl)−1,2,3triazole-5-carboxylate (MHTC) ∗

The inhibition efficiency values were determined using polarization curves at 298 K after 30 min of immersion. Table 4 Electrochemical parameters and performance ability of steel corrosion in 1 M HCl with no and through addition of different concentrations of MHTC at 298 K. Medium

Conc. [M]

Rs ( cm²)

Q (μ −1 sn cm−2 )

ndl

Rp ( cm²)

χ × 10−3

Cdl (μf/cm²)

ηEIS (%)

HCl MHTC

1 10−3 10−4 10−5 10−6

1.22 1.33 1.80 1.53 1.32

315.1 88.05 117.3 126.3 255.9

0.820 0.831 0.815 0.809 0.811

34.85 278.2 241.1 218.3 68.34

1.11 5.19 2.19 4.46 5.29

114.1 45.76 58.32 65.98 93.14

— 87 85 84 49

where Y0 represents the CPE constant, j represents the imaginary root and ω represents the angular frequency. n represents the diffusion effect index. The values Cdl are acquired via Eq. (14) as following [37]:



Cdl = Q × R1p−n

1 / n

(14)

From Table 3, we can clearly find that the Rp values significantly increase with increasing concentrations of MHTC, which reveals that MHTC molecules adsorption on the surface of CS effectively suppress the charge transfer process. In addition, the Cdl values manifest a decreasing tendency with increasing MHTC concentrations. The values of Cdl according to the Helmholtz model as follows [37,38]:

Cdl =

ε0 × ε d

×S

(15)

ɛ0 represent

where ɛand double layer local dielectric constant and air dielectric constant, respectively. The S is the copper electrode area, which exposed to the test solution. The d represents thickness of double layer. In Table 4, the Cd1 values and the concentrations of corrosion inhibitor show the opposite trend. This event is due to the substitute of water (H2 O) molecules on the CS surface by MHTC molecules. On the one hand, MHTC molecules have lower dielectric constants than water [37]. On the other hand, MHTC molecules are also larger than H2 O molecules. The values of double electric layer thickness increases and the local dielectric constant value of the electric double layer decreases when the water(H2 O) molecules on the CS surface are replaced by MHTC. Therefore, the increase of inhibitor(MHTC) concentrations will lead to the decrease of Cdl values. 3.3. ANOVA statistical tests Analysis of variance (ANOVA) shows that the difference existing between inhibition efficiencies obtained by PDP, EIS and WL for 10−3 —10−6 M is the statistically significant level of α = 0.05 and the ANOVA results are given in Table 5. From this

10

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242 Table 5 ANOVA for inhibition efficiency of MHTC in 1 M HCl (at 95% confidence level). Source of Variation

Sum of squares

Degree of freedom

F

P-value

F crit

Between groups Within groups Total

75.5 2251.5 2327

2 9 11

0.150

0.862

4.256

Table 6 Temperature effect on electrochemical quantities and proportion inhibition of CS in 1 M HCl alone and in the existence of 1 mM MHTC. Medium

Temp (K)

-Ecorr (mV/SCE)

icor (μA cm−2 )

-β c (mV dec−1 )

ηTafel (%)

Blank

298 308 318 328 298 308 318 328

454.4 451.1 454.8 464.3 439.1 458.7 446.5 476.1

590.8 666.8 982.4 2368.5 54.6 113.3 178.5 399.8

101.1 104.9 111.1 101.0 101.1 119.0 139.8 157.5

— — — — 91 83 82 83

MHTC

Fig. 7. Influence of the temperature on the i = f (E)curves of steel in 1 M HCl.

table, it is clear that the p-values (0.862) are > 0.05. From p-values, it is concluded that there is a significant difference in inhibition efficiencies obtained by increasing concentration. 3.3.1. Influence of temperature The deterioration characteristics of CS immersed in 1 M HCl solution not including and including 1 mM of MHTC were inspected in the four temperatures(298, 308, 318 and 328 K). Results of PDP measurements performed to assess the temperature factor on the dissolution characteristic of CS in the corrosive medium with and without TMHTC are presented in Table 6 and Figs 7,8. The influence of temperature on the inhibition process is a complex phenomenon because of the numerous competing factors that affect the kinetics of inhibitor adsorption/desorption, electrochemical reactions, and diffusion of the reactive species. For example, rise in temperature may alter the chemical behavior of inhibitor molecules such that the electron densities at the centers of the molecules increase and improves molecular adsorption. The data indicate that the icor increased with increases in the four temperatures(298, 308, 318 and 328 K)for the blank and the inhibited solutions. Indeed, the Arrhenius law was utilized to calculate the value of the activation energy in the nonexistence and in the existence of the 1 mM inhibitor [39]:

 E  a

icor = Aexp −

RT

(16)

R the molar gas constant, Ea is the activation energy, and T the absolute temperature and Ais the Arrhenius preexponential coefficient. A plot of Ln icor as a function of the inverse of the temperature in Kelvin (Fig. 9) gives -Ea /R as the slope from which the value of Ea was extracted. The calculated Ea values are listed in Table 6.

