Corrosion inhibition effect of spiropyrimidinethiones on mild steel in 15% HCl solution: insight from electrochemical and quantum studies

M. Yadav*, R. R. Sinha, Sumit Kumar and T. K. Sarkar
Department of Applied Chemistry, Indian School of Mines, Dhanbad 826004, India. E-mail: yadav_drmahendra@yahoo.co.in; Fax: +91-326-2296563; Tel: +91-326-2235428

Received 21st July 2015 , Accepted 12th August 2015

First published on 13th August 2015


Abstract

The effect of synthesized spiropyrimidinethiones, namely, 6′-(4-methoxyphenyl)-1′-phenyl-2′-thioxo-2′,3′-dihydro-1′H-spiro[indoline-3,4′-pyrimidine]-2-one (MPTS) and 6′-(4-methoxyphenyl)-1′-phenyl-2′-thioxo-2′,3′-dihydro-1′H-spiro[indoline-3,4′-pyrimidine]-2-one (CPTS) on the corrosion of mild steel in 15% HCl solution was investigated by using weight loss and electrochemical methods. Both inhibitors act as mixed inhibitors and their adsorption on mild steel obeyed Langmuir's adsorption isotherm. The potential of zero charge (EPZC) for the mild steel was determined by electrochemical impedance spectroscopy methods. Scanning electron microscopy, energy dispersion X-ray spectroscopy, FTIR, X-ray photoelectron spectroscopy (XPS) and atomic force microscopy were used to characterize the surface morphology of the uninhibited and inhibited mild steel specimens. Density Functional Theory (DFT) was employed for theoretical calculations.


1. Introduction

Mild steel is extensively used as a construction material in various industries due to its excellent mechanical properties and low cost, but its poor corrosion resistance to acids, restrains its utility. Hydrochloric acid solutions are commonly used for pickling, industrial acid cleaning, acid de-scaling, and in oil well acidifying processes. Acidization of petroleum oil wells for enhancing oil production is commonly brought about by forcing 15% HCl solution through steel tubing into the well to open up near-bore channels in the formation and, hence, to increase the flow of oil. Because of the aggressiveness of the acid solutions, mild steel corrodes severely during these processes, particularly with the use of hydrochloric acid, which results in terrible waste of both resources and money. A corrosion inhibitor is often added to acid solutions for minimizing the effect of corrosion on metal during these processes. The selection of appropriate inhibitors mainly depends on its economic availability, environmental side effects, concentration, temperature, type of acid, the presence of dissolved inorganic and/or organic substances even in minor amounts and, of course, on the type of metallic material supposed to be protected.1

Heterocyclic organic compounds containing sulfur, phosphorus, oxygen, nitrogen and aromatic rings are reported in literature as most effective corrosion inhibitor for the metals in acidic medium.2–9 Some spiro compounds10 have been reported as good corrosion inhibitor for Al in NaCl solution, and exhibit different inhibition performance with the difference in substituent groups and substituent positions on the aromatic rings. The inhibitor, MPTS and CPTS contains indoline and pyrimidine rings as two anchoring sites suitable for bonding with the metal surface. So, it was expected that MPTS and CPTS would work as good corrosion inhibitor for mild steel in 15% hydrochloric acid solution.

Although, the protection mechanism of organic corrosion inhibitors has not been clearly understood,11 it is generally accepted that the organic inhibitors decrease the corrosion rate by adsorbing on the metal/solution interface and blocking the active sites by displacing water molecules and forming a compact barrier film on the metal surface.12–14 The adsorption process depends upon the nature and surface charge of the metal, the type of aggressive media, the structure of the inhibitor and the nature of its interaction with the metal surface.15 Research activities in present situations are geared towards developing the cheap, non-toxic and environment friendly corrosion inhibitors. Theoretical calculations using Density Functional Theory (DFT) are being used to explain the mechanism of corrosion inhibition.16–18

The objective of the present work was to evaluate the corrosion parameters of two selected inhibitors (MPTS and CPTS) in mild steel/inhibitor/15% HCl solution system by means of weight loss measurement, potentiodynamic polarization, and electrochemical impedance spectroscopy (EIS) techniques. The choice of these compounds is based on molecular structure considerations, i.e., the number of active centers and type of substituent's present in these compounds. Both the compounds have the same number of active centers but they only differ in the substituent type (OCH3 in MPTS and Cl in CPTS) at the para position of phenyl group and hence if a difference in the inhibition efficiency is observed, it should be predominately attributed to a difference in the electronic effect of the substituent type. Quantum chemical calculations were performed in order to determine the correlation between the corrosion inhibition properties of the investigated spiropyrimidinethiones and their molecular structures.

2. Material and methods

2.1. Synthesis of inhibitors

The studied inhibitors, namely, 6′-(4-methoxyphenyl)-1′-phenyl-2′-thioxo-2′,3′-dihydro-1′H-spiro[indoline-3,4′-pyrimidine]-2-one (MPTS) and 6′-(4-methoxyphenyl)-1′-phenyl-2′-thioxo-2′,3′-dihydro-1′H-spiro[indoline-3,4′-pyrimidine]-2-one (CPTS) were synthesized according to a previously reported experimental procedure19 as shown in Scheme 1. The purity of the synthesized compounds was checked by thin layer chromatography (TLC). The structure of the MPTS and CPTS was confirmed by elemental analysis, FTIR, and 1H NMR spectroscopy.
image file: c5ra14406j-s1.tif
Scheme 1 Synthetic route and structure of MPTS and CPTS.
2.1.1. MPTS. Yield: 58%; mp 187; analytical data calculated for C24H19N3SO2; C, 69.73; H, 4.60; N, 10.17; S, 7.75. Found: C, 69.12; H, 4.58%; N, 10.05; S, 7.48.

1H NMR (300 MHz, DMSO-d6) δ ppm: δ 8.12–8.28 (m, 13H, Ar-H), 6.64 (s, 1H, CH), 9.85 (s, 2H, 2NH), 4.68 (s, 3H, OCH3).

IR (KBr) λmax (cm−1): 1510 (C[double bond, length as m-dash]S), 1660 (C[double bond, length as m-dash]O), 3240 (NH), 1160 (–OCH3).

2.1.2. CPTS. Yield: 62%; mp 242; analytical data calculated for C23H16N3SOCl; C, 66.19; H, 3.84; N, 10.47; S, 7.67. Found: C, 65.96; H, 3.82%; N, 10.36; S, 7.57.

1H NMR (300 MHz, DMSO-d6) δ ppm: δ 8.24–8.46 (m, 13H, Ar-H), 6.68 (s, 1H, CH), 9.78 (s, 2H, 2NH).

IR (KBr) λmax (cm−1): 1500 (C[double bond, length as m-dash]S), 1620 (C[double bond, length as m-dash]O), 3250 (NH).

2.2. Mild steel specimens

Mild steel specimens of size 6.0 cm × 2.5 cm × 0.1 cm and 1.0 cm × 1.0 cm × 0.1 cm with composition (w%) C, 0.12; Mn, 0.11; Cu, 0.01; Si, 0.02; Sn, 0.01; P, 0.02; Ni, 0.02 and remainder iron were employed for the weight loss and electrochemical studies, respectively. The specimens were abraded using emery papers of different grit sizes up to 1200 grit, polished with Al2O3 (1 μm and then 0.3 μm particle size), washed with tap water followed by distilled water, degreased with acetone, dried and stored in desiccator.

2.3. Test solution

The test solution (15% HCl solution) was prepared by dilution of AR grade HCl with double distilled water. The volume of test solution was 250 mL for weight loss study 150 mL for electrochemical study. The concentration range of inhibitors employed was 25 to 150 ppm (mg L−1).

2.4. Weight loss measurements

Weight loss measurements were performed in the absence and presence of different concentrations (25 to 150 ppm) of inhibitors at different temperatures (303–333 K) according to the standard methods.20 The corrosion rate (CR), inhibition efficiency (η%) and surface coverage (θ) were determined by following equations:21
 
image file: c5ra14406j-t1.tif(1)
where, W = weight loss (g), A = area of specimen (cm2) exposed in acidic solution, t = exposure time (h), and D = density of mild steel (g cm−3).
 
image file: c5ra14406j-t2.tif(2)
 
image file: c5ra14406j-t3.tif(3)
where, CR0 and CRi are corrosion rate in absence and presence of inhibitors.

2.5. Electrochemical methods

The electrochemical studies were conducted in a conventional three-electrode cell consisting of mild steel sample of 1 cm2 exposed area as working electrode, a platinum counter electrode and a saturated calomel electrode (SCE) as reference electrode, using CH electrochemical workstation (Model No. CHI 760D, manufactured by CH Instruments, Austin, USA) at 303 K. Potentiodynamic polarization curves were obtained at a scan rate of 1 mV s−1 in the potential range −250 to +250 mV vs. SCE at open circuit potential. The linear Tafel segments of anodic and cathodic curves were extrapolated to obtain corrosion current densities (icorr). The percentage inhibition efficiency (η%), was calculated using the equation
 
image file: c5ra14406j-t4.tif(4)
where, i0corr and icorr are the values of corrosion current density in the absence and presence of inhibitor, respectively.

Electrochemical impedance measurements were carried out using AC signals of amplitude 10 mV peak to peak at the open circuit potential in the frequency range of 100 kHz to 10 mHz. The experimental impedance data were fitted to appropriate equivalent circuit using ZSimpWin.3.21 software. The inhibition efficiency (η%) was calculated from charge transfer resistance values obtained from impedance measurements using the following relation

 
image file: c5ra14406j-t5.tif(5)
where Rct(inh) and Rct are charge transfer resistance in presence and absence of inhibitor, respectively. The values of double layer capacitance (Cdl) were calculated from charge transfer resistance and CPE parameters (Y0 and n) using the expression22
 
Cdl = (Y0Rct1−n)1/n (6)
where Y0 is CPE constant and n is CPE exponent. The value of n represents the deviation from the ideal behavior and it lies between 0 and 1.

2.6. FTIR spectrum analysis

The FTIR spectrum of the pure compound and film formed on the surface of N80 steel samples was recorded on Perkin Elmer FTIR, (Spectrum-2000) spectrophotometer.

2.7. Scanning electron microscopic and energy dispersive X-ray spectroscopy analysis

For surface morphological study of the uninhibited and inhibited mild steel samples, SEM and EDX images were recorded using the instrument HITACHI S3400N.

2.8. Atomic force microscopy

The AFM images of polished, uninhibited and inhibited mild steel specimens were carried out using a Nanosurf Easyscan2 instrument.

2.9. XPS analysis

The chemical composition of adsorbed film of inhibitors on mild steel was detected by XPS (VSW Spectrometer) employing Al Kα (1486.6 eV) as the incident radiation source and the binding energy of C 1s (284.6 eV) was used as a reference. XPS analysis was performed without argon sputtering of the specimens.

2.10. Quantum chemical study

Complete geometrical optimization of the investigated inhibitors were performed using Density Functional Theory (DFT) with the Becke's three parameter exchange functional along with the Lee–Yang–Parr nonlocal correlation functional (B3LYP) with 6-31G (d, p) basis set implemented in Gaussian 03 program package.23,24 Theoretical parameters such as the energy of the highest occupied and lowest unoccupied molecular orbital (EHOMO and ELUMO), energy gap (ΔE), dipole moment (μ) absolute electronegativity (χ), global hardness (γ) and softness (σ), and fraction of electrons transferred (ΔN) were determined.

3. Results and discussion

3.1. Weight loss measurements

3.1.1. Effect of inhibitor concentration and temperature. Corrosion parameters for mild steel in 15% HCl solution were determined from weight loss measurement data in the absence and presence of different concentrations (25–250 ppm) of the studied inhibitors at different temperatures (303–333 K) and listed in Table 1. Inspection of the data in Table 1, reveals that inhibition efficiency increases with increasing the concentration of each inhibitor up to 150 ppm concentration and after 150 ppm concentration remains almost constant therefore 150 ppm was taken as optimum concentration and the concentration range was taken as 25 ppm to 150 ppm for other studies. The inhibition efficiency of both inhibitors decreases with an increase in temperature. Such behavior can be interpreted on the basis that the inhibitors exert their action by adsorbing themselves on the mild steel surface and an increase in temperature resulted in desorption of some adsorbed inhibitor molecules leading to a decrease in the inhibition efficiency.25 It is evident from Table 1 that both inhibitors are good inhibitors even at the concentration as low as 25 ppm. The inhibition efficiency of MPTS and CPTS at 150 ppm concentration was found to be 94.8% and 93.4% respectively, while it was 67.4% and 64.7% respectively at 25 ppm concentration at 303 K (Table 1). The inhibition efficiency of MPTS was found to be better than CPTS at all concentrations and temperatures. The inhibitors MPTS and CPTS have nearly same size and number of active centers, the difference between the two inhibitors is that MPTS contains –OCH3 substituent on phenyl ring whereas CPTS contains –Cl substituent on phenyl ring present in these inhibitors. Thus, the difference in corrosion inhibition efficiency between MPTS and CPTS is due to different nature of these substituent's (–OCH3 and –Cl). The delocalized π-electron density at phenyl ring in case of MPTS increased due to electron donating nature of methoxy (–OCH3) substituent whereas the electron density on phenyl ring in case of CPTS decreased due to electron withdrawing nature of chloro (–Cl) substituent. The higher delocalized π-electron density at phenyl ring in case of MPTS as compared to CPTS facilitate greater adsorption of MPTS on mild steel surface than CPTS, leading to higher inhibition efficiency of MPTS than CPTS. The same effect of substituent has been reported in literature by some researchers.26
Table 1 Corrosion parameters of mild steel in 15% HCl solution in the presence and absence of inhibitors at different temperatures, obtained from weight loss measurements
Conc. (ppm) 303 K 313 K 323 K 333 K
CR (mm per year) θ η% CR (mm per year) θ η% CR (mm per year) θ η% CR (mm per year) θ η%
Blank 20.20 34.55 55.47 92.69
[thin space (1/6-em)]
MPTS
25 6.58 0.674 67.4 11.91 0.655 65.5 20.87 0.623 62.3 38.16 0.588 58.8
50 3.97 0.803 80.3 7.34 0.787 78.7 13.12 0.763 76.3 24.96 0.730 73.0
75 2.41 0.880 88.0 4.58 0.867 86.7 8.60 0.844 84.4 17.24 0.814 81.4
100 1.66 0.917 91.7 3.20 0.907 90.7 6.27 0.886 88.6 13.22 0.857 85.7
150 1.05 0.948 94.8 2.29 0.933 93.3 4.71 0.915 91.4 10.32 0.888 88.8
200 0.97 0.952 95.2 2.28 0.934 93.4 4.16 0.915 92.5 10.01 0.892 89.2
250 0.91 0.955 95.5 2.21 0.936 93.6 4.05 0.917 92.7 9.82 0.894 89.4
[thin space (1/6-em)]
CPTS
25 7.11 0.647 64.7 12.84 0.628 62.8 22.23 0.599 59.9 40.54 0.562 56.2
50 4.34 0.784 78.4 8.04 0.767 76.7 14.40 0.740 74.0 27.18 0.706 70.6
75 2.75 0.863 86.3 5.17 0.850 85.0 9.67 0.825 82.5 18.99 0.795 79.5
100 1.95 0.903 90.3 3.74 0.891 89.1 7.19 0.870 87.0 14.54 0.843 84.3
150 1.31 0.934 93.4 2.75 0.920 92.0 5.35 0.900 90.0 11.57 0.875 87.5
200   0.946 94.6   0.934 93.4   0.903 90.3   0.877 87.7
250   0.948 94.8   0.936 93.6   0.905 90.5   0.879 87.9