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

11

Fig. 8. Influence of the temperature on the i = f (E)curves of steel in 1 M medium in the presence of 1 mM of MHTC.

Fig. 9. Arrhenius plots of CS in 1 M HCl through and with no 1 mM of MHTC.

To obtain the enthalpy Ha and entropy Sa an alternative formulation of the Arrhenius equation is used [40].



icor =

RT S a exp Nh R





exp −

Ha RT



(17)

where N is the Avogadro’s number, h is the Planck’s constant. Fig. 10 shows a plot of Ln (icor /T) vs. 1/T for the blank and inhibited solution. The (Ha + Sa ) values were predictable and given in Table 7. Examination of this table reveals that the value of Ha for the dissolution reaction of CS in 1 M HCl is positive, reflecting the endothermic nature of the process dissolution of CS. In the existence of MHTC, the value of Ha is higher (33.9 kJ mol−1 ) than that in its absence (49.5 kJ mol−1 ). This suggests that the dissolution of CS is slow in the presence of the triazole derivative [41]. The positive sign of Ha reflects the endothermic nature of the dissolution process of CS metal. On the other hand, the value of the activation energy (Ea ) is higher than its enthalpy analogue (Ha ) signifying that the corrosion process would involve a gaseous reaction, which is the proton reduction reaction (H+ ) [42]. Concerning the activation entropy (Sa ), the analysis of the results obtained in Table 7 shows that Sa increases in the presence of MHTC. The increase in Sa

12

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 10. Transition state plots of Ln (icor /T) versus 10 0 0/T in the nonexistence and the existence of 10−3 M MHTC. Table 7 Activation parameters, Ea , Ha and Sa of the dissolution of CS in1 M HCl in the nonexistence and the existence of 1 mM of MHTC. Medium

Ea (kJ mol−1 )

Ha (kJ mol−1 )

Sa (J mol−1 K−1 )

Blank MHTC

36.5 52.2

33.9 49.5

−79.0 −45.4

Table 8 Thermodynamic adsorption parameters of MHTC on CS surface at 298 K. Isotherms

Linear forms

Curves

Langmuir

Cinh

Cinh

Temkin

θ = 1f LnKads + 1f LnCinh

θ = f (Ln(Cinh ))

FloryHuggins

Ln ( Cθ ) = Ln (Kads ) + a × Ln (1 − θ )

Ln( Cθ ) = f (Ln(1 − θ ))

Frumkin

Ln(Cinh × 1−θ θ ) = −LnKads − 2 × f × θ

Ln(Cinh × 1−θ θ ) = f (θ )

El-Awady

Lnθ (1 − θ )−1 = LnK + yLnC

Lnθ (1 − θ )−1 = f (LnC )

θ =

1 Kads

+ Cinh

inh

θ = f (Cinh )

inh

Parameters

Values

R2 Kads (L mol−1 ) Slope G◦ads (kJ/mol) R2 Kads (L mol−1 ) Slope (1/f) F R2 Kads (L mol−1 ) Slope (a) R2 Kads (L mol−1 ) Slope (−2f) R2 Kads (L mol−1 ) Slope (y)

1 623,092 1.09 −43.00 0.83683 3.79.1012 0.043 23.025 0.63512 529,664 1.948 0.9352 1.92 109 −12.308 0.8787 77.62 0.2633

after addition of MHTC, is commonly interpreted as an augment of the disorder for the duration of the transformation of the reagents into activated complexes. This behavior can also be explained by the process of replacing the water(H2 O) molecules during the adsorption of MHTC molecules on the surface of the steel and thus the increase of Sa is attributed to the increase of the entropy of the solvent (water) [43]. 3.4. Isotherm of adsorption In order to all-sided comprehend the adsorption behavior of MHTCmolecules on CSsurface in HCl, the potentiodynamic polarization (θ = η Tafel / 100) data were used to fit using Langmuir, Temkin, Flory-Huggins, Frumkin, and El-Awady isothermal equations. These fitting data indicate that the correlation coefficient (R2 ) values of the Langmuir adsorption isotherm equation are significantly larger than other adsorption isotherms and is equal to 1 (Fig. 11 andTable 8). It was found that the most suitable adsorption is the Langmuir adsorption isotherm.

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 11. Langmuir (a), Flory-Huggins (b), Temkin (c), Frumkin (d), El-Awady (e) adsorption isotherm of CS in 1 M HCl containing MHTC at 298 K.

13

14

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 12. SEM images of carbon steel: (a) steel alone, (b) in the nonexistence and (c) in the existence of 1 mM of MHTC.

The Langmuir adsorption isotherm equation can be expressed as following [44,45]:

Cinh

θ

=

1 + Cinh Kads

(18)

Where θ can be obtained via inhibition efficiency and it represents the surface coverage. Kads represents the equilibrium constant of adsorption. C represents MHTC concentrations. According to Eq. (18), C is the horizontal axis and C/θ is the vertical axis. Then 1/K is the intercept as shown in Fig. 11. In order to study the adsorption type of MHTC on the CS surface. G◦ads is defined as following [46–48]:

G◦ads = −RT Ln(55.5Kads )

(19)

where T and R retain the meanings as defined in Eq. (13); 55.5 is the concentration of water in mol/dm3 . The value of G◦ads calculated for MHTC adsorption on CS is estimated to be −43 kJ/mol which is in the range of values reported in the corrosion literature for chemical adsorption [49]. Hence, in this work it is plausible to say that MHTC adsorbs on CS surface by chemisorption mechanism involving sharing of electrons or transfer of electrons from MHTC molecules to the partially filled d-orbital of Fe atoms at the surface. The bigger value of Kads and the negative value of G◦ads indicate a strong interaction between MHTC molecules and CS surface.

3.5. Scanning electron microscopic (SEM) analysis The SEM pictures in Fig. 12 depict the samples before and after exposing to 1 M HCl in the nonexistence and existence of 1 mM of MHTC. The smooth morphology of steel surface after abrasion is quite obvious in Fig. 12(a). After 6 h of immersion in the inhibitor-free solution, a severely damaged and coarse morphology due to the rapid and aggressive corrosion attack of steel in studied environment is observed (Fig. 12(b)). The corrosive attack was minimized in the system containing 1 mM MHTC (Fig. 12c); the steel surface remains relatively un-attacked probably due to the formation of adsorbed MHTC film which protected the metal from corrosion.

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

15

Fig. 13. UV−visible spectra of 1 M HCl solution containing 1 mM of MHTC before and after After 6 h of CS immersion.

Fig. 14. Optimized molecular structures, HOMO and LUMO of MHTC.

3.6. UV-visible spectroscopy Spectrometry analyses were carried to gain further insights into steel-inhibitor interactions. The UV- Visible adsorptions spectral of acidic solutions of triazole derivatives based on MHTC before and after CS immersion are shown in Fig. 13. The electron absorption spectrum of MHTC (Fig. 13) before immersion has an absorption bung in the range 200–260 nm attributed to the π -π ∗ transitions relating the entire electronic system of the aromatic, followed by another 341 nm band is shifted to a higher value (332 nm) in the case of CS + MHTC through an increase in absorbance of 0.325 to 1.148, suggesting that interaction between the MHTC molecule and the Fe2+ ions in the solution, the modify in the place of the absorption maximum (λmax ) and / or the variation of the absorbance value signify the formation of a complex between the two species in solution as indicated in the literature [50,51].

3.7. Quantum chemical calculations 3.7.1. Optimization, distribution of electron density and molecular electrostatic potential The hypothesis of DFT is a vital technique for concentrate the reactivity of the atoms of an inhibitor [44]. In this sense we will take after an all around organized methodology utilizing a few extremely exact descriptors. The principal factor is the densities HOMO and LUMO; this is appearing in Fig. 14 with the electrostatic contour and the optimized molecule of MHTC. As indicated by this figure it is certain that the electron density HOMO is situated on the whole sub-atomic surface of triazole cycle, while the LUMO density is circulated around the aromatic cycles of quinoline. The molecular electrostatic potential (MEP) distribution and the contour of the electrostatic potential surface (EPS) are presented in Fig. 15. From this figure the maximum and minimum in the MEP values mapped of MHTC are presented by colors Such as red represents negative regions (nucleophilic attack), then blue detects positive regions (electrophilic attack). The estimations of the EHOMO and ELUMO energies are figured utilizing DFT by Gaussian 09 program. From these qualities, we can extricate the distinctive structural quantum parameters of MHTC, for example, Egap, dipole moment (μ), electronegativity (χ ), overall hardness (η), chemical softness (σ ), the number of electrons transferred from the inhibitor to the surface of the metal (N110 ), and these parameters are listed inTable 9.

16

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Fig. 15. Contour of the electrostatic potential surface and electrostatic potential map around the molecule of MHTC. (For interpretation of the references to color in this figure, the reader is referred to the web version of this article.) Table 9 Calculated quantum parameters of MHTC inhibitor. Parameters

ELUMO (eV)

EHOMO (eV)

E (eV)

μ (D)

η (eV)

σ (eV-1)

IE (eV)

χ (eV)

N110

TE (u a)

MHTC

−1.613

−5.766

4.153

4.940

2.202

0.481

5.766

3.816

0.228

−1040.759

Table 10 Values of the Fukui function consideringNatural Population analysis (NPA) of MHTC molecule calculated at the B3LYP / 6–31 G (d,p). Atoms

f k+

f k−

C1 C2 N3 C4 C5 C6 C7 C8 C9 C10 O11 C12 N13 N14 N15 C16 C17 N18 C29 O30 O31 C32

0.022 0.048 0.092 0.012 −0.009 0.104 0.042 0.034 0.028 0.040 0.055 −0.002 −0.015 −0.006 0.026 0.015 −0.007 0.015 0.008 0.026 −0.013 −0.010