3.1.2. Effect of exposure period. The variation of the inhibition efficiency for MPTS and CPTS at optimum concentration (150 ppm) with immersion time is shown in Fig. 1. The corrosion parameters such as corrosion rate and corrosion inhibition efficiencies obtained for MPTS and CPTS at optimum concentration (150 ppm) and different exposure periods (6–120 h) are shown in Table 2. It is clear from the Fig. 1 that inhibition efficiency of MPTS and CPTS decreases with increase in exposure time (6–120 h) at 303 K due to the partial desorption of adsorbed inhibitors on increasing the exposure time.27
image file: c5ra14406j-f1.tif
Fig. 1 Variation of inhibition efficiency with exposure period in presence of 150 ppm concentration of MPTS and CPTS at 303 K.
Table 2 Effect of time on corrosion rate and inhibition efficiency without and with 150 ppm of MPTS and CPTS at 303 K
Exposure period (h) Blank MPTS CPTS
CR (mm per year) CR (mm per year) η (%) CR (mm per year) η (%)
6 20.2 1.05 94.8 1.31 93.5
24 109.4 8.75 92.0 9.62 91.2
48 157.9 15.15 90.4 16.57 89.5
72 182.6 20.63 88.7 24.46 86.6
96 202.4 27.12 86.6 31.97 84.2
120 222.6 35.17 84.2 39.40 82.3


3.1.3. Thermodynamic and activation parameters. The apparent activation energy (Ea) for dissolution of mild steel in 15% HCl was calculated by using the Arrhenius equation:
 
image file: c5ra14406j-t6.tif(7)
where R is the molar gas constant (8.314 J K−1 mol−1), T is the absolute temperature (K) and A is the Arrhenius pre-exponential factor. Fig. 2(a and b) presents the Arrhenius plot of log[thin space (1/6-em)]CR against 1/T for the corrosion of mild steel in 15% HCl solution in the absence and presence of inhibitors (MPTS and CPTS) at concentrations ranging from 25 to 150 ppm. From Fig. 1(a and b), the activation energy was calculated using the expression Ea = −(slope) × 2.303R. The calculated values of Ea are summarized in Table 3. It is evident from Table 3 that the values of the apparent activation energy for the inhibited solutions were higher than that for the uninhibited solution and continued to increase on increasing the concentration of inhibitor. This explains that the energy barrier of corrosion reaction increases with the concentration of inhibitor.28

image file: c5ra14406j-f2.tif
Fig. 2 Arrhenius plots of log[thin space (1/6-em)]CR versus 1000/T−1 for mild steel corrosion in 15% HCl solution (a) MPTS (b) CPTS.
Table 3 Activation parameter for mild steel in 15% HCl solution in the absence and presence of inhibitors obtained from weight loss measurements
Inhibitor Concentration (ppm) Ea (kJ mol−1) ΔH* (kJ mol−1) ΔS* (J mol−1 K−1)
Blank 42.34 39.70 −89.20
MPTS 25 48.96 46.32 −76.75
50 51.15 48.51 −73.31
75 54.80 52.16 −66.03
100 58.15 55.51 −58.18
150 65.59 60.95 −43.82
CPTS 25 48.42 45.79 −77.90
50 51.05 48.43 −73.88
75 53.88 51.24 −67.94
100 56.04 53.40 −63.73
150 60.42 57.78 −52.44


From the thermodynamic and kinetic point of view, the unimolecular reactions is characterized by following equation:

 
Ea − ΔH* = RT (8)

As observed from Table 3, for both inhibitors Ea > ΔH* by a value which approximately equal to RT. Hence, mild steel sample corrodes in 15% HCl solutions either in absence or presence of different concentrations of the studied inhibitors by a unimolecular reaction.

The values of standard enthalpy of activation (ΔH*) and standard entropy of activation (ΔS*) for the formation of the activation complex in the transition state was calculated by using the transition state equation:

 
image file: c5ra14406j-t7.tif(9)
where, h is Planck's constant and N is the Avogadro number, respectively.

A plot of log(CR/T) against 1/T (Fig. 3(a and b)) gave straight lines with a slope of −ΔH*/2.303R and an intercept of [log(R/Nh) + (ΔS*/2.303R)], from which the activation thermodynamic parameters ΔH* and ΔS* were calculated, as listed in Table 3. The negative value of ΔS* for both inhibitors indicates that the formation of the activated complex in the rate determining step represents an association rather than a dissociation step, meaning that a decrease in disorder takes place during the course of the transition from reactants to activated complex.29


image file: c5ra14406j-f3.tif
Fig. 3 Transition state plots of log[thin space (1/6-em)]CR/T versus 1000/T−1 for mild steel in 15% HCl solution at different concentrations (a) MPTS (b) CPTS.
3.1.4. Adsorption isotherm. The surface coverage (θ) for different concentrations of both studied inhibitors in 15% hydrochloric acid obtained from weight loss method was tested graphically for fitting a suitable adsorption isotherm. Plotting Cinh/θ vs. Cinh yielded a straight line [Fig. 4] with a correlation coefficient (R2) and slope values as given in Table 4 at different temperatures. The correlation coefficient and slope values for both inhibitors as shown in Table 3 are near to unity indicating that the adsorption of these inhibitors obey the Langmuir adsorption isotherm represented by the following equation.
 
image file: c5ra14406j-t8.tif(10)
where, Cinh is the inhibitor concentration and Kads is the equilibrium constant for adsorption–desorption process. From the intercept of Fig. 3, the values of Kads were calculated and represented in Table 4. Large values of Kads obtained for both studied inhibitors imply more efficient adsorption and hence better corrosion inhibition efficiency. Using the values of Kads, the values of ΔG0ads was calculated by using the equation:
 
ΔG0ads = −RT[thin space (1/6-em)]ln(55.5Kads) (11)
where R is the gas constant and T is the absolute temperature (K). The value of 55.5 is the concentration of water in solution in mol L−1. Calculated values of ΔG0ads are listed in Table 4. Normally, a value of ΔG0ads around −20 kJ mol−1 or less negative is assumed for electrostatic interactions between inhibitor molecules and the charged metal surface (physisorption), and a value around −40 kJ mol−1 or more negative is indicative of charge sharing or transferring from organic species to the metal surface to form a coordinate bond (chemisorption).30 Thus, the negative values of ΔG0ads in the range of −40 kJ mol−1 to −30 kJ mol−1, indicate that the adsorption mechanism of inhibitor on mild is a combination of both physisorption and chemisorptions. The calculated ΔG0ads values for MPTS and CPTS were found in the range of −36.2 to −39.0 and −36.1 to −38.8 kJ mol−1, respectively, at different temperatures (303–333 K). These values are between the threshold values for physical adsorption and chemical adsorption, indicating that the adsorption process of these inhibitors at mild steel surface involve both the physical as well as chemical adsorption. More negative value of ΔG0ads for MPTS indicates a stronger interaction between the inhibitor molecules and the metal surface. It is apparent from the Table 3 that, the negative value of ΔG0ads decreased on increasing the temperatures, suggesting that the adsorption of inhibitor molecules on mild steel surface is not favorable with increasing experimental temperature, indicating that physisorption has the major contribution while chemisorption has the minor contribution in the adsorption process. Similar conclusion was also reported by Ozcan,31 who studied the use of cystine as a corrosion inhibitor on mild steel in sulfuric acid.

image file: c5ra14406j-f4.tif
Fig. 4 Langmuir plots of (Cinh/θ) versus Cinh for (a) MPTS (b) CPTS.
Table 4 Thermodynamic parameters of adsorption for studied inhibitors on mild steel in 15% HCl solution at different temperatures
Inhibitor Temperature (K) Kads (M−1) ΔG0ads (kJ mol−1) R2 ΔH0ads (kJ mol−1) ΔS0ads (J mol−1 K−1)
MPTS 303 K 2.4 × 104 −39.0 0.999    
313 K 2.7 × 104 −38.2 0.996    
323 K 2.9 × 104 −37.1 0.997 −67.83 −95
333 K 3.2 × 104 −36.2 0.994    
CPTS 303 K 2.2 × 104 −38.8 0.993    
313 K 2.5 × 104 −38.0 0.996 −66.41 −91
323 K 2.7 × 104 −37.0 0.997    
333 K 3.0 × 104 −36.1 0.998    


The enthalpy and entropy changes for the adsorption of inhibitors (MPTS and CPTS) on mild steel surface were determined from the basic thermodynamic equation:32

 
ΔG0ads = ΔH0adsTΔS0ads (12)
where ΔH0ads and ΔS0ads are the standard enthalpy and entropy changes of adsorption process, respectively. The plot of ΔG0ads versus T was linear (Fig. 5) with the slope equal to −ΔS0ads and intercept of ΔH0ads. The calculated values of ΔH0ads and ΔS0ads are listed in Table 4. The ΔHads values are negative for both studied inhibitors, suggests that the adsorption of inhibitor's molecules on mild steel surface is an exothermic process. It has been reported in literature that an endothermic adsorption process (ΔHads > 0) is due to chemisorption while an exothermic adsorption process (ΔHads < 0) may be attributed to physisorption, chemisorption or a mixture of both.33 In an exothermic process, physisorption is distinguished from chemisorption by considering the absolute value of ΔHads, it is lower than 40 kJ mol−1 for the physisorption process whereas for chemisorption process approaches 100 kJ mol−1. In the present study, Hads values for MPTS and CPTS are −67.83 kJ mol−1 and −66.41 kJ mol−1 respectively, which are larger than the common physical adsorption heat, but smaller than the common chemical adsorption heat, emphasizing that mixed type (both physical and chemical) adsorption take place. The similar results were also reported by other authors.34–36


image file: c5ra14406j-f5.tif
Fig. 5 The relationship between ΔG0ads and temperature.

The negative value of ΔS0ads obtained for adsorption of both the inhibitors as shown in Table 4 suggested that before the adsorption of inhibitor's molecules on the mild steel surface, inhibitor molecules might freely move in the bulk solution, but with the progress in the adsorption, inhibitor molecules were orderly adsorbed on the mild steel surface, as a result a decrease in entropy is observed.37 From the thermodynamic principles, since the adsorption is an exothermic process, it must be accompanied by a decrease of entropy.38

3.2. Electrochemical studies

3.2.1. Polarization studies. The potentiodynamic polarization curves of mild steel in 15% HCl solution in the absence and presence of various concentrations of MPTS and CPTS at 303 K are shown in Fig. 6(a and b), respectively. It is clear from the Fig. 6(a and b), that both, the anodic metal dissolution and cathodic hydrogen evolution reactions are inhibited after the addition of inhibitors to the 15% HCl solution. The inhibition of these reactions are more pronounced with the increasing inhibitor concentrations, results increase in inhibition efficiency. The inhibitor molecules are first adsorbed on the mild steel surface and blocking the available reaction sites.39 The surface coverage increases with increase in inhibitor concentrations. The corrosion current densities were calculated by extrapolation of linear parts of anodic and cathodic curves to the point of intersection. The electrochemical corrosion parameters such as corrosion potential (Ecorr), anodic Tafel slope (βa), cathodic Tafel slope (βc), corrosion current density (icorr) and percentage inhibition efficiency (η%) obtained from these curves are given in Table 5. It is clear from Table 4 that both the anodic and cathodic current density values are considerably decreased in inhibited solutions when compared with the uninhibited solution. The observed phenomenon may be the result of covering of adsorbed inhibitor molecules on the mild steel surface and decreasing dissolution of mild steel. The minor shift in Ecorr value (15 mV) towards negative direction in presence of inhibitor as compared to the Ecorr value in absence of inhibitor, indicating that MPTS and CPTS acts as mixed type inhibitor with predominant control of cathodic reaction.40
image file: c5ra14406j-f6.tif
Fig. 6 Potentiodynamic polarization curves for mild steel in 15% HCl solution in the presence and absence of inhibitor (a) MPTS (b) CPTS at 303 K.
Table 5 Electrochemical parameters determined from polarization measurements for mild steel in 15% HCl solution in the presence and absence of different concentrations of inhibitors at 303 K
Conc. (ppm) Ecorr (mV vs. SCE) βa (mV per decade) βc (mV per decade) icorr (μA cm−2) η%
Blank −495 93 142 568
[thin space (1/6-em)]
MPTS
25 −503 73 124 169.4 70.1
50 −498 86 131 107.7 81.0
75 −500 91 148 59.6 89.5
100 −505 62 162 46.5 91.8
150 −505 77 152 29.5 94.8
[thin space (1/6-em)]
CPTS
25 −502 94 138 186.3 67.2
50 −504 74 114 113.6 80.0
75 −508 58 157 67.6 88.1
100 −510 78 129 51.1 91.0
150 −500 85 140 39.9 92.9


3.2.2. EIS studies. Electrochemical impedance measurements were undertaken to provide information on the kinetics of the electrochemical processes at the mild steel/acid interface. Nyquist plots for mild steel corrosion in 15% HCl solution in the absence and presence of different concentrations (25–150 ppm) of the inhibitors MPTS and CPTS are shown in Fig. 7(a and b), respectively. The Nyquist plots for both inhibitors show single semicircles over the frequency range studied, corresponding to one time constant. The capacitive loops are not perfect semicircles, because of non-homogeneity and roughness of the mild steel surface.41 The impedance spectra were analyzed by fitting the experimental data to the equivalent circuit model shown in Fig. 8, which is a parallel combination of the charge transfer resistance (Rct) and the constant phase element (CPE), both in series with the solution resistance (Rs). This type of electrochemical equivalent circuit was reported previously to model the iron/acid interface.42 The CPE is used in place of a capacitor to compensate deviations from ideal dielectric behavior arising from the inhomogeneous nature of the electrode surfaces. The impedance of the CPE is given by
 