0.021 0.021 0.021 0.019 0.006 0.008 0.031 0.038 0.015 0.025 0.078 −0.005 0.010 0.039 0.062 0.013 0.049 0.083 0.018 0.051 0.002 0.054

According to the frontier molecular orbital (FMO) theory, the formation of a transition state is due to an interaction between the frontier orbital’s HOMO and LUMO [52]. A high value of (EHOMO = - 5.766 eV) show an augment for the electron(e− ) donor and which implies a better inhibitory activity with a growth of the adsorption of MHTC on a metal surface similarly, a lower value of (ELUMO =−1.613 eV) show the capacity to accept an electron(e− ) from the molecule. The adsorption capacity of MHTC to the metal surface rise with rising of EHOMO and decrement of ELUMO . Likewise, a lesser value of (E = 4.153 eV) is linked with a high chemical reactivity and consequently important performance capacity similarly, the values of electronegativity and hardness were very small as 3.689 eV and 2.076 eV respectively signifying that the MHTC has a very important electron donating capacity. This finding was further supported by a elevated value of global softness (σ ) and electron transfer (N110 ). Otherwise the strong tendency of MHTC to substitute water(H2 O) molecules from the metallic surface due to a elevated value of dipole moment (4.940 D) as compared to the dipole moment of water(H2 O) (1.85 D).

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

17

Table 11 Different values of the quantum parameters of the protonated and no-protonated MHTC molecule. Inhibitors

ELUMO (eV)

EHOMO (eV)

࢞Egap (eV)

࢞N110

ET (u a)

MHTC MHTC (N18) H+ MHTC (N3) H+ MHTC (N14) H+

−1.614 −5.812 −6.512 −6.120

−6.019 −8.332 8.310 −9.173

4.405 2.520 1.798 3.053

2.049 −0.893 −1.441 −0.925

−1040.759 −1041.115 −1041.141 −1041.129

Table 12 Outputs and descriptors for the lowest adsorption configurations for Fe (110)/ MHTC /500 H2 O/5 (Cl− , H3 O+ ) systems. (All values in Kcal/mol). Structures Fe Fe Fe Fe Fe Fe Fe Fe Fe

(1 (1 (1 (1 (1 (1 (1 (1 (1

1 1 1 1 1 1 1 1 1

0) 0) 0) 0) 0) 0) 0) 0) 0)

− − − − − − − − −

1 2 3 4 5 6 7 8 9

Total energy

Adsorption energy

Rigid adsorption energy

Deformation energy

MHTC: dEad /dNi

H3 O+ : dEad /dNi

water: dEad /dNi

−9666.20 −9648.99 −9636.53 −9636.02 −9627.84 −9626.51 −9624.81 −9622.81 −9620.66

−9814.90 −9797.69 −9785.23 −9784.72 −9776.548 −9775.214 −9773.51 −9771.51 −9769.36

−10 0 06.40 −9990.21 −9974.21 −9976.77 −9962.33 −9955.86 −9962.39 −9965.93 −9953.64

191.49 192.52 188.98 192.05 185.79 180.64 188.88 194.41 184.27

−445.94 −444.49 −445.17 −442.79 −458.50 −460.48 −443.37 −444.58 −460.73

−159.29 −160.71 −167.28 −167.27 −171.03 −160.97 −167.26 −167.05 −166.51

−7.22 −6.29 −5.93 −6.54 −5.83 −6.66 −5.64 −6.89 −5.73

Cl− : dEad /dNi −3.85 −3.97 −1.72 −4.06 −3.47 −2.56 −2.00 −3.72 −3.02

Fig. 16. Top and side views of the best adsorption of the MHTC on the Fe (110) surface in acidic solution.

The values of the Fukui functions for a nucleophilic and electrophilic attack are given for the MHTC in Table 10. Analysis of the results obtained in Table 8 shows that the MHTC molecule has the highest values of f k+ , these last ones are localized on the atoms. N3, C6 and O11. These cities can be accepted electrons. While, the highest values of f k− shows that atoms O11, C17,N18,O30 and C32 are able to give electrons which leads to the increase of MHTC compound adsorption capacity on the metal surface. We conclude that the results obtained by the local reactivity (IF) study confirm the results obtained by the total electron density and the Mulliken atomic charge.

18

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

Scheme 1. synthesis of triazole derivative based on 8-hydroxyquinoline (MHTC).