ZCPE = Y0−1()n (13)
where Y0 and n stand for the CPE constant and exponent, respectively, j = (−1)1/2 is an imaginary number, and ω is the angular frequency in rad s−1 (ω = 2πf), where f is the frequency in Hz. The electrochemical parameters such as solution resistance (Rs), charge transfer resistance (Rct) and CPE constants (Y0 and n) obtained from fitting the experimental data of Nyquist plots in the equivalent circuit shown in Fig. 8 are presented in Table 6. The data shown in Table 5 reveal that the value of Rct increases with addition of inhibitors as compared to the blank solution, the increase in Rct value is attributed to the formation of a protective film at the metal/solution interface. The Cdl value decreases on increasing the concentration of both the inhibitors, indicating the decrease in local dielectric constant and/or to an increase in the thickness of the electrical double layer, suggesting that the inhibitor molecules are adsorbed at the metal/solution interface.43

image file: c5ra14406j-f7.tif
Fig. 7 Nyquist plot for mild steel in 15% HCl solution containing various concentrations of (a) MPTS (b) CPTS at 303 K.

image file: c5ra14406j-f8.tif
Fig. 8 Equivalent circuit applied for fitting of the impedance spectra.
Table 6 Electrochemical parameters determined from electrochemical impedance measurements for mild steel in 15% HCl solution in the presence and absence of different concentrations of inhibitors at 303 K
Conc. (ppm) Rs (Ω cm2) Rct (Ω cm2) Y0 (μF cm−2) n Cdl (μF cm2) η%
Blank 0.93 20 573 0.845 252
[thin space (1/6-em)]
MPTS
25 0.76 66 237 0.865 123.8 69.7
50 0.54 112 157 0.872 86.7 82.1
75 0.69 161 122 0.878 70.6 87.5
100 0.73 230 91 0.893 57.2 91.3
150 0.82 327 58 0.923 41.6 93.8
[thin space (1/6-em)]
CPTS
25 0.82 62 268 0.852 131.5 67.7
50 0.63 98 183 0.861 95.6 79.6
75 0.69 148 145 0.867 80.4 86.5
100 0.58 217 113 0.872 65.5 90.7
150 0.74 296 69 0.912 47.4 93.2


The Bode phase angle plots (Fig. 9(a and b)) show single maximum (one time constant) at intermediate frequencies, broadening of this maximum in presence of inhibitors accounts for the formation of a protective layer on the electrode surface. Fig. 9(a and b) shows that the impedance value in the presence of both inhibitors is larger than in absence of inhibitors and the value of impedance increases on increasing the concentration of both studied inhibitors. These mean that the corrosion rate is reduced in presence of the inhibitors and continued to decreasing on increasing the concentration of inhibitors.


image file: c5ra14406j-f9.tif
Fig. 9 Bode (log[thin space (1/6-em)]f vs. log|Z|) and phase angle (log[thin space (1/6-em)]f vs. −phase) plots of impedance spectra for mild steel in 15% HCl solution in the presence and absence of inhibitor (a) MPTS (b) CPTS at 303 K.

Electrochemical results (η%) are in good agreement with the results (η%) obtained by weight loss experiments.

3.2.3. Potential of zero charge (EPZC). The phenomenon of adsorption is influenced by the chemical structure of the inhibitor, the charge or dipole moment of the inhibitor molecules and the charge of the metal surface. The potential of zero charge (EPZC) of mild steel in 15% HCl solution in presence of inhibitors was determined by using electrochemical impedance spectroscopy method. The minimum on the double layer capacitance (Cdl) versus applied voltage (E) plots (Fig. 10) is considered as the value of EPZC of the mild steel. The surface charge of the metal can be defined by the position of the open circuit potential (Eocp) with respect to the EPZC. The net surface charge of the mild steel at the open circuit potential was evaluated according to the equation:
 
Er = EocpEPZC (14)
where Er is the Antropov's “rational” corrosion potential.44 The positive and negative values of Er suggesting the positive and negative net charge on the mild steel surface, respectively, at open circuit potential. Table 7 shows the values of Eocp, EPZC and Er for both inhibitors at 150 ppm concentration. The positive value of Er for both the inhibitors with respect to EPZC, as shown in Table 7, suggesting that mild steel surface is positively charged at open circuit potential in presence of 150 ppm of these inhibitors. Since mild steel surface is positively charged therefore anions (Cl ions) in aqueous hydrochloric acid solution gets adsorbed on the mild steel surface. After the adsorption of the Cl ions, the mild steel surface becomes negatively charged. Hence, the protonated positively charged form of inhibitor (MPTS+ and MPTS+) adsorbed on the surface of mild steel via electrostatic interactions with the Cl ions already adsorbed on mild steel surface. Other sides, the inhibitor molecules may also adsorbed on the mild steel surface via the unshared pair of electrons present on N, O and S atoms.

image file: c5ra14406j-f10.tif
Fig. 10 The plot of Cdl vs. applied potential for mild steel in 15% HCl solution containing 150 ppm of MPTS and CPTS at 303 K.
Table 7 Values of Eocp, EPZC and Er recorded for mild steel in 15% HCl solution in absence and presence of 150 ppm of MPTS and CPTS at 303 °C
Inhibitor Eocp (mV vs. SCE) EPZC (mV vs. SCE) Er (mV vs. SCE)
MPTS −508 −520 12
CPTS −500 −505 05


3.3. Analysis of FTIR spectra

The FTIR spectra of pure inhibitors and reflectance absorption FTIR spectra of the inhibited mild steel samples were recorded and represented in Fig. 11(a and b). The pure MPTS shows the IR bands around 3240, 1660 and 1510 cm−1 due to presence of –NH, C[double bond, length as m-dash]O and C[double bond, length as m-dash]S, respectively, whereas pure CPTS shows the IR band around 3250, 1620 and 1500 cm−1 due to presence of –NH, C[double bond, length as m-dash]O and C[double bond, length as m-dash]S group, respectively. The reflectance absorption FTIR spectra of the exposed specimens in MPTS and CPTS show bands around 3180, 1640, 1460 and 3250, 1620, 1500 cm−1 respectively due to presence of –NH, C[double bond, length as m-dash]O and C[double bond, length as m-dash]S group, respectively. The shift in peak position of –NH, C[double bond, length as m-dash]O and C[double bond, length as m-dash]S in reflectance absorption FTIR spectra of both the inhibitors as compared to their pure compounds spectra, indicating the involvement of these groups in adsorption process of inhibitors with the mild steel surface.
image file: c5ra14406j-f11.tif
Fig. 11 FTIR spectrum of the pure inhibitors and film formed on the surface of N80 steel specimen.

3.4. Scanning electron microscopy

SEM images of mild steel specimen before and after immersion in 15% HCl solution as well as in presence of 150 ppm of MPTS and CPTS are shown in Fig. 12(a–d). The morphology of the mild steel specimen before immersion (Fig. 12(a)) is very smooth and shows no corrosion while mild steel specimen immersed in 15% HCl solution in the absence of inhibitor (Fig. 12(b)) is very rough and the surface is badly damaged due to metal dissolution. However, the presence of 150 ppm of MPTS and CPTS suppresses the rate of corrosion and surface damage has been diminished considerably (Fig. 12(c and d)) as compared to the blank solution (Fig. 12(b)), suggesting formation of a protective inhibitor film at the mild steel surface.
image file: c5ra14406j-f12.tif
Fig. 12 SEM images of mild steel in 15% HCl solution after 6 h immersion at 303 K. (a) Before immersion (polished), (b) after immersion without inhibitor (c) after immersion with 150 ppm MPTS (d) after immersion with 150 ppm CPTS.

3.5. Energy dispersive X-ray spectroscopy

Energy dispersive X-ray analysis (EDX) technique was employed in order to get information about the composition of the surface of the mild steel sample in the absence and presence of inhibitors in 15% HCl solution. The results of EDX spectra are shown in Fig. 13(a–d). The EDX spectra of uninhibited mild steel specimen contains the peaks corresponding to the element present in mild steel whereas inhibited mild steel contains the peaks corresponding to all the elements present in the inhibitor molecules, indicating the adsorption of inhibitor molecules at the surface of mild steel. The percentage atomic content of various elements on the surface of mild steel specimen before immersion, after immersion and after inhibition was determined by EDX and summarized in Table 8. The percentage atomic content of Fe for mild steel after immersion in 15% HCl solution is 83.12%, and those for mild steel immersed in presence of 150 ppm of MPTS and CPTS are 72.48% and 76.38%, respectively. The suppressed percentage atomic content of Fe in inhibited specimen as compared with the uninhibited mild steel specimen is due to the inhibitory film formed on the mild steel surface.
image file: c5ra14406j-f13.tif
Fig. 13 EDX spectra of mild steel specimens (a) before immersion (b) after immersion without inhibitor (c) after immersion with 150 ppm MPTS (d) after immersion with 150 ppm CPTS.
Table 8 Percentage atomic contents of elements obtained from EDX spectra
Inhibitors Fe C S Mn Cl N O
Mild steel 85.26 12.46 0.46
Blank HCl 83.12 15.68 0.28 2.29 6.36
MPTS 72.48 19.26 1.84 0.16 0.34 5.36 15.24
CPTS 76.38 19.84 1.72 0.12 0.64 5.48 14.12


3.6. Atomic force microscopy

The three-dimensional AFM images of polished (before immersion), uninhibited and inhibited mild steel specimen are shown in Fig. 14(a–d). The average roughness of polished mild steel specimen (Fig. 14(a)) and mild steel specimen in 15% HCl solution without inhibitor (Fig. 14(b)) were found as 25 and 450 nm. It is clearly shown in Fig. 14(b) that mild steel sample is badly damaged due to the acid attack on surface. However, in presence of optimum concentration (150 ppm) of MPTS and CPTS as shown in Fig. 14(c and d), the average roughness were reduced to 130 and 180 nm, respectively. The lower value of roughness for MPTS than CPTS reveals that MPTS protects the mild steel surface more efficiently than CPTS in 15% HCl solution.
image file: c5ra14406j-f14.tif
Fig. 14 AFM micrograph of mild steel surface (a) before immersion (b) after immersion without inhibitor (c) after immersion with 150 ppm MPTS (d) after immersion with 150 ppm CPTS at 303 K.

3.7. XPS analysis

XPS analysis was carried out in order to investigate the composition of the adsorbed films on the steel surfaces in the presence of inhibitors. The XPS spectra were obtained for the mild steel surfaces taken out after 24 h immersion in 15% HCl solution containing 150 ppm of MPTS and CPTS separately. The nature of both the XPS spectra were almost similar therefore here we are discussing only the XPS spectra of MPTS protected mild steel surface. The XPS spectra, obtained for MPTS (C 1s, Fe 2p, O 1s, S 2p, N 1s) is shown in Fig. 15(a–e). These spectra show complex forms which have been assigned to the corresponding species through a deconvolution fitting procedure using the XPS Peak-Fit 4.1 software.
image file: c5ra14406j-f15.tif
Fig. 15 The XPS deconvoluted profiles for C 1s, Fe 2p, O 1s, S 2p and N 1s for mild steel after 24 h immersion period in 15% HCl solution in the presence of 150 ppm of MPTS at 303 K.

The C 1s spectrum of protected mild steel in presence of MPTS are fitted into three peaks (284.6, 287.2 and 288.6 eV) (Fig. 15(a)). The peak with characteristic binding energy of 284.6 eV can be attributed to the C–C and C–H aromatic bonds. The peak at 287.2 eV may be assigned to the carbon atoms bonded to nitrogen. The peak at 288.6 eV may be ascribed to the carbon atom of the C–O and C–S bond.45

The Fe 2p spectrum of MPTS adsorbed films (Fig. 15(b)) exhibit two peaks at 711.4 and 715.2 eV for Fe 2p3/2 and Fe 2p1/2, respectively which are attributed to ferric compound such as Fe2O3 and FeOOH.40 The formation of a stable and insoluble layer of Fe2O3 and FeOOH reduces ionic diffusion and thus improves the corrosion resistance of the mild steel in hydrochloric acid solution. The peak, observed at 715.4 eV is characteristic of Fe(II) species.45

The deconvolution of the O 1s spectrum of MPTS may be fitted into three main peaks shown in Fig. 15(c). The peak at 530.4 eV can be assigned to O2−, bonded with Fe3+ in the Fe2O3 oxides.48 The peak located at 532.4 eV is ascribed to OH, as in FeOOH. The small peak at 533.6 eV may be assigned to the C–O bond and of adsorbed water with mild steel.45

The S 2p XPS spectra for MPTS protected mild steel surface displayed three peaks (Fig. 15(d)) at 161.4 eV, 164.4 eV and 169.2 eV assigned as sulfide species, S–C bond and S–Fe bound respectively.45

The N 1s XPS spectra for MPTS protected mild steel surface displayed three peaks (Fig. 15(e)). The first peak located at 399.6 is attributed to C–N bond. The second peak at 401.2 eV is assigned to nitrogen atoms of indoline and/or pyrimidine ring coordinated with the steel surface (N–Fe). The last peak at 401.6 eV is attributed to positively charged nitrogen, and could be related to protonated nitrogen atoms in indoline and/or pyrimidine ring.45

From these observations it is concluded that the XPS results support the presence of adsorbed MPTS on the mild steel surface.

3.8. Quantum chemical analysis

In order to study the effect of molecular structure on the inhibition efficiency, quantum chemical calculations were performed by using DFT and all the calculations were carried out with the help of complete geometry optimization. Optimized structure, EHOMO and ELUMO of MPTS and CPTS are shown in Fig. 16(a and b). For the calculations of quantum chemical parameters the following equations were used:46
 
image file: c5ra14406j-t9.tif(15)
 
image file: c5ra14406j-t10.tif(16)

image file: c5ra14406j-f16.tif
Fig. 16 The optimized structure (left), HOMO (center) and LUMO (right) distribution for molecules (a) MPTS (b) CPTS [atom legend: white = H; grey = C; blue = N; red = O; yellow = S; green = Cl].

The inverse of the global hardness is designated as the softness, σ as follows:

 
image file: c5ra14406j-t11.tif(17)
where, hardness and softness measures the stability and reactivity of a molecule. Soft molecules are considered to be more reactive than hard ones because they can offer electron to acceptors easily. For the simplest transfer of electrons, adsorption could occur at the part of the molecule where σ which is a local property, has the highest value.45 Thus, the inhibitor having maximum value of softness adsorbed strongly at the surface of mild steel and shows maximum inhibition efficiency.