3.7.2. Influence of protonation on the main structural parameters of the neutral MHTC inhibitory molecule In HCl medium, the MHTC molecule is very likely to be protonated at the sites, namely the nitrogen (N) and oxygen (O) atoms. But in this study we have chosen the N3, N14 and N18 atoms as the most favorable sites for protonation. To determine the most available site for protonation for these nitrogen atoms, we have made several attempts at protonation, each time we protonate an atom. The calculation results are shown in Table 11. The analysis of the results in Table 11 shows that after the protonation of inhibitory molecule (MHTC) in the different sites of the nitrogen atoms, the EHOMO values of protonated MHTC are shifted to more negative values compared to that of the value of d EHOMO of neutral MHTC. This suggests that the protonated forms of an inhibitory molecule have a capacity to accept higher electrons than the neutral form [53]. While, the number of transferred electrons (N110 ) of the protonated studied molecule (MHTC) was decreased and negative compared to the non-protonated form, which means that electron donation from MHTC protonated on the surface metal is no longer possible [53]. These results probably indicate that the interaction between this protonated molecule (MHTC) and the metal surface of iron is electrostatic in nature. The minimum values of N110 and ET of MHTC (N3) H+ indicate that this protonated form is more favorable for protonation than the other protonated forms of MHTC (N14) H+ and MHTC (N18) H+ . 3.8. MD simulations The results of the calculations are grouped in the Table 12. Fig. 16 depicts the best adsorption configuration of the inhibitor tested on the Fe (110). This figure shows that the molecule MHTC adsorbs on the metal surface in a parallel manner. The MHTC has five nitrogens. These are more reactive with the iron atoms. There are other active sites like oxygen atoms and the carbon atoms belong to the aromatic rings which are favored the adsorption phenomenon. Table 1 brings together the different values of energies calculated by the Adsorption Locator for Fe (110)/ MHTC /500 H2 O/5 (Cl− , H3 O+ ) systems. In Table 1, calculated total energy, adsorption energy, rigid adsorption energy and deformation energy values are tabulated. The calculated minimum interaction energy value for MHTC inhibitor is −9814.904 kcal/mol. This value corresponds to the best configuration of the molecule (MHTC). 4. Conclusion The conclusions drawn on the basis of the results obtained in this work are as follow: • MHTC is effective in retarding corrosion of CS in the temperature range of 298–328 K. • The adsorption of MHTC can be explained with Langmuir adsorption isotherm and chemisorption is the dominant mode of interaction. • MHTC behaves as a typical mixed type corrosion inhibitor • Scanning electronic microscopy (SEM) confirms the performance of MHTC against corrosion. • The UV-visible studies clearly reveal the formation of Fe-MHTC complex which may be also responsible for the observed inhibition. • The theoretical calculations show that the MHTC having a high donor power and a low energy gap has a good performance ability. • Analysis of atomic sites show that the MHTC studied contains highly reactive centers that are the origin of their inhibitory efficiencies. • Molecular dynamics simulations show the great interaction between MHTC and the metal surface. Supplementary material Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cdc.2019.100242. References [1] M. Lagrenee, B. Mernari, M. Bouanis, M. Traisnel, F. Bentiss, Study of the mechanism and inhibiting 3,5-bis(4-methylthiophenyl)-4H-1,2,4-triazole on mild steel corrosion in acidic media, Corros. Sci. 44 (3) (2002) 573–588.

efficiency

of

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

19

[2] V.V. Torres, R.S. Amado, C.F. De Sá, T.L. Fernandez, C.A. da Silva Riehl, A.G. Torres, E. D’Elia, Inhibitory action of aqueous coffee ground extracts on the corrosion of carbon steel in HCl solution, Corros. Sci. 53 (7) (2011) 2385–2392. [3] A. Dehghani, G. Bahlakeh, B. Ramezanzadeh, A detailed electrochemical/theoretical exploration of the aqueous Chinese gooseberry fruit shell extract as a green and cheap corrosion inhibitor for mild steel in acidic solution, J. Mol. Liq. 282 (2019) 366–384. [4] A. Zarrouk, H. Zarrok, Y. Ramli, M. Bouachrine, B. Hammouti, A. Sahibed-dine, F. Bentiss, Inhibitive properties, adsorption and theoretical study of 3,7-dimethyl-1-(prop-2-yn-1-yl)quinoxalin-2(1H)-one as efficient corrosion inhibitor for carbon steel in hydrochloric acid solution, J. Mol. Liq. 222 (2016) 239–252. [5] A. El yaktini, A. Lachiri, M. El Faydy, F. Benhiba, H. Zarrok, M. El Azzouzi, M. Zertoubi, M. Azzi, B. Lakhrissi, A. Zarrouk, Inhibitor effect of new azomethine derivative containing an 8-hydroxyquinoline moiety on corrosion behavior of mild carbon steel in acidic media, Int. J. Corros. Scale Inhib. 7 (4) (2018) 609–632. [6] H.J. Habeeb, H.M. Luaibi, R.M. Dakhil, A.A.H. Kadhum, A.A. Al-Amiery, T.S. Gaaz, Development of new corrosion inhibitor tested on mild steel supported by electrochemical stud, Results Phys. 8 (2018) 1260–1267. [7] F.K. Khaled, The inhibition of benzimidazole derivatives on corrosion of iron in 1 M HCl solutions, Electrochim. Acta 48 (2003) 2493–2503. [8] A. Popova, M. Christov, S. Raicheva, E. Sokolova, Adsorption and inhibitive properties of benzimidazole derivatives in acid mild steel corrosion, Corros. Sci. 46 (2004) 1333–1350. [9] M. El Faydy, M. Galai, A. El Assyry, A. Tazouti, R. Touir, B. Lakhrissi, A. Zarrouk, Experimental investigation on the corrosion inhibition of carbon steel by 5-(chloromethyl)-8-quinolinol hydrochloride in hydrochloric acid solution, J. Mol. Liq. 219 (2016) 396–404. [10] H. Lgaz, R. Salghi, M. Larouj, M. Elfaydy, S. Jodeh, Z. Rouifi, B. Lakhrissi, H. Oudda, Experimental, Theoretical and Monte Carlo simulation of quinoline derivative as effective corrosion inhibitor for mild steel in 1 M HCl, J. Mater. Environ. Sci. 7 (2016) 4471–4488. [11] M.H.O. Ahmed, A.A. Al-Amiery, Y.K. Al-Majedy, A.A.H. Kadhum, A.B. Mohamad, T.S. Gaaz, Synthesis and characterization of a novel organic corrosion inhibitor for mild steel in 1 M hydrochloric acid, Results Phys. 8 (2018) 728–733. [12] B. Himmi, B. Messnaoui, S. Kitane, A. Eddaif, A. Alaoui, A. Bouklouz, M. Soufiaoui, Study of Zn (II) extraction by 5-azidomethyl-8-hydroxyquinoline: Experiment and modelling, Hydrometallurgy 93 (1) (2008) 39–44. [13] Z. Rouifi, M. El Faydy, H. About, F. Benhiba, H. Ramsis, M. Boudalia, H. Zarrok, R. Touir, M. El M’Rabet, H. Oudda, A. Guenbour, B. Lakhrissi, Electrochemical and theoretical studies of adsorption and corrosion inhibition of ethyl 5-amino-1-((8-hydroxyquinolin-5-yl) methyl)-1H-1,2,3-triazole-4-carboxylate on carbon steel in acidic solution, J. Mater. Environ. Sci. 9 (2) (2018) 453–465. [14] K. Boumhara, F. Bentiss, M. Tabyaoui, J. Costa, J.M. Desjobert, A. Bellaouchou, A. Guenbour, B. Hammouti, S.S. Al-Deyab, Use of Artemisia Mesatlantica Essential Oil as Green Corrosion Inhibitor for Mild Steel in 1 M Hydrochloric Acid Solution, Int. J. Electrochem. Sci. 9 (2014) 1187–1206. [15] H. Zarrok, R. Salghi, A. Zarrouk, B. Hammouti, H. Oudda, Lh. Bazzi, L. Bammou, S.S. Al-Deyab, Investigation of the Inhibition Effect of N-1-Naphthylethylenediamine Dihydrochloride Monomethanolate on the C38 Steel Corrosion in 0.5M H2SO4, Der Pharm. Chem. 4 (1) (2012) 407–416. [16] K. Abderrahim, I. Selatnia, A. Sid, P. Mosset, 1,2-bis(4-chlorobenzylidene)Azine as new and effective corrosion inhibitor for copper in 0.1 N HCl: A combined experimental and theoretical approach, Chem. Phys. Lett. 707 (2018) 117–128. [17] M. Rbaa, M. Galai, M.EL Faydy, Y. El Kacimi, M. Ebn Touhami, A. Zarrouk, B. Lakhriss, Synthesis, inhibition effects and thermodynamic studies of novel substituted quinoline son the corrosion of mild steel in 1 M HCl solution, J. Mater. Environ. Sci. 8 (10) (2017) 3529–3549. [18] T. Laabaissi, H. Lgaz, H. Oudda, F. Benhiba, H. Zarrok, A. Zarrouk, A. El Midaoui, B. Lakhrissi, R. Touir, Comparative study of corrosion inhibition effect of benzodiazepine derivative on the carbon steel surface in 2.0 M H3PO4 and 1.0 M HCl mediums: Electrochemical, theoretical and Monte Carlo simulations studies, J. Mater. Environ. Sci. 8 (2017) 1054–1067. [19] C. Verma, I.B. Obot, I. Bahadur, E.