The fraction of electrons transferred (ΔN) was calculated by using the equation:46

 
image file: c5ra14406j-t12.tif(18)
where a theoretical value of χFe ≈ 4.06 eV is taken for iron46 and γFe = 0 is taken assuming that I = A for bulk metals.47

The quantum chemical parameters such as the energy of the highest occupied molecular orbital (EHOMO), the energy of the lowest unoccupied molecular orbital (ELUMO), energy gap ΔEE = ELUMOEHOMO), ΔE1E1 = ELUMO(inhibitor)EHOMO(Fe)], ΔE2E2 = ELUMO(Fe)EHOMO(inhibitor)], dipole moment (μ), absolute electronegativity (χ), global hardness (γ) and softness (σ), and fraction of electrons transferred (ΔN) were calculated and summarized in Table 9.

Table 9 Quantum chemical parameters for both inhibitors
Inhibitor EHOMO (eV) ELUMO (eV) ΔE (eV) ΔE1 (eV) ΔE2 (eV) μ (D) γ (eV) σ (eV−1) χ (eV) ΔN
Fe −5.075 −1.747  
MPTS −4.884 −1.176 3.707 4.899 3.137 6.718 1.854 0.539 3.030 0.278
CPTS −4.998 −1.146 3.852 3.929 3.251 6.034 1.926 0.519 3.072 0.256


The reactivity of a chemical species can be defined in terms of frontier orbitals: the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO).44 According to the frontier molecular orbital (FMO) theory of chemical reactivity, the formation of a transition state is due to interaction between HOMO and LUMO of reacting species. The smaller the orbital energy gap (ΔE) between the participating HOMO and LUMO, the stronger is the interactions between two reacting species.48 The results listed in Table 9 show that the interaction between the HOMO of the inhibitor and the LUMO of the Fe atom, as represented by ΔE2E2 = ELUMO(Fe)EHOMO(inhibitor)], is stronger than that between the HOMO of the Fe atom and the LUMO of the inhibitor ΔE1E1 = ELUMO(inhibitor)EHOMO(Fe)]. In principle, the interaction between the HOMO of the inhibitor and the LUMO of the Fe atom, ΔE2, should be dominated by the interaction between the HOMO of the Fe atom and the LUMO of the inhibitor, ΔE1, for good inhibitive property.49 Adsorption of the inhibitor on the mild steel surface can occur on the basis of donor–acceptor interactions between the lone-pair electron of the heteroatom present in the inhibitors and the vacant d orbital of the mild steel surface Fe atoms. The inhibitors with high values of EHOMO have a tendency to donate electrons to appropriate acceptors with low-energy, empty molecular orbital.50,51 The calculated value of EHOMO for MPTS (−4.884 eV) and CPTS (−4.998 eV) being higher than that for Fe (−5.075 eV) indicates that both inhibitors have a tendency to donate electrons to vacant d orbital of Fe. On the other hand, a lower value of ELUMO (−1.747 eV) for Fe than that for MPTS (−1.176 eV) and CPTS (−1.146 eV) favors Fe to accept electrons. The lower values of ΔE2 as compared to ΔE1 for both the inhibitors suggesting that more donation of electrons occur from inhibitor to metal. The higher value of EHOMO and lower value of ΔE2 for MPTS as compared to CPTS, indicating that MPTS is better inhibitor than CPTS.

The higher value of dipole moment of MPTS and CPTS as shown in Table 9, clearly suggest that both inhibitors are polar compounds and can easily, donate electrons to form strong dπ–pπ bond.52 According to HSAB theory hard acids prefer to co-ordinate to hard bases and soft acid to soft bases. Fe is considered as soft acid and will co-ordinate strongly to molecule having maximum softness (σ) value. The inhibitor MPTS, which has the highest value of σ (0.278 eV), than CPTS has the higher inhibition efficiency than CPTS (Table 8). Generally, ΔN shows inhibition efficiency resulting from electrons transferred from the inhibitor molecule to the iron atom. According to Lukovits et al.,53 if the value of ΔN is less than 3.6, the inhibition efficiency increases with increasing electron-donating ability of the inhibitor at the metal surface. An improvement in electron-releasing power was shown by replacing one hydrogen atom of phenyl ring by electron-donating substituent (−OCH3 group) as in case of MPTS, which improved the inhibition efficiency but at the other hand a diminishing effect has been observed by electron-withdrawing substituent (−Cl) as in case of CPTS.

Fig. 16(a and b), reveal that the HOMO location in both inhibitors is mostly distributed in vicinity of the pyrimidine rings and –C[double bond, length as m-dash]S, C[double bond, length as m-dash]O groups. The LUMO location in both inhibitors is mostly distributed in vicinity of the pyrimidine, indoline and phenyl rings and –C[double bond, length as m-dash]S, C[double bond, length as m-dash]O group. These indicate the reactive sites of the inhibitors for interaction between inhibitor molecules and mild steel surface.

3.9. Mechanism of inhibition

Corrosion inhibition of mild steel in hydrochloric acid solution by MPTS and CPTS can be explained on the basis of molecular adsorption. These compounds inhibit corrosion by controlling both anodic as well as cathodic reactions. In acidic solutions these inhibitors exist as protonated species. In both inhibitors nitrogen atoms present in the molecules can be easily protonated in acidic solution and convert into quaternary compounds. These protonated species adsorbed on the cathodic sites of the mild steel and decrease the evolution of hydrogen. The adsorption on anodic site occurs through π-electrons of aromatic rings and lone pair of electrons of nitrogen, sulphur and oxygen atoms which decrease the anodic dissolution of mild steel. The inhibitors MPTS and CPTS are expected to get adsorbed on surface of mild steel through the lone pairs of electrons on N, S and O atoms present in these inhibitors. The schematic illustration of different modes of adsorption on metal/acid interface is shown in Fig. 17.
image file: c5ra14406j-f17.tif
Fig. 17 The schematic illustration of different modes of adsorption on metal/acid interface.

4. Conclusions

(1) Both the synthesized spiro compounds are good corrosion inhibitor for the corrosion of mild steel in 15% HCl solutions and inhibition efficiency of MPTS is better than CPTS.

(2) The potentiodynamic polarization studies showed that both the studied inhibitors are mixed type inhibitor.

(3) EIS measurements showed that charge transfer resistance (Rct) increases and double layer capacitance (Cdl) decreases in presence of inhibitors, suggested the adsorption of the inhibitor molecules on the surface of mild steel.

(4) The results obtained from SEM, AFM and Langmuir adsorption isotherm suggested that the mechanism of corrosion inhibition is occurring mainly through adsorption process.

(5) Quantum chemical results of MPTS and CPTS showed good correlation with the experimental results.

5. Acknowledgments

The authors acknowledge the financial supports of the CSIR New Delhi and Indian School of Mines, Dhanbad, India.