-S.M. Sherif, E.E. Ebenso, Choline based ionic liquids as sustainable corrosion inhibitors on mild steel surface in acidic medium: Gravimetric, electrochemical, surface morphology, DFT and Monte Carlo simulation studies, App. Surf. Sci. 457 (2018) 134–149. [20] Y. Feng, S. Chen, W. Guo, Y. Zhang, G. Liu, Inhibition of iron corrosion by 5,10,15,20-tetraphenylporphyrin and 5,10,15,20-tetra-(4-chlorophenyl)porphyrin adlayers in 0.5 M H2SO4 solutions, J. Electroanal. Chem. 602 (2007) 115–122. [21] M. El Faydy, R. Touir, M. Ebn Touhami, A. Zarrouk, C. Jama, B. Lakhrissi, L.O. Olasunkanmi, E.E. Ebenso, F. Bentiss, Corrosion inhibition performance of newly synthesized 5-alkoxymethyl-8-hydroxyquinoline derivatives for carbon steel in 1 M HCl solution: experimental, DFT and Monte Carlo simulation studies, Phys. Chem. Chem. Phys. 20 (2018) 20167–20187. [22] R.G. Pearson, Absolute electronegativity and hardness: application to inorganic chemistry, Inorg. Chem. 27 (1988) 734–740. [23] J.J. Fu, S.N. Li, Y. Wang, X.D. Liu, Computational and electrochemical studies on the corrosion inhibition of mild steel by L-cysteine and its derivatives, J. Mater. Sci. 46 (10) (2011) 3550–3559. [24] M. Bouanis, M. Tourabi, A. Nyassi, A. Zarrouk, C. Jama, F. Bentiss, Corrosion inhibition performance of 2,5-bis(4-dimethylaminophenyl)-1,3,4-oxadiazole for carbon steel in HCl solution: Gravimetric, electrochemical and XPS studies, Appl. Surf. Sci. 389 (2016) 952–966. [25] F. Bentiss, M. Bouanis, B. Mernari, M. Traisnel, H. Vezin, M. Lagrenée, Understanding the adsorption of 4H-1,2,4-triazole derivatives on mild steel surface in molar hydrochloric acid, Appl. Surf. Sci. 253 (2007) 3696–3704. [26] Q. Ma, S. Qi, X. He, Y. Tang, G. Lu, 1,2,3-Triazole derivatives as corrosion inhibitors for mild steel in acidic medium: Experimental and computational chemistry studies, Corros. Sci. 129 (2017) 91–101. [27] C.M. Fernandes, L.X. Alvarez, N. Escarpini dos Santos, A.C.M. Barrios, E.A. Ponzio, Green synthesis of 1-benzyl-4-phenyl-1H-1,2,3-triazole, its application as corrosion inhibitor for mild steel in acidic medium and new approach of classical electrochemical analyses, Corros. Sci. 149 (2019) 185–194. [28] H.-L. Wang, R.-B. Liu, J. Xin, Inhibiting effects of some mercapto-triazole derivatives on the corrosion of mild steel in 1.0 M HCl medium, Corros. Sci. 46 (2004) 2455–2466. [29] Q. Deng, N.-N. Ding, X.-L. Wei, L. Cai, X.-P. He, Y.-T. Long, G.-R. Chen, K. Chen, Identification of diverse 1,2,3-triazole-connected benzyl glycosideserine/threonine conjugates as potent corrosion inhibitors for mild steel in HCl, Corros. Sci. 64 (2012) 64–73. [30] S. Zhang, Z. Tao, W. Li, B. Hou, The effect of some triazole derivatives as inhibitors for the corrosion of mild steel in 1 M hydrochloric acid, Appl. Surf. Sci. 255 (2009) 6757–6763. [31] W. Li, Q. He, C. Pei, B. Hou, Experimental and theoretical investigation of the adsorption behaviour of new triazole derivatives as inhibitors for mild steel corrosion in acid media, Electrochim. Acta 52 (2007) 6386–6394. [32] Y. Elkhotfi, I. Forsal, E.M. Rakib, B. Mernari, Evaluation of the inhibitor effect of new class triazole derivatives on the corrosion of ordinary steel in hydrochloric acid solution, Der Pharma Chem 8 (15) (2016) 160–170. [33] X. Zheng, S. Zhang, W. Li, M. Gong, L. Yin, Experimental and theoretical studies of two imidazolium-based ionic liquids as inhibitors for mild steel in sulfuric acid solution, Corros. Sci. 95 (2015) 168–179. [34] M.A. Hegazy, A.A. Nazeer, K. Shalabi, Electrochemical studies on the inhibition behavior of copper corrosion in pickling acid using quaternary ammonium salts, J. Mol. Liq. 209 (2015) 419–427. [35] Y. Qiang, S. Zhang, S. Xu, L. Yin, The effect of 5-nitroindazole as an inhibitor for the corrosion of copper in a 3.0% NaCl solution, RSC Adv. 5 (2015) 63866–63873. [36] L.L. Liao, S. Mo, H.Q. Luo, N.B. Li, Corrosion protection for mild steel by extract from the waste of lychee fruit in HCl solution: Experimental and theoretical studies, J. Colloid Interface Sci. 520 (2018) 41–49. [37] H. Zarrok, A. Zarrouk, R. Salghi, M. Ebn Touhami, H. Oudda, B. Hammouti, R. Touir, F. Bentiss, S.S. Al-Deyab, Corrosion Inhibition of C38 Steel in Acidic Medium Using N-1 Naphthylethylenediamine Dihydrochloride Monomethanolate, Int. J. Electrochem. Sci. 8 (2013) 6014–6032.

20

Z. Rouifi, F. Benhiba and M.E. Faydy et al. / Chemical Data Collections 22 (2019) 100242

[38] A. Ehsani, M.G. Mahjani, M. Hosseini, R. Safari, R. Moshrefi, H.M. Shiri, Evaluation of Thymus vulgaris plant extract as an eco-friendly corrosion inhibitor for stainless steel 304 in acidic solution by means of electrochemical impedance spectroscopy, electrochemical noise analysis and density functional theory, J. Colloid Interface Sci. 490 (2017) 444–451. [39] M. Tourabi, A. Sahibed-dine, A. Zarrouk, I.B. Obot, B. Hammouti, F. Bentiss, A. Nahlé, 3,5-Diaryl-4-amino-1,2,4-triazole derivatives as effective corrosion inhibitors for mild steel in hydrochloric acid solution: Correlation between anti-corrosion activity and chemical structure, Prot. Met. Phys. Chem. 53 (3) (2017) 548–559. [40] F. Zhang, Y. Tang, Z. Cao, W. Jing, Z. Wu, Y. Chen, Performance and theoretical study on corrosion inhibition of 2-(4-pyridyl)-benzimidazole for mild steel in hydrochloric acid, Corros. Sci. 61 (2012) 1–9. [41] N.M. Guan, L. Xueming, L. Fei, Synergistic inhibition between o-phenanthroline and chloride ion on cold rolled steel corrosion in phosphoric acid, Mater. Chem. Phys. 86 (2004) 59–68. [42] O. Fergachi, F. Benhiba, M. Rbaa, M. Ouakki, M. Galai, R. Touir, B. Lakhrissi, H. Oudda, M.E. Touhami, Touhami, Corrosion Inhibition of Ordinary Steel in 5.0 M HCl Medium by Benzimidazole Derivatives: Electrochemical, UVVisible Spectrometry, and DFT Calculations, J. Bio Tribo-Corros. 5 (2019), doi:10.1007/s40735- 018- 0215- 3. [43] B. Ateya, B.E. El-Anadouli, F.M. El-Nizamy, The adsorption of thiourea on mild steel, Corros. Sci. 24 (1984) 509–515. [44] L.L. Liao, S. Mo, J.L. Lei, H.Q. Luo, N.B. Li, Application of a cosmetic additive as an eco-friendly inhibitor for mild steel corrosion in HCl solution, J. Colloid Interface Sci. 474 (2016) 68–77. [45] X. Ma, X. Jiang, S. Xia, M. Shan, X. Li, L. Yu, Q. Tang, New corrosion inhibitor acrylamide methyl ether for mild steel in 1 M HCl, Appl. Surf. Sci. 371 (2016) 248–257. [46] H. Tayebi, H. Bourazmi, B. Himmi, A. El Assyry, Y. Ramli, A. Zarrouk, A. Geunbour, B. Hammouti, E.E. Ebenso, An electrochemical and theoretical evaluation of new quinoline derivative as a corrosion inhibitor for carbon steel in HCl solutions, Der Pharm. Lett. 6 (6) (2014) 20–34. [47] A. Singh, Y. Lin, K.R. Ansari, M.A. Quraishi, E.E. Ebenso, S. Chen, W. Liu, Electrochemical and surface studies of some Porphines as corrosion inhibitor for J55 steel in sweet corrosion environment, Appl. Surf. Sci. 359 (2015) 331–339. [48] N. El Hamdani, R. Fdil, M. Tourabi, C. Jama, F. Bentiss, Alkaloids extract of Retama monosperma (L.) Boiss. seeds used as novel eco-friendly inhibitor for carbon steel corrosion in 1 M HCl solution: Electrochemical and surface studies, Appl. Surf. Sci. 357 (2015) 1294–1305. [49] H. Tayebi, H. Bourazmi, B. Himmi, A. El Assyry, Y. Ramli, A. Zarrouk, A. Geunbour, B. Hammouti, Combined electrochemical and quantum chemical study of new quinoxaline derivative as corrosion inhibitor for carbon steel in acidic media, Der Pharm. Chem. 6 (5) (2014) 220–234. [50] Y. Abboud, A. Abourriche, T. Saffaj, M. Berrada, A novel azo dye, 8-quinolinol-5-azoantipyrine as corrosion inhibitor for mild steel in acidic media, Desalination 237 (1–3) (2009) 175–189. [51] M. Gopiraman, N. Selvakumaran, D. Kesavan, Adsorption and corrosion inhibition behaviour of N (phenylcarbamothioyl) benzamide on mild steel in acidic medium, Prog. Org. Coat. 73 (1) (2012) 104–111. [52] R. Hsissou, F. Benhiba, M. Khudhair, M. Berradi, A. Mahsoune, H. Oudda, A. El Harfi, I.B. Obot, A. Zarrouk, Investigation and comparative study of the quantum molecular descriptors derived from the theoretical modeling and Monte Carlo simulation of two new macromolecular polyepoxide architectures TGEEBA and HGEMDA, J. King Saud Univ. Sci. (2018), doi:10.1016/j.jksus.2018.10.008. [53] Y. Feng, S. Chen, Q. Guo, Y. Zhang, Inhibition of iron corrosion by 5,10,15,20-tetraphenylporphyrin and 5,10,15,20-tetra-(4-chlorophenyl)porphyrin adlayers in 0.5 M H2SO4 solutions, J. Electroanal. Chem. 602 (2007) 115–122.