References

  1. E. A. Noor and A. H. Al-Moubaraki, Mater. Chem. Phys., 2008, 110, 145 CrossRef CAS PubMed.
  2. L. M. Vracar and D. M. Drazic, Corros. Sci., 2002, 44, 1669 CrossRef CAS.
  3. M. Behpour, S. M. Ghoreishi, N. Soltani and M. Salavati-Niasari, Corros. Sci., 2009, 51, 1073 CrossRef CAS PubMed.
  4. Y. M. Tang, X. Y. Yang, W. Z. Yang, R. Wan, Y. Z. Chen and X. S. Yin, Corros. Sci., 2010, 52, 1801 CrossRef CAS PubMed.
  5. M. Abdallah, Corros. Sci., 2002, 44, 717 CrossRef CAS.
  6. S. L. Granese, B. M. Rosales, C. Oviedo and J. O. Zerbino, Corros. Sci., 1992, 33, 1439 CrossRef CAS.
  7. G. Subramaniam, K. Balasubramaniam and P. Shridhar, Corros. Sci., 1990, 30, 1019 CrossRef CAS.
  8. S. N. Raicheva, B. V. Aleksiev and E. I. Sokolova, Corros. Sci., 1993, 34, 343 CrossRef CAS.
  9. M. Yadav, D. Behera, S. Kumar and R. R. Sinha, Ind. Eng. Chem. Res., 2013, 52, 6318 CrossRef CAS.
  10. L. J. Berchmans, S. V. Iyer, V. Sivan and M. A. Quraishi, Anti-Corros. Methods Mater., 2001, 48, 376 CrossRef CAS.
  11. A. Kokalj, Electrochim. Acta, 2010, 56, 745 CrossRef CAS PubMed.
  12. M. A. Deyab, Corros. Sci., 2007, 49, 2315 CrossRef CAS PubMed.
  13. S. A. Abd El-Maksoud and A. S. Fouda, Mater. Chem. Phys., 2005, 93, 84 CrossRef CAS PubMed.
  14. K. F. Khaled, Electrochim. Acta, 2003, 48, 2493 CrossRef CAS.
  15. R. Solmaz, G. Kardaş, M. Çulha, B. Yazıcı and M. Erbil, Electrochim. Acta, 2008, 53, 5941 CrossRef CAS PubMed.
  16. G. Gece and S. Bilgic, Corros. Sci., 2010, 52, 3304 CrossRef CAS PubMed.
  17. J. Radilla, G. E. Negron-Silva, M. Palomar-Pardave, M. Romero-Romo and M. Galvan, Electrochim. Acta, 2013, 112, 577 CrossRef CAS PubMed.
  18. T. Arslan, F. Kandemirli, E. E. Ebenso, I. Love and H. Alemu, Corros. Sci., 2009, 51, 35 CrossRef CAS PubMed.
  19. M. N. Ibrahim, M. F. El-Messmary and M. G. A. Elarfi, E-J. Chem., 2010, 7(1), 55 CrossRef CAS PubMed.
  20. ASTM, G 31–72, American Society for Testing and Materials, Philadelphia, PA, 1990.
  21. A. Dhanapala, S. Rajendra Boopathy and V. Balasubramanian, J. Alloys Compd., 2012, 523, 49 CrossRef PubMed.
  22. M. Lebrini, F. Robert, A. Lecante and C. Roos, Corros. Sci., 2011, 53, 687 CrossRef CAS PubMed.
  23. C. Lee, W. Yang and R. G. Parr, Phys. Rev. B: Condens. Matter Mater. Phys., 1988, 37, 785 CrossRef CAS.
  24. A. D. Becke, J. Chem. Phys., 1993, 98, 1372 CrossRef CAS PubMed.
  25. S. S. Abdel-Rehim, M. A. M. Ibrahim and K. F. Khaled, J. Appl. Electrochem., 1999, 29, 593 CrossRef.
  26. F. Bentiss, B. Mernari, N. Chaibi, M. Traisnel, H. Vezin and M. Lagrenée, Corros. Sci., 2002, 44, 2271 CrossRef CAS.
  27. M. Palomar-Pardavé, M. Romero-Romo, H. Herrera-Hernández, M. A. Abreu-Quijano, N. V. Likhanova, J. Uruchurtu and J. M. Juárez-García, Corros. Sci., 2012, 54, 231 CrossRef PubMed.
  28. A. Popova, E. Sokolova, S. Raicheva and M. Chritov, Corros. Sci., 2003, 45, 33 CrossRef.
  29. S. V. Ramesh and V. Adhikari, Bull. Mater. Sci., 2008, 31, 699 CrossRef.
  30. A. Espinoza-Vázquez, G. E. Negrón-Silva, D. Angeles-Beltrán, H. Herrera-Hernández, M. Romero-Romo and M. Palomar-Pardavé, Int. J. Electrochem. Sci., 2014, 9, 493 Search PubMed.
  31. M. Ozcan, J. Solid State Electrochem., 2008, 12, 1653 CrossRef CAS.
  32. A. M. Badiea and K. N. Mohana, Corros. Sci., 2009, 51, 2231 CrossRef CAS PubMed.
  33. F. Bentiss, M. Lebrini and M. Lagrenee, Corros. Sci., 2005, 47, 2915 CrossRef CAS PubMed.
  34. L. Tang, G. Mu and G. Liu, Corros. Sci., 2003, 45, 2251 CrossRef CAS.
  35. X. Li and G. Mu, Appl. Surf. Sci., 2005, 252, 1254 CrossRef CAS PubMed.
  36. M. Benabdellah, R. Touzani, A. Dafali, M. Hammouti and S. El Kadiri, Mater. Lett., 2007, 61, 1197 CrossRef CAS PubMed.
  37. G. Mu, X. Li and G. Liu, Corros. Sci., 2005, 47, 1932 CrossRef CAS PubMed.
  38. J. M. Thomas and W. J. Thomas, Introduction to the Principles of Heterogeneous Catalysis, Academic Press, London, 5th edn, 1981, p. 14 Search PubMed.
  39. X. Wang, H. Yang and F. Wang, Corros. Sci., 2011, 53, 113 CrossRef CAS PubMed.
  40. E. S. Ferreira, C. Giancomlli, F. C. Giacomlli and A. Spinelli, Mater. Chem. Phys., 2004, 83, 129 CrossRef CAS PubMed.
  41. S. Ramesh and S. Rajeswari, Electrochim. Acta, 2004, 49, 811 CrossRef CAS PubMed.
  42. H. A. Sorkhabi, B. Shaabani and D. Seifzadeh, Electrochim. Acta, 2005, 50, 3446 CrossRef PubMed.
  43. S. Banerjee, V. Srivastava and M. M. Singh, Corros. Sci., 2012, 59, 35 CrossRef CAS PubMed.
  44. B. El Mehdi, B. Mernari, M. Traisnel and F. Bentiss, Mater. Chem. Phys., 2002, 77, 489 CrossRef.
  45. S. Ciampi and V. Di Castro, Surf. Sci., 1995, 331, 294 Search PubMed.
  46. R. G. Pearson, Inorg. Chem., 1988, 27, 734 CrossRef CAS.
  47. M. J. S. Dewar and W. Thie, J. Am. Chem. Soc., 1977, 99, 4899 CrossRef CAS.
  48. P. Mourya, S. Banerjee, R. B. Rastogi and M. M. Singh, Ind. Eng. Chem. Res., 2013, 52, 12733 CrossRef CAS.
  49. C. Ogretir, B. Mihci and G. Bereket, THEOCHEM, 1999, 488, 223 CrossRef CAS.
  50. W. Huang, Y. Tan, B. Chen, J. Dong and X. Wang, Tribol. Int., 2003, 36, 163 CrossRef CAS.
  51. E. E. Ebenso, M. M. Kabanda, L. C. Murulana, A. K. Singh and S. K. Shukla, Ind. Eng. Chem. Res., 2012, 51, 12940 CrossRef CAS.
  52. I. B. Obot and N. O. Obi-Egbedi, Mater. Chem. Phys., 2010, 122, 325 CrossRef CAS PubMed.
  53. I. Lukovits, E. Klaman and F. Zucchi, Corrosion, 2001, 57, 3 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2015
Click here to see how this site uses Cookies. View our privacy policy